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
1311.3185
1
1311
2013-11-13T16:05:05
Accumulation of motile elongated micro-organisms in turbulence
[ "physics.bio-ph", "physics.flu-dyn" ]
We study the effect of turbulence on marine life by performing numerical simulations of motile microorganisms, modelled as prolate spheroids, in isotropic homogeneous turbulence. We show that the clustering and patchiness observed in laminar flows, linear shear and vortex flows, are significantly reduced in a three-dimensional turbulent flow mainly because of the complex topology; elongated micro-orgamisms show some level of clustering in the case of swimmers without any preferential alignment whereas spherical swimmers remain uniformly distributed. Micro-organisms with one preferential swimming direction (e.g. gyrotaxis) still show significant clustering if spherical in shape, whereas prolate swimmers remain more uniformly distributed. Due to their large sensitivity to the local shear, these elongated swimmers react slower to the action of vorticity and gravity and therefore do not have time to accumulate in a turbulent flow. These results show how purely hydrodynamic effects can alter the ecology of microorganisms that can vary their shape and their preferential orientation.
physics.bio-ph
physics
1 Accumulation of motile elongated micro-organisms in turbulence C A I J U A N Z H A N1 , G A E T A N O S A R D I N A1 2 , E N K E L E I D A L U S H I3 A N D L U C A B R A N D T1 † 1Linn´e Flow Centre and SeRC (Swedish e-Science Research Centre), KTH Mechanics, SE-100 44, Stockholm, Sweden 2Facolt´a di Ingegneria, Architettura e Scienze Motorie, UKE Universit´a Kore di Enna, 94100 Enna, Italy 3 School of Engineering, Brown University, 182 Hope Street, Providence, Rhode Island 02912, USA (Received 14 November 2013) We study the effect of turbulence on marine life by performing numerical simulations of motile microorganisms, modelled as prolate spheroids, in isotropic homogeneous tur- bulence. We show that the clustering and patchiness observed in laminar flows, linear shear and vortex flows, are significantly reduced in a three-dimensional turbulent flow mainly because of the complex topology; elongated micro-orgamisms show some level of clustering in the case of swimmers without any preferential alignment whereas spherical swimmers remain uniformly distributed. Micro-organisms with one preferential swimming direction (e.g. gyrotaxis) still show significant clustering if spherical in shape, whereas prolate swimmers remain more uniformly distributed. Due to their large sensitivity to the local shear, these elongated swimmers react slower to the action of vorticity and gravity and therefore do not have time to accumulate in a turbulent flow. These results show how purely hydrodynamic effects can alter the ecology of microorganisms that can vary their shape and their preferential orientation. Key words: 1. Introduction The macroscopic phenomena of marine landscape are influenced by the interactions between the flow and the motility of bacteria and phytoplankton. The effect of turbu- lence on marine life is therefore a key research question that also has relevance on the understanding of the consequences of climate changes. Microorganisms concentrate in the turbulent regions close to the surface and to the sea bed where the level of turbu- lence is also affected by external factors. The motion of an individual microorganism is determined by its swimming and by the advection of the fluid, where vorticity and rate of strain re-orient it, and by the response to external stimuli and biases such as nutri- ent concentration, gravity and light. Depending on the external stimulus, the behavior is categorized for example as geotaxis (Adams & Paul 1999), phototaxis (Martin 1983), gyrotaxis (Kessler 1985) and chemotaxis (Adler et al. 1974), etc. Among the different biases, we consider gyrotaxis. This bias results from the combi- nation of a viscous torque on the cell body, caused by the flow shear, and a gravitational † Email address for correspondence: [email protected] 2 C. J. Zhan, G. Sardina, E. Lushi and L. Brandt torque, arising from an asymmetric distribution of mass within the organism (bottom- heaviness). This induces an accumulation of cells heavier then water at the free surface and the occurrence of a bioconvective instability that develops from an initially uniform suspension without any background flow (Pedley & Kessler 1992, 1993). Recent studies consider the interactions between a complex flow and swimming mi- croorganisms. Cellular laminar flows are used by Torney & Neufeld (2007) to study the aggregation of self-propelled particles without taxis: these authors show that the parti- cles concentrate around chaotic tra jectories. Clustering is more pronounced for prolate swimmers and higher swimming speeds when particles escape from regular elliptic re- gions. In this particular case, spherical particles cannot enter these regions if initially outside of them. The accumulation of swimming microorganisms in chaotic regions of a fluid flow can be advantageous for fast dispersion. Khurana, Blawzdziewicz & Ouellette (2011) consider a two-dimensional chaotic vortical flow and show that swimming does not necessarily lead to enhanced particle transport. For small but finite values of the swimming speed, particles can be trapped for very long time near the boundaries be- tween chaotic and regular flow region. In a later study, Khurana & Ouellette (2012) add stochastic terms to the swimmer equation of motion and notice that suppression of trans- port (trapping) will increase dramatically with the addition of these random motions and rotational stochasticity and, more significantly, with elongated particles. At higher swimming speeds, elongated swimmers tend to be attracted to the stable manifolds of hyperbolic fixed points, leading to increased transport relative to swimming spheres. The coupling of gyrotactic particles and a layer with higher shear is studied by Durham, Kessler & Stocker (2009) who demonstrated that motility and shear are re- sponsible for the formations of intense thin layers of phytoplankton by gyrotactic trap- ping. These layers appear in regions characterized by a vertical gradient of the horizontal velocity that exceeds a critical threshold value: cells cluster in thin layers and tumble end over end. The coupling between the flow and the swimmers’ motility may lead to the formation of macroscopic flow features in the more general case of more complex flows. Durham, Climent & Stocker (2011) investigated how the gyrotactic motile microorgan- isms aggregate in a steady TGV flow (Taylor-Green Vortex) and suggested that the patchiness regimes can be characterized by two non-dimensional parameters: the swim- ming speed relative to a characteristic fluid speed and the magnitude of the gyrotactic torque. In this paper, we will consider the behavior of gyrotactic and non-gyrotactic swimmers in a three-dimensional turbulent flow, as turbulence characterizes the life of microbes in water supply systems, ocean and bioreactors. As in previous studies, we approximate swimmers as prolate spheroids. The statistics of non-swimming ellipsoidal particles in the turbulent flow can be found in Parsa et al. (2012) and references therein. These studies show that the rotation rate is influenced by alignment, particles orientations become correlated with the velocity gradient tensor, and the alignment depends strongly on the particle shape. Lewis (2003) proposes a Fokker-Planck model for the orientation of motile spheroids in homogeneous isotropic turbulence and shows that fluctuations in orientation manifest themselves as an increase of the effective rotary diffusivity. Thorn & Bearon (2010) considered the long-time tra jectories of deterministic and stochastic swimmers in shear flows. These authors simulated stochastic gyrotactic micro-organisms in synthetic turbulence and demonstrate quantitatively the transition between swimming-dominated drift and turbulence-dominated diffusion as a function of the kinetic energy dissipation rate. DeLillo, Boffetta & Cencini (2012) describe the spatial distribution of gyrotactic spherical microorganisms transported by three-dimensional turbulence flows generated by DNS (Direct Numerical Simulations). They show that coupling gyrotactic motion to Accumulation of motile micro-organisms in turbulence 3 turbulent flow produces small-scale patchiness (smaller than the Kolmogorov scale) in the swimmer distribution. Durham (2012); Durham et al. (2013) also examined whether gyrotaxis can generate cell patchiness in turbulent flow using experiments and numerical simulations. They found that accumulation in downwelling regions is the dominant means of aggregation also in turbulent flow and shows that patchiness is not significantly affected by the Taylor Reynolds number Reλ . Croze et al. (2013) studied dispersion of gyrotactic algae in turbulent channel flow and quantify the increased diffusivity induced by the turbulent fluctuations. In this paper, we document how clustering of prolate swimmers without taxis, observed in simple flows, is destroyed by turbulence. We then consider gyrotactic microorganisms of different shapes and confirm that bottom-heavy microorganisms accumulate in down- welling flows. This explains how settling larvae changing the offset of the centers of buoyancy and of gravity can preferentially accumulate in updrafts, favorable for disper- sal, or downdrafts, favorable for settlement, thus exploiting the hydrodynamics of the vorticity near the sea bed (Grunbaum & Strathmann 2003). We finally show that clus- tering is most evident for spherical shapes. This suggests that micro-organisms like the dinoflagellate Ceratocorys horrida, able to reversibly change its morphology in response to variations of the ambient flow (Zirbel, Veron & Latz 2002), can exploit hydrodynamics effects to increase or decrease encounter rates in an active way. 2. Problem Formulation The flow velocities are solution of the Navier-Stokes equation 2.1. Governing equations ∂u ∂ t + (u · ∇) u = −∇p + 1 Re ∇2u + f , (2.1) ∇ · u = 0. Here, u is the fluid velocity, p is the pressure and f is the external large-scale forcing needed to keep homogeneous isotropic turbulence in three-periodic domain. Re is the Reynolds number. The swimmers in our simulation are prolate spheroids advected by the local velocity while moving with constant speed us (2.2) dx dt = u + usp, (2.3) where the versor p defines the orientation of the swimmers. Assuming inertialess mo- tion, the angular velocity of the organisms is determined by the balance of viscous and gyrotactic torques (Pedley & Kessler 1992) 1 2B 1 2 dp dt = ω × p + α [I − pp] · E · p. [k − (k · p) p] + In the above equation, E is symmetric part of the deformation tensor and ω is the vorticity vector. α = (cid:0)AR2 − 1(cid:1) / (cid:0)AR2 + 1(cid:1) defines the eccentricity of the spheroids and AR the corresponding aspect ratio (the ratio of the ma jor to the minor axis). B is the characteristic time a perturbed cell takes to return to the vertical orientation, k, the preferred swimming direction (e.g. gyrotaxis). When there is no preferred direction, B → ∞ and the first term on the right hand side vanishes. (2.4) 4 C. J. Zhan, G. Sardina, E. Lushi and L. Brandt 2.2. Numerical methodology The numerical data set has been obtained from a Direct Numerical Simulation (DNS) us- ing a classical pseudo-spectral method coupled with a Lagrangian solver for the swimming micro-organisms. For the fluid phase, the Navier-Stokes equations have been integrated in a three-periodic domain of length LD = 2π using a Fourier spectral method with the nonlinear terms de-aliased by the 3/2 rule. The solution is advanced in time using a third- order low-storage Runge-Kutta method; specifically, the nonlinear terms are computed using an Adam-Bashforth like approximation while the diffusive terms are analytically integrated (Rogallo 1981). A random forcing is applied isotropically to the first shell of wave vectors, with fixed amplitude f0 , constant in time and uniformly distributed in phase and directions (Vincent & Meneguzzi 1991). We use a resolution of 192 Fourier modes in each of the three directions that correspond to a grid size in physical space of 2883 collocation points due to the de-aliasing. The ratio between the highest resolved wave number kmax and the Kolmogorov scale kη is kmax/kη = 1.73, which is within the usual accepted range to ensure a stable code and a good resolution (Pope 2000). The simulation has a Taylor Reynolds number Reλ = 150. To compute statistics, we analyze about 160 configurations, stored with an interval corresponding to half of the turbulence integral time scale (for each case considered). For the dispersed phase, we use the point particle approximation as typical micro- organisms are smaller than the flow Kolmogorov length. The same Runge-Kutta temporal integration used for the carrier phase is adopted for the swimmers. The integration of Eq. (2.4) for the orientation is performed using the quaternion formulation. A second-order interpolation scheme is used to compute the flow velocity, vorticity and velocity gradients at particle position with a high enough resolution to capture the flow velocity gradients. The swimmers are initially uniformly distributed in the domain and reach a statistically steady state after a short transient. Several simulations have been performed by keeping fixed the properties of the turbulence and changing the parameters of the swimming organisms, i.e. aspect ratio AR , swimming velocity and the reorientation time B . A list of the different simulations is reported in Table 1. In our simulation, the number of s = us/uη is the swimming velocity made non- particles is 200,000. In the following, V ∗ dimensional with the Kolmogorov velocity scale and ωrms the root mean square of the vorticity fluctuations. 3. Results 3.1. Non-gyrotactic swimmers We first study the behavior of micro-organisms who do not possess a preferential swim- ming direction. These have been shown to accumulate and get trapped in simple vortical flows (Torney & Neufeld 2007; Khurana et al. 2011). To quantify patchiness (clustering) in fully three-dimensional isotropic turbulent flows, we use the radial pair distribution function (RDF), sometimes also called correlation function. This is defined as g (r) = 1 dNr 1 4πr2 dr n0 where, n0 = 0.5Np (Np − 1) /V0 is the density of pairs in the whole volume V0 . Np is the total number of particles in the domain and Nr is the number of pairs at distance r. The RDF measures the probability to find a particle pair at a given radial distance normalized by the values of a uniform distribution. An indicator of patchiness is also the scaling exponent of RDF at small separations. , (3.1) Accumulation of motile micro-organisms in turbulence 5 AR V ∗ s Bωrms 1 3 9 1 1 0.5, 1, 3, 5, ∞ 1, ∞ 0 ∞ 0.5 ∞ 1 0.2, 0.5, 1, 1.5, 2, 2.5, 3, 5, 10, 30, ∞ 2 ∞ 5 ∞ ∞ 1 1, ∞ Table 1. Parameters defining the swimmers in the different simulations presented here. These evolve in a homogeneous isotropic turbulent flow with Reλ = 150. Bωrms = ∞ indicates no gyrotaxis. (a) 1.25 ) r ( F D P 1.2 1.15 1.1 1.05 1 0.95 100 Vs*=0 Vs*=1 Vs*=0.5 Vs*=2 Vs*=5 (b) 1.15 1.1 ) r ( F D P 1.05 1 Vs*=0,AR=9 Vs*=1,AR=1 Vs*=1,AR=3 Vs*=1,AR=9 Vs*=1,AR=∞ 102 0.95 100 r/η r/η 102 Figure 1. Radial distribution function of particle pair for (a) swimmers of same aspect ratio AR = 9 and different swimming speeds. (b) same swimming speed V ∗ s =1 and different aspect ratios. Results for swimmers of different shape and different swimming speed are presented in figure 1. Clustering, the value of the RDF at small separations, is relatively weak and flow visualizations indeed show an almost uniform distribution. The slope of the RDF at r → 0 s = 5, is larger for higher swimming speeds and elongated particles: swimmers with V ∗ AR = 9 exhibit maximum accumulation. As expected, the populations characterized by zero swimming speed or spherical shape do not present any clustering as their velocity field is divergence free. In the first case they are advected as passive tracers, whereas in the second case the swimming orientations are uniformly distributed. Large swimming velocities are able to counteract the dispersive effect of turbulence (Croze et al. 2013) and more elongated particles are associated to a potentially compressible velocity field, being sensitive to the background shear. In summary, the accumulation is very weak and from our observations we can conclude that in three-dimensional turbulence patchiness exceeds that of a Poisson distribution only for elongated cells, not for spherical ones. This fact could be quite important for the swimmers’ ecology and deserves further analysis. C. J. Zhan, G. Sardina, E. Lushi and L. Brandt (b) 1.6 1 Vs=0,AR=9 AR=1 AR=3 AR=9 AR=∞ −0.5 0 cos θ2 0.5 Vs=0,AR=9 AR=1 AR=3 AR=9 AR=∞ 6 (a) F D P 0.9 0.8 0.7 0.6 0.5 0.4 −1 (c) 1.4 F D P 1.2 1 0.8 0.6 0.4 0.2 0 −1 Vs=0,AR=9 AR=1 AR=3 AR=9 AR=∞ −0.5 0 cos θ1 0.5 1 Vs=0,AR=9 AR=1 AR=3 AR=9 AR=∞ −0.5 0 cos θ3 0.5 1 F D P 1.4 1.2 1 0.8 0.6 0.4 0.2 −1 (d) 4 F D P 3.5 3 2.5 2 1.5 1 0.5 0 −1 −0.5 0 cos ψ 0.5 1 Figure 2. PDF of orientation of swimmers with V ∗ s =1 with respect to (a), (b) and (c) the three eigenvectors of the deformation tensor; and (d) the local vorticity vector By tracking the swimmers in a frozen (time-independent) velocity field we also regis- tered a significant decrease of the cell accumulations, when compared to laminar flows. In particular, the value of the RDF at small r increases only from 1.1 to 1.2 when remov- ing the flow unsteadiness for the simulation with AR = 9. We therefore conclude that the complex three-dimensional flow topology is the main responsible of the lack of any significant clustering reported above. In figure 2, we aim to understand the relation between particle orientation and the underlying flow field when varying the micro-organism shape. In figure 2(a),(b) and (c), we show the orientation with respect to the three eigendirections of the strain tensor, λ1 > λ2 > λ3 . Here, the angles between these eigendirections and the cell orientation are denoted as θ1 , θ2 and θ3 . ψ represents the angle between the orientation and vorticity vector (figure 2d). The spherical swimmers do not show a preferential orientation with strain or vorticity as expected by the lack of any accumulation. The pdfs show peaks when prolate swimmers are parallel to the eigendirections associated to λ1 and λ2 . An increasing probability of parallel alignment with the first eigendirection of the strain is seen at larger aspect ratio. The swimming direction is more likely to be normal to the third eigendirection of the strain as shown in figure 2(c). Interestingly, figure 2(d) shows that the strongest tendency is to align with the local vorticity vector, which is the dominant effect. This finding resembles previous observations in oscillatory flows of dense fiber suspensions (Franceschini et al. 2011). Non-motile fibers are more likely to align with the vorticity vector than motile swimmers, whereas it is less likely to find them parallel to the second eigenvector of the strain tensor. The distributions of orientation Accumulation of motile micro-organisms in turbulence AR=1 AR=3 AR=9 AR=∞ −slope 1 0.5 0 0 (b) 3 2.5 ) r ( F D P 2 1.5 5 AR 10 Bω rms=0.5 Bω rms=1 Bω rms=1.5 Bω rms=2 Bω rms=3 Bω rms=5 −slope 1 0.5 0 0 Bω rms 5 7 102 1 100 101 r/η 102 AR=9 AR=1 (a) 3 ) r ( F D P 2.5 2 1.5 1 0.5 100 (c) 1.25 1.2 1.15 1.1 1.05 ) r ( F D P 1 100 101 r/η Bω =0.2 rms Bω =0.5 rms Bω =1 rms Bω =1.5 rms Bω =2 rms Bω =5 rms Bω =10 rms Bω =30 rms −slope 0.15 0.1 0.05 0 0 (d) 10 20 Bω rms 30 e p o l s − 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 101 r/η 102 1 2 Bω rms 3 4 Figure 3. Radial distribution function (RDF) for gyrotactic swimmers. (a) Bωrms = 1 and different aspect ratios as indicated. (b) Aspect ratio AR = 1 and (c) Aspect ratio AR = 9 and gyrotactic torque Bωrms as indicated in the legend for the same swimming speed V ∗ s =1. The inset displays the scaling exponent at small separations. (d) Close-up of the scaling exponent at small separations for the values of Bωrms with maximum clustering. The slope of the RDF is estimated by fitting with an ordinary least square method a power law in the range r/η=[0.2:2]. with respect to the vorticity vector appears to weakly vary with the aspect ratio for AR > 1. 3.2. Gyrotactic swimmers Here we study bottom-heavy swimmers that tend to align with gravity and swim up- wards and focus on the effect of the micro-orgamism shape. These type of swimmers are known to accumulate in regions of negative vertical velocity, in downwelling flows, also in the turbulent regime when spherical (Kessler 1985; Durham 2012; DeLillo et al. 2012; Croze et al. 2013). The main findings of this work are reported in figure 3: For gyrotac- tic swimmers of different shape clustering decreases when the aspect ratio of swimmers is increasing. In other words, spherical particles having a preferential swimming direc- tion exhibit the most significant aggregation. Results in figure 3(a) are obtained with fixed value of Bωrms , a non-dimensional parameter measuring the ratio of the alignment timescale to the rotation timescale induced by vorticity (Durham 2012). We show in figure 3(b) and (c), and more clearly in the close-up in figure 3(d), that the maximal clustering of prolate swimmers, yet significantly weaker than that observed for spherical microorganisms, is for Bωrms ≈ 2 whereas it is about 1 for spherical swim- mers. The curve of the scaling exponent versus Bωrms is more flat in the case of prolate swimmers, whereas a distinct peak is evident for swimmers with AR = 1. Low values of Bωrms indicate short re-orientation times and strong torques. In this case, the swimmers tend to align to the vertical direction and swim upwards: an initial uniform distribution will tend to remain as such. For weak gyrotaxis, large values of Bωrms , particles tend to behave as shown above for non-gyrotactic swimmers and the accumulation is not rele- 8 (a) 1 F D P (c) F D P 0.9 0.8 0.7 0.6 0.5 0.4 −1 1 0.8 0.6 0.4 0.2 0 −1 C. J. Zhan, G. Sardina, E. Lushi and L. Brandt (b) Bω rms=0.2 Bω rms=3 Bω rms=∞ −0.5 0 cos θ1 0.5 1 F D P (d) F D P −0.5 0 cos θ3 0.5 1 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 −1 40 35 30 25 20 15 10 5 0 −1 Bω rms=0.2 Bω rms=3 Bω rms=∞ −0.5 0 cos θ2 0.5 1 F D P 4 3 2 1 0 −1 −0.5 0 cos ψ 0.5 1 −0.5 0 p⋅ k 0.5 1 Figure 4. PDF of orientation for gyrotactic swimmers of aspect ratio AR = 9, swimming speed s =1, with respect to: (a), (b) and (c) the three eigenvectors of the deformation tensor; (d) the V ∗ local vorticity vector (inset) and the direction of gravity. vant. Small-scale clustering is therefore occurring for intermediate values of Bωrms and these optimal values shifts towards longer time scales for prolate swimmers. Comparing figures 3(a) and 1(a), we note that elongated cells exhibit the weakest clustering, how- ever, they alone have some accumulation when gyrotaxis is turned off. To gain further insight, we examined the behavior of gyrotactic swimmers in laminar uniform shear flows by solving numerically Eq. (2.4) for different values of the aspect ratio AR. As discussed below, the appearance of Jeffrey-like orbits at larger values of Bωrms , i.e. weak reorien- tations, in the case of prolate swimmers explains the shift in the values of Bωrms where the largest clustering is observed. To further document the decrease of the clustering for prolate swimmers we display in figures 4 and 5 the mean orientation of the gyrotactic micro-organisms with respect to the underlying flow field. We keep AR = 9 and swimming speed the same and investigate the influence of Bωrms in figure 4. The results for weak gyrotaxis, large values of Bωrms , resemble those in figure 2, whereas the orientation with respect to the strain and vorticity fields tends to be more uniform for strong gyrotaxis, where the cell behavior is dominated by gravity (see 4d). In all cases, peaks of the PDFs occur when the orientation is at 0 and 180 degrees with respect to the direction associated to λ1 . The distribution of the angles between the swimming direction and the second direction of the strain tensor becomes asymmetric when increasing the strength of the swimming bias. As for non- gyrotactic swimmers, the orientation of particles is more likely to be normal to the third eigendirection of the strain; this tendency increases as Bωrms increases. In the inset of Accumulation of motile micro-organisms in turbulence (a) 1 (b) 1.2 9 F D P (c) F D P 0.9 0.8 0.7 0.6 0.5 0.4 −1 1 0.8 0.6 0.4 0.2 0 −1 AR=1 & Bω rms=1 AR=9 & Bω rms=2 −0.5 0 cos θ1 0.5 1 −0.5 0 cos θ3 0.5 1 F D P (d) F D P 1 0.8 0.6 0.4 0.2 −1 10 8 6 4 2 0 −1 AR=1 & Bω rms=1 AR=9 & Bω rms=2 −0.5 2 1.5 F D P 1 0.5 0 −1 0 cos θ2 0.5 1 −0.5 0 cos ψ 0.5 1 −0.5 0 p⋅ k 0.5 1 Figure 5. PDF of orientation for the most accumulating gyrotactic swimmers, AR = 1, 9, swimming speed V ∗ s =1, with respect to: (a), (b) and (c) the three eigenvectors of the deformation tensor; (d) the local vorticity vector (inset) and the direction of gravity. figure 4d) we see that the alignment with the vorticity vector also increases as the stability of the cell relative to that of the flow increases. Finally we consider the relative orientation with respect to the direction of gravity. As expected, particles tend to align more in the vertical direction as Bωrms decreases. As mentioned above, for strong gyrotaxis swimmers tend to move upwards keeping a uniform distribution, while for lower magnitudes of the gyrotactic torque they are mainly oriented by the flow vorticity and strain. In figure 5, we investigate the alignments of gyrotactic swimmers for the values of Bωrms giving the maximal clustering, Bωrms = 2 for AR = 9, and Bωrms = 1 for AR = 1. Prolate swimmers align with the direction of strain defined by λ1 whereas spherical particles are more likely to be normal to it. Conversely, prolate are orthogonal to the third eigendirection while spherical swimmers tend to be more uniformly distributed. Further, prolate swimmers tend to align to the vorticity vector more than their spherical counterpart. The shape of the swimmers also affect the orientation with respect to the gravity vector with the spherical swimmers more likely to swim parallel to gravity at given orientation time. Recalling that the spherical swimmers have larger aggregation than those elongated, we find the orientation of spherical particles to be more strongly dominated by gravity and less by the local shear. To quantify the possible ecological implications of clustering induced by biased swim- ming in turbulence, we estimate the collision rate of micro swimmers. To this end, we follow the approach adopted in the study of inertial particles. The particle collision rate is proportional to clustering (RDF) at a distance corresponding approximatively to one 10 (a) 0.8 0.6 e p o l S − 0.4 0.2 0 0 0 C. J. Zhan, G. Sardina, E. Lushi and L. Brandt (b) 2 0.2 1 1.5 S N n o i s i l l o c / n o i s i l l o 0.5 c 0.15 e p o l S − 0.1 0.05 2 1.5 S N n o i s i l l o c / n o i s i l l o 0.5 c 1 5 5 10 10 15 15 Bω rms 20 20 25 25 0 30 30 0 0 0 5 5 10 10 20 20 25 25 0 30 30 15 15 Bω rms Figure 6. Collision rate and slope of the RDF at small separations versus the non-dimensional re-orientation time Bωrms for swimming speed V ∗ s =1. (a) Aspect ratio AR = 1 and (b) Aspect ratio AR = 9. particle diameter and to the mean particle concentration weighted by the probability of particles having negative (approaching) relative velocity (see Sundaram & Collins 1997; Gualtieri et al. 2012; Sardina et al. 2012): (3.2) Nc = πc2 σ2g0 (σ) hδvp (σ) δvp < 0i. In the expression above Nc is the total collision rate, c the mean concentration, g0 the RDF, σ the length scale characteristic of the collision and hδvp (σ) δvp < 0i the spherical average of the mean relative velocity of two microorganisms. A similar relation is adopted in Kjørboe & Visser (1999); Kiørboe (2008), where a uniform organism distribution is assumed. This observable, used for inertial particles in turbulence where accumulation occurs, enables us to take into account the combined effect of clustering and swimming. The collision rates computed for the different values of the gyrotactic torque and two values of the cell aspect ratio are reported in figure 6. In the plot, the collision rates are divided by the value pertaining the case of non-swimming organisms without gyrotaxis (denoted as collision NS); the value σ = 0.01η is used to define a distance of approach. The data for these low values of separation have been extrapolated by assuming a power law for the collision rate: Nc ∝ g0 (σ)σ2 hδvp (σ) δvp < 0i is a power law at small scales since σ2 is a power law by definition and hδvp (σ) δvp < 0i is known to be well approximated by a power law (classical Kolmogorov theory) in the viscous range. Results obtained with values to σ = 0.5η and thus without extrapolation do not show significant qualitative differences. The slope of the RDF is also reported in the figure to display the peak in clustering for the same configurations. The data show that the collision rates increases with Bωrms reaching values about twice as large and 50% larger than those for passive tracers It can be seen in the figure that the magnitude of the gyrotactic torque is the most important factor affecting the collision rate, more than the clustering. Surprisingly, the most clustered microorganisms show a reduced level of collisions. This can be explained by the fact that for low values of Bωrms the microorganisms, although close to each other, tend to swim in the same direction and thus the collision rate decreases due to the low values attained by the factor hδvp (σ) δvp < 0i. Our results therefore suggest that other (sensorial) mechanisms should be active at the micro scale, sub-Kolmogorov scale, to enhance encounters between those microorganisms that turbulent motions may have brought close to each other. Accumulation of motile micro-organisms in turbulence 11 3.3. Equilibrium orientation in laminar flows To try to explain the results presented above for a fully three-dimensional time-dependent turbulent flow we examine the solution of the equation governing the swimmer orienta- tion, eq. (2.4), for different values of the aspect ratio AR. An analytical steady-state solu- tion can be found for the case of spherical swimmers and is discussed in Pedley & Kessler (1987); Thorn & Bearon (2010). The cells have a stable orientation parallel to gravity for low values of BΩ, where Ω is the vorticity of the background laminar flow. This regime equilibrium rotate towards the direction of the vorticity vector for large values of BΩ. A stable solution exists for all values of the non-dimensional parameter BΩ if the vorticity vector is not orthogonal to the the gravitational force, or equivalently, the gravitational torque is not parallel to the vorticity vector. In this case, an equilibrium solution exists only if BΩ < 1; above this value the cells tumble (the limit for infinite B, no gyrotaxis, being the Jeffrey orbit). Here, we compute numerical solutions of the time-dependent eq. (2.4) and determine the effect of the swimmer shape on the equilibrium orientation, as well as on the appear- ance of tumbling. We first analyze the case of homogenous shear flow in the plane parallel to gravity, with gravity in the direction of the velocity gradient (ux = Ωz ). Figure 7(a) reports the period of tumbling versus BΩ for the 4 different values of the aspect ratio considered in the turbulent cases above, whereas 7(b) displays the angle at equilibrium in the x − z plane, with x the flow direction and z the direction of gravity. Note that the two plots are complementary as either tumbling or an equilibrium angle is observed for each value of BΩ considered. As swimmers are more and more elongated, the tumbling motion is observed to start at larger values of BΩ, that is for long reorientation times. The period of tumbling decreases and approaches that of the Jeffrey orbit as BΩ → ∞. Below the threshold value for the occurrence of tumbling, the equilibrium angle increases from zero, swimmers almost aligned with the flow, to 90o as BΩ → 0 , e.g. swimmers par- allel to gravity for strong gyrotaxis. The figure clearly shows that the transition between gravity-dominated swimming and tumbling-motions shifts to larger BΩ for elongated swimmers. Next we consider the case of vorticity and gravity vector forming an angle of 45o , as in Thorn & Bearon (2010). In this case, no tumbling is observed and the cells tend to orient parallel to the gravitational force for small values of BΩ, whereas they are aligned with the vorticity vector for large BΩ. This is displayed in figure 8(a) where we report the scalar product between the cell orientation at the final steady state and the flow vorticity, with ±√2/2 indicating the direction of gravity. Again, we note that the tran- sition between gravity-dominated swimming and flow-dominated orientation, the regime where the largest clustering is observed in turbulence, is shifted to large values of BΩ for elongated swimmers. While this accounts for the shift of the maximum accumulation to larger values of Bωrms in the turbulent case documented above, it does not help to explain the reduced clustering. To this aim, we report in figure 8(b) the time evolution of the cell orientation for different initial orientations and the values (AR, BΩ) = (1, 0.625) and (AR, BΩ) = (9, 1.66) right during the transition between gravity-dominated swim- ming and vorticity-dominated swimming (same final orientation). Elongated swimmers, AR = 9, react more slowly and approach the equilibrium position at a later time, t ≈ 20, whereas spherical swimmers reach their final equilibrium at t ≈ 10, where these values correspond approximatively to the mean values computed by varying the initial orienta- tion. The trend shown in the figure (longer transients for elongated cells) is observed in the range 0.25 < BΩ < 1.7 for the spherical swimmers and 0.63 < BΩ < 2.5 for those with AR = 9. This range of values is consistent with the values of Bωrms where max- 12 C. J. Zhan, G. Sardina, E. Lushi and L. Brandt AR=1 AR=3 AR=5 AR=9 (a) T Ω 50 45 40 35 30 25 20 15 10 (b) Θ 90 80 70 60 50 40 30 20 10 AR=1 AR=3 AR=5 AR=9 5 100 102 101 100 BΩ BΩ Figure 7. Orientation and tumbling of gyrotactic cells in homogenous shear with gravity in the direction of the velocity gradient. (a) Period of tumbling versus BΩ, with Ω the shear flow vorticity. (b) Equilibrium angle Θ in the x − z plane, with x the flow direction and z the direction of gravity. 10−1 102 101 103 0 10−2 (a) −0.7 −0.75 −0.8 p · Ω −0.85 −0.9 −0.95 AR=1 AR=9 (b) Θ 180 160 140 120 100 80 60 40 20 −1 10−2 AR=9 AR=1 102 101 10−1 100 tΩ BΩ Figure 8. Orientation of gyrotactic cells in homogenous shear with gravity inclined by 45o with respect to the vorticity vector. (a) Equilibrium angle expressed as the scalar product between the regime solution and the vorticity vector p · Ω. (b) Time evolution of the cell orientation for different initial orientations and values of BΩ during the transition between gravity-dominated swimming and vorticity-dominated swimming (same final orientation). 20 10 15 25 0 0 5 imum clustering is observed in turbulence. This would imply that in a time-dependent flow elongated swimmers are more unlikely to reach their equilibrium position and their orientation would be determined to a larger extent by the initial condition. As a conse- quence they would not be able to accumulate like spherical swimmers. 4. Summary and Discussion In this paper we consider direct numerical simulations of the dynamics of non-spherical swimmers in realistic turbulent flows. We demonstrate how the clustering and trapping of swimmers without a preferential alignment observed in simple cellular or vortical flows (e.g. Torney & Neufeld 2007; Khurana et al. 2011) is significantly reduced in a three- dimensional time-dependent flow and exceeds that of a Poisson distribution for elongated cells, whereas spherical cells remain uniformly distributed. In addition, we expand on recent studies of small-scale patchiness in the distribution of gyrotactic micro-organisms in vortical or turbulent flows. We confirm that bottom- Accumulation of motile micro-organisms in turbulence 13 heavy cells tend to accumulate in downwelling flows, a fact exploited by settling lar- vae changing the offset of the centers of buoyancy and of gravity to preferentially ac- cumulate in updrafts, favorable for dispersal, or downdrafts, favorable for settlement (Grunbaum & Strathmann 2003). In particular, we investigate how the gyrotactic clus- tering phenomenon in turbulence is modified by the elongation of the ellipsoidal swim- mers. The parameters used in this study are in a realistic range and can also be replicated in the laboratory: for typical marine micro-organisms B = 1 − 6s, us = 100 − 200µm/s, thus us/uη ∈ [0.02 : 0.4], Bωrms ∈ [0.1, 50] (Jumars et al. 2009; DeLillo et al. 2012). The shapes of the marine organisms however vary and many phytoplankton species are not spherical, thus the clustering of such organisms in turbulent flows is affected by their individual geometry. All other conditions equal, we find that clustering is highest for spherical gyrotactic swimmers and decreases the more elongated the swimmers. The orientation of these latter micro-organisms undergoes longer transient phases, and appear to be more often a consequence of the local shear rate rather than of the local gravity field in a fluctuating flow; this explains the decreased accumulation. Our finding confirms that microorganisms that can actively change their shape, such as Ceratocorys horrida (Zirbel et al. 2002), have an active control mechanism to alter their distribution and favor encounters or uptake. Acknowledgements Computer time provided by SNIC (Swedish National Infrastructure for Computing) is gratefully acknowledged. The present work is supported by the Swedish Research Council (VR), by the Linn´e Flow Centre, KTH, and CSC (Chinese Scholarship Council). REFERENCE S Adams, C. F. & Paul, A. J. 1999 Phototaxis and geotaxis of light-adapted zoeae of the golden king crab lithodes aequispinus (anomura: Lithodidae) in the laboratory. Journal of Crustacean Biology pp. 106–110. Adler, J., Tso, W. W. et al. 1974 “Decision”-making in bacteria: chemotactic response of escherichia coli to conflicting stimuli. Science (New York, NY) 184 (143), 1292. Croze, O. A., Sardina, G., Ahmed, M., Bees, M. A. & Brandt, L. 2013 Dispersion of swimming algae in laminar and turbulent channel flows: consequences for photobioreactors. Journal of The Royal Society Interface 10 (81). DeLillo, F., Boffetta, G. & Cencini, M. 2012 Clustering of gyrotactic microorganisms in turbulent flows. arXiv preprint arXiv:1206.2570 . Durham, W. M., Climent, E., Barry, M., DeLillo, F., Boffetta, G., Cencini, M. & Stocker, R. 2013 Turbulence drives microscale patches of motile phytoplankton. Nature Communiactions 4 (2148), 1:7. Durham, W. M., Climent, E. & Stocker, R. 2011 Gyrotaxis in a steady vortical flow. Phys. Rev. Lett. 106 (23), 238102. Durham, W. M., Kessler, J. O. & Stocker, R. 2009 Disruption of vertical motility by shear triggers formation of thin phytoplankton layers. Science 323 (5917), 1067–1070. Durham, W. M. K. 2012 Phytoplankton in flow. PhD thesis, Massachusetts Institute of Tech- nology. Franceschini, A., Filippidi, E., Guazzelli, E. & Pine, D. J. 2011 Transverse alignment of fibers in a periodically sheared suspension: An absorbing phase transition with a slowly varying control parameter. Phys. Rev. Lett. 107, 250603. Grunbaum, D. & Strathmann, R. R. 2003 Form, performance and trade-offs in swimming and stability of armed larvae. Journal of Marine Research 61, 659–691. Gualtieri, P, Picano, F, Sardina, G & Casciola, CM 2012 Statistics of particle pair 14 C. J. Zhan, G. Sardina, E. Lushi and L. Brandt relative velocity in the homogeneous shear flow. Physica D: Nonlinear Phenomena 241 (3), 245–250. Jumars, P. A., Trowbridge, J. H., Boss, E. & Karp-Boss, L. 2009 Turbulence-plankton interactions: a new cartoon. Marine Ecology 30 (2), 133–150. Kessler, J.O. 1985 Hydrodynamic focusing of motile algal cells. Nature 313, 218–220. Khurana, N., Blawzdziewicz, J. & Ouellette, N. T. 2011 Reduced transport of swimming particles in chaotic flow due to hydrodynamic trapping. Phys. Rev. Lett. 106 (19), 198104. Khurana, N. & Ouellette, N. T. 2012 Interactions between active particles and dynamical structures in chaotic flow. Phys. Fluids 24 (9), 091902–091902. Kiørboe, Thomas 2008 A mechanistic approach to plankton ecology . Princeton University Press. Kjørboe, T. & Visser, A. W. 1999 Predator and prey perception in copepods due to hy- dromechanical signals. Mar. Ecol. Prog. Ser. 179, 81–95. Lewis, D. M. 2003 The orientation of gyrotactic spheroidal micro-organisms in a homogeneous isotropic turbulent flow. Proceedings of the Royal Society of London. Series A: Mathemat- ical, Physical and Engineering Sciences 459 (2033), 1293–1323. Martin, E. A. 1983 Macmil lan Dictionary of Life Sciences.. Macmillan Press Ltd. Parsa, S., Calzavarini, E., Toschi, F. & Voth, G. A. 2012 Rotation rate of rods in turbulent fluid flow. Phys. Rev. Lett. 109, 134501. Pedley, T. J. & Kessler, J. O. 1987 The orientation of spheroidal microorganisms swimming in a flow field. Proc. R. Soc. London, Ser. B 231, 47. Pedley, T. J. & Kessler, J. O. 1992 Hydrodynamic phenomena in suspensions of swimming microorganisms. Annu. Review Fluid Mechanics 24 (1), 313–358. Pedley, T. J. & Kessler, J. O. 1993 Bioconvection. Science Progress 76, 105–105. Pope, S. B. 2000 Turbulent flows . Cambridge university press. Rogallo, R. S. 1981 Numerical experiments in homogeneous turbulence , , vol. 81315. National Aeronautics and Space Administration. Sardina, G, Schlatter, P., Brandt, L., Picano, F & Casciola, CM 2012 Wall accumu- lation and spatial localization in particle-laden wall flows. J. Fluid Mech. 1 (1), 1–29. Sundaram, S. & Collins, L R 1997 Collision statistics in an isotropic particle-laden turbulent suspension. part 1. direct numerical simulations. J. Fluid Mech. 335 (75), 109. Thorn, Graeme J & Bearon, Rachel N 2010 Transport of spherical gyrotactic organisms in general three-dimensional flow fields. Phys. Fluids 22, 041902. Torney, C. & Neufeld, Z. 2007 Transport and aggregation of self-propelled particles in fluid flows. Phys. Rev. Lett. 99 (7), 78101. Vincent, A. & Meneguzzi, M. 1991 The spatial structure and statistical properties of homo- geneous turbulence. J. Fluid Mech. 225 (1), 1–20. Zirbel, M. J., Veron, F. & Latz, M. I. 2002 The reversible effect of flow on the morphology of ceratocorys horrida (peridiniales, dinophyta)*. Journal of Phycology 36 (1), 46–58.
1006.5161
1
1006
2010-06-26T19:48:35
A hopping mechanism for cargo transport by molecular motors in crowded microtubules
[ "physics.bio-ph" ]
Most models designed to study the bidirectional movement of cargos as they are driven by molecular motors rely on the idea that motors of different polarities can be coordinated by external agents if arranged into a motor-cargo complex to perform the necessary work [gross04]. Although these models have provided us with important insights into these phenomena, there are still many unanswered questions regarding the mechanisms through which the movement of the complex takes place on crowded microtubules. For example (i) how does cargo-binding affect motor motility? and in connection with that - (ii) how does the presence of other motors (and also other cargos) on the microtubule affect the motility of the motor-cargo complex? We discuss these questions from a different perspective. The movement of a cargo is conceived here as a hopping process resulting from the transference of cargo between neighboring motors. In the light of this, we examine the conditions under which cargo might display bidirectional movement even if directed by motors of a single polarity. The global properties of the model in the long-time regime are obtained by mapping the dynamics of the collection of interacting motors and cargos into an asymmetric simple exclusion process (ASEP) which can be resolved using the matrix ansatz introduced by Derrida [Derrida et al].
physics.bio-ph
physics
A hopping mechanism for cargo transport by molecular motors in crowded microtubules. Carla Goldman Departamento de F´ısica Geral - Instituto de F´ısica Universidade de Sao Paulo CP 66318 05315-970 Sao Paulo, Brazil. May 2010 Abstract Most models designed to study the bidirectional movement of cargos as they are driven by molecular motors rely on the idea that motors of different polarities can be coordinated by external agents if arranged into a motor-cargo complex to perform the necessary work [1]. Although these models have provided us with important insights into these phenomena, there are still many unanswered questions regarding the mechanisms through which the movement of the complex takes place on crowded microtubules. For example (i) how does cargo-binding affect motor motility? and in connection with that - (ii) how does the presence of other motors (and also other cargos) on the microtubule affect the motility of the motor-cargo complex? We discuss these questions from a different perspective. The movement of a cargo is conceived here as a hopping process resulting from the transference of cargo between neighboring motors. In the light of this, we examine the conditions under which cargo might display bidirectional movement even if directed by motors of a single polarity. The global properties of the model in the long-time regime are obtained by mapping the dynamics of the collection of interacting motors and cargos into an asymmetric simple exclusion process (ASEP) which can be resolved using the matrix ansatz introduced by Derrida [23]. 1 keywords - intracellular transport by molecular motors; bidirectional movement of cargo, traffic jam on microtubules; ASEP models. 1 Introduction Research interest in the origins of the long-range bidirectional movement of particles (organelles, vesicles, nutrients) driven by molecular motors is motivated by fundamental questions concerning the nature of interactions between motors and their cargos as transport processes take place. A current explanation for the phenomenon relies on the idea that motors of different polarities act coordinately on the same particle at different times. If, however, they act in parallel, the bidirec- tional movement would reflect dominance of one or another kind of motor achieved by a tug-of-war mechanism [1], [2], [3], [4], [5]. An important question that remains in this context concerns the mechanisms that would promote such coordination [6]. Alternatives to the coordination or tug-of- war models in the literature arise from the possibility of attributing the phenomenon to a dynamic role of the microtubules [7] or to a mechanical coupling between different motors [8]. A general difficulty encountered within any of these views is related to the presence of other particles (including other motors) on the microtubule at a given time that are not directly involved with the transfer process. These other particles are expected to impose restrictions on motility and performance of the motors that are directly interacting with cargo at that time [9]. Contrarily to these expectations, however, data from observations of beads driven by kinesins in steady-state conditions indicate that the number of long length runs of such beads increases significantly as the density of motors at the microtubule increases, although their velocities remain essentially unaltered within a wide range of motor concentrations [10], [11]. Thus, the reality of traffic jam in crowded microtubules still challenges the current view of long-range cargo transport that presupposes an effective and controllable movement of the motor(s) arranged into a motor-cargo complex. This, of course, requires a certain degree of stability of motor-cargo interactions and motor processivity. Our intention here is to discuss these problems from a different perspective by bringing into this scenario the model introduced in [12] to examine cargo transport as a hopping process. According 2 to that, motors and cargos would not assemble into complexes to put transport into effect. On the contrary, each motor would function as an active overpass for cargo to step over to a neighboring motor. In this case, the long-range movement of cargo is envisaged as a sequence of these elementary (short-range) steps either forwards or backwards. In [12] we examined the conditions under which this may happen, accounting for the fact that motor motility is affected by the interactions with other motors and with cargos on the microtubule. There, we considered the presence of a collection of interacting motors, all of them presenting the same polarity (kinesins may be thought of as prototypes) and a single cargo. Here, we examine whether it is possible to explain in a similar context the origin of the observed bidirectional movement displayed by cargos. The particular mechanism we propose to substantiate the hopping differs from that suggested in [12]. It keeps, however, the same general ideas of the original. As it will be explained below, we view the hopping of cargo between motors as an effect of thermal fluctuations undergone by motor tails. The flexibility of the tails may promote contact and, eventually, exchange of cargo between neighboring motors. As in [12], the model dynamics is mapped into an asymmetric simple exclusion process (ASEP) [13], [14], [15] whose stationary properties are resolved explicitly in the limit of very large systems. Other ASEP models have already been considered in the literature to study the conditions for motor jamming in the absence of cargo [9], [16], [17]. Our model is conceived to account explicitly for changes in the dynamics of the motors that at a certain instant of time are interacting with cargos. The model is reviewed here in order to include a second cargo in the system, still keeping the presence of motors of a single polarity. We believe that this approaches more realistic situations in which the simultaneous presence of many cargos and motors on the same microtubule must be the prevailing situation [7]. We show that under these conditions, a cargo may be able to execute long-range bidirectional movement as it moves over clusters of motors assembled either at its back end or at the back end of the cargo in front. One may recognize in this a possibility for explaining the origins of self-regulation in intracellular transport since it has been suggested in the last few years that signaling pathways involved in intracellular traffic regulation can be performed simply 3 by the presence of cargos at the microtubule [18]. We then speculate that the passage of cargos on microtubules does not get blocked by motor jamming. On the contrary, jamming operates as an allied process to promote long runs of cargos across motor clusters. In this case, the density of motors on the microtubule can be identified as an element of control in intracellular transport since it directly affects the conditions for jamming. It is worth mentioning that the model developed here does not rule out other possibilities, such as the tug-of-war or competition models. What we suggest is that the presence of motors of different polarities may not be essential to explain the origin of the bidirectional movement. The hopping mechanism is presented in Sec.2. The kinetic properties of the extended version are developed in Sec.3, considering the presence of two cargos. In Sec.4 we present our results. Additional remarks and conclusions are in Sec.5. 2 An alternative for cargo driving The stochastic model in [12] formulated in a lattice describes the dynamics of motors and cargos accounting for (i) steric interactions among different particles moving on the same microtubule; (ii) the presence of motors of a single polarity; (iii) the fact that cargos do not move if not driven by motors. The crucial point is in item (iii) because it requires a specific model for motor-cargo dynamics as transportation takes place. We offer here a slightly different view from that in [12] keeping however the reliance on the ability of motors to transfer cargo. One way by which this may be achieved is sketched in Figure (1). The stepping of cargo would be accomplished as it is released from a motor to which it is attached at a certain instant of time and then get attached to another (neighboring) motor either at the left or at the right -- see Figure (1a). This process, like the one discussed in [12], relies strongly on the flexibility of the motor´s tail. The idea was inspired by experimental results suggesting a dynamic role of the kinesin´s coiled-coil segment in the process [19], [20] and also by data indicating that under load, kinesin motors display an oscillatory movement [21] [22]. Here, we think of these oscillations as signaling fluctuations in the position of motor´s tail, not necessarily 4 being correlated with displacement of its center of mass. If this is the case, such oscillations would promote contact between neighboring motors favoring cargo exchange. Accordingly, long-range displacements of cargo would reflect a hopping process extended over many neighboring motors which may be accomplished if these motors get jammed into clusters for sufficient long periods of time. Notice that the whole mechanism does not require special stability of cargo-motor binding. On the contrary, the transfer of cargo by a motor would be ease by a loose attachment between them. The model in [12] was resolved explicitly considering the presence of a single cargo in the system. The averages for the quantities of interest were determined in steady-state conditions. We showed there that the long-range displacements of the cargo would occur predominantly in the backwards direction, i.e. in opposition to the direction of the movement of the considered motors. We shall show here that the same dynamics may lead cargo to display bidirectional movement if the system contains at least one more cargo interfering with the movement of the motors. To examine the properties of the system containing two cargos and an arbitrary number of motors, we map it into the same ASEP as in [12] whose dynamics can also reproduce the scheme in Figure(1). 2.1 Movement of motors and cargos: the ASEP model with two cargos We consider a one-dimensional lattice with M sites, representing the microtubule with periodic boundary conditions. This system contains N motors and a number K of other particles - the cargos - that interact with motors in order to move. Each site can be occupied by a motor or by a motor attached to a cargo (see Fig.1(a)), otherwise it is empty. The total number of sites that remain unoccupied is G = M − N − K > 0. Here, we analyze the long-time behavior of this system for K = 2 and determine the average cargo velocity as a function of the parameters. The results indicate conditions for cargos to perform a type of long-range movement that share the characteristics of the observed bidirectional movement. The map of the dynamics shown in Figure.1 into the considered ASEP is carried out as follows. 5 First, each site is identified by its position j = 1, 2, ...N at the lattice. Then, to each of these sites is associated a variable σj that assumes integer values 0, 1 or 2 such that σj = 0 if the site j is empty, σj = 1 if it is occupied by a motor; or σj = 2 if it is occupied by a motor attached to a cargo. With these, a configuration C of the lattice is specified by the set {σ1σ2... σN}. The dynamics of the ASEP that reproduces the elementary steps in Fig.1 can now be defined. For this, consider that at each time interval dt a pair of consecutive sites, say j and j + 1 are selected at random. The occupancy of these two sites is then switched according to the following rules (a) 10 → 01 with rate k, (b) 12 → 21 with rate w, (c) 21 → 12 with rate p, probability kdt probability wdt probability pdt (1) where the pair (j, j+1) is represented by the values of the corresponding site variables (σj, σj+1).The parameters k, w and p are the assigned probabilities per unit time (rates) for occurrence of the processes indicated. Process (a) describes the possibility for a motor (kinesin) that carries no cargo to step forward to a neighboring empty site (Figure 1b). Processes (b) and (c) account for the switching of the cargo between two neighboring motors. This accounts either for backward (b) or for forward steps (c). Notice that the dynamics conserves the number of particles of type-1 as well as those of type-2. In order to investigate the long-time dynamics of a cargo resulting from these elementary steps we use the matrix ansatz introduced by Derrida [23], [15]. The idea is to represent the probability PN,M (C) of a configuration C of the system with N sites and M particles of type-1 as a trace over a product of N non-commuting matrices, each specifying the corresponding site occupancy: where PN,M (C) = 1 ZN,M T r N (δσi,1D + δσi,2A + δσi,0E) Yi=1 N ZN,M =X{σi} T r (δσi,1D + δσi,2A + δσi,0E) Yi=1 (2) (3) is the normalization. The sum runs over all configurations for whichPN K = 2. In this product, a site i is represented by a matrix D if it is occupied by a motor (σi = 1) i δσi,1 = M andPN i δσi,2 = or by a matrix A if occupied by a motor with a cargo (σi = 2); if the site is empty it is represented 6 by a matrix E (σi = 0). In order to calculate averages over these configurations in the stationary state, it is necessary at first to find the algebra that must be satisfied by these matrices such that the probabilities defined in (2) satisfy the stationary conditions [14], PN,M (C ′)Γ(C ′ → C) − PN,M (C)Γ(C → C ′) = 0 (4) XC ′ where the sum extends over all configurations of M motors distributed over N − K lattice sites. Observe that the nonzero terms on the LHS of the above equation are those for which configurations C and C ′ differ from each other at most by the positions of a pair of consecutive sites, which can be reversed by any of the elementary processes defined by the dynamics in (1). In this case, each factor Γ(C ′ → C) (or Γ(C → C ′)) must be replaced by the rate w, k or p for the corresponding elementary process that brings C back from C ′ (or C ′ from C ). The algebra corresponding to the ASEP defined by the dynamics (1) has been presented in ([12]) for K = 1 : with DA − xAD = E − D DE = E EA = E EE = E x = k + p w (5) (6) Here, we shall use this same algebra to evaluate the traces over products of matrices D, A and E that appear in calculating averages over the quantities that characterize the movement of a cargo. Before proceeding, however, a few remarks are in order. 2.2 Traffic profile in the system with two cargos The model with two cargos is not ergodic. The dynamics preserves the number of empty spaces in each of the two partitions defined by the initial positions of the two cargos in the system with 7 periodic boundary conditions. In this case, all configurations C and C´ that satisfy equation (4) must share the number of empty spaces in each of the partitions. Moreover, configurations in which the empty spaces are all concentrated in one of the two partitions must be excluded, for these do not satisfy (4) with the algebra (5). We treat the initial conditions (IC ), namely the number of empty spaces - h in one of the partitions and G−h in the other, as random variables. This artifact shall account for the uncertainty one has in experimental data regarding the relative positions of the particles, and also for effects of random processes that are not explicitly described by the present model such as motor binding and unbinding at the microtubule. For computing averages, we shall account first for all possible configurations at fixed h and then average the results over h. The procedure is further specified observing that (a) because there is no reason to favor any initial configuration, we may consider that h is uniformly distributed and (b) in analogy with a situation of equilibrium, we take the average annealing as the averages over particle configurations are performed in parallel with average over h. The measure PN,M (C(h)) of a configuration C(h) of a subset-h is written as the trace over a product of matrices A, D and E that satisfy the algebra in (5): PN,M (C(h)) = 1 Z (h) N,M T r(Dp1Ek1Dp2Ek2...DpkEkk ADpk+1Ekk+1... (7) ...DppEkp ADpp+1Ekp+1...DpN −1EkN −1DpN EkN ) In the expression above, each pi is a binary variable such that pi ∈ {0, 1}, i = 1, 2, ...N and ki = 1 − pi satisfying k1 + k2 + ... + kk + kp+1 + ... + kN −1 + kN = h and kk+2 + kk+3 + ... + kp = G − h The normalization Z (h) N,M = W (h) 2−2 = W (h) 02−02 12−12 + W (h) 02−12 + W (h) 12−02 + W (h) 8 (8) (9) is conveniently expressed in terms of the weights W (h) (σi−1σiσi+1....)−(σj−1σjσj+1....). These are defined as the sum over the traces corresponding to the configurations that belong to the subset h for which the occupation of the sites ... i − 1, i, i + 1... and ...j − 1, j, j + 1... in the n-tuples are fixed and specified by the values of the corresponding site variables (σi−1, σi, σi+1, ...) and (σj−1, σj, σj+1, ...), respectively. The PN,M (C(h)) defined above must satisfy the stationary conditions PN,M (C ′ (h))Γ(C ′ (h) → C(h)) − PN,M (C(h))Γ(C(h) → C ′ (h)) = 0 (10) XC ′ (h) where the sum extends over all configurations C ′ (h) that belong to the subset h. The transition rates Γ(C ′ (h) → C(h)) lead configurations C ′ (h) into configurations C(h). 2.3 The average velocity of a cargo Consistently with the above definitions we represent the average value < v(h) > of the velocity of any of the two cargos at fixed h as < v(h) > = 1 Z (h) N,M npW (h) (21)−(2) − wW (h) (12)−(2)o (11) The configurations associated with W (h) (21)−(2) are such that the specified cargo has one motor at its right side that allows it to move one step to the right at a rate p. Similarly, in the configurations associated to W (h) (12)−(2) there is a motor at the left side of this cargo that allows it to move one step to the left at rate w. In both types of configurations the neighborhood of the other cargo is not specified. It shall be convenient to subdivide the above sum into sums over configurations having the same trace. This is achieved by specifying in (11) the occupation of the sites that precede both cargos. For this, < v(h) > is rewritten as < v(h) > = (121)−(12) + W (h) (121)−(02) + W (h) (021)−(12) + W (h) 1 Z (h) N,M nphW (h) (021)−(02)i (12)−(12) + W (h) −whW (h) (12)−(02)io 9 (12) To proceed in the evaluation of (12) it is also convenient to replace the site variables {σi} by block variables {mi} and {qi} i = 1, 2...k, that assume integer values to represent, respectively, sequences of motors and empty sites in a configuration C(h). With these, the sum over configurations that contribute to W (h) (121)−(12) for example, can be expressed as W (h) (121)−(12) = X{qi} ′X{mi} ′tr(Dm1Eq1Dm2Eq2... ...DmkEqkDmk+1ADmk+2Eqk+2...DmpEqpDmp+1A) (13) with mk+1, mk+2, mp+1 ≥ 1. Here, Dmi (or Eqi) indicates a product of mi (or qi) matrices D (or E). The symbol on the summation signals indicates that these are restricted to the configurations that satisfy the constraints in (8) for 0 < h < G. All the traces in the RHS of (12) can now be reduced with the aid of the algebra in (5). For this, we use the identity DKAE = xKAE (14) that also follows directly from (5) [12]. The results are quoted as follows W (h) W (h) W (h) (12)−(12) = Xconf (121)−(12) = Xconf (12)−(02) = Xconf (021)−(12) = Xconf (121)−(02) = Xconf (021)−(12) = Xconf (02)−(02) = Xconf W (h) W (h) W (h) W (h) ′tr(EADmiEDmj A) ′tr(EDmiAEDmj EA) ′tr(EDmiADmk EDmj A) ′tr(EDmiAEDmk EDmj AEDmk+1) = Xconf = Xconf = Xconf = Xconf = Xconf = Xconf = Xconf ′tr(EADmiEDmj A) ′tr(EDmiADmj EA) ′tr(EAEA) ′xmixmj tr(E) ′xmixmj tr(E) ′xmitr(E) ′xmitr(E) ′xmitr(E) ′xmitr(E) ′tr(E) (15) The above expressions are independent of h. The only dependence on h in the evaluation of the weights comes from the multiplicity of the configurations. Configurations for which h = 0 or h = G contribute with a factor 1/N with respect to the contributions from all other configurations that result in the same trace. 10 2.4 Average over the random variable We now take the average of < v(h) > over all realizations of h that assumes an integer value within the interval h ∈ [1, G − 1] for G > 2, with equal probability. This is performed here as < v >= Ph{pW(21)−(2) − wW(12)−(2)} Ph ZN,<M > ≡ {pW (21)−(2) − wW (12)−(2)} Z N,<M > (16) that corresponds to the average annealing in analogy to a situation of equilibrium. The averaged quantities are indicated by the bars over the corresponding symbols representing the weights W and normalization Z. Now, observe that because the traces do not depend on h, then the sums in the above expression, both in the numerator and in the denominator, account for all possible configurations of arbitrary sequences of empty and occupied sites, keeping fixed just the n-tuples indicated in each term. With this, the restrictions imposed on the sums in (15) are removed. 2.5 Sum over configurations We estimate the number of configurations that contribute to W(σi−1σiσi+1....)−(σj−1σj σj+1....) for a given n-tuple (σi−1σiσi+1....) − (σj−1σjσj+1....) by fixing the relative position of the cargos ξ and counting for all possible sequences of 0′s and 1′s. We then sum over ξ observing the invariance of the trace under cyclic transformations. The results are compiled below. 11 (N − mj − mi − 3) 2 M − mj − mi (cid:19)xmixmj tr(E) (cid:18)N − mj − mi − 4 (N − mj − mi − 3) 2 M − mj − mi (cid:19)xmixmj tr(E) (cid:18)N − mj − mi − 5 Xmi=1 (a) W (12)−(12) ≃ M −2 Xmj=1 M −mj Xmi=1 M −2 M −mj M M −2 M −1 (b) W (121)−(12) (d) W (021)−(12) ≃ (c) W (12)−(02) ≃ Xmj=1 Xmi=1(cid:18)N − mi − 4 Xmi=1(cid:18)N − mi − 5 Xmi=1(cid:18)N − mi − 5 M − 1(cid:19) [(N − 4)] tr(E) (f ) W (021)−(02) ≃ (cid:18)N − 5 M (cid:19)tr(E) 2(cid:18)N − 4 (g) W (02)−(02) ≃ 1 (e) W (121)−(02) ≃ M − mi (cid:19) [(N − mi − 3) xmi] tr(E) M − mi − 1(cid:19) [(N − mi − 4)xmi] tr(E) M − mi − 1(cid:19) [(N − mi − 4)xmi] tr(E) (17) Variables mi and mj indicate the number of possible consecutive motors at the left of the cargos in each of these configurations contributing to a given W(σi−1σiσi+1....)−(σj−1σjσj+1....). The sums over mi and mj are estimated here in the limit of very large systems for which N → ∞ and M → ∞ keeping the motor density M/N → ρ finite within the range 0 < ρ < 1. In this limit, the sums converge to integrals and these integrals can be evaluated using Laplace´s asymptotic method. Consider, for instance, the sum (a) in (17). We use Stirling´s formula N! ∼ √2πN N N e−N to approximate the factorials involving the variables N and M and define the new variables [24] z ≡ mi/N and 12 y ≡ mj/N. (18) z and y assume continuous values in this limit so that the referred sum converges to the integral e−N (1−ρ) ln(1−ρ)Z ρ 0 dyZ (ρ−y) 0 dzr 1 − y − z ρ − y − z eN h(y,z) xN zxN y (1 − y − z)3 . (19) The function h(y, z) in the expression above depends only on the sum y + z : W 12−12 ∼ N 3 (1 − ρ)4 2p2πN (1 − ρ) h(y, z) = h(y + z) = [1 − (y + z)] ln[1 − (y + z)] − [ρ − (y + z)] ln[ρ − (y + z)]. (20) Thus, by defining ν ≡ y + z, (19) can be rewritten as W12−12 ∼ Xconf N 3 (1 − ρ)4 2p2πN (1 − ρ) e−N (1−ρ) ln(1−ρ)Z ρ 0 dyZ ρ y dνr 1 − ν ρ − ν eN f2 (1 − ν)3 where f2 = h(ν) + ν ln x (21) (22) In order to apply Laplace´s method for estimating the above integral, it is convenient to change the order of the integration observing that Z ρ 0 dyZ ρ y dν(·) = Z ρ 0 dνZ ν 0 dy(·) With this change, the integral in y becomes trivial and the double integral in (21) reduces to I =Z ρ 0 νr 1 − ν ρ − ν 1 (1 − ν)3 eN f2(ν)dν (23) (24) which can be estimated by its maximum at large N . For this, notice that f2(ν) has a maximum at Thus, if ν max = 1 − xρ 1 − x . (A) xρ < 1 (25) (26) the condition ν max < ρ can not be satisfied and the maximum contribution to the integral in (24) comes from the extremum of the interval at ν = 0 which is a local maximum. The result is [25], I A,C f2 ∼(cid:20)(1 − ρ) N 2(cid:21) 1 N 2 1 √ρ exp[−Nρ ln ρ] [ln(xρ)]2 . (27) 13 If however, then ν max is localized inside the integration interval so that the integral in (24) is estimated as [25] (B) xρ > 1 (28) I B f2 ∼(cid:20)(1 − ρ) N 2(cid:21) (x − 1)2 (x)3/2 1 (1 − ρ)5/2 1 − x r 2π (1 − xρ) N expnNhln x − (1 − ρ) ln (x−1) (1−ρ)io . (29) We use this same procedure to estimate all the remaining terms in expression (16). We merely quote the results below, making some extra comments when necessary. For estimating the sum indicated as (b) in (17) we notice that the difference Xmj,mi the expression in the RHS of (b) and the sum in (a) is ∆mj ,mi between Xmj ,mi ∆mj ,mi ≡ Xconf W121−12 −Xconf (N − M)2 M −1 = − 2 W12−12 M −mj Xmj =1 Xmi=1 (cid:18)N − mj − mi − 3 M − mj − mi (cid:19) xmj xmi N − mj − mi (30) Applying Laplace´s method to the resulting integrals after taking the thermodynamic limit, it gives Xmj ,mi ∆mj ,mi ∼ −" N 3(1 − ρ)5 2p2πN(1 − ρ) e−N (1−ρ) ln(1−ρ)# 1 N 2 1 √ρ exp (−Nρ ln ρ) (ln xρ)2 , xρ < 1 ×   expnNhln x − (1 − ρ) ln (x−1) (1−ρ)io , xρ > 1 (31) or −√ρ (x)5/2 (1 − xρ) (1 − ρ)7/2r 2π (1 − x)2 N 14 The sum over configurations that contribute to W12−02 in (17) - (c) is estimated through the asymptotic behavior of a single integral, which gives W12−02 ∼ Xconf e−N (1−ρ) ln(1−ρ) (32) exp (−Nρ ln ρ) ln xρ , xρ < 1 (1 − x)2 (1 − ρ)5/2 exp(cid:26)N(cid:20)ln x − (1 − ρ) ln (1 − ρ)(cid:21)(cid:27)r 2π (x − 1) N , xρ > 1 1 N N 2(1 − ρ)4 p2πN(1 − ρ)  1 √ρ  (x)3/2 or 1 × Analogous procedures are used to estimate the sums over configurations of the kind W021−12 and W121−02 in (17) - (d) and (e),which coincide in this limit: e−N (1−ρ) ln(1−ρ) (33) (1 − x)2 (1 − ρ)5/2 expnNhln x − (1 − ρ) ln (x−1) (1−ρ)ior 2π N xρ < 1 , xρ > 1 For the remaining sums indicated in (17) - (f ) and (g), it is sufficient to estimate the relevant W121−02 ∼ N 2(1 − ρ)4 p2πN(1 − ρ) exp (−Nρ ln ρ) ln xρ , Xconf W021−12 ∼ Xconf   × √ρ 1 N or 1 (x)5/2 contributions to O(√N ) which are W02−02 ∼ Xconf N 2 (1 − ρ)4 p2πN(1 − ρ)ρ and W021−02 ∼ N Xconf (1 − ρ)4ρ p2πN(1 − ρ)ρ using the estimates above. 15 e−N [(1−ρ) ln(1−ρ)+ρ ln ρ] (34) e−N [(1−ρ) ln(1−ρ)+ρ ln ρ] (35) In the following, we analyze the results for the average velocity of a cargo obtained in this limit 3 Results The average velocity of a cargo in the system of interacting motors and cargos that obey the ASEP dynamics set in (1) can now be analyzed observing the differences in the expressions obtained above for the integrals in each of the asymptotic regions limited by the range of the product ρx of the two variables ρ and x. Such differences lead to distinct behaviors for < v > characterizing different phases of the system that, in turn, reflect the differences in the distribution of motors along the considered microtubule. A a fixed value of x such that x > 1 and for (a) 0 < ρ ≤ 1/x < v > is obtained from the behavior of the integrals for ρx ≤ 1, resulting in < v >∼ (p − w) + ln ρx [2pρln ρx + (2pρ − w)] − p(1 − ρ) 1 + 4 ln ρx + [ln ρx]2 Within the complementary region in which < v > is determined from the behavior of the integrals for ρx ≥ 1; we find (b) 1/x ≤ ρ ≤ 1 < v >∼ − k x (36) (37) (38) (39) Notice that for any x < 1 the condition (a) ρx < 1 is always satisfied so that the results for < v > are given in this case by (37) within the entire interval 0 < ρ < 1. Thus, for small values of x the system does not exhibits phase transitions. The behavior of < v > is shown in Fig.2 for the whole range of motor density ρ, at fixed k = 1 and w = 3 and at various values of x (6). We notice in these results that for varying ρ and for x slightly above 1, the average velocity of the cargo changes sign. This means that at steady state, which may be achieved at sufficiently 16 short times after an eventual change in motor density at the microtubule, cargos may adjust and change their direction of propagation moving across motor clusters. 17 4 Discussions and additional remarks The mechanism for cargo transfer envisaged here is equivalent to a hopping process in which the associated rates depend on site occupation. Because motors move and their movement is affected by the presence of the cargos and all other motors on the microtubule, the long-time dynamics of the system must be examined globally. Our results indicate that the existence of mutual interactions and the fact that many cargos are allowed to coexist at the microtubule are determinant for reproducing in this context the characteristics of the bidirectional movement. We show that within a certain range of motor density a cargo in this system executes long-range displacements in both directions. We may argue then that long-range cargo transfer is facilitated by traffic and specifically, by the assembly of motors into clusters, which characterizes traffic jam. The presence of the other cargos in the system is essential for this to occur as they function as additional obstacles that interfere in the motor density profile. Each cargo induces aggregation of motors at its back end. In turn, this provides the conditions for cargo to execute long-range displacements either backwards, over the aggregate assembled at its back end , as well as forwards, over the aggregate assembled at the back end of the cargo in front. As it was originally formulated the model does not account for the possibility that a motor with one or more attached cargos may move as well. In fact, this is the only mechanism that is usually employed to describe cargo transport and it is the basis for coordination or tug-of-war models. Also, for simplicity we have considered interactions of a cargo with a single motor at a time. Should a set of motors be allowed to interact with the cargo to participate in the transfer process then the map into the ASEP would need to be modified accordingly. We are currently working on these possibilities by including into the dynamics (1) a process of the kind 20 → 02 that recovers ergodicity of the model [26] We should emphasize that the occurrence of the long-range bidirectional movement as a conse- quence of the hopping processes devised here may happen by the action of motors of just one kind possessing a well defined polarity. Changes in motor density and related traffic profile suffice as a mechanism to control cargo direction and the size of the runs determined essentially by the extent 18 of motor clusters at jamming conditions. This offers a rather straightforward explanation for the data mentioned above suggesting that the number of long run-lengths performed by the observed beads increases significantly as the density of motors at the microtubule increases [11]. In addition, the results presented here indicate that, for sufficiently high values of the motor density for which ρ > 1/x, at the point where the model displays a phase transition, cargos would perform a uniform movement (on average) since their velocities become independent of ρ. Such behavior has also been observed in the same set of experiments. As noticed by Ma and Chisholm [6], "little is known regarding motor traffic and how it correlates with the movement of cargo". Here, we offer a possibility based on the idea that the transport does not require the action of an external agent to coordinate the process, or a tug-of-war mechanism or even the existence of a mechanical coupling between two kinds of motors as proposed more recently [8]. Instead, it suggests that such coordination can be achieved by collective effects on the course of the dynamics as the system "self-organizes" so that it presents characteristics that reflect an internal (and global) order that does not have its origin in the characteristics of the external medium. It is then possible that the necessary transport in cells is accomplished just by adjusting the density of motors at the microtubule. Accordingly, the presence of processive motors of different polarities that are normally required to explain the movement of a putative motor-cargo complex would not be necessary. Transportation here is based on a mechanism that requires formation of clusters of motors, not necessarily on their ability to travel along long distances. Acknowledgements I would like to thank Domingos H.U. Marchetti for very helpful and enthusiastic discussions regarding the conditions imposed by Fubini´s theorem and the procedure used here to estimate the double integrals; and also Elisa T. Sena for pointing to me many of the difficulties in the initial developments of this work. I thank Scott Hines for kindly editing a preliminary version of the text. 19 This work had integral support from Funda¸cao de Amparo a Pesquisa do Estado de Sao Paulo (FAPESP) - Brazil. References [1] S. P. Gross, Hither and yon: a review of bidirectional microtubule-based transport, Phys. Biol., 1, R1-R11 (2004). [2] M. A. Welte, Bidirectional Transport Along Microtubules, Curr. Biol. 14, R525-R537 (2004). [3] K.B. Zeldovich, J-F, Joanny, J. Prost, Motor proteins transporting cargos, Eur. Phys. J. E 17, 155-163 (2005). [4] R. Mallick, S. P. Gross, Molecular Motors: Strategies to Get Along, Curr. Biol., 23 R971-R982 (2004); [5] S.Klumpp, M. J. Muller, R. Lipowsky, Cooperative transport by small teams of molecular motors, Biophys. Rev. and Lett, 1, 353-361 (2006). [6] S. Ma, R. L. Chisholm, Cytoplasmatic dynein-associated structures move bidirectionally in vivo, J. Cell Sci. 115, 1453-1460 (2002). [7] I. M. Kulic, A. E. X. Brown, H. Kim, C. Kural, B. Blehm, P. R. Selvin, P. C. Nelson, V. I. Gelfand, The role of microtubule movement in bidirectional organelle transport, Proc. Natl. Acad. Sci. USA 105, 10011-10016 (2008). [8] S. Ally, A. G. Larson, K. Barlan, S. E. Rice, V. I. Gelfand, Opposite-polarity motors activate one another to trigger cargo transport in live cells, J. Cell Biol. 187, 1071 - 1082 (2009). [9] S.Klumpp, R. Lipowsky, Traffic of Molecular Motors Through Tube-Like Compartments, J. Stat. Phys. 113, 233 -268 (2003). [10] A. Seitz, T. Surrey, Processive movement of single kinesinss on crowded microtubules visualized using quantum dots, EMBO J. 25, 267 - 277 (2006). 20 [11] J. Beeg, S. Klumpp, R. Dimova, R. S. Graci`a, E. Unger, R. Lipowsky, Transport of beads by several kinesin motors, Biophys. J 94, 538-541 (2008). [12] C. Goldman, E. T. Sena, The dynamics of cargo driven by molecular motors in the context of an asymmetric simple exclusion process, Physica A 388, 3455-3464 (2009). [13] P.A. Ferrari, C. Kipnis, E. Saada, Microscopic Structure of Travelling Waves in the Asymmetric Simple Exclusion Process, Ann. Prob. 19, 226 - 244 (1991). [14] B. Derrida, S.A. Janowsky, J.L. Lebowitz, E.R. Speer, Exact solution of the totally asymmetric simple exclusion process: shock profiles, J. Stat. Phys. 73, 813 - 842 (1993). [15] R.A. Blythe, M.R. Evans, Nonequilibrium steady states of matrix-product form: a solver's guide, J. Phys. A: Math Theor. 40, R333-R441 (2007). [16] A. Parmeggiani, T.Franosch, E. Frey, Totally Asymmetric Simple Exclusion Process with Langmuir Kinetics, Phys. Rev. E70 046101-1 -- 046101-20 (2004). [17] Chowdhury, D. Traffic Flow of Interacting Self-Driven Particles: Rails and Trail, Vehicles and Vesicles. Physica Scripta, T106 (2003) 13 -- 18. [18] K. J. Verhey, J. W. Hammond, Traffic control: regulation of kinesin motors, Nat. Rev. Mol. Cell Biol. 10, 765 - 777 (2009). [19] M. Imanishi, N. F. Endres, A. Gennerich, R. D. Vale, Autoinhibition regulates the motility of the C.elegans intraflagellar transport motor OSM-3, J. Cell Biol. 174, 931-937 (2006). [20] A.Yildiz, M. Tomishige, A. Gennerich, R.D. Vale, Intramolecular strain coordinates kinesin stepping behavior along microtubules, Cell 134, 1030-1041(2008). [21] M. Nishiyama, H. Higuchi, T. Yanagida, Chemomechanical coupling of the forward and back- ward steps of single kinesin molecules. Nat. Cell Biol. 4, 790-797 (2002). [22] Y. Ishii, Y. Taniguchi, M. Iwaki, T. Yanagida, thermal fluctuations biased for directional motion in molecular motors, BioSystems 93, 34-38 (2008). 21 [23] B. Derrida, M.R. Evans, The asymmetric exclusion model: exact results through a matrix ap- proach, Nonequilibrium statistical mechanics in one dimension, University Press, UK, Chapt. 14, 277-304 (1997); B. Derrida, M.R. Evans, V. Hakim, V. Pasquier, An exact solution of a 1D asymmetric exclusion model using a matrix formulation, J. Phys.A26, 1493-1517 (1993). [24] D.H.U. Marchetti, P.A.F. da Veiga, T.R.Hurd, The 1/N-expansion as a perturbation about the mean field theory: a one-dimensional fermion model, Comm. Math. Phys. 179, 623-646 (1996). [25] J. D. Murray, Asymptotic Analysis, Springer (1984). [26] L. W. Rossi, T. Santos, C. Goldman, work in progress; [27] K. Kawaguchi, S. Uemura , S. Ishiwata, Equilibrium and transition between single- and double- headed binding of kinesin as revealed by single-molecule mechanics, Biophys. J. 84, 1103 - 1113 (2004); [28] J. Howard, Mechanics of Motor Proteins and the Cytoskeleton, Sinauer Associates Inc., 2001. 22 Figure Caption Figure 1 - Dynamics of motors and cargos. (a) Cargo transfer. It happens here through a mechanism of hopping between neighbor motors. Due to the flexibility of the tail, the attached cargo may display small oscillations leading to the possibility of it being caught either by the motor at its left or by the motor at its right. The corresponding processes 12 → 21 or 21 → 12 are represented in the figure. (b) The step of a motor. The time spent by the motor with the two heads attached to the microtubule is much larger than the time it spends with just one of the heads attached [27], as a part of the "hand-over-hand" mechanism proposed to explain the kinetics of two-headed motor proteins [28]. Occupation of a site by a motor occurs here whenever it is occupied by the two heads of a motor. The motor step is then represented as 10 → 01 which is indicated in the figure. Figure 2 - The average velocity of a cargo for k = 1 and w = 3 as a function of the density ρ of motors at the microtubule, for various values of parameter x. 23 This figure "Figure_1a.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/1006.5161v1 This figure "Figure_1b.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/1006.5161v1 This figure "Figure_2.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/1006.5161v1
1608.05433
1
1608
2016-08-18T21:09:08
Comment on "Enhanced Diffusion of Enzymes that Catalyze Exothermic Reactions" by R.Golestanian
[ "physics.bio-ph" ]
We (Riedel et al. Nature 2015), as well as others, have showed that some enzymes exhibit enhanced diffusion when active. In a recent PRL, (Golestanian, PRL 2015, arXiv:1508.03219) R.Golestanian theoretically examines a number of possible explanations for this phenomenon and concludes that "collective heating" is the best candidate to account for the observed diffusion coefficient increase. Here we present evidence showing that collective heating cannot possibly apply to our experiments.
physics.bio-ph
physics
Comment on R. Golestanian's "Enhanced Diffusion of Enzymes that Catalyze Exothermic Reactions" K. Tsekouras1, C.Riedel2, R.Gabizon2, S.Marqusee2,3, S.Pressé1,4∗ and C.Bustamante2,3,5,6,7,8∗ 1Department of Physics, IUPUI Indianapolis IN 46202 USA 2California Institute for Quantitative Biosciences, QB3, UC Berkeley, CA 94720 USA 3Department of Molecular and Cell Biology, UC Berkeley, CA 94720, USA 4Department of Cellular & Integrative Physiology, IU School of Medicine, Indianapolis IN 46202 USA 5Jason L. Choy Laboratory of Single-Molecule Biophysics and Department of Physics, UC Berkeley, CA 94720, USA 6Department of Chemistry, UC Berkeley, CA 94720, USA 7Howard Hughes Medical Institute, UC Berkeley, CA 94720, USA 8Kavli Energy Nano Sciences Institute, UC Berkeley and Lawrence Berkeley National Laboratory, CA 94720, USA. ∗corresponding author We [1] as well as others [2 -- 4] have shown that some active enzymes exhibit enhanced diffusion. In the process of building a model for this effect, we eliminated alternative explanations through ex- perimental controls; see Ref. [1] and SI. In a recent (2015) PRL by R. Golestanian [5], the author theoretically examines multiple explana- tions for increased diffusion of enzymes that cat- alyze highly exothermic reactions. One of these is collective heating (CH) that attributes the ac- tive enzyme's enhanced diffusion to a tempera- ture rise of the buffer caused by the accumulation of reaction heat in the center of the reaction con- tainer. Golestanian concludes that CH, is the best candidate explanation for our results recently pub- lished in Nature [1]. Here we present evidence to counter this claim. Our controls rule out the possibility that global or local heating of the solvent surrounding the en- zyme in our experiments is responsible for en- hanced enzyme diffusion, as predicted from a heat capacity calculation [1]. Briefly, assuming perfect thermal isolation, we cal- culate the expected maximum temperature rise, ∆T , anywhere within the container, assuming to- tal substrate depletion for our most exothermic en- zyme, catalase. Importantly, substrate is not re- plenished in our experiments. Using [C(H2O2)] = 25mM as the concentration of substrate (which for catalase is hydrogen peroxide), Cp = 4.18 · 103J/(KL) as the water heat capacity and ∆H = 100kJ/mol as the enthalpy of the reaction, we get: ∆T = [C(H2O2)]∆H/Cp ≈ 0.6K. (1) Interpolating the 300.6K viscosity from experi- mental values [6] and using the Stokes-Einstein's relation yields this ratio of enhanced (T = 300.6K) to unenhanced (T = 300K) diffusion coefficient: D′ D = kT ′ 6πη′r 6πηr kT → D′ = D T ′η T η′ ≈ 1.015D. (2) the spatially averaged Note that 300.6K is not temperature rise within the container. Instead it is the maximum temperature rise possible at any point inside the experimental container irrespec- tive of the speed with which the reaction happens. Since substrate is depleted and not replenished, no more heat can ever be produced within the container. What is more, this temperature rise is totally insufficient to explain the 25% diffusion co- efficient rise for catalase. If, instead, we choose to calculate the temperature rise expected to occur during our experiment, we employ catalase's concentration ([Ccat] = 1nM) and catalytic rate (k = 2.5 · 104s−1 that we mea- sured at 300K at saturation) plus the experiment duration (t = 30s), and the number of catalase's catalytic sites (n = 4), to find: ∆T = ∆Q Cp = nkt[Ccat]∆H Cp ≈ 0.072K (3) where once more we assumed no heat loss. In- terpolating the 300.072K viscosity from experi- mental values [6] as before and using again Eq.2 yields this ratio of enhanced (T = 300.072K) to unenhanced (T = 300K) diffusion coefficient: D′ = D T ′η If we take the Arrhe- nius rate enhancement into consideration, we get T η′ ≈ 1.00112D. all enzymes would experience the same heating and exhibit size, not activity, dependent diffusion; so CH cannot apply to those experiments either. Summary: Collective heating cannot explain our [1] and other [4] experiments on enhanced en- zyme diffusion. This said, collective heating could very well apply in systems where many orders of magnitude more heat is produced. Acknowledgements We thank A. Sen, G. King, D. Makarov, A. Lee and K. Ghosh for help- ful discussions. We acknowledge support from NIH grants R01-GM0325543 (C.B.) and R01- GM05945 (S.M.), the US DoE, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering under contract no. DE-AC02- 05CH11231 (C.B.), the NSF grants MCB-1412259 (S.P.) and MCB-1122225 (S.M.), the Human Fron- tier Science Program (C.R) and the Burroughs- Wellcome Fund (S.P). [1] C. Riedel, R. Gabizon, C. A. M. Wil- son, K. Hamadani, K. Tsekouras, S. Mar- qusee, S. Pressé, and C. Bustamante, Nature 517, 227 (2015). [2] S. Sengupta, K. K. Dey, H. S. Muddana, T. Tabouil- and A. Sen, lot, M. E. Journal of the American Chemical Society 135, 1406 (2013). Ibele, P. J. Butler, [3] H. S. Muddana, Sen, A. Mallouk, Journal of the American Chemical Society 132, 2110 (2010). and P. S. Sengupta, T. E. Butler, J. [4] K. K. Dey, S. Das, M. F. Poyton, S. Sen- and A. Sen, gupta, P. J. Butler, P. S. Cremer, ACS Nano 8, 11941 (2014). [5] R. Golestanian, Physical Review Letters 115, 108102 (2015). [6] J. C. Crittenden, R. R. Trussell, D. W. and G. Tchobanoglous, Hand, K. J. Howe, "MWH Water Treatment Principles and Design. Third Edition - Springer," (2012). ≈ 1.00118D. Diffusion coefficient increases of D′ 0.0112% or 0.00118% are also nowhere near the 25% observed for active labeled catalase. Our measurements were taken during a window of just 30s, starting 10s after the enzyme was added to the solvent-substrate mix. In all cases, unavoid- able heat loss further reduces these small tem- perature changes. In order to obtain a larger diffusion coefficient en- hancement, Golestanian ignores substrate deple- tion and assumes steady state (and thus replen- In our experiments, since ishment of substrate). we never replenish substrate, the only steady state that can be reached is when substrate has been depleted. We add finally that CH cannot explain why la- beled active urease shows a larger diffusion in- crease than labeled inactive urease placed in the same container as unlabeled active catalase, de- spite active catalase producing more heat than ac- tive urease as seen in Fig. 3 of the Ref. [1] SI. Nor can CH explain why labeled inactive urease in the same container as unlabeled active catalase shows smaller diffusion rises than labeled active catalase, despite the fact that in identical contain- ers the same amounts of active catalase produce identical amounts of heat. In addition, a simple (Gaussian) likelihood ratio test reveals that our control (Ext. Data Fig.3 [1]) is 22× more likely to show no trend (fixed urease diffusion at 31.5µm2/s) than a 25% increase (as would be expected in the presence of active cata- lase had Golestanian's CH hypothesis held for our experiment). Finally, ignoring our own experiments altogether, Dey et al. [4] show enzymes in the same device separating out in physical space on the basis of catalytic activity. If CH held for these experiments, and since all enzymes share the same container, 2
1512.01476
3
1512
2018-02-06T20:36:37
Spatial fluctuations at vertices of epithelial layers: quantification of regulation by Rho pathway
[ "physics.bio-ph", "cond-mat.soft", "cond-mat.stat-mech" ]
In living matter, shape fluctuations induced by acto-myosin are usually studied in vitro via reconstituted gels, whose properties are controlled by changing the concentrations of actin, myosin and cross-linkers. Such an approach deliberately avoids to consider the complexity of biochemical signaling inherent to living systems. Acto-myosin activity inside living cells is mainly regulated by the Rho signaling pathway which is composed of multiple layers of coupled activators and inhibitors. We investigate how such a pathway controls the dynamics of confluent epithelial tissues by tracking the displacements of the junction points between cells. Using a phenomenological model to analyze the vertex fluctuations, we rationalize the effects of different Rho signaling targets on the emergent tissue activity by quantifying the effective diffusion coefficient, the persistence time and persistence length of the fluctuations. Our results reveal an unanticipated correlation between layers of activation/inhibition and spatial fluctuations within tissues. Overall, this work connects the regulation via biochemical signaling with mesoscopic spatial fluctuations, with potential application to the study of structural rearrangements in epithelial tissues.
physics.bio-ph
physics
Biophysical Journal Volume: 00 Month Year 1 -- 8 1 Spatial Fluctuations at Vertices of Epithelial Layers: Quantification of Regulation by Rho Pathway ´E. Fodor,1,2,∗ V. Mehandia,3,4,5,∗ J. Comelles,3,4 R. Thiagarajan,3,4 N. S. Gov6 P. Visco,2 F. van Wijland,2 D. Riveline3,4 8 1 0 2 b e F 6 ] h p - o i b . s c i s y h p [ 3 v 6 7 4 1 0 . 2 1 5 1 : v i X r a 1DAMTP, Centre for Mathematical Sciences, University of Cambridge, Cambridge, United Kingdom; 2Laboratoire Mati`ere et Syst`emes UMR 7057 CNRS/P7, Universit´e Paris Diderot, Paris cedex 13, France; 3Laboratory of Cell Physics, ISIS/IGBMC, Universit´e de Strasbourg and CNRS (UMR 7006), Strasbourg, France; 4Development and Stem Cells Program, IGBMC, CNRS (UMR 7104), INSERM (U964), Universit´e de Strasbourg, Illkirch, France; 5School of Mechanical, Materials and Energy Engineering, Indian Institute of Technology, Ropar, India; and 6Department of Chemical Physics, Weizmann Institute of Science, Rehovot, Israel Abstract In living matter, shape fluctuations induced by acto-myosin are usually studied in vitro via reconstituted gels, whose proper- ties are controlled by changing the concentrations of actin, myosin and cross-linkers. Such an approach deliberately avoids to consider the complexity of biochemical signaling inherent to living systems. Acto-myosin activity inside living cells is mainly regulated by the Rho signaling pathway which is composed of multiple layers of coupled activators and inhibitors. We investigate how such a pathway controls the dynamics of confluent epithelial tissues by tracking the displacements of the junction points between cells. Using a phenomenological model to analyze the vertex fluctuations, we rationalize the effects of different Rho signaling targets on the emergent tissue activity by quantifying the effective diffusion coefficient, the persistence time and persistence length of the fluctuations. Our results reveal an unanticipated correlation between layers of activation/inhibition and spatial fluctuations within tissues. Overall, this work connects the regulation via biochemical sig- naling with mesoscopic spatial fluctuations, with potential application to the study of structural rearrangements in epithelial tissues. ∗These authors contributed equally to this work. INTRODUCTION Changes in shapes of cells and tissues are mediated by the acto-myosin cytoskeleton. To reproduce the dynamics of this network, minimal systems made of actin filaments, myosin motors and cross-linkers are synthetized in vitro (1 -- 5). The mechanics and dynamics of such active gels are controlled by varying the concentration of their various com- ponents. Activity of each component is monitored by adding some inhibitor drugs, and/or by tuning the ATP concen- tration of the system. Recent experimental evidence have shown the relevance of this approach to investigate the role of motors and cross-linkers in the emerging properties of the network (2, 6). These studies are based on tracking the motion of tracers injected in active gels: analyzing the spon- taneous fluctuations of such tracers enables one to extract information about the activity of internal motors. In multicellular systems, such as tissues of develop- ing embryos, acto-myosin drives the morphogenesis: dra- matic rearrangements leading to the formation of distinct organs (7, 8). This remodelling is mainly under the con- trol of intracellular activity, which powers spatial fluctua- tions (9), and intercellular interactions mediated by adhesion between neighboring cells (10, 11). In contrast to synthetic gels, the internal regulation of the cellular acto-myosin activ- ity is more complex in vivo. Therefore, extending the in vitro approach, based on controlling externally the activity of each specific component, to in vivo situations requires new strategies. The Rho signaling pathway is known to regulate the acto- myosin activity in living cells (12). It also controls cell-cell junctions (13) and the elasticity of stress fibers (14). Such a pathway can be viewed as a series of activators and inhibitors installing a hierarchy of potential targets (15, 16). Activa- tions and inhibitions controlled by each target are such that anticipating their net effects on the tissue fluctuations, pow- ered by acto-myosin activity, remains a challenge (17, 18). In that respect, the inherent complexity of internal activity in vivo calls for new experiments and quantitative analysis © 2013 The Authors 0006-3495/08/09/2624/12 $2.00 2 to bridge the biochemical signaling of the Rho pathway with the emerging tissue dynamics. In this paper, we explore the regulation of active fluc- tuations by the Rho pathway in epithelial monolayers. We measure these fluctuations by tracking tricellular junctions or vertices over time. In contrast with active gels, our analy- sis of internal fluctuations does not require to inject external tracers. Based on a phenomenological model, we quantify key parameters of junction activity: their effective diffusion coefficient, as well as the persistence time and persistence length of spatial fluctuations. We report modifications of these parameters for various targets along the signaling path- way. These results support that, for the inhibitions that we considered, the active fluctuations of the vertices are reduced when going downstream in the Rho pathway inhibition. MATERIALS AND METHODS Experiments were performed with MDCK II cells stably expressing E-cadherin GFP (Nelson Lab.). We culture cells in DMEM containing 10% Fetal Calf Serum (FCS) and antibiotics. We replate them on glass coverslips (CS) of 25 mm diameter for live cell imaging. When the cell mono- layer covered 70% of the CS area, we firmly place the sam- ple at the bottom of a custom made metallic holder. For acquisition, we change the medium to L15, 10% FCS and antibiotics. We use the following inhibitors from myosin up to Rho at optimal concentrations following the manufac- turer recommendations: inhibition of acto-myosin by ML-7 (Sigma-Aldrich, 10 µM), inhibition of Rho kinase (ROCK) by Y-27632 (Sigma-Aldrich, 10 µM), and inhibition of Rho by C3 Transferase (Cytoskeleton, 0.04 µM). Note that the use of blebbistatin to inhibit directly myosins yielded some detrimental effects, leading us to rather use ML-7 instead. For observation, we use a motorized inverted micro- scope (Nikon Eclipse Ti), equipped with a 12 bit CCD camera (Photometric CoolSNAP HQ2). The setup is tem- perature controlled at 37◦C (Life Imaging Services). We check with fluorescent beads (4 µm, TetraSpeck, Invitro- gen Molecular Probes) grafted on CS surface that no drift appears during 24 hours of live imaging after 2 hour sta- bilization. We take pictures of the monolayer every 5 min during the next 8 hours with multiple z-stacks 1 µm apart. They span 3 µm depth of the cell monolayer. We merge the z-stacks into one image by using the maximum intensity pro- jection. We then extract vertex positions from the sequence of merged images by manually clicking in each frame as long as they are visible. The procedure was validated in its precision through automatised detection as well. For each condition, we check that the average cell area was always about 180 ± 15 µm2, and we consider more than 20 vertices for at least 3 biological repeats. Biophysical Journal 00(00) 1 -- 8 RESULTS Vertex tracking and inhibitors in the Rho pathway We use Madin Darby Canine Kidney (MDCK) cells sta- bly transfected with E-cadherin fused with the Green Flu- orescent Protein (GFP) as a paradigm for epithelial tis- sues dynamics (19). This allows us to study live cells while interacting with each other. We seek to identify spa- tial points primarily involved in tissue transformations. The meeting points between three cells are involved in exchanges between neighbouring cells, thus serving as hallmark of tis- sue dynamics (11, 20). Some specific proteins, such as tri- cellulin, are known to accumulate at this point in cell culture. Besides, vertices have also attracted the attention of develop- mental biologists which exhibit accumulation of proteins at these specific points (21). Vertex dynamics are driven both by thermal fluctua- tions and by active fluctuations powered by some inter- nal nonequilibrium processes such as motor-induced forces, actin polymerization and cell-cell adhesion. The myosin-II motors are localized in dense contractile units present in the apical surface of the tissue (22, 23) [Figs. 1(a-c)]. The active forces lead to large displacements of the vertices dis- tinct from the thermal fluctuations of smaller amplitude. We focus here on large displacements which do not lead to any topological transitions in the tissue [Figs. 1(d-e)]. The Rho signaling pathway controls the acto-myosin activity inside cells, namely the forces induced by myosin and/or by actin polymerization (see Fig. 9 in (16)). In this study, we focus on the downstream targets that affect myosin to establish the principles for the validity and relevance of our framework. Upstream and downstream targets install a hierarchy in the activation of myosin. We specifically inhibit the following targets: Rho, Rho kinase (ROCK) and myosin-II [Fig. 1(f)]. Each inhibitor is specific to its target and incubated at the optimal concentrations for its inhibition. We used standard concentration values already utilized in other cell biology studies for many cell types including MDCK (24 -- 27). Alto- gether, we probe four conditions on the same system by considering untreated cells as a control. To demonstrate that the chosen inhibitors and their con- centration specifically act on the contractile state of cells, we investigate their effect on both architecture and contrac- tile forces within tissues by measuring single cell area and myosin cluster area in each condition. The distribution of single cell area is not strongly affected in the myosin-II inhibitor case compared with control, as apparent from the mean value in Fig. 1(g). The distribution gets modified in the Rho kinase and Rho inhibitor cases, with mean values being slightly reduced and increased, respectively. In con- trast, the polygonicity distribution remains approximately the same for all conditions at different times, supporting that tissue architecture is barely affected by external inhibitors (see Fig. S1 in the Supporting Material). Myosin clusters have been shown to be relevant read- outs to assess the contractile state of cells (28, 29). Along this line, measurements of cluster characteristics are infor- mative. In our case, the mean density value of myosin clus- ters for control is larger than in the inhibited cases (see Figure 1: Study of vertex fluctuations. (a) Actin (red) and myosin (green) structures at the apical surface of a MDCK cell (scale bar 3 µm). The myosin is concentrated in dense contractile units (yellow arrow) referred to as myosin clus- ters. (b) Actin structure alone. (c) Myosin structure alone. (d) We visualise MDCK cell monolayer by GFP E-cadherin (scale bar 30 µm). (Inset) We identify the meeting points between three cells as the privileged point for our analy- sis (scale bar 4 µm). (e) Extraction of a typical transition between two locally stable positions in vertex trajectory (total time 8 hours). (f) Simplified diagram of the Rho path- way installing an order relation in myosin activation, as pre- sented in (16); in red the specific inhibitors and their targets. (g) Area of individual cells in each condition. C: Control; My: Myosin inhibitor; Rk: Rho kinase (ROCK) inhibitor; R: Rho inhibitor. Number of experiments × number of cells = C: 4×235; My: 4×246; Rk: 3×249; R: 3×148. (h) Area of myosin clusters. Number of experiments × number of clus- ters = C: 2 × 312; My: 2 × 186; Rk: 2 × 197; R: 2 × 257. Statistical analysis with one-way ANOVA test: ns (non sig- nificant) p > 0.05, * p < 0.05, ** p < 0.01, *** p < 0.001 (see the Supporting Material). 3 Fig. S2 in the Supporting Material), suggesting a decrease in force generation for the same level of myosin per cell. In addition, the area of each myosin cluster is smaller in con- trol than in other conditions [Fig. 1(h)]. Besides, this area increases the more downstream along the Rho pathway inhi- bition, suggesting a relaxation of the myosin pool in the apical side, consistently with the notion that myosin-induced forces are reduced. Altogether, our analysis of clusters con- firms that we are acting on the contractile state of cells. Since inhibitors are specific and used at their optimal con- centrations, these measurements support the validity of our experimental approach. Statistics of vertex displacement: inhibitors affect spatial fluctuations Our goal is to investigate how the emergent fluctuations of the tissue are regulated by the Rho pathway. To this aim, we first demonstrate that the inhibitors in the pathway affect these fluctuations by extracting the statistics of displace- ments from the vertex trajectories. This allows us to assess the existence of a direct link between biochemical signaling and mechanical fluctuations. In our analysis, we do not con- sider neither spatial inhomogeneities nor topological transi- tions that occur in the tissues. In that respect, we measure vertex trajectories in the absence of neighboring cell divi- sion, by tracking them as long as they are visible until a maximum of 8 hours. We compute the projected one-dimensional mean square displacement (MSD) within the four different conditions [Fig. 2(a)]. For each condition, the short time MSD exhibits a power-law behavior with exponent close to 0.7 over about one decade. Interestingly, a subdiffusive behavior was reported for the dynamics of vertices in the endoplasmic reticulum (30). The large time MSD depends on conditions and exhibits a behavior that, for simplicity, we have char- acterized by a power-law. The corresponding exponent is typically larger than 1, except for myosin inhibitor where it is smaller. The crossover between the two regimes appears between 20 and 60 min. Fluctuations are reduced in the myosin inhibited case, which has the lowest MSD, and they are enhanced for the Rho inhibitor, where the long time MSD is the largest. We also explore the full statistics of vertex dis- placement by measuring the probability distribution function (PDF) for each condition, as shown in Figs. 2(b-e). At short time the PDF is Gaussian, while it exhibits broader tails at large time. These tails reveal large displacements of the ver- tex, and were already observed for tracer particles in active gels (31) and living cells (32 -- 36). They are more pronounced in the Rho inhibitor case, as a signature of larger fluctuations, possibly due to directed motion events. Biophysical Journal 00(00) 1 -- 8 4 Figure 2: Statistics of vertex displacement. (a) Mean square displacement as a function of time in four conditions: con- trol (black), myosin inhibitor (blue), Rho kinase inhibitor (orange), and Rho inhibitor (red). The corresponding best fitting curves are in solid lines. The blue and red dashed lines report the large time behaviors. (b-e) Distribution of displacement for the four conditions at three times: 5 (•), 25 (+), and 60 min (◦). Exponential tails appear at long times as a consequence of directed motion events in vertex dynamics. Results of simulated dynamics are in solid lines. Phenomenological model of vertex dynamics: tran- sient confinements and large displacements To quantitatively discriminate between the effects of the different inhibitors, we analyze our measurements with a nonequilibrium model previously introduced to describe tracer fluctuations inside living cells (36). This model is not aimed at describing any specific process that produces the active fluctuations, it rather formulates a general framework that allows one to quantify non-equilibrium forces and fluc- tuations. We regard the vertex as a virtual particle which dynamics is prescribed by two coupled equations: (i) an equilibrium diffusion of the vertex in a cage, modelled as an harmonic potential of stiffness k -- the displacement is driven by a Gaussian white noise of variance 2γkBT with a drag force of coefficient γ; (ii) a non-Gaussian colored diffusion Biophysical Journal 00(00) 1 -- 8 equation for the center of the cage, mimicking nonequilib- rium activity as a run-and-tumble dynamics. Inspired by the large ballistic-like displacements that we observe in experi- mental trajectories, we model this active noise as a two-state Poisson process: the cage has a constant velocity v in a ran- dom uniformly sampled two-dimensional direction during a random persistence time of average τ, and it remains fixed during a random quiescence time of mean τ0. We understand the confinement as an elastic mechanical stress resulting from cells surrounding each vertex, and the nonequilibrium motion of the cage as an active stress. The effect of this active stress is to reorganize the structure of the monolayer, and therefore to spatially redistribute the elastic mechanical stress. In the absence of activity, this model predicts a short time diffusion, and then a large time plateau expressing the elastic confinement. Such dynamics is entirely under the control of equilibrium thermal fluctuations. In an active sys- tem, nonequilibrium processes enhance vertex displacement via the cage motion, yielding a free diffusion of the ver- tex with coefficient DA = (vτ )2/[2(τ + τ0)]. The large time dynamics is fully determined by the active parame- ters {v, τ, τ0}, whereas thermal fluctuations control the short times via {k, γ, T}. A sub-diffusive transient regime appears between the two diffusions, as a crossover towards a plateau, and a super-diffusive regime can also precede the large time diffusion, as a signature of the ballistic motion involved in the active noise. In such a case, thermal effects are negligi- ble at times larger than τc = (cid:112)τ kBT /(kDA), a timescale quantifying the transition from the short time equilibrium- like dynamics to the large time active diffusion. Simulated trajectories exhibit clusters of similar size accounting for the transient confinement of the vertex. Occasionally large dis- placements of order vτ appear. The vertices do not only fluc- tuate around a local equilibrium position, they also undergo rapid directed jumps (compare Fig. 1(e) and Fig. 3(a)). In contrast to previous works which describe the many- body dynamics of cells in the tissue in a more complete framework (37 -- 39), we do not explicitly account for inter- actions between neighbouring vertices. As a result of such interactions, the cells experience an intermittent dynam- ics alternating between fluctuations of small amplitude and rapid large displacements. As in glassy systems, the jumps appear through collective rearranging regions (40), thus con- tributing to the non-Gaussian fluctuations experienced by the vertices. Even though non-Gaussian fluctuations are present independently of topological transitions, the existence of a quantitative connexion between collective rearrangements and such transitions is still an open question (41). Our approach consists in reducing the dynamics of a large num- ber of interacting vertices into the dynamics of a single vertex embedded in an effective background that describes the mean-field effect of the surrounding system. Within our model, interactions are embodied by both the elastic confine- ment and the active source of fluctuations leading to large displacements. We argue below that our phenomenological approach is sufficient to capture the vertex dynamics, since it provides a framework to decipher the effects of the Rho pathway inhibitions on this dynamics. Order relation in the active diffusion coefficient and the persistence time We fit the MSD data with our analytic prediction to estimate the parameters characterizing nonequilibrium activity. Our fits convincingly capture the transition from sub-diffusive to super-diffusive like behaviors [Fig. 2(a)]. Within our model, these behaviors correspond to cross-over regimes between the short and large time diffusion. We extract from the best fits a single set of passive and active parameters for each condition. Our estimate for τc can be compared with Figure 3: Active parameters of vertex fluctuations. (a) Typ- ical trajectory obtained from simulations of the vertex dynamics in control condition (scale bar 1 µm). Isotropic "blobs" reveal equilibrium-like transient confinement during a typical time τc (dashed blue box), and large displacements of order vτ occur due to nonequilibrium activity (orange arrows). (b) Best fit values of the active diffusion coeffi- cient, (c) the persistence time, (d) the persistence length, and (e) the energy dissipation rate. Statistical analysis with t-test: ns (non significant) p > 0.05, * p < 0.05, ** p < 0.01, *** p < 0.001 (see the Supporting Material). 5 timescales quantifying the transition from elastic to fluid- like behavior of the material (see Tab. S1 in the Supporting Material), which is of the same order as the Maxwell time reported in three-dimensional cell agregates, i.e. about 30- 40 min (9, 42 -- 44). We report clear quantitative variations of both the active diffusion coefficient DA and the persis- tence time τ for all conditions [Figs. 3(b,c)]. This suggests that our model, based on separating purely active fluctua- tions from equilibrium thermal ones, is a reliable framework to capture the effects of our inhibitors on tissue dynamics. Note that passive parameters, such as the relaxation time scale τr = γ/k reported in Tab. S1 (see the Supporting Material), have also different values between the conditions. This reflects the effects of inhibitors on tissue mechanics, showing that inhibitors also affects the characteristics of pas- sive fluctuations. In this respect, DA and τ are the parame- ters that characterize only the active contribution to vertex fluctuations, which is the main focus of our study. Strikingly, DA and τ are larger for Rho inhibitor than for Rho kinase inhibitor, and than for direct myosin inhibi- tion [Figs. 3(b,c)]. The more upstream the inhibition along the pathway, the larger the amplitude of fluctuations and the more persistent the ensuing displacement. The myosin inhi- bition leads to the smallest DA and τ values, suggesting that the mesoscopic activity of vertices is strongly affected. The Rho kinase target directly activates the myosin, but it also inhibits the myosin light chain phosphatase (MLCP) which in turn inhibits the myosin. Therefore, the result of Rho kinase inhibition on myosin can not be anticipated a priori. Our analysis shows that activity of vertices is less affected than for the myosin inhibitor case: the corresponding value of DA for Rho kinase inhibitor is close to the one for con- trol, which suggests a compensation between activation and de-activation of myosin. Order relation confirmed by persistence lengths To gain further insight into the active component of the dynamics, we compare the displacement PDF extracted from the simulated trajectories of vertex dynamics with experi- mental distributions. The distribution at short time is Gaus- sian and entirely controlled by the passive parameters: the simulations with or without active fluctuations, where we use passive parameters estimated from fits of MSD data, give the same results at short times (see yellow curves in Figs. 2(b- e)). Including the active component for the dynamics leaves us with one remaining free parameter: the average persis- tence length vτ of large displacements. The short time Gaus- sian remains unchanged, whereas exponential tails develop at large times in the simulated PDF. The tails are more pro- nounced as time increases, while the central Gaussian part barely changes. We adjust the vτ value by matching the tails appearing in numerical results and experimental data. Biophysical Journal 00(00) 1 -- 8 vτTimeτcConfinement1µm(a)CMyRkR01234567DA(10−2µm2/min)∗∗∗∗∗∗∗∗∗(b)ActivediffusioncoefficientCMyRkR04080120160τ(min)∗ns∗∗(c)PersistencetimeCMyRkR01234J/kBT(s−1)nsnsns(e)EnergydissipationrateCMyRkR024681012vτ(µm)∗∗∗nsns(d)Persistencelength 6 The simulated PDFs compare very well with experi- ments at large times, showing that our simulations repro- duce the evolution of experimental distributions at all times [Figs. 2(b-e)]. This is one of the main success of our analy- sis: the phenomenological model on which relies the quan- tification of internal activity is able to capture the strong non-Gaussian tails of the distribution with only one free parameter. This supports the underlying picture that vertex dynamics essentially alternates between transient confine- ment and directed motions. The order of magnitude of the extracted mean persistence length vτ is consistent with our measurements [Fig. 3(d)]. We report again the same order relation within the Rho pathway, i.e. an increase of vτ from myosin inhibitor to Rho kinase inhibitor, and from Rho kinase inhibitor to Rho inhibitor, as a signature of enhanced directed motion. Dissipation is constant for all inhibitors A major asset of our model is that it allows us to predict the mean rate of energy dissipated by the vertex dynamics in its surrounding environment. It is defined as the differ- ence between the power injected by the fluctuating ther- mal force and the one that the moving vertex dissipates via the drag force: J = (cid:10) x(γ x − √ 2γkBT ξ)(cid:11), where x is the vertex velocity, and ξ is a zero-mean Gaussian white noise (45, 46). It vanishes for systems in a thermodynamic equilibrium state. The active nonequilibrium fluctuations lead to a non zero dissipation rate: J = kDA/(1+τ /τr) (47). This rate of dissipated energy reflects the excess power injected by nonequilibrium internal activity driving the large displacements of vertices. When computing the dissipation rate in the four con- ditions, it appears as approximately constant [Fig. 3(e)], in contrast with the order relation found for active parameters [Figs. 3(b-d)]. This supports that the same amount of energy is dissipated by the vertex large displacements, though the features of such displacements differ between conditions. Given that J depends both on parameters of active fluc- tuations {DA, τ} and on parameters of passive mechanics {γ, k}, our result suggests that there may be an underlying coupling between mechanical properties of the tissue and its nonequilibrium fluctuations. A possible interpretation is that the nonequilibrium processes at the origin of active fluctua- tions, such as forces induced by myosin and by actin poly- merization, might also affect the tissue mechanics in such a way that the dissipation rate remains unchanged over all the conditions. In that respect, we observe that the relaxation time τr is increased for Rho inhibitor case with respect to oth- ers (see Tab. S1 in the Supporting Material), as also observed for the persistence time τ [Fig. 3(c)]. Biophysical Journal 00(00) 1 -- 8 DISCUSSION The parameters of vertex fluctuations reveal a correlation between the Rho pathway hierarchy and the junction active fluctuations: the higher up the Rho pathway is the inhibi- tion, the lower is the effect on decreasing the fluctuations of the vertices. Such a relation highlights the usefulness of our methodology to probe quantitatively how signal- ing pathways control emergent fluctuations. In that respect, our approach bridges biochemical signaling pathways with mechanical fluctuations in vivo. It could be used to char- acterize quantitatively the effect on fluctuations of other inhibitions acting on different signaling pathways. Our analysis is based on a phenomenological model which deliberately avoids a detailed description of the many processes operating at the subcellular scale. It is also dis- tinct from other models which consider explicit interac- tions between neighbouring cells. In that respect, our results do not rely neither on the microscopic details of activity- induced forces nor on the form of interactions within the tissue. We rather postulate an effective vertex dynamics by explicitly distinguishing thermal and active fluctuations. As a result, any underlying mechanism which could rational- ize the relation between pathway inhibition and mechanical fluctuations is out of the scope of our approach. Yet, the relation between biochemical signaling and active fluctua- tions that we quantified in tissues could help future studies to understand the precise relation between the subcellular cytoskeleton dynamics and the emergent vertex fluctuations. In that respect, a possible extension of the model could con- sist in considering more complex activation/inhibition in the Rho pathway and their connection with contractile forces in the cell (17, 18). In order to test the robustness of our model, one could measure the response of vertices to an external perturba- tion (48). The intracellular mechanics is viscoelastic in a large variety of living systems. Our phenomenological model has already been extended to account for a complex rheology (49). It would be interesting to determine whether the relation between fluctuations and pathway inhibitions is similar when including viscoelastic effects. Moreover, applying an external potential to confine a vertex, one could extract work from the vertex fluctuations by varying in time the potential parameters. Our framework allows one to pre- dict the details of the extracted work as a function of the active fluctuation characteristics (47). Confronting such pre- dictions with experimental results would provide another test for the validity of our approach. The response in living systems is not related to the spon- taneous fluctuations, in contrast to equilibrium where such a relation is given by the fluctuation-dissipation theorem (FDT). Therefore, by comparing the response with sponta- neous fluctuations of tissues, one could quantify the depar- ture from the FDT, namely the deviation of the dynamics from equilibrium. Moreover, recent progress in nonequilib- rium statistical mechanics have shed light on the direct rela- tion between the rate of energy dissipated by the internal nonequilibrium processes and the violation of the FDT (50). This has already led to access to the dissipation rate in some biological contexts (49, 51). In that respect, compar- ing response and fluctuations will allow one to propose an alternative quantification of dissipation rates in epithelial tissues. The Rho pathway is conserved across species, suggest- ing that regulations of activity in epithelial monolayers may share common pathways in a large variety of tissues and organisms (12). Our approach, based on a quantitative anal- ysis of vertex fluctuations, could serve as a novel framework to decipher the complex regulation of spatial fluctuations by the Rho pathway in other model systems. In that respect, it could be used to analyze vertex fluctuations in developing embryos, such as in Drosophila or in C. elegans, where inter- nal reorganizations are driven by spontaneous topological transitions (7, 8). Author contributions V.M., J.C., R.T., and D.R. designed and performed the exper- iments. ´E.F., N.S.G., P.V., and F.v.W. designed the model. All authors contributed to analyzing the data and writing the manuscript. Acknowledgements We thank F. Graner for helpful discussions. We acknowl- edge J. W. Nelson for sending the MDCK cell lines. We also thank N. Maggartou for extraction of data and lab. members for discussions. This work is supported by CNRS, FRC and University of Strasbourg. N.S.G. is the incumbent of the Lee and William Abramowitz Professorial Chair of Biophysics. SUPPORTING MATERIAL An online supplement to this article can be found by visiting BJ Online at http://www.biophysj.org. References 1. Backouche, F., L. Haviv, D. Groswasser, and A. Bernheim- Groswasser, 2006. Active gels: dynamics of patterning and self-organization. Physical Biology 3:264. 2. Soares e Silva, M., B. Stuhrmann, T. Betz, and G. H. Koen- derink, 2014. Time-resolved microrheology of actively remod- eling actomyosin networks. New J. Phys. 16:075010. 3. Blanchoin, L., R. Boujemaa-Paterski, C. Sykes, and J. Plas- tino, 2014. Actin dynamics, architecture, and mechanics in cell motility. Physiological reviews 94:235 -- 263. 4. Murrell, M., P. W. Oakes, M. Lenz, and M. L. Gardel, 2015. 7 Forcing cells into shape: the mechanics of actomyosin contrac- tility. Nature Reviews Molecular Cell Biology 16:486 -- 498. 5. Schuppler, M., F. C. Keber, M. Kroger, and A. R. Bausch, 2016. Boundaries steer the contraction of active gels. Nature Communications 7:13120. 6. Soares e Silva, M., M. Depken, B. Stuhrmann, M. Korsten, F. C. MacKintosh, and G. H. Koenderink, 2011. Active mul- tistage coarsening of actin networks driven by myosin motors. Proc. Natl. Acad. Sci. USA 108:9408 -- 9413. 7. Rauzi, M., P. Verant, T. Lecuit, and P.-F. Lenne, 2008. Nature and anisotropy of cortical forces orienting Drosophila tissue morphogenesis. Nature Cell Biology 10:1401 -- 1410. 8. Guillot, C., and T. Lecuit, 2013. Mechanics of Epithelial Tissue Homeostasis and Morphogenesis. Science 340:1185 -- 1189. 9. Marmottant, P., A. Mgharbel, J. Kafer, B. Audren, J.-P. Rieu, J.-C. Vial, B. van der Sanden, A. F. M. Mar´ee, F. Graner, and H. Delanoe-Ayari, 2009. The role of fluctuations and stress on the effective viscosity of cell aggregates. Proc. Natl. Acad. Sci. USA 106:17271 -- 17275. 10. Cavey, M., M. Rauzi, P.-F. Lenne, and T. Lecuit, 2008. A two-tiered mechanism for stabilization and immobilization of E-cadherin. Nature 453:751 -- 756. 11. Bardet, P.-L., B. Guirao, C. Paoletti, F. Serman, V. Lopold, F. Bosveld, Y. Goya, V. Mirouse, F. Graner, and Y. Bellaıche, 2013. PTEN Controls Junction Lengthening and Stability dur- ing Cell Rearrangement in Epithelial Tissue. Developmental Cell 25:534 -- 546. 12. Etienne-Manneville, S., and A. Hall, 2002. Rho GTPases in cell biology. Nature 420:629 -- 635. 13. Reyes, C., M. Jin, E. Breznau, R. Espino, R. Delgado-Gonzalo, A. Goryachev, and A. Miller, 2014. Anillin Regulates Cell- Cell Junction Integrity by Organizing Junctional Accumula- tion of Rho-GTP and Actomyosin. Current Biology 24:1263 -- 1270. 14. Oakes, P. W., E. Wagner, C. A. Brand, D. Probst, M. Linke, U. S. Schwarz, M. Glotzer, and M. L. Gardel, 2016. Opto- genetic control of RhoA reveals zyxin-mediated elasticity of stress fibres. Nature Communications 8:15817. 15. Gibson, M. C., A. B. Patel, R. Nagpal, and N. Perrimon, 2006. The emergence of geometric order in proliferating metazoan epithelia. Nature 442:1038 -- 1041. 16. Riveline, D., E. Zamir, N. Q. Balaban, U. S. Schwarz, T. Ishizaki, S. Narumiya, Z. Kamb, B. Geiger, and A. D. Ber- shadsky, 2001. Focal Contacts as Mechanosensors. J. Cell Biol. 153:1175 -- 1186. 17. Besser, A., and U. S. Schwarz, 2007. Coupling biochemistry and mechanics in cell adhesion: a model for inhomogeneous stress fiber contraction. New J. Phys. 9:425. 18. Bement, W. M., M. Leda, A. M. Moe, A. M. Kita, M. E. Lar- son, A. E. Golding, C. Pfeuti, K.-C. Su, A. L. Miller, A. B. Goryachev, and G. von Dassow, 2008. Activatorinhibitor coupling between Rho signalling and actin assembly makes the cell cortex an excitable medium. Nature Cell Biology 17:1471 -- 1483. 19. Adams, C. L., Y. Chen, S. J. Smith, and W. J. Nelson, 1998. Mechanisms of Epithelial Cell-Cell Adhesion and Cell Com- paction Revealed by High-resolution Tracking of E-Cadherin- Green Fluorescent Protein. J. Cell Biol. 142:1105 -- 1119. Biophysical Journal 00(00) 1 -- 8 8 20. Salomon, J., C. Gaston, J. Magescas, B. D. D. Canioni, L. Sengmanivong, A. Mayeux, G. Michaux, F. Campeotto, J. Lemale, J. Viala, F. Poirier, N. Minc, J. Schmitz, N. Brousse, B. Ladoux, O. Goulet, and D. Delacour, 2017. Contractile forces at tricellular contacts modulate epithelial organization and monolayer integrity. Nature Materials 8. 21. Lye, C. M., H. W. Naylor, and B. Sanson, 2014. Subcellular localisations of the CPTI collection of YFP-tagged proteins in Drosophila embryos. Development 141:4006 -- 4017. 22. Rauzi, M., P.-F. Lenne, and T. Lecuit, 2010. Planar polar- ized actomyosin contractile flows control epithelial junction remodelling. Nature 468:1110 -- 1114. 23. Klingner, C., A. V.Cherian, J. Fels, P. M. Diesinger, R. Auf- I. M. Toli´c- schnaiter, N. Maghelli, T. Keil, G. Beck, Nørrelykke, M. Bathe, and R. Wedlich-Soldner, 2014. RhoA and RhoC have distinct roles in migration and invasion by acting through different targets. J. Cell Biol. 207:107 -- 121. 24. Jin, Y., S. J. Atkinson, J. A. Marrs, and P. J. Gallagher, 2001. Myosin II Light Chain Phosphorylation Regulates Membrane Localization and Apoptotic Signaling of Tumor Necrosis Factor Receptor-1. J. Biol. Chem. 276:30342 -- 30349. 25. Reffay, M., M. C. Parrini, O. Cochet-Escartin, B. Ladoux, A. Buguin, S. Coscoy, F. Amblard, J. Camonis, and P. Sil- berzan, 2014. Interplay of RhoA and mechanical forces in collective cell migration driven by leader cells. Nature Cell Biology 16:217 -- 223. 26. Sorce, B., C. Escobedo, Y. Toyoda, M. P. Stewart, C. J. Cat- tin, R. Newton, I. Banerjee, A. Stettler, B. Roska, S. Eaton, A. A. Hyman, A. Hierlemann, and D. J. Muller, 2015. Mitotic cells contract actomyosin cortex and generate pres- sure to round against or escape epithelial confinement. Nature Communications 6:8872. 27. Imai, M., K. Furusawa, T. Mizutani, K. Kawabata, and H. Haga, 2015. Three-dimensional morphogenesis of MDCK cells induced by cellular contractile forces on a viscous sub- strate. Scientific Reports 5:14208. 28. Munjal, A., and T. Lecuit, 2014. Actomyosin networks and tissue morphogenesis. Development 141:1789 -- 1793. 29. Wollrab, V., R. Thiagarajan, A. Wald, K. Kruse, and D. Rive- line, 2016. Still and rotating myosin clusters determine cytoki- netic ring constriction. Nature Communications 7:11860. 30. Nixon-Abell, J., C. J. Obara, A. V. Weigel, D. Li, W. R. Legant, C. S. Xu, H. A. Pasolli, K. Harvey, H. F. Hess, E. Betzig, C. Blackstone, and J. Lippincott-Schwartz, 2016. Increased spatiotemporal resolution reveals highly dynamic dense tubular matrices in the peripheral ER. Science 354. 31. Toyota, T., D. A. Head, C. F. Schmidt, and D. Mizuno, 2011. Non-Gaussian athermal fluctuations in active gels. Soft Matter 7:3234 -- 3239. 32. Park, Y., C. A. Best, T. Auth, N. S. Gov, S. A. Safran, G. Popescu, S. Suresh, and M. S. Feld, 2010. Metabolic remodeling of the human red blood cell membrane. Proc. Natl. Acad. Sci. USA 107:1289 -- 1294. 33. Ben-Isaac, E., Y. Park, G. Popescu, F. L. H. Brown, N. S. Gov, and Y. Shokef, 2011. Effective Temperature of Red-Blood- Cell Membrane Fluctuations. Phys. Rev. Lett. 106:238103. 34. Trepat, X., M. R. Wasserman, T. E. Angelini, E. Millet, D. A. Weitz, J. P. Butler, and J. J. Fredberg, 2009. Physical forces during collective cell migration. Nature Physics 5:426 -- 430. Biophysical Journal 00(00) 1 -- 8 35. Bursac, P., G. Lenormand, B. Fabry, M. Oliver, D. Weitz, V. Viasnoff, J. Butler, and J. J. Fredberg, 2005. Cytoskele- tal remodelling and slow dynamics in the living cell. Nature Materials 4:557 -- 561. 36. Fodor, ´E., M. Guo, N. S. Gov, P. Visco, D. A. Weitz, and F. van Wijland, 2015. Activity-driven fluctuations in living cells. EPL (Europhysics Letters) 110:48005. 37. Farhadifar, R., J.-C. Roper, B. Aigouy, S. Eaton, and F. Julicher, 2007. The Influence of Cell Mechanics, Cell-Cell Interactions, and Proliferation on Epithelial Packing. Current Biology 17:2095 -- 2104. 38. Kafer, J., T. Hayashi, A. F. M. Mar´ee, R. W. Carthew, and F. Graner, 2007. Cell adhesion and cortex contractility deter- mine cell patterning in the Drosophila retina. Proc. Natl. Acad. Sci. USA 104:18549 -- 18554. 39. Bi, D., J. H. Lopez, J. M. Schwarz, and M. L. Manning, 2015. A density-independent rigidity transition in biological tissues. Nature Physics 11:1074 -- 1079. 40. Bi, D., X. Yang, M. C. Marchetti, and M. L. Manning, 2016. Motility-Driven Glass and Jamming Transitions in Biological Tissues. Phys. Rev. X 6:021011. 41. Bi, D., J. H. Lopez, J. M. Schwarz, and M. L. Manning, 2014. Energy barriers and cell migration in densely packed tissues. Soft Matter 10:1885 -- 1890. 42. Forgacs, G., R. A. Foty, Y. Shafrir, and M. S. Steinberg, 1998. Viscoelastic Properties of Living Embryonic Tissues: a Quantitative Study. Biophys. J. 74:2227 -- 2234. 43. Guevorkian, K., M.-J. Colbert, M. Durth, S. Dufour, and F. Brochard-Wyart, 2010. Aspiration of Biological Viscoelas- tic Drops. Phys. Rev. Lett. 104:218101. 44. Stirbat, T. V., A. Mgharbel, S. Bodennec, K. Ferri, H. C. Mer- tani, J.-P. Rieu, and H. Delanoe-Ayari, 2013. Fine Tuning of Tissues' Viscosity and Surface Tension through Contractility Suggests a New Role for α-Catenin. PLoS ONE 8:e52554. 45. Sekimoto, K., and S. Sasa, 1997. Complementarity Relation for Irreversible Process Derived from Stochastic Energetics. J. Phys. Soc. Jpn. 66:3326 -- 3328. 46. Sekimoto, K., 1997. Kinetic Characterization of Heat Bath and the Energetics of Thermal Ratchet Models. J. Phys. Soc. Jpn. 66:1234 -- 1237. 47. Fodor, ´E., K. Kanazawa, H. Hayakawa, P. Visco, and F. van Wijland, 2014. Energetics of active fluctuations in living cells. Phys. Rev. E 90:042724. 48. Harris, A. R., L. Peter, J. Bellis, B. Baum, A. J. Kabla, and G. T. Charras, 2012. Characterizing the mechanics of cultured cell monolayers. Proc. Natl. Acad. Sci. USA 109:16449 -- 16454. 49. Fodor, ´E., W. W. Ahmed, M. Almonacid, M. Bussonnier, N. S. Gov, M.-H. Verlhac, T. Betz, P. Visco, and F. van Wijland, 2016. Nonequilibrium dissipation in living oocytes. EPL (Europhysics Letters) 116:30008. 50. Harada, T., and S.-i. Sasa, 2005. Equality Connecting Energy Dissipation with a Violation of the Fluctuation-Response Rela- tion. Phys. Rev. Lett. 95:130602. 51. Toyabe, S., T. Okamoto, T. Watanabe-Nakayama, H. Take- tani, S. Kudo, and E. Muneyuki, 2010. Nonequilibrium Ener- getics of a Single F1-ATPase Molecule. Phys. Rev. Lett. 104:198103.
1703.02922
2
1703
2017-03-09T15:21:02
Dynamic transition from $\alpha$-helices to $\beta$-sheets in polypeptide superhelices
[ "physics.bio-ph", "cond-mat.mtrl-sci", "cond-mat.soft", "physics.chem-ph" ]
We carried out dynamic force manipulations $in$ $silico$ on a variety of superhelical protein fragments from myosin, chemotaxis receptor, vimentin, fibrin, and phenylalanine zippers that vary in size and topology of their $\alpha$-helical packing. When stretched along the superhelical axis, all superhelices show elastic, plastic, and inelastic elongation regimes, and undergo a dynamic transition from the $\alpha$-helices to the $\beta$-sheets, which marks the onset of plastic deformation. Using Abeyaratne-Knowles formulation of phase transitions, we developed a theory to model mechanical and kinetic properties of protein superhelices under mechanical non-equilibrium conditions and to map their energy landscapes. The theory was validated by comparing the simulated and theoretical force-strain spectra. Scaling laws for the elastic force and the force for $\alpha$-to-$\beta$ transition to plastic deformation can be used to rationally design new materials of required mechanical strength with desired balance between stiffness and plasticity.
physics.bio-ph
physics
Dynamic transition from α-helices to β-sheets in polypeptide superhelices Kirill A. Minin,1 Artem Zhmurov,1 Kenneth A. Marx,2 Prashant K. Purohit,3 and Valeri Barsegov2, 1 1Moscow Institute of Physics and Technology, Dolgoprudny, 141701, Russia 2Department of Chemistry, University of Massachusetts, Lowell, MA 01854, USA 3Department of Mechanical Engineering and Applied Mechanics, University of Pennsylvania, Philadelphia, PA 19104, USA We carried out dynamic force manipulations in silico on a variety of superhelical protein fragments from myosin, chemotaxis receptor, vimentin, fibrin, and phenylalanine zippers that vary in size and topology of their α-helical packing. When stretched along the superhelical axis, all superhelices show elastic, plastic, and inelastic elongation regimes, and undergo a dynamic transition from the α-helices to the β-sheets, which marks the onset of plastic deformation. Using Abeyaratne-Knowles formulation of phase transitions, we developed a theory to model mechanical and kinetic proper- ties of protein superhelices under mechanical non-equilibrium conditions and to map their energy landscapes. The theory was validated by comparing the simulated and theoretical force-strain spec- tra. Scaling laws for the elastic force and the force for α-to-β transition to plastic deformation can be used to rationally design new materials of required mechanical strength with desired balance between stiffness and plasticity. In 1953 coiled-coils were proposed independently by Crick and Pauling as structures comprised of supercoiled α-helical segments. Since then, coiled-coils have been recognized as ubiquitous, critically important, highly sta- ble biomechanical structures, occuring either at the tis- sue level (hair, blood clots, etc.), in individual cellular structures (intracellular cytoskeleton, flagella, etc.), or as components of individual proteins. Many well stud- ied proteins performing mechanical functions, utilize the superhelical coiled-coil architecture, including ones in the present study: muscle proteins (myosin), interme- diate filaments (vimentin), blood clots and thrombi (fib- rin), chemotaxis (chemotaxis receptors), cellular trans- port (kinesin), and bacterial adhesion (protein tetrapra- chion) [1]. Recently, the unique superhelical symmetry of coiled-coils has inspired the design of new materials [2]: short supercoils [3], long and thick fibers [4], nan- otubes [5], spherical cages [6] and synthetic virions [7]. In this study, we solve a longstandng problem, provid- ing a theoretical basis for understanding different coiled- coils' stability and dynamic properties when undergoing the pulling force induced α-helix to β-sheet transition. We combined dynamic force manipulations in silico with theoretical modeling and found that all systems studied, with from two-to-five helices forming parallel and anti- parallel supercoil architectures, uniformly undergo three force induced extension regimes including a remarkable plastic phase transition from all α-helices to all β-sheets. The quantitative agreement between the theory and sim- ulations allows for a new approach to rationally design coiled-coils with specific mechanical properties into novel biomaterials applications. We used the atomic models (see Supplemental Mate- rial, SM) of myosin, vimentin, fibrin, bacterial chemo- taxis receptor and phenylalanine zippers (PDB entries: 2FXO [8], 1GK4 [9], 3GHG [10], 1QU7 [11], 2GUV and 2GUS [12] respectively). (i) Myosin II contains a double- stranded parallel coiled-coil [13]. Upon muscle contrac- tion, tension is transfered along the myosin tail [13] (ex- perimental force data are available for myosin [14, 15]). (ii) Vimentin, in intermediate filaments in cells [16, 17], helps determine their resistance to mechanical factors [18]. Vimentin's structure contains an α-helical rod do- main, which can be divided into several double-helical parallel coiled-coil segments [18]. (iii) Fibrin forms the fibrous network of a blood clot that stops bleeding [19]. Triple-helical parallel coiled-coils create the unique visco- elastic properties of fibrin [20, 21]. (iv) Bacterial chemo- taxis receptor is responsible for signal transduction across cell membranes [11]. The cytoplasmic domain contains two double-stranded anti-parallel coiled-coils forming a four-stranded superhelix. (v) Phenylalanine zippers, arti- ficial four-to-five stranded parallel supercoils, are promis- ing biomaterials with tunable properties [5, 12]. Force spectroscopy in silico: We employed all-atom Molecular Dynamics (MD) simulations using the Sol- vent Accessible Surface Area (SASA) model with CHARMM19 unified hydrogen force-field [22] imple- mented on a GPU (see SM) [21]. Protein models were constructed using the CHARMM program [23]. We used a lower damping coefficient γ=0.15 ps−1 vs. γ=50 ps−1 for ambient water at 300K for more efficient sampling of conformational space [24]. To mimic experimental con- ditions, we implemented the pulling plane with harmoni- cally attached tagged residues at one end of the molecule and the resting plane with constrained residues at the other (Fig. 1a). The pulling plane was connected to a vir- tual cantilever moving with a velocity vf =104−106µm/s and ramping up the force f =rf t with a loading rate rf =ksvf =10−3−10−1N/s (ks=100 pN/nm is the can- tilever spring). The α-to-β transition: All superhelices undergo the force-driven transition from α-helices to β-sheets regard- less of the number of helices and parallel or anti-parallel architecture. In Fig. 1, we display the force-strain (f ε) curves (and structure snapshots) for bacterial receptor and myosin which are reminiscent of experimental force extension profiles [14, 15]. The plateau force for myosin arXiv:1703.02922v2 [physics.bio-ph] 9 Mar 2017 2 more, in the α-state H-bonds are all intramolecular while β-sheets form due to intermolecular H-bonds linking par- allel or antiparallel β-strands. We profiled the probabil- ities of finding a system in the α-state pα and β-state pβ using dihedral angles and H-bonds. We calculated the relative amount of intrachain bonds pα=Nintra/NH and interchain bonds pβ=Ninter/NH (NH -total number of H-bonds), and the relative amounts of residues in the α-region pα=Nα/Nφψ and β-sheet region pβ=Nβ/Nφψ (Nφψ-total number of φ, ψ-angles). The profiles of pα and pβ displayed in Fig. 2 for bacterial receptor and myosin show the following: (i) the α-helical (β-strand) content decreases (increases) with time t (and force f =rf t); (ii) the transformation from α-helices to β-sheets is a two- state transition; and (iii) this transition is accompanied by redistribution of φ, ψ-angles (α→β; Fig. 1b,d insets) and reconfiguration of H-bonds (from intra- to interchain H-bonds; snapshots in Fig. 2). We obtained similar re- sults for the vimentin, fibrin and Phe-zippers (Fig. S2). Hence, the H-bonds and φ, ψ-angles can be used as molec- ular signatures to characterize the α-to-β transition in proteins with superhelical symmetry. Two-state model: Under the constant-force conditions (force-clamp), the α-to-β transition can be described using a two-state model. The sigmoidal extension- force phase diagram for myosin and fibrin coiled-coils in Fig. S4 is divided into the α-phase and β-phase [21]. The α-state can be modeled as an entropic spring with energy G=kα∆X 2 α/2, where kα is the spring constant and ∆Xα=f /kα is extension. The β-state can be described by a wormlike chain [26] with en- ergy Gβ=(3kBT /2lβ)R [∂n(s)/∂s]2ds, where lβ is the A persistence length, and T is the temperature. pulling force stretches the α-helices by a fractional extension yα(f )=∆Xα(f )/Lα, where Lα is the maxi- mal extension in the α-state, and lowers the energy barrier ∆G(f ) by increasing the transition probabil- ity pβ(f )/pα(f )=exp [−∆G(f )/kBT ]. The force in- duces elongation in the β-states, yβ(f )=∆Xβ(f )/Lβ, where Lβ is the maximal extension in the β-states. Here, yβ(f )=1−{ξ(q)1/3+[4q/3 − 1]/ξ(q)1/3}−1, where ξ(q)=2+{4−[(4/3)q−1]3}1/2 and q=f lβ/kBT . The total extension is ∆X(f )=pαyαLα+pβyβLβ [21]. Continuum theory: The α-to-β transition in the non- equilibrium regime of time-dependent force can be de- scribed using the Abeyaratne-Knowles formulation of phase transitions [27, 28]. The displacement of a ma- terial point at reference position x at time t is given by u(x, t)=X(x, t)−x. The end at x=0 is fixed, and u(0, t)=0 for all t. At the other end x=L, u(L, t)=∆X(t), where ∆X(t)=vf t>0 (<0) when the sample is loaded (unloaded). The α-phase is described by a stretch Γα(f )=Xα(f )/L, f <f∗αβ, where Xα and f∗αβ are the end- to-end distance and critical force for the α-phase. We set L=Lα throughout. At f =fαβ, the β-phase nucleates and f∗αβ>fαβ>f0, where f0 is the Maxwell force at which the free energies per residue for the two phases become equal. The β-phase is described by a stretch Γβ(f )=Xβ(f )/L, FIG. 1. The α-to-β transition. a: pulling setup with pulling and resting planes. Transition is shown for the chemotaxis receptor (b, c) and myosin (d, e). The f ε-curves (b, d) show- ing elastic, plastic, and inelastic regimes of superhelices' un- folding, overlaid with the continuum theoretical curves. In- sets in b and d are Ramachandran plots of dihedral (φ,ψ) angles in the two states. Snapshots (c,e) from similarly num- bered regions in f ε-curves show unfolding progress from ini- tial α-structure (0), to mixed α/β-structures (1−3), to final β- structure (4). Snapshots 3−4 show exotic spirals of β-sheets. f∗∼100 pN (Fig. 1d) is higher than the experimental value of ∼40 pN [15] due to 104−105-fold faster pulling speeds used in silico. Results obtained for vimentin, fib- rin and phenylalanine zippers are in Fig. S1. The f ε- curves reflect the three regimes of dynamic mechanical response of the superhelices to an applied force: i) elas- tic regime I of coiled-coil elongation at low strain (f∼ε; ε<0.2), characterized by a linear growth of f with ε; ii) plastic transition regime II of protein unfolding at in- termediate strain (f =const; 0.2<ε<0.9), where the α-to- β transition occurs; and iii) inelastic regime III of non- linear elongation of the β-structure at high strain (f∼ε2; ε>0.9). In the α-to-β transition regime, the coiled-coils unwound and underwent a large 80−90% elongation. The α-to-β transition nucleated at both ends of the molecule, and two phase boundaries propagated towards the cen- ter (snapshots 1 -- 2, Fig. 1c,e). All of these features were observed irrespective of the number and architecture of α-helices. Dihedral angles (φ,ψ) are sensitive to changes in pro- tein secondary structure [25]. This is reflected in migra- tion of φ, ψ-angles from the α-region to the β-region in the Ramachandran plots (insets to Fig. 1b,d). Further- 3 Application to α-to-β transition: Calculating the f ε-curves requires a kinetic relation and a nucleation criterion. By treating the α-phase as an entropic spring and the β-phase as a wormlike chain, we ob- tain Γα(f ) = f /κα+1 and Γ0α(f )=1/κα for the α-phase, is obtained by inverting the wormlike chain formula for the force f vs. extension yβ). The α-to-β transformation strain becomes: where κα=kα/Lα, and Γβ(f )=(Lβ/Lα)(1−pkBT /4lβf ) and Γ0β(f )=(Lβ/4Lα)pkBT /lβf 3 for the β-phase (Γβ(f ) γ(f ) = (Lβ/Lα)(cid:18)1 −qkBT /4lβf(cid:19) − f /κα − 1, and γ0(f )=(Lβ/4Lα)pkBT /lβf 3−1/κα. To obtain the kinetic relation (Eq. (2)), we use the Arrhenius rates for forward and backward transitions kαβ and kβα [29], (5) pα = Φ(f ) = A(cid:20)e− ∆Gαβ (f ) kB T − e− ∆Gβα (f ) kB T (cid:21) (6) where A is an attempt frequency, and ∆Gαβ(f ) = 0−f zαβ, ∆Gβα(f ) = 0−αβ +f zβα (7) are the energy barriers. In Eq. (7), 0 and αβ are the force-free energy barrier and energy difference (per residue length) between the α- and β-states, respectively, and zαβ and zβα are transition distances (see the inset to Fig. 3a). The detailed balance is given by: = e− ∆Gαβ (f )−∆Gβα (f ) kB T = e− αβ−f (zαβ +zβα ) kB T (8) pβ pα Setting pα=pβ in Eq. (8), we obtain the Maxwell force: f0 = αβ/(zαβ + zβα) (9) Initially, the continuum is in the α-phase. Before the nucleation of the phase boundary, the force in the linear elastic regime I (Fig. 1b,d) is given by f (t) = καvf t/Lα = κα∆Xα(t)/Lα (10) The nucleation criterion requires that when f =fαβ, ∆X=∆Xαβ=fαβL/κα, a phase boundary nucleates at x=0 and x=L. The force is governed by Eq. (3) with the initial condition f =fαβ. The phase boundaries move through the continuum and convert all the material into the β-phase. This marks the onset of the plastic transi- tion regime II, which corresponds to the force plateau in f ε-curves (Fig. 1b,d; see SM): f∗ =(cid:18) zαβ kBT − Lα κα(Lβ − Lα)(cid:19)−1(cid:20)ln vf A (Lβ − Lα) + 0 kBT(cid:21) (11) Using the structures generated for the mixed α+β-phase, we run long 10 µs simulations for myosin and fibrin with pulling force gradually quenched to zero (with same rf ). The reverse β→α transition was not observed. Hence, FIG. 2. Dynamics of pα and pβ for chemotaxis receptor (a) and myosin tail (b), described using φ, ψ-angles and H-bonds. Compared are curves from MD simulations with green curves from continuum theory. We used intervals −80◦<φ<−48◦ and −59◦<ψ<−27◦ for the α-phase, and −150◦<φ<−90◦ and 90◦<ψ<150◦ for the β-phase. [25]. For H-bonds, we used a 3A cut-off distance between hydrogen donor (D) and acceptor (A) atoms and a 20◦ cut-off for the D-H. . . A bond angle. f∗βα<f <∞, where Xβ(f ) and f∗βα are the end-to-end dis- tance and critical force in the β-phase. A transformation strain is γ(f )=Γβ(f )−Γα(f ), f∗αβ≥f≥f∗βα. The equation of motion for the 1D continuum (assum- ing negligible drag and inertia forces) is ∂f /∂x=0, so that f (x, t) is constant for 0<x<L. If f∗αβ<f <f∗βα, a mixture of the α- and β-phases is possible, and the total extension is given by X(L, t) = L+∆X(t) = L [Γα(f (t))pα(t) + Γβ(f (t))pβ(t)] (1) A kinetic relation expressed in terms of the force f (t) describes the evolution of pα, i.e. pα = Φ(f (t)), (2) where Φ(f ) is a material property. By differentiating Eq. (1) and eliminating pα using Eq. (2), we obtain: [g(f ) − γ0(f )(1 + ∆X/L)] f + γ(f )vf /L = γ2(f )Φ(f ) (3) where g(f )=Γα(f )Γ0β(f )−Γ0α(f )Γβ(f ). From Eq. (3) a force plateau ( f =0) forms at f =f∗ if either γ(f∗) = Γβ(f∗) − Γα(f∗) = 0, or vf /L = γ(f∗)Φ(f∗). (4) Because the height of force plateau depends on rf and, hence, is strain-rate (vf /L) dependent, Eq. (4) defines the force plateau height. 4 be readily estimated. Since zαβ and zβα barely change with f (Fig. 3a), αβ is obtained using the Maxwell force (Eq. (9)), and 0 is obtained using either of Eqs. (7). A is estimated by fitting the profiles of pα and pβ=1−pα shown in Fig. 2 with Eq. (6). All paramter values are presented in Table I. We used the values of Lα, κα, Lβ, zαβ, 0 and A from Table I as the input and Eqs. (10)-(12) to fit the sim- ulated f ε-curves for all the protein superhelices at all pulling speeds (in Eqs. (10) and (12), ε(t)=∆X(t)/L is the time-dependent strain). This enabled us to refine the parameter values and estimate lβ (shown in squared brackets in Table I). The results of fitting for chemo- taxis receptor and myosin presented in Fig. 1b,d and in Fig. S1 for the other superhelices show excellent agree- ment between simulated and theoretical f ε-curves, which validates our theory. The refined final values of model pa- rameters from the fit do not differ much from the initial input, which points to the internal consistency between the results of theory and simulations. We used these pa- rameters and Eq. (6) to reconstruct the profiles pα and pβ. The Fig. 2 results show that, although H-bond re- distribution and dihedral angles' migration capture the α-to-β transition, the H-bond based estimation of pα and pβ results in a better agreement with the simulations. Scaling laws: The theory predicts that the stiffness of superhelices κα/Lα is inversely proportional to their length Lα (Eqs. (10)), so shorter (longer) superhelices are stiffer (less stiff). It also predicts a logarithmic scal- ing of the critical force f∗ with the pulling speed vf (loading rate rf =ksvf ; Eq. (11)). We tested this pre- diction by performing a fit of the f∗ vs. ln vf profiles for superhelices from chemotaxis receptor and myosin with Eq. (11) borrowing parameter values from Table I. The Fig. 3b results show excellent agreement between the predicted and simulated values of f∗ extracted from the f ε-spectra (Figs. 1 and S1), which further validates our theory. Also, f∗∼Nh -- the number of α-helices form- ing a superhelix (Nh=2−5; Table I). The profiles of f∗ vs. Nh in Fig. 3c show a roughly additive contribution to the mechanical strength from α-helices, which weakly cooperate to sustain the stress. For vf =105µm/s, f∗ in- creases from 103 pN for double-helical myosin tail and vimentin (∼ 52 pN per helix) to 187 pN for the triple- helical fibrin coiled-coil (∼62 pN per helix), to 268 pN for four-stranded Phe-zipper (∼67 pN per helix), and to 383 pN for Phe-zipper with five α-helices (∼77 pN per helix). A lower f∗=232 pN for the four-stranded chemo- taxis receptor (∼58 pN per helix) is due to the antipar- allel arrangement implying that parallel arrangement of α-helices provides higher mechanical strength. The developed theory can be used to accurately describe dynamic transitions in wild-type and syn- thetic superhelical polypeptides under mechanical non- equilibrium conditions and model their force-strain spec- tra from dynamic force experiments. The theory can be used to probe mechanical and kinetic characteristics of any coiled-coil superhelical polypeptide and to map out FIG. 3. a: Free energy G(x) for myosin superhelix at different forces. Inset shows a histogram of projected lengths P (x) of myosin residues for f =90 pN and energy landscape G(x) with model parameters (Eqs. (7)-(9)). b: Plot of f∗ vs. ln vf (data points) and theoretical curves of f∗ predicted using Eq. (11); slopes and y-intercepts were calculated using parameters from Table I. c: Plot of f∗ vs. Nh-number of α-helices and a linear fit with y=aNh (a=88 and 103 for vf =104 and 105µm/s). superhelices extended and β-sheets formed permanently, i.e. the α→β-transition marks the onset of plasticity in protein superhelices. Stretching of the β-phase continues in the inelastic regime III (Fig. 1b,d) and the force is given by the wormlike chain formula: f (t) = (kBT /4lβ) [1 − (Lα/Lβ)(1 + vf t/L)]−2 = (kBT /4lβ) [1 − (Lα/Lβ)(1 + ∆Xβ(t)/L)]−2 (12) Force spectra and energy landscape: Parameters ac- cessed directly is MD simulations are the end-to-end dis- tances in the α-phase Lα; the slope of (elastic) portion of an f ε-curve κα; the maximum extension in the β-phase Lβ; and f0≈f∗ (lβ is obtained using a wormlike chain fit to the inelastic part of a f ε-curve; see Fig. 1). The other parameters, zαβ, zβα, αβ, and 0, are estimated by mapping the free-energy landscape. We employed a mean-field approach, G(x)=−kBT ln P (x), to profile G as a function of projection of the average residue length (unit length) along the superhelical axis x (reaction co- ordinate). An example of P (x) for myosin is in the inset to Fig. 3a (see Fig. S3a for P (x) and G(x) for other su- perhelices). For G(x) sampled at f≈0, zαβ and zβα can TABLE I. Mechanical, thermodynamic and kinetic parameters for protein superhelices from myosin, chemotaxis receptor, vimentin, fibrin, and phenylalanine zippers differing in the number of α-helices Nh: Lα, κα and zαβ (α-phase), Lβ, lβ and zβα (β-phase), f0, αβ, 0 and A. Standard deviations (not shown) are ≤3% of the average parameter values obtained directly from MD simulations or by using the energy landscapes (in parentheses), or by performing a fit of Eqs. (10) -- (12) to the simulated f ε-curves in Figs. 1 and S1 (in squared brackets). Superhelical protein Nh Lα, nm κα, pN lβ, Lβ, nm nm zαβ, nm zβα, nm f0, pN αβ, kBT 0, kBT 5 A, µs−1 [0.5] [0.2] [1.0] [0.4] [0.2] [0.1] fragments Myosin 2 18.1[18.1] 2169[2170] [0.38] 46[43] (0.13)[0.13] (0.07)[0.07] 103(90)[90] (4.4)[4.4] (3.3)[3.3] (8.3)[8.3] (5.3)[5.3] Chemotaxis receptor 4 13.6[13.6] 2367[2370] [0.33] 34[32] (0.11)[0.11] (0.08)[0.08] 2 11.5[11.5] 1375[1375] [0.36] 28[27] (0.14)[0.14] (0.06)[0.06] (4.3)[4.3] (3.6)[3.6] 3 16.2[16.2] 1135[1135] [0.35] 41[38] (0.11)[0.11] (0.09)[0.09] 187(160)[160] (7.8)[7.8] (4.6)[4.6] (9.3)[9.3] (5.8)[5.8] (17)[17] (9.4)[9.4] 6.2[6.2] 2644[2645] [0.36] 15[14] (0.11)[0.11] (0.08)[0.08] 7.8[7.8] 1793[1795] [0.35] 20[19] (0.11)[0.11] (0.09)[0.09] 4-strand. Phe-zipper 4 5 Phe-zipper Vimentin Fibrin 232[176] 103[90] 268[200] 383[350] their free-energy landscapes. The slope and y-intercept of the line of critical force f∗ vs. ln vf (Fig. 3b) can be used to estimate the critical distance zαβ and the force- free barrier height 0 (SM). Scaling laws for the elastic force f vs.length L, and f∗ vs. loading rate rf , and the dependence of f∗ on L and number of helices Nh provide a powerful new method to rationally design novel syn- thetic biomaterials with the required mechanical strength and balance between stiffness and plasticity. Acknowledgments: This work was supported by NSF (grant DMR1505662 to VB and PKP), American Heart Association (grant-in-aid 13GRNT16960013 to VB), and Russian Foundation for Basic Research (grants 15-37- 21027, 15-01-06721 to AZ). [1] A. Lupas and J. Bassler, Trends Bioch. Sci. 42, 130 [15] D. Root, V. Yadavalli, J. Forbes, and K. Wang, Biophys. (2016). [2] R. Quinlan, E. Bromley, and E. Pohl, Curr. Opin. Cell Biol. 32, 131 (2015). [3] K. M. Arndt, J. N. Pelletier, K. M. Muller, A. Pluckthun, and T. Alber, Structure 10, 1235 (2002). [4] S. Potekhin, T. Melnik, V. Popov, N. Lanina, A. Vaz- ina, P. Rigler, A. Verdini, G. Corradin, and A. Kajava, Chem. Biol. 8, 1025 (2001). [5] N. C. Burgess, T. H. Sharp, F. Thomas, C. W. Wood, A. R. Thomson, N. R. Zaccai, R. L. Brady, L. C. Ser- pell, and D. N. Woolfson, J. Am. Chem. Soc. 137, 10554 (2015). [6] J. M. Fletcher, R. L. Harniman, F. R. H. Barnes, A. L. Boyle, A. Collins, J. Mantell, T. H. Sharp, M. Antog- nozzi, P. J. Booth, N. Linden, M. J. Miles, R. B. Ses- sions, P. Verkade, and D. N. Woolfson, Science 340, 595 (2013). [7] J. Noble, E. De Santis, J. Ravi, B. Lamarre, V. Castel- letto, J. Mantell, S. Ray, and M. Ryadnov, J. Am. Chem. Soc 138, 12202 (2016). [8] W. Blankenfeldt, N. Thoma, J. Wray, M. Gautel, and I. Schlichting, Proc. Natl. Acad. Sci. USA 103, 17713 (2006). [9] S. Strelkov, H. Herrmann, N. Geisler, T. Wedig, R. Zim- belmann, U. Aebi, and P. Burkhard, EMBO J. 21, 1255 (2002). [10] J. M. Kollman, L. Pandi, M. R. Sawaya, M. Riley, and J. 90, 2852 (2006). [16] J. Eriksson, T. Dechat, B. Grin, B. Helfand, M. Mendez, H.-M. Pallari, and R. D. Goldman, J. Clin. Invest. 119, 1763 (2009). [17] D. A. Fletcher and R. D. Mullins, Nature 463, 485 (2010). [18] H. Herrmann, H. Bar, L. Kreplak, S. V. Strelkov, and U. Aebi, Nat. Rev. Mol. Cell Bio. 8, 562 (2007). [19] J. W. Weisel, in Fibrous Proteins: Coiled-Coils, Collagen and Elastomers, Advances in Protein Chemistry, Vol. 70, edited by D. A. D. Parry and J. M. Squire (Academic Press, 2005) pp. 247 -- 299. [20] A. Zhmurov, A. E. X. Brown, R. I. Litvinov, R. I. Dima, and V. Barsegov, Structure 19, 1615 J. W. Weisel, (2011). [21] A. Zhmurov, O. Kononova, R. Litvinov, R. Dima, V. Barsegov, and J. Weisel, J. Am. Chem. Soc. 134, 20396 (2012). [22] P. Ferrara, J. Apostolakis, and A. Caflisch, Proteins 46, 24 (2004). [23] B. Brooks, C. Brooks, A. MacKerell, L. Nilsson, R. Pe- trella, B. Roux, Y. Won, G. Archontis, C. Bartels, S. Boresch, et al., J. Comput. Chem. 30, 1545 (2009). [24] S. Falkovich, I. Neelov, and A. Darinskii, Polym. Sci. Ser. A 52, 662 (2010). [25] R. Srinivasan and G. Rose, Proteins 22, 81 (1995). [26] V. Barsegov, D. Klimov, and D. Thirumalai, Biophys. R. F. Doolittle, Biochemistry 48, 3877 (2009). J. 90, 3827 (2006). [11] K. Kim, H. Yokota, and S.-H. Kim, Nature 400, 787 [27] R. Abeyaratne and J. Knowles, Evolution of phase tran- (1999). sitions (Cambridge University Press, New York, 2006). [12] J. Liu, Q. Zheng, Y. Deng, N. Kallenbach, and M. Lu, [28] R. Raj and P. Purohit, EPL-Europhys. Lett. 91, 28003 J. Mol. Biol. 361, 168 (2006). (2010). [13] H. Warrick and J. Spudich, Ann. Rev. Cell Biol. 3, 379 [29] R. Raj and P. Purohit, J. Mech. Phys. Solids 59, 2044 (1987). (2011). [14] I. Schwaiger, C. Sattler, D. Hostetter, and M. Rief, Nat. Mater. 1, 232 (2002). Dynamic transition from α-helices to β-sheets in polypeptide superhelices Kirill A. Minin,1 Artem Zhmurov,1 Kenneth A. Marx,2 Prashant K. Purohit,3 and Valeri Barsegov2, 1 1Moscow Institute of Physics and Technology, Dolgoprudny, 141701, Russia 2Department of Chemistry, University of Massachusetts, Lowell, MA 01854, USA 3Department of Mechanical Engineering and Applied Mechanics, University of Pennsylvania, Philadelphia, PA 19104, USA Supplemental Material S1. MODEL SYSTEMS In this study, we have selected several coiled-coils frag- ments from various proteins (Table S1). The idea was to cover all known variants of the coiled coil geometries, including both parallel and anti-parallel arrangements of the α-helices in supercoils and differing numbers of α- helices Nh=2−5 (see Fig. S5 and Table I in the Main Text). We deliberately selected proteins whose function has a mechanical nature. Also, we have added a cou- ple of synthetic coiled-coils since these are very promis- ing systems in terms of material design applications. All the superhelical protein fragments used in this work are summarized in Table S1. Structure snapshots for all su- perhelices are presented in Figs. S1 and S3. Myosin is a muscle protein containing long double- stranded coiled-coil in its tail fragment [1]. Upon mus- cle contraction, the head of the myosin moves along the actin filament and the resulting mechanical tension is transferred along the tail [1]. In this study, we used the structure of the part of the human myosin II tail (PDB structure 2FXO [2]). This structure contains two chains with residues Gly835-Lys963 and Ser831-Leu961, respec- tively, that form a double-stranded homodimeric parallel coiled-coil. There is also a good amount of experimen- tal data available for myosin coil, including the force- extension curves from single-molecule AFM experiments [3, 4]. Myosin coiled-coils can be classified as leucine zip- pers, with around 50% of amino-acids on the positions a and d being leucine residues (Fig. S5 and Section S5). The second most popular amino acid at these positions is methionine (around 25%). Vimentin is an elementary building block of interme- diate filaments from connective tissue. It is also found in other types of mesodermal tissue [5]. Intermediate fila- ments, along with microtubules and actin, are involved in the construction of the cytoskeleton [6]. Vimentin plays a significant role in fastening organelles and maintain- ing their positions in the cytoplasm. While most of the intermediate filaments retain their structure, vimentin filaments are dynamic, which is important when the cell changes its shape. Vimentin provides strength to the cells, thereby helping to determine their resistance to me- chanical stress and to maintain their integrity [7]. The mechanical function of the vimentin makes it a natural system for our study. Similarly to many other interme- diate filaments, the structure of vimentin can be divided into three domains -- non-α-helical head and tail do- mains and an α-helical rod domain. While head and tail domains are responsible for the lateral aggregation, the mechanical footprint of vimentin is solely determined by the α-helical coiled-coil of the rod domain [7]. This do- main can be divided into several coiled-coils segments. The atomistic structure of some of these segments was determined by X-Ray crystallography [8]. In this study, we used the structure of the 2B segment from human vimentin, namely residues Cys328 to Gly406 in each of two α-helical strands (PDB structure 1GK4 [8]). To- gether, these helices form a parallel two-stranded coiled- coil, mostly stabilized by leucine residues on a and d positions (see Fig. S5 and Section S5). Fibrinogen is an abundant blood protein. Upon con- version to fibrin it forms the fibrous network called fibrin gel [9, 10]. This network acts as a scaffold for blood clot formation, capturing platelets and other blood cells to form a plug that stops bleeding. It is hypothesized that both globular parts and coiled-coils are responsible for the unique mechanical properties of fibrin [11 -- 14]. In this study, we used the structure of the entire human fibrino- gen molecule (PDB structure 3GHG [15]) and isolated the coiled-coils region, which contain residues αCys49- Cys161, βCys80-Cys193 and γCys23-Cys135. The fib- rinogen coiled-coil is a parallel heterotrimeric left-handed supercoil (Fig. S5 and Section S5). An interesting feature of this particular structure is the presence of the plasmin binding site in the center -- a slight distortion in the α-helical structure which also forms a flexible hinge [16]. The bacterial chemotaxis receptors are trans- membrane receptors with a simple signaling pathway re- sponsible for signal recognition and transduction across membranes [17]. In contrast to many mammalian recep- tors, whose signaling mechanism involves forming pro- tein oligomers upon ligand binding, the chemotaxis re- ceptors are dimeric even in the absence of their ligands, and their signaling does not depend on an equilibrium between monomeric and dimeric states. Bacterial chemo- taxis receptors are composed of a ligand-binding domain, a transmembrane domain consisting of two helices TM1 and TM2, and a cytoplasmic domain. All known bacte- rial chemotaxis receptors have a highly conserved cyto- plasmic domain, which unites signals from different lig- arXiv:1703.02922v2 [physics.bio-ph] 9 Mar 2017 and domains into a single signaling pathway [18]. The cytoplasmic domain contains two identical chains that make a "U-turn" at the end that is furthest from the membrane. These two chains form a two double-stranded anti-parallel coiled coil that are wrapped around one an- other to form a four-stranded supercoil (Fig. S5 and Sec- tion S5). In this work we use the part of this structure that is bordered by the residues Arg294-Ala489 for the first and Ala300-Glu479 for the second chain [18]. Phenylalanine zippers are artificial proteins that re- semble five- or four stranded parallel left-handed coiled coils (PDB structures 2GUV and 2GUS [19]). As the name suggests, the hydrophobic a and d positions are occupied by phenylalanine residues (Fig. S5 and Sec- tion S5). In the coiled-coils these residues form a hy- drophobic core, interlocking in a "knobs-into-holes" pat- tern. Interestingly, changing of only one Phe27 to a Met27 residue favors the formation of tetramer (PDB structure 2GUS) instead of the pentamer (PDB struc- ture 2GUV). Although these coiled-coils are synthetic, the zippers they form are very promising building blocks for design of new materials, including those with tun- able mechanical properties [20 -- 22]. In this work, we used the entire homopentameric (homotetrameric) struc- tures, which involve residues Ser1-Arg56 in each of the five (four) chains. S2. MD SIMULATIONS Pulling simulations: In our computer-based mod- eling, we utilized our in-house code for the all-atom MD simulations in implicit water [12] fully implemented on a Graphics Processing Units (GPUs). We used the SASA model of implicit solvation in conjunction with the CHARMM19 unified hydrogen force field [23]. To speedup simulations, all the computational subroutines were fully implemented on a GPU. This enabled us to attain long 1−10µs-timescales in reasonable wall- clock time. All the model systems were built using the CHARMM program [24]. The systems were energy- minimized using the steepest descent algorithm and grad- ually heated up to T =300K in a span of 300 ps. We used lower damping with a smaller damping coefficient γ=0.15 ps−1 compared to γ=50 ps−1 for ambient water at T =300 K. In effect, lower damping (viscosity) allows for faster and more effivient sampling of the conforma- tional phase space [25 -- 27]. A similar approach was used, e.g., in Ref. [25]. Protein superhelices are expected to unwind in the course of mechanical unfolding [12]. The process of un- winding is accompanied by rotational motion of certain atoms in the plane perpendicular to the pulling axis (superhelical symmetry axis). In the standard Steered Molecular Dynamics (SMD) approach, there is a desig- nated set of fixed atoms and a set of atoms that are pulled. The pulling force is determined by the Hooks Law with a virtual spring attached to the pulled atoms. 2 The other end of the spring moves with a constant veloc- ity mimicking the cantilever motion in the AFM experi- ment. There are problems with this approach when it is applied to the coiled-coil proteins. First, the fixed atoms cannot move relative to one another, which restricts their rotational motion [4]. Second, the force is applied to all α-helical chains independently, which might cause some chains' sliding past one another. Force protocol: To overcome this problem and to en- sure an equal distribution of the force load, while also al- lowing the rotational motion of atoms around the pulling axis, we have implemented the following setup. We intro- duced two planes that are perpendicular to the pulling direction. One of the planes was fixed, representing a constrained end of a protein system. The other plane was pulled using the external force (see Fig. 1a in the Main Text). The selected constrained and tagged atoms for each superhelical system were attached to the cor- responding plane via harmonic springs. The forces on the springs were acting only in the longitudinal direction perpendicular to the planes, thus allowing atoms to move freely in the transverse directions. The pulled plane was connected to a virtual cantilever bead that moved with the constant velocity vf , ramping up the force f =rf t with the force-loading rate rf =ksvf . We solved numerically the overdamped Langevin equa- tion of motion for the pulled plane with the timestep τpl=100×τ =100 f s, where τ =1 f s is the timestep used in the numerical integration of the Langevin equations of motion for the atoms. The cantilever spring constant ks was set to 100 pN/nm in all the simulations described in this work. For each model system, we performed 5 simu- lation runs for each of the pulling speeds vf =104, 105 and 106µm/s (force-ramp simulations). We also performed the constant-force pulling experiments (force clamp sim- ulations) using the following force values for the myosin tail: f =20, 40, 60, 70, 80, 90, 100, 120, 130, 140, 160, 180 and 200 pN, and for fibrin: f =40, 60, 80, 100, 120, 140, 160, 180, 200, 220, 240, 260 and 300 pN. S3. DERIVATION OF EQUATION (11) The dynamic force-ramp experiments in silico pro- duced force plateaus that are mostly flat, albeit a bac- terial receptor produces a slightly downward sloping plateau. When we plug in parameters κα, Lβ/Lα, etc., from Table 1 in the Main Text into the transformation strain,γ(f∗)=Γβ(f∗)−Γα(f∗)=0, we obtain much lower values of the plateau forces than those obtained in the all-atom MD simulations. Hence, the plateau forces f∗ conform to Eq. (4) in the Main Text, and also imply that f∗ is a solution of the following algebraic equation: 0 − αβ + f∗zβα 0 − f∗zαβ A(cid:20)exp(− Lα 1 −s kBT ×" Lβ kBT ) − exp(− f∗ 4lβf∗! − κα − 1# = )(cid:21) (1) kBT vf L . The meaning of all model parameters entering Eq. (1) is described in the Main Text (see also Table I in the Main Text). For large values of f∗, the second exponential in the first square brackets in Eq. (1) becomes negligible compared to the first exponential, and so the second ex- 4lβ f∗ tends to zero as f∗ increases, and so we can neglect this term in compar- ison with the term f∗/κα in the second square bracket. Hence, we obtain from Eq. (1) the following approximate equation for f∗: ponential can be neglected. Also,q kB T exp [− 0 − f∗zαβ kBT ](cid:20) Lβ Lα − f∗ κα − 1(cid:21) = vf AL (2) By taking the logarithm on both sides of Eq. (2) above we obtain the following equation: 0 − f∗zαβ kBT − + ln(cid:18) Lβ Lα − f∗ κα − 1(cid:19) = ln vf AL . (3) f∗Lα Assuming that κα(Lβ−Lα) < 1, based on the results of MD simulations, we can expand the logarithmic term on the left-hand-side of the above Eq. (3) in Taylor series. By retaining the linear terms in the Taylor expansion, we obtain: f∗(cid:18) zαβ kBT − Lα κα(Lβ − Lα)(cid:19) = ln vf A (Lβ − Lα) + 0 kBT , (4) Solving Eq. (4) for the plateau force f∗, we obtain the final expression for the calculation of f∗, which reads: f∗ =(cid:18) zαβ kBT − Lα κα(Lβ − Lα)(cid:19)−1(cid:20)ln vf A (Lβ − Lα) (5) Eq. (5), which is Eq. (11) of the Main Text, shows that the dependence of f∗ on ln vf can be recast as f∗=a ln vf +b, with the slope + 0 kBT(cid:21) and the y-intercept Lα kBT − a =(cid:18) zαβ b = a(cid:20) 0 κα(Lβ − Lα)(cid:19)−1 kBT − ln (A (Lβ − Lα))(cid:21) (6) (7) 3 the Boltzmann transformation, G(x)=−kBT ln P (x). For constant force (force-clamp) MD simulations, we use the last 100 ns of trajectories to build the distributions. Us- ing these simulations, we devided each trajectory into three parts, i.e. when the system is in the α-state, in the mixed α+β-state, and in the β-state. Estimating model parameters: There are several parameters that can be estimated either directly using the MD simulation output or using the energy landscape reconstruction described above. The length of the sys- tem in the α-state Lα is an equilibrium distance between the two ends of the system. The spring constant κα is the tangent of the initial (elastic) part of the force-strain curve (in regime I) per residue length (unit length; see Fig. 1 and Fig. S1). The maximum length in the β-state Lβ and the persistence length lβ are computed from the worm-like chain fit to the non-linear part of the force- strain curve (regime III). To estimate the distance be- tween the α-state minimum and the peak of the energy barrier, zαβ, and the distance between the peak of the barrier and the β-state minimum on the energy land- scape, zβα, we first note that the positions of the minima and the barrier do not shift much when the force increases (see Fig. 3 and Fig. S3). Hence, we can use the landscape constructed for f > 0 by selecting the landscape for the system, e.g. in the mixed α+β-state. The positions of the minima and the energy barrier are clearly visible on these landscapes, which allows for the estimation of the values for zαβ and zβα (Fig. 3 and Fig. S3). The energy landscape parameters can be estimated us- ing the following approach. The Maxwell force, f0, is the force at which the depths of the α-state and β-state minima are equal. Out of 13 constant force values used for myosin, f =90 pN gives the closest values of energies at the minima, which are different by less than 0.1kBT (see Fig. S4 and the inset to Fig. 3). Hence, we take f0≈90 pN for myosin. Similarly estimated, the value of f0 for fibrin coiled-coils comes to f0≈160 pN (Figs. S3 and S4). The values for αβ are computed from f0 us- ing Eq. (9) in the Main Text. To obtain the values of 0, we also use the energy landscape profiles obtained for the forces at which both the α-state and β-state are populated. Next, We measure the height of the barrier and then use the first Eq. (7) in the Main Text, which reads: ∆Gαβ(f )=0−f zαβ to compute 0. Finally, the attempt frequency A is estimated as a fitting parameter (see Eq. (6) in the Main text). S4. DATA ANALYSES AND MODELING Constructing free-energy landscapes: To build the energy landscape, we use the projection of the length of a single amino-acid residue as a reaction coordinate. For each model system, we collected the distribution P (x) of the projections of residue lengths along the pulling axis x (superhelical symmetry axis) and per- formed a plot the histogram using all structural snap- shots saved. The landscape is then obtained by using S5. ARCHITECTURE OF THE COILED COILS A typical turn per amino acid residue in the α-helix is around 100◦. Seven residues correspond to a 700◦ rota- tion which is 20◦ less than the 720◦ rotation for two com- plete turns. Hence, one can construct a seven-residue re- peat that will lag 20◦ behind for each two turns. Since the α-helices are right-handed, the aforementioned under- turn would give a left-handed stretch of residues of the surface of a straight α-helix. In a coiled coil, when each α-helix is twisted, this stretch would form a straight line along which two (of more) α-helices interact with each other, forming a left-handed supercoil [28]. The seven-residue (or heptad) repeat that is usually denoted as a-b-c-d-e-f-g (Fig. S5), is a core feature of a so-called PV ("Peptide Velcro") hypothesis [28, 29]. This hypothesis separates residues in the heptad by their po- sitions using the following assumptions. The residues a and d in the repeat are the residues that form hydropho- bic core of the coiled coil. For example, in the double- stranded coiled coil the residues a and d from one of the α-helices form hydrophobic contacts with residues a0 and d0 from the heptad repeat a0-b0-c0-d0-e0-f0-g0 of the other α-helix. The residues e and g are usually polar with long side chains, and they can further stabilize the coiled coils via the electrostatic interactions. Residues b, c and f are hydrophilic since they form the exterior of the coiled coil (Fig. S5). 4 [1] H. Warrick and J. Spudich, Ann. Rev. Cell Biol. 3, 379 R. F. Doolittle, Biochemistry 48, 3877 (2009). (1987). [2] W. Blankenfeldt, N. Thoma, J. Wray, M. Gautel, and I. Schlichting, Proc. Natl. Acad. Sci. USA 103, 17713 (2006). [16] S. Kohler, F. Schmid, and G. Settanni, PLoS Comput. Biol. 11, e1004346 (2015). [17] G. H. Wadhams and J. P. Armitage, Nat. Rev. Mol. Cell Bio. 5, 1024 (2004). [3] I. Schwaiger, C. Sattler, D. Hostetter, and M. Rief, Nat. [18] K. Kim, H. Yokota, and S.-H. Kim, Nature 400, 787 Mater. 1, 232 (2002). (1999). [4] D. Root, V. Yadavalli, J. Forbes, and K. Wang, Biophys. [19] J. Liu, Q. Zheng, Y. Deng, N. Kallenbach, and M. Lu, J. 90, 2852 (2006). [5] J. Eriksson, T. Dechat, B. Grin, B. Helfand, M. Mendez, H.-M. Pallari, and R. D. Goldman, J. Clin. Invest. 119, 1763 (2009). [6] D. A. Fletcher and R. D. Mullins, Nature 463, 485 (2010). [7] H. Herrmann, H. Bar, L. Kreplak, S. V. Strelkov, and U. Aebi, Nat. Rev. Mol. Cell Bio. 8, 562 (2007). [8] S. Strelkov, H. Herrmann, N. Geisler, T. Wedig, R. Zim- belmann, U. Aebi, and P. Burkhard, EMBO J. 21, 1255 (2002). J. Mol. Biol. 361, 168 (2006). [20] W. Kohn and R. Hodges, Trends Biotechnol. 16, 379 (1998). [21] S. Potekhin, T. Melnik, V. Popov, N. Lanina, A. Vaz- ina, P. Rigler, A. Verdini, G. Corradin, and A. Kajava, Chem. Biol. 8, 1025 (2001). [22] N. C. Burgess, T. H. Sharp, F. Thomas, C. W. Wood, A. R. Thomson, N. R. Zaccai, R. L. Brady, L. C. Ser- pell, and D. N. Woolfson, J. Am. Chem. Soc. 137, 10554 (2015). [23] P. Ferrara, J. Apostolakis, and A. Caflisch, Proteins 46, [9] J. Ferry, Proc. Natl. Acad. Sci. USA 38, 566 (1952). 24 (2004). [10] J. W. Weisel, in Fibrous Proteins: Coiled-Coils, Collagen and Elastomers, Advances in Protein Chemistry, Vol. 70, edited by D. A. D. Parry and J. M. Squire (Academic Press, 2005) pp. 247 -- 299. [24] B. Brooks, C. Brooks, A. MacKerell, L. Nilsson, R. Pe- trella, B. Roux, Y. Won, G. Archontis, C. Bartels, S. Boresch, et al., J. Comput. Chem. 30, 1545 (2009). [25] S. Falkovich, I. Neelov, and A. Darinskii, Polym. Sci. [11] A. Zhmurov, A. E. X. Brown, R. I. Litvinov, R. I. Dima, and V. Barsegov, Structure 19, 1615 J. W. Weisel, (2011). [12] A. Zhmurov, O. Kononova, R. Litvinov, R. Dima, V. Barsegov, and J. Weisel, J. Am. Chem. Soc. 134, 20396 (2012). [13] J. W. Weisel, Science 320, 456 (2008). [14] A. E. X. Brown, R. I. Litvinov, D. E. Discher, and J. W. Weisel, Biophys J. 92, L39 (2007). Ser. A 52, 662 (2010). [26] X. Liang, H. Kuhn, and M. Frank-Kamenetskii, Biophys. J. 90, 2877 (2006). [27] D. West, P. Olmsted, and E. Paci, J. Chem. Phys. 124, 154909 (2006). [28] P. B. Harbury, J. J. Plecs, B. Tidor, T. Alber, and P. S. Kim, Science 282, 1462 (1998). [29] K. M. Arndt, J. N. Pelletier, K. M. Muller, A. Pluckthun, and T. Alber, Structure 10, 1235 (2002). [15] J. M. Kollman, L. Pandi, M. R. Sawaya, M. Riley, and [30] J. M. Mason and K. M. Arndt, ChemBioChem 5, 170 (2004). TABLE S1. Structure and topology of superhelical proteins: the PDB entries (four-character codes) of the atomic structural models available from the Protein Data Bank; Nh is the number of α-helices formulating the superhelical structure; the vertical arrows show the topology and mutual orientation of the α-helices in space; stretches of amino acids provide information about the identity of α-helical residues formulating the superhelical architecture (taken from PDB files). 5 PDB Nh Topology System 2FXO 2 Myosin tail 1GK4 2 Vimentin Fibrinogen coiled-coils 3GHG 3 Four-stranded Phe-zipper 2GUS 4 1QU7 4 Bacterial receptor Phenylalanine zipper 2GUV 5 ↑↑ ↑↑ ↑↑↑ ↑↑↑↑ ↑↓↑↓ ↑↑↑↑↑ Residues 835 -- 963, 831 -- 961 328 -- 406, 328 -- 406 Homo-dimer Homo-dimer α49 -- 161, β80 -- 193, γ23 -- 135 Hetero-trimer Homo-tetramer 294 -- 489, 300-479 (13-54)×4 (1-56)×5 Homo-dimer Homo-pentamer 6 FIG. S1. Force-strain profiles (f ε-curves) for all superhelical protein fragments studied: myosin (a), bacterial chemo- taxis receptor (b), vimantin (c), fibrin (d), four-stranded (e) and five-stranded (f) phenylalanine zippers. Detailed information about all the model systems is provided in Table S1. The f ε-curves were obtained for various pulling speeds vf =104, 105 and 106 µm/s (color code is presented in panel a). The dynamic force protocol is described in the Main Text. All the simulated f ε-curves show good agreement with the theoretical predictions obtained using contin- uum phase transition theory (green curves) described in the Main Text. All the model parameters are accumulated in Table I in the Main Text). FIG. S2. Dynamics of the populations of the α-state pα and the β-state pβ for all superhelical protein fragments studied: myosin (a), bacterial chemotaxis receptor (b), vimentin (c), fibrin (d), four-stranded (e) and five-stranded (f) phenylalanine zippers. Detailed information about all the model systems is provided in Table S1. The populations were obtained using the hydrogen bond (H-bond) based estimation (bottom panels) and the dihedral angles' based estimation (top panels) as described in the Main Text. (see also Fig. 2 in the Main text). Compared are the colorized curves from MD simulations (color code is explained in the graphs) and green (solid and dashed) curves obtained with continuum theory described in the Main Text. For the H-bonds (D-H. . . A), we used the 3A cut-off for the distance between hydrogen donor atom D and acceptor atom A and the 20◦ cut-off for the D-H. . . A bond angle. For the dihedral angles, we used intervals −80◦<φ<−48◦ and −59◦<ψ<−27◦ for the α-phase, and −150◦<φ<−90◦ and 90◦<ψ<150◦ for the β-phase. FIG. S3. Free-energy profiles as functions of the projection of amino acid residue length along the direction of pulling (energy landscapes) for all superhelical protein fragments studied: myosin (a), bacterial chemotaxis receptor (b), vimentin (c), fibrin (d), four-stranded (e) and five-stranded (f) phenylalanine zippers. Detailed information about all the model systems is provided in Table S1. For each model system, the landscapes are shown for the following values of constant pulling force: low force f(cid:28)f0 from the 0−60 pN (single left minimum corresponding to the α-state; red curves); Maxwell force f≈f0 (two minima with the left (right) minimum corresponding to the α-state (β-state); black curves; see Table I for values of f0 for all systems); and large force f(cid:29)f0 (single right minimum corresponding to the β-state; blue curves). The insets show the corresponding normalized histograms constructed based on the MD simulation output. FIG. S4. The α-to-β structural transition in myosin superhelix (a) and fibrin superhelix (b) reflected in the sigmoidal profile of the average extension h∆Xi (with standard deviations) versus constant pulling force f . The theoretical curves X(f ) (solid red line) was used to perform a fit to the elongation data from the dynamic force-clamp measurements in silico (black squares) as described in the Main Text. Shown are snapshots of the α-phase, β-phase and mixed α+β-phase. The phase diagrams were constructed using the simulation output from five 200 ns long steered MD simulations (for each force value) for myosin and for fibrin. Theoretical curves were calculated using the analytically tractable two-state model from our earlier study [12]. FIG. S5. Schematic representation of the α-helices in superhelical proteins. In a heptad repeat of a left-handed coiled coil, residues a and d are hydrophobic. These residues form a hydrophobic core of the supercoil both in parallel and anti-parallel arrangements. The coiled coil is additionally stabilized by charged residues e and g. In a parallel supercoil, residue a is close to residue a0 and residue d is close to d0. In an anti-parallel supercoil, residue a forms a strong contact with residue d0, and residue d forms a strong contact with residue a0. Supercoils with more than two α-helices can be constructed in a similar fashion. For more information, see Refs. [28, 30]. 7 FIG. S1. Force-strain profiles (f ε-curves) for all superhelical protein fragments studied: myosin (a), bacterial chemotaxis receptor (b), vimantin (c), fibrin (d), four-stranded (e) and five-stranded (f) phenylalanine zippers. Detailed information about all the model systems is provided in Table S1. The f ε-curves were obtained for various pulling speeds vf =104, 105 and 106 µm/s (color code is presented in panel a). The dynamic force protocol is described in the Main Text. All the simulated f ε-curves show good agreement with the theoretical predictions obtained using continuum phase transition theory (green curves) described in the Main Text. All the model parameters are accumulated in Table I in the Main Text). 8 FIG. S2. Dynamics of the populations of the α-state pα and the β-state pβ for all superhelical protein fragments studied: myosin (a), bacterial chemotaxis receptor (b), vimentin (c), fibrin (d), four-stranded (e) and five-stranded (f) phenylalanine zippers. Detailed information about all the model systems is provided in Table S1. The populations were obtained using the hydrogen bond (H-bond) based estimation (bottom panels) and the dihedral angles' based estimation (top panels) as described in the Main Text. (see also Fig. 2 in the Main text). Compared are the colorized curves from MD simulations (color code is explained in the graphs) and green (solid and dashed) curves obtained with continuum theory described in the Main Text. For the H-bonds (D-H. . . A), we used the 3A cut-off for the distance between hydrogen donor atom D and acceptor atom A and the 20◦ cut-off for the D-H. . . A bond angle. For the dihedral angles, we used intervals −80◦<φ<−48◦ and −59◦<ψ<−27◦ for the α-phase, and −150◦<φ<−90◦ and 90◦<ψ<150◦ for the β-phase. 9 FIG. S3. Free-energy profiles as functions of the projection of amino acid residue length along the direction of pulling (energy landscapes) for all superhelical protein fragments studied: myosin (a), bacterial chemotaxis receptor (b), vimentin (c), fibrin (d), four-stranded (e) and five-stranded (f) phenylalanine zippers. Detailed information about all the model systems is provided in Table S1. For each model system, the landscapes are shown for the following values of constant pulling force: low force f(cid:28)f0 from the 0−60 pN (single left minimum corresponding to the α-state; red curves); Maxwell force f≈f0 (two minima with the left (right) minimum corresponding to the α-state (β-state); black curves; see Table I for values of f0 for all systems); and large force f(cid:29)f0 (single right minimum corresponding to the β-state; blue curves). The insets show the corresponding normalized histograms constructed based on the MD simulation output. 10 FIG. S4. The α-to-β structural transition in myosin superhelix (a) and fibrin superhelix (b) reflected in the sigmoidal profile of the average extension h∆Xi (with standard deviations) versus constant pulling force f . The theoretical curves X(f ) (solid red line) was used to perform a fit to the elongation data from the dynamic force-clamp measurements in silico (black squares) as described in the Main Text. Shown are snapshots of the α-phase, β-phase and mixed α+β-phase. The phase diagrams were constructed using the simulation output from five 200 ns long steered MD simulations (for each force value) for myosin and for fibrin. Theoretical curves were calculated using the analytically tractable two-state model from our earlier study [12]. 11 FIG. S5. Schematic representation of the α-helices in superhelical proteins. In a heptad repeat of a left-handed coiled coil, residues a and d are hydrophobic. These residues form a hydrophobic core of the supercoil both in parallel and anti-parallel arrangements. The coiled coil is additionally stabilized by charged residues e and g. In a parallel supercoil, residue a is close to residue a0 and residue d is close to d0. In an anti-parallel supercoil, residue a forms a strong contact with residue d0, and residue d forms a strong contact with residue a0. Supercoils with more than two α-helices can be constructed in a similar fashion. For more information, see Refs. [28, 30].
1105.1004
2
1105
2011-08-03T11:12:18
Growth and shortening of microtubules: a two-state model approach
[ "physics.bio-ph", "q-bio.SC" ]
In this study, a two-state mechanochemical model is presented to describe the dynamic instability of microtubules (MTs) in cells. The MTs switches between two states, assembly state and disassembly state. In assembly state, the growth of MTs includes two processes: free GTP-tubulin binding to the tip of protofilament (PF) and conformation change of PF, during which the first tubulin unit which curls outwards is rearranged into MT surface using the energy released from the hydrolysis of GTP in the penultimate tubulin unit. In disassembly state, the shortening of MTs includes also two processes, the release of GDP-tibulin from the tip of PF and one new tubulin unit curls out of the MT surface. Switches between these two states, which are usually called rescue and catastrophe, happen stochastically with external force dependent rates. Using this two-state model with parameters obtained by fitting the recent experimental data, detailed properties of MT growth are obtained, we find that MT is mainly in assembly state, its mean growth velocity increases with external force and GTP-tubulin concentration, MT will shorten in average without external force. To know more about the external force and GTP-tubulin concentration dependent properties of MT growth, and for the sake of the future experimental verification of this two-state model, eleven {\it critical forces} are defined and numerically discussed.
physics.bio-ph
physics
Growth and shortening of microtubules: a two-state model approach Laboratory of Mathematics for Nonlinear Science, Centre for Computational System Biology, School of Mathematical Sciences, Fudan University, Shanghai 200433, China. (Dated: October 24, 2018) Yunxin Zhang∗ In this study, a two-state mechanochemical model is presented to describe the dynamic insta- bility of microtubules (MTs) in cells. The MTs switches between two states, assembly state and disassembly state. In assembly state, the growth of MTs includes two processes: free GTP-tubulin binding to the tip of protofilament (PF) and conformation change of PF, during which the first tubulin unit which curls outwards is rearranged into MT surface using the energy released from the hydrolysis of GTP in the penultimate tubulin unit. In disassembly state, the shortening of MTs includes also two processes, the release of GDP-tibulin from the tip of PF and one new tubulin unit curls out of the MT surface. Switches between these two states, which are usually called rescue and catastrophe, happen stochastically with external force dependent rates. Using this two-state model with parameters obtained by fitting the recent experimental data, detailed properties of MT growth are obtained, we find that MT is mainly in assembly state, its mean growth velocity increases with external force and GTP-tubulin concentration, MT will shorten in average without external force. To know more about the external force and GTP-tubulin concentration dependent properties of MT growth, and for the sake of the future experimental verification of this two-state model, eleven critical forces are defined and numerically discussed. I. INTRODUCTION In eukaryotic cells, microtubules (MTs) serve as tracks for motor proteins [1 -- 6], give shape to cells, and form rigid cores of organelles [7 -- 10]. They also play essen- tial roles in the chromosome segregation [11 -- 18]. During cell division, MTs in spindle constantly grow and shorten by addition and loss of enzyme tubulin (GTPase) from their tips, then the attached duplicated chromosomes are stretched apart (through two kinetochores) from one an- other by the opposing forces (produced by MTs based on different spindles). Recently, many theoretical mod- els have been designed to understand the roles of MTs during chromosome segregation [19 -- 25]. One essential point to understand how MTs help chromosome segrega- tion during cell division is to know the mechanism of MT growth and shortening. In this study, inspired by the mechanochemical model for molecular motors [26], the GTP-cap model and catch bonds model for MT [8, 27], a two-state mechanochemical model will be presented. Electron microscopy indicates MT is composed of n parallel protofilaments (PFs, usually 12 ≤ n ≤ 15 and n = 13 is used in this study) which form a hollow cylinder [7, 8, 28]. Each PF is a filament that made of head-to-tail ∗Email: [email protected] associated αβ heterodimers. At the tip of MT, PFs curl outward from the MT cylinder surface. The tip might be in shrinking GDP-cap state or growing GTP-cap state. In contrast to the tip in shrinking GDP state, the growing GTP tip is fairly straight. Or in other words, in GTP-cap state, the angle between the curled out segment of PFs and MT surface is less than that in GDP state. In this study, we will only consider the growth and shortening of one single PF, and assume that each step of growth and shortening of one PF contributes to L (nm) of the growth and shortening of the whole MT. Intuitively, L = L1/n with L1 the length of one αβ hetrodimer. In the numerical calculations, L = 8 nm/13 ≈0.615 nm is used [19, 29]. Our two-state mechanochemical model for PF growth and shortening is schematically depicted in Fig. 1(a), and mathematically described by a two-line Markov chain in Fig. 1(b). In this model, PF stochastically switches be- tween two states: assembly state and disassembly state. During assembly state, PF grows through two processes, (i) 1 → 2: free GTP-tubulin binding process with GTP- tubulin concentration (denoted by [Tubulin]) dependent rate constant k1, and (ii) 2 → 1: PF conformation change process, during which, using energy released from GTP hydrolysis, the curled PF segment is straightened with one PF unit (i.e. one αβ heterodimer) rearranged into the MT surface, i.e. to parallel the MT axis ap- proximately. During disassembly state, each step of PF shortening includes also two processes, (i') 2′ ← 1′: dis- association of GDP-tubulin from PF tip to environment and (ii') 1′ ← 2′: one new PF unit curls out from the MT surface (during which phosphate is released from the tip tubulin unit simultaneously). The two-state model presented here can be regarded as a generalization of the one employed by B. Akiyoshi et al to explain their experimental data [27], which can be depicted by Fig. 2(a) [our corresponding generalized two-state model including bead detachment from MT is depicted in Fig. 2(b), see Sec. II.A for detailed discus- sion]. The reasons that we prefer to use this general- ized model are that, the simple model of B. Akiyoshi et al cannot fit the measured attachment lifetime of bead on MT well (see Fig. 4(a) in [27]), and moreover, the measurements in [30 -- 32] indicate that the rate of catas- trophe, i.e. transition from elongation to shortening, de- pendent on GTP-tubulin concentration of the solution. However, for the simple model depicted in Fig. 2(a), the catastrophe rate kc is independent of GTP-tubulin con- centration (it is biochemically reasonable to assume that the elongation rate k1 depend on GTP-tubulin concentra- tion, k1 = k0 1[Tubulin], but with no reasons to write kc as as a function of [Tubulin]). We will see from the Results section that, for our generalized model, the catastrophe rate does change with [Tubulin], since GTP-tubulin con- centration will change the probabilities of PF in states 1 and 2, and consequently change the transition rate from assembly state to disassembly state. At the same time, for the simple model depicted in Fig. 2(a), the distribu- tion of catastrophe time is an exponential. However, the experimental measurement under a particular situation indicates this distribution is clearly not an exponential [32] [44]. It should be pointed out, although our model presented here looks more complex, there are only two more parameters than the one depicted in Fig. 2(a) [45]. The organization of this paper is as follows. The two- state mechanochemical model will be presented and theo- retically studied in the next section, and then in Sec. III, based on the model parameters obtained by fitting the experimental data mainly obtained in [27], properties of MT growth and shortening are numerically studied, in- cluding its external force and GTP-tubulin concentration dependent growth and shortening speeds, mean dwell times in assembly and disassembly state, mean growth or shortening length before the bead, used in experiments 2 FIG. 1: Schematic depiction of the two-state mechanochem- ical model of protofilament (PF) growth and shortening (a) and its corresponding two-line Markov chain (b). In assem- bly state, the growth of PF accomplished by two processes, one GTP-tubulin binds to the tip of PF (with GTP-tubulin concentration dependent rate k1) and one PF unit rearranges into the MT surface (with external force dependent rate k2). The energy used in the second process comes from the GTP hydrolysis in the penultimate tubulin unit. One tubulin unit binding to the tip of PF is assumed to be equivalent to L (nm) growth of the whole MT (L = 0.615 nm is used in this study [19, 29]). Similarly, in disassembly state, the shortening of PF also includes two processes, one PF unit detaches from the tip of PF and one new PF unit curls out the MT surface. In this depiction, the same as in [25], a segment 5 dimers in length is assumed to curls out from the MT surface. to apply external force, detachment from MT. To know more properties about the MT dynamics, eleven critical forces (detailed definitions will be given in Sec. III) are also numerically discussed in Sec. III. Finally, Sec IV includes conclusions and remarks. II. TWO-STATE MECHANOCHEMICAL MODEL OF PROTOFILAMENT As the schematic depiction in Fig. 1, PF might be in two states, assembly (growth) state and disassem- bly (shortening) state. Each of the two states includes two sub-states, denoted by 1, 2 and 1′, 2′ respectively. Let p1, p2 be the probabilities that PF in assembly sub- states 1 and 2 respectively, and ρ1, ρ2 be the probabili- ties that the PF in disassembly sub-states 1′ and 2′, then p1, p2, ρ1, ρ2 are governed by the following master equa- tion dp1/dt =k2p2 − k1p1 + krρ1 − kcp1, dp2/dt =k1p1 − k2p2, dρ1/dt =k4ρ2 − k3ρ1 − krρ1 + kcp1, dρ2/dt =k3ρ1 − k4ρ2. (1) Where k1 is the rate of GTP-tubulin binding to the tip of PF, k2 is the rate of PF realignment with one new unit lying into the MT surface, k3 is the dissociation rate of GDP-tubulin from the tip of PF, and k4 is the rate of curling out of one tubulin unit from the MT surface (with Pi release simultaneously). The steady state solution of Eq. (1) is p1 = (cid:20)1 + k1 k2 + kc kr (cid:18)1 + k3 k4(cid:19)(cid:21)−1 , p2 = k1 k2 p1, ρ1 = kc kr p1, ρ2 = k3 k4 kc kr p1. (2) One can easily show that the mean steady state velocity of MT growth or shortening is [33, 34] V = (k2p2 − k4ρ2)L = (k1 − k3kc/kr) p1L, (3) where L is the effective step size of MT growth corre- sponding to one step growth of one PF, and V < 0 means MT is shortening in long time average with speed −V . Let ¯p1, ¯p2 be the probabilities that PF in sub-state 1 and sub-state 2 respectively, provided the PF is in as- sembly state, then ¯p1, ¯p2 satisfy d¯p1/dt =k2 ¯p2 − k1 ¯p1 = −d¯p2/dt. (4) One can easily get that, at steady state, the mean growth speed of MT with a PF in assembly state is Vg = k2 ¯p2L = k1k2L k1 + k2 . (5) Similarly, the mean shortening speed of MT with a PF in disassembly state is Vs = k3k4L k3 + k4 . (6) A. Modified model according to experiments: including bead detachment from MT To know the model parameters ki, i = 1, · · · , 4 and kc, kr, we need to fit the model with experimental data. In recent experiments [27], Akiyoshi et al attached a bead prepared with kinetochore particles to the growing end of 3 FIG. 2: (a) Schematic depiction of the two-state model used by B. Akiyoshi et al in [27]. In which, both the assembly and disassembly of PF are assumed to include only one pro- cess, described by rates k1 and k2 respectively. (b) Modified mechanochemical model with bead detachment. In the ex- periments of [27], Akiyoshi et al attached a bead prepared with kinetochore to the growing end of MTs, and measured not only the force dependent mean growing and shortening speeds, switch rates between assembly and disassembly states (i.e. rates of rescue and catastrophe), but also the mean life- time of the bead on MTs, and the rates of bead detachment during assembly and disassembly states respectively. There- fore, to get the model parameters and know more properties of MT growth and shortening, this modified model is used in this study . The main difference between these two mod- els is that, the rate of detachment from assembly state and the rate of catastrophe in model (b) depend on GTP-tubulin concentration [Tubulin], but they do not in model (a). MTs, and constant tension was applied to bead using a servo-controlled laser trap. In their experiments, not only the force dependent mean growth and shortening speeds of MTs, the rates of rescue and catastrophe, but also the force dependent mean lifetime, during which the bead is keeping attachment to MT, and mean detachment rates of the bead during assembly and disassembly states are measured. Therefore, to fit these experimental data, the model depicted in Fig. 1 should be modified to include the bead detachment processes [see Fig. 2(b)]. For the model depicted in Fig. 2(b), the formulations of mean growth velocity V , mean growth and shortening speeds Vg and Vs are the same as in Eqs. (3) and (5) (6). In the following, we will get the expression of mean lifetime of the bead on MTs. Let T1, T2, T1′, T2′ be the mean first passage times (MFPTs) of a bead to detach- ment with initial sub-states 1, 2, 1′ and 2′ respectively, then T1, T2, T1′, T2′ satisfy [35 -- 37] Let Tc1 and Tc2 be the MFPTs of MT to catastrophe from sub-states 1 and 2 respectively, then Tc1 and Tc2 satisfy [see Fig. 2(b)] Tc1 = 1 k1 + kc + k1 k1 + kc Tc2, Tc2 = 1 k2 + Tc1. (15) 4 T1 = T1′ = T2 = 1 k1 + ka + kc 1 kr + k3 + kd 1 k2 + T1, Then the mean lifetime can be obtained as follows T = p1T1 + p2T2 + ρ1T1′ + ρ2T2′, + + k1 k1 + ka + kc k3 kr + k3 + kd T2 + T2′ + kc k1 + ka + kc kr kr + k3 + kd T2′ = 1 k4 + T1′. T1′, The mean rate of catastrophe can be obtained by Kc = 1/Tc = 1/(¯p1Tc1 + ¯p2Tc2). The explicit expression can be obtained by replace ka with kc in Eq. (13) T1, (7) (8) Kc = (k1 + k2)k2kc (k1 + k2)2 + k1kc Similarly, the mean rate of rescue is Kr = (k3 + k4)k4kr (k3 + k4)2 + k3kr . . (16) (17) where p1, p2, ρ1, ρ2 can be obtained by formulations in Eq. (2). In assembly state, let Ta1 and Ta2 be the MFPTs to detachment of the bead initially at sub-states 1 and 2 respectively, then Ta1, Ta2 satisfy Ta1 = 1 k1 + ka + k1 k1 + ka Ta2, Ta2 = 1 k2 + Ta1. (9) One can easily show Ta1 = k1 + k2 k2ka , Ta2 = k1 + ka + k2 k2ka . (10) Therefore, the MFPT to detachment of the bead in as- sembly state is Ta = ¯p1Ta1 + ¯p2Ta2, where the steady state probabilities ¯p1 = k2 k1 + k2 , ¯p2 = k1 k1 + k2 , (11) (12) are obtained from Eq. (4). The mean detachment rate during assembly can then be obtained by Ka = 1/Ta = 1/(¯p1Ta1 + ¯p2Ta2), i.e., B. Force and GTP-tubulin concentration dependence of the transition rates From the experimental data in [27] [or see Fig. 3], one sees some transition rates in our model should depend on the external force. Since the processes 1 → 2 and 2′ ← 1′ are accomplished by binding tubulin unit to and releasing tubulin unit from the tip of PF [see Fig. 1 and 2(b)], we assume that k1 and k3 are force independent. Similar as the methods demonstrated in the models of molecular motors [26, 38] and models for adhesive of cells to cells [39], the external force F dependence of rates k2, k4, ka, kd, kr, kc are assumed to be the following forms kl = k0 l eF Lδl/kB T , l = 2, 4, a, d, r, c. (18) Hereafter, the external froce F is positive if it points to the direction of MT growth. Meanwhile, the rate k1 should depend on the concen- tration of free GTP-tubulin in solution. Similar as the method in [26], we simply assume k1 = k0 1[Tubulin]. Ka = (k1 + k2)k2ka (k1 + k2)2 + k1ka . (13) C. Critical forces of MT growth Similarly, the mean detachment rate during disassembly can be obtained as follows Kd =1/Td = 1/(¯ρ1Td1 + ¯ρ2Td2) = (k3 + k4)k4kd (k3 + k4)2 + k3kd , (14) with steady state probabilities ¯ρ1 = k4/(k3 + k4), ¯ρ2 = k3/(k3 + k4). For the sake of the better understanding of external force F dependent properties of MTs and the experimen- tal verification of the two-state model, in the following, we will define altogether eleven critical forces. Corre- sponding numerical results will be presented in the next section. (1) Critical Force Fc1: under which Vg(Fc1) = Vs(Fc1), i.e. the average speeds of assembly and disassembly are the same. From Eqs. (5) (6) one sees Fc1 satisfies k1k2(Fc1)[k3 + k4(Fc1)] = k3k4(Fc1)[k1 + k2(Fc1)]. (19) (2) Critical Force Fc2: under which the mean veloc- ity of MT growth is vanished. Formulation (3) gives k1kr(Fc2) = k3kc(Fc2), i.e. Fc2 = kBT (δr − δc)L = kBT (δr − δc)L ln ln k3k0 c k1k0 r k3k0 c r [Tubulin] k0 1k0 . (20) (3) Critical Force Fc3: under which p1+p2 = ρ1 +ρ2, i.e., the probabilities that MTs in assembly and disassembly states are the same. From expressions in Eq. (2), one easily sees Fc3 satisfies kr(Fc3)k4(Fc3)[k1 + k2(Fc3)] =kc(Fc3)k2(Fc3)[k3 + k4(Fc3)]. (21) (4) Critical Force Fc4: under which the detachment rates during assembly and disassembly states are the same. In view of formulations (13) and (14), one can get Fc4 by Ka(Fc4) = Kd(Fc4). (5) Critical Force Fc5: under which the mean dwell times of MT in assembly and disassembly states are the same. Let Tg1 and Tg2 be the MFPTs of bead to detachment or catastrophe of MT with initial sub-states 1 and 2 re- spectively, then Tg1, Tg2 satisfy (see Fig. 2(b) and Refs. [35, 36]) + k1 k1 + ka + kc Tg2, (22) Tg1 = Tg2 = 1 k1 + ka + kc 1 k2 + Tg1. Its solution is Tg1 = k1 + k2 k2(ka + kc) , Tg2 = k1 + ka + kc + k2 k2(ka + kc) . (23) The mean dwell time of MT in assembly (or growth) state is then Tg = ¯p1Tg1 + ¯p2Tg2 = (k1 + k2)2 + k1(ka + kc) k2(k1 + k2)(ka + kc) . (24) Similarly, the mean dwell time of MT in disassembly (or shortening) state can be obtained as follows Ts = (k3 + k4)2 + k3(kd + kr) k4(k3 + k4)(kd + kr) . (25) The critical force Fc5 can then be obtained by Tg(Fc5) = Ts(Fc5). 5 (6) Critical Force Fc6: under which the mean lifetime of the bead on MT attains its maximum, i.e. T (Fc6) = maxF T (F ) with T given by formulation (8). (7) Critical Force Fc7: under which the mean growth length of MT attains its maximum. The mean growth length of MT can be obtained by l+ = V T with V, T satisfy formulations (3) and (8) respectively. (8) Critical Force Fc8: under which the mean shortening length of MT attains its maximum. The mean shortening length of MT can be obtained by l− = −V T with V, T satisfy formulations (3) and (8) respectively. (9) Critical Force Fc9: the rates of catastrophe and res- cue are the same, i.e. Kc(Fc9) = Kr(Fc9) [see Eqs. (16) and (17)]. Under critical force Fc9, the average switch time between growth and shortening, 1/Kc and 1/Kr, are the same. It is to say that the mean dura- tion for each growth and each shortening period are the same. (10) Critical Force Fc10: under which VgTg = VsTs. Here VgTg =: lg is the mean growth length before bead detach- ment or catastrophe, and VsTs =: ls is the mean short- ening length before bead detachment or rescue. The for- mulations of Vg, Vs and Tg, Ts are in Eqs. (5) (6) and (24) (25). (11) Critical Force Fc11: under which Vg/Kc = Vs/Kr. Here Vg/Kc =: l∗ g is the mean growth length before catas- trophe, and Vs/Kr = l∗ s is the mean shortening length before rescue. The formulations of Kc, Kr are in Eqs. (16) (17). i.e. It needs to be clarified that, the definitions for Fc1, Fc2, Fc3, Fc9, Fc11 are unrelated to bead detachment, but the definitions for Fc4, Fc5, Fc6, Fc7, Fc8, Fc10 do. Therefore the values of Fc1, Fc2, Fc3, Fc9, Fc11 obtained in this theoretical study can be verified by various experi- mental methods as in [13, 30 -- 32, 40 -- 42], but the values of Fc4, Fc5, Fc6, Fc7, Fc8, Fc10 can only be verified by similar experimental method as in [27]. For the sake of conve- nience, and based on the above definitions and numerical calculations in Sec. III (see Figs. 7 and 8), basic prop- erties of the eleven critical forces Fci are listed in Tab. I. Meanwhile, the main symbols used in this study are listed in Tab. II. III. RESULTS In order to discuss the properties of MT growth and shortening, the model parameters, i.e. k0 i , δi for i = 2, 4, a, d, r, c should be firstly obtained. By fit- 1, k3 and k0 TABLE I: Basic properties of the critical forces as defined in Sec. II.C, see also Figs. 7 and 8. Ka < Kd Tg > Ts F < Fci Vg < Vs V < 0 F > Fci Vg > Vs V > 0 F = Fci Vg = Vs V = 0 i 1 2 3 p1 + p2 < ρ1 + ρ2 p1 + p2 = ρ1 + ρ2 p1 + p2 > ρ1 + ρ2 4 5 6 7 8 9 10 11 T < maxF T l+ < maxF l+ l− < maxF l− T < maxF T l+ < maxF l+ l− < maxF l− T = maxF T l+ = maxF l+ l− = maxF l− Ka = Kd Tg = Ts Ka > Kd Tg < Ts lg < ls l∗ g < l∗ s lg > ls l∗ g > l∗ s Kr > Kc Kr = Kc lg = ls l∗ g = l∗ s Kr < Kc TABLE II: The main symbols and their expressions (or defi- nitions) used in this study. Symbol V Vg Vs T Ka Kd Ta Td Kc Kr pi, ρi Tg Ts lg, ls g, l∗ l∗ s l+, l− kl Definitions Biophysical meaning mean velocity of MTs growth speed of MTs shortening speed of MTs Eq. (3) Eq. (5) Eq. (6) Eq. (8) Eq. (13) bead detachment rate (disassembly) Eq. (14) bead detachment rate (assembly) mean lifetime of bead time to detachment (assembly ) time to detachment (assembly ) catastrophe rate rescue rate probability mean growth time mean shortening time lg = VgTg, ls = VsTs g = Vg/Kc, l∗ l∗ l+ = V T, l− = −V T s = Vs/Kr rate constants [Fig. 2(2)] 1/Ka 1/Kd Eq. (16) Eq. (17) Eq. (2) Eq. (24) Eq. (25) see Fc10 see Fc11 see Fc7,8 Eq. (18) ting the expressions of Vg, Vs, T, Ka, Kd, Kc, Kr, which are given in Eqs. (5) (6) (8) (13) (14) (16) (17) respec- tively, to the experimental data mainly measured in [27], these parameter values are obtained (see Fig. 3 and Tab. III, the fitting methods are discussed in [46]). The data corresponding to zero external force in Figs 3(b) and 3(c) [the two black dots on vertical axis] are obtained by fit- ting the corresponding measurement in [30] with a con- stant [see the two lines in Figs. 4(a) and 4(b)], since as implied by our model, the rates of MT shortening and res- cue are independent of GTP-tubulin concentration. All the following calculations will be based on the parameters listed in Tab. III. The curve in Fig. 4(b) is the theoret- 6 TABLE III: Model parameters obtained by fitting the expres- sions in (5) (6) (8) (13) (14) (16) (17) to the experimental data mainly measured in [27]. In the fitting, kBT = 4.12 pN·nm and effective step size L = 0.615 nm are used [19, 29, 43]. The fitting results are plotted in Fig. 3. Parameter value Parameter k0 1 k3 k0 r k0 a δg δr δa 103 102 101 100 ) 1 − r h ( e t a r t n e m h c a t e d ) 1 − r h ( e t a r h c t i w S (c) 10−1 0 103 102 101 100 10−1 0 5.8 × 105 s−1µM−1 1.0 × 108 s−1 7.4 × 103 s−1 41.9 s−1 0.68 3.71 1.77 (a) Disassembly 2 4 k0 k0 k0 c k0 d δs δc δd (b) 103 102 ) 1 − s m n ( Assembly 4 8 Force (pN) 12 16 Rescue Catastrophe 2 4 Force (pN) 6 8 (d) 101 d e e p S 100 0 ) n m i ( e m i t e f i L 70 60 50 40 30 20 10 0 0 Value 9.3 s−1 571.6 s−1 1.8 × 103 s−1 1.4 × 104 s−1 -2.88 -2.96 0.33 Shortening Growth 4 8 Force (pN) 12 16 4 8 Force (pN) 12 16 FIG. 3: Theoretical results of the two-state model [see Fig. 2(b)] and experimental data obtained by Akiyoshi et al [27]: The detachment rates are obtained by formulations (13) (14), the speeds are obtained by formulations (5) (6), the switch rates are obtained by formulations (16) (17), and the lifetime is obtained by formulation (8). The model parameters used in the theoretical calculations are listed in Tab. III. The two black dots on vertical axis of (b) and (c) are obtained by averaging the data in [30] [see the lines in Fig. 4(a) and Fig. 4(b)]. ical prediction of GTP-tubulin concentration dependent catastrophe rate Kc by formulation (16). Compared with the experimental data measured in [30, 32], these predic- tion looks satisfactory [47]. From Fig. 5(a), one can see that, the MT is mainly in assembly state. Further calculations indicate that the ra- tio of probabilities in assembly state to disassembly state, i.e. (p1 + p2)/(ρ1 + ρ2), increases exponentially with ex- ternal force F [see Fig. 6(a)]. In experiments of Akiyoshi at al [27], the external force F is applied to MT through 100 ) n m i p1 p2 ρ 1 ρ 2 7 Tg Ts 4 8 Force (pN) 12 16 4 8 Force (pN) 12 16 ( e m T i l l e w D n a e M ) m n ( h t g n e L h t w o r G n a e M 10 1 0.3 (b) 0.05 0 x 104 (d) 3 2.5 2 1.5 1 0.5 0 −0.5 −1 0 ρ , ρ , 1 2 2 p , p 1 100 10−10 (a) 0 ) 1 − s m n ( h t w o r g f o y t i c o e v n a e M l (c) 30 20 10 0 −10 −20 0 4 8 Force (pN) 12 16 4 8 Force (pN) 12 16 FIG. 5: Properties of MT growth and shortening obtained by the two-state model [see Fig. 2(b)] with model parame- ters listed in Tab. III: (a) Under external force, the MT is mainly in assembly state, both the probabilities p1 and p2 of MT in assembly sub-states 1 and 2 increase with force but with p2 ≫ p1. During disassembly state, the MT is mainly in sub-state 2′, ρ2 > ρ1. Here p1, p2, ρ1, ρ2 are calculated by formulations in (2). (b) The dwell time Tg of MT in assembly state is always larger than that in disassembly state (denoted by Ts) for external force less than 16 pN [see formulations (24) (25) for Tg and Ts]. Similar as the mean lifetime of bead at- tachment to MT [see Fig. 3(d)], both Tg and Ts increase firstly and then decrease with external force. (c) The mean velocity of MT growth [see formulation (3)] increases with external force monotonically, where the negative velocity means MT shortens its length in long time average, though the curves in (a) and (b) imply that the MT is mainly in assembly state. (d) The mean growth length before bead detachment increases firstly and then deceases with external force. Here the mean growth length is obtained by mean growth velocity of MT multiplied by mean lifetime of the bead, i.e. V T [see formu- lation (3) for V and formulation (8) for T ]. tration, which can be verified by the data in [30], see Fig. 4), the eleven critical forces defined in the pre- vious section also depend on GTP-tubulin concentra- tion. For convenience, in our calculations (the results are plotted in Figs. 7 and 8), [Tubulin]=1 means the free GTP-tubulin concentration is the same as the one used by Akiyoshi et al [27]. From Figs. 7 and 8, one can see, the critical forces Fci for i = 4, 5, 6 in- crease, but others decrease with GTP-tubulin concen- tration [Tubulin]. For high GTP-tubulin concentration, Fc2 ≈ Fc11 and Fc3 ≈ Fc9 since for kc ≪ k1, equations Vg/Kc = Vs/Kr and Kc = Kr can be well approximated FIG. 4: GTP-tubulin concentration dependent data measured by Walker et al [30]: (a) Shortening speed of MT and their average value. (b) Switch rates of MT between assembly and disassembly, where the curve is obtained by our theoretical model using the parameter listed in Tab. III (see formulation (16) with k1 = k0 1[Tubulin]), the solid squares are experimen- tal data from [32]. a bead attached to its growing tip. Fig. 5(b) indicates that, for F ≤ 16 pN, the mean dwell time of MT in as- sembly state before bead detachment is larger than that in disassembly state. Although the MT is mainly in as- sembly state, its mean growth velocity is negative under small external force [Fig. 5(c)], since for such cases, the shortening speed is greatly larger than the growth speed [see Fig. 3(b)]. But, Fig. 5(c) indicates the mean veloc- ity of MT growth always increases with external force. Similar as the mean growth velocity, the mean growth length of MT before bead detachment might be negative [i.e. MT shortens its length in long time average, see Fig. 5(d)], though the MT spends most of its time in assem- bly state [Fig. 5(b)]. Similar as the mean lifetime [Fig. 3(d)], the mean growth length of MT also has a global maximum for external force [Fig. 5(d)]. As we have mentioned in the Introduction, the chromosome segrega- tion is accomplished by the tensile force generated during MTs disassembly, Fig. 5 tells us the critical force of one MT disassembly is about 1.2 pN under the present exper- imental environment [27]. In Fig. 6(b), the mean growth length lg, l∗ s which are given in the definitions of critical force Fc10, Fc11 are also plotted as functions of external force. One can easily see that lg ≤ l∗ s since the mean dwell time of MT in assembly state Tg ≤ 1/Kc and mean dwell time in disassembly state Ts ≤ 1/Kr. But for large external force, ls ≈ l∗ s since, for such cases, MTs leave disassembly state mainly by rescue. g and mean shortening length ls, l∗ g, and ls ≤ l∗ Since the assembly of MT depends on free GTP- tubulin concentration (in our model, the simple relation k1 = k0 1[Tubulin] is used, and the disassembly process is assumed to be independent of GTP-tubulin concen- (a) 11 10 9 8 e c r o F l a c i t i r C 7 10−6 (c) 20 18 16 14 s e c r o F l a c i t i r C 12 10−6 Fc1 Fc3 Fc9 10−5 [Tubulin] Fc2 Fc11 10−5 [Tubulin] 16 (b) s e c r o F l a c i t i r C 14 12 10−4 10 10−6 4 (d) e c r o F l a c i t i r C 3.8 3.6 3.4 10−5 [Tubulin] 10−4 3.2 10−3 10−2 [Tubulin] 8 10−4 Fc6 10−1 FIG. 8: Plots of critical forces Fci for i = 1, 2, 3, 6, 9, 11 under low GTP-tubulin concentration. The meaning of [Tubulin] is the same as described in the caption of Fig. 7. [Tubulin]. Therefore, the critical force Fc1 decreases with GTP-tubulin concentration [Tubulin] [see Fig. 8(a)]. But for high [Tubulin], critical force Fc1 is almost a constant [see Fig. 7(a)] since, for saturating concentration, the growth speed Vg tends to a constant [see Eq. (5) and Fig. 9(a)]. The decrease of critical force Fc2 with [Tubu- lin] can be easily seen from expression (20) [see Fig. 7(b)]. The decrease of critical forces Fc1, Fc2 implies that low GTP-tubulin concentration might be helpful to chromo- some segregation. From expressions in (2) one can verify (p1 + p2)/(ρ1 + ρ2) = krk4[k1 + k2]/kck2[k3 + k4]. So (p1 + p2)/(ρ1 + ρ2) increases linearly with [Tubulin] [see Fig. 10(a)]. At the same time, δr + δs > 0, δg > 0 and δg +δc < 0, δs < 0 (see Tab. III) imply (p1 +p2)/(ρ1 +ρ2) also increases with external force F [see Fig. 6(a)]. Therefore, the critical force Fc3 decreases with [Tubulin] [see Fig. 7(c)]. Since the detachment rate Ka increases and detach- ment rate Kd decreases with external force F [see Fig. 3(a)], and Ka increases with but Kd is independent of [Tubulin] [see Eqs. (13) (14)], the critical force Fc4 in- creases with [Tubulin] [see Fig. 7(a)]. The increase of critical force Fc5 indicates MTs will spend more time in assembly state at high GTP-tubulin concentration [see Tab. I and Figs. 7(a) and 5(b)]. The increase of critical force Fc6 [see Fig. 8(d)] implies, the peak of the lifetime- force curve as plotted in Fig. 3(d) will move rightwards as the increase of [Tubulin], but with a upper bound around 4 pN [see Figs. 7(d) and 9(d)]. Similarly, the decrease of critical force Fc7 [see Fig. 7(d)] means, the peak of the mean growth length-force curve will move leftwards as FIG. 6: (a) The ratio of probability p1 + p2 that MT in as- sembly state to probability ρ1 + ρ2 that MT in disassembly state increases exponentially with the external force, and un- der positive external force, the MT mainly stays in assem- bly state, although the assembly speed might be much lower compared with the disassembly speed [see Fig. 3(b)]. (b) The mean growth length lg, l∗ s of MT in one assembly and disassembly period. The difference between lg, ls and l∗ s , the bead at- tached to the tip of MT, through which the external force is applied to MT, is assumed to keep attachment to MT, or the external force just applied by other methods [30 -- 32, 41], so the MT can only leave assembly state by catastrophe and leave disassembly state by rescue. s is that, in the calculation of l∗ g and shortening length ls, l∗ g, l∗ g, l∗ (a) s e c r o F l a c i t i r C 20 15 10 10−1 −2 −4 −6 −8 −10 −12 (c) 10−1 s e c r o F l a c i t i r C 100 [Tubulin] 101 Fc1 Fc4 Fc5 Fc3 Fc8 Fc9 s e c r o F l a c i t i r C 4 2 0 −2 −4 10−1 5.5 5 4.5 4 s e c r o F l a c i t i r C 100 [Tubulin] 101 (d) 3.5 10−1 Fc2 Fc10 Fc11 100 [Tubulin] (b) 101 Fc6 Fc7 100 [Tubulin] 101 FIG. 7: Critical forces as defined in Sec. II.C, in which [Tubu- lin]=1 means the concentration of GTP-tubulin is the same as the one used in the experiments of Akiyoshi et al [27]. For better understand the curves for Fci, see Tab. I. by k1Kr = k3kc and krk4(k1 + k2) = kck2(k3 + k4) [48]. Since the force distribution factors δg > 0, δs < 0 (see Tab. III), from Eqs. (5) (6) one can easily show that the growth speed Vg increases but the shortening speed Vs decreases with external force F . Therefore, Vg < Vs if F < Fc1 (see Tab. I). Eqs. (5) (6) also indicate that the growth speed Vg increases with but the shortening speed Vs is independent of GTP-tubulin concentration the increase of [Tubulin], and with lower bound around 4.44 pN. Finally, critical forces Fc8, Fc9, Fc10, Fc11 all decrease with [Tubulin]. It may need to say that, in Ref. [27], only experimental data for positive force cases are measured, and similar experimental methods as used in Refs. [32, 41] might be employed to apply negative force to MTs. At the same time, the mechanism of MT growth and shortening under negative external force cases might be completely different from that under positive external force cases, so for the results of critical forces plotted in Fig. 7 which have negative values, experimental verifica- tion should be firstly done before further analysis. To better understand the GTP-tubulin concentration [Tubulin] dependent properties of MT assembly and dis- assembly, more figures are plotted in Figs. 9 and 10. One can see that the mean lifetime T , ratio (p1 +p2)/(ρ1 +ρ2), and mean growth length lg, l∗ g all increase linearly with [Tubulin] (from the corresponding formulations, one can easily see that the mean shortening speed Vs, and mean shortening length ls, l∗ s are all independent of [Tubulin]). The mean velocity V and mean growth speed Vg also increase with [Tubulin], but tend to a external force F dependent constant [one can verify this limit constant is k2L = k0 2L exp(F δg/kBT ). For such cases, the MT stays mainly in sub-state 2, i.e., p2 ≈ 1, see Fig. 5(a)]. The mean growth length V T does not change monotonically with external force [see Figs. 5(a) and 9(c)] but increases with [Tubulin] for high GTP-tubulin concentration cases. IV. CONCLUDING REMARKS In this study, a two-state mechanochemical model of microtubulin (MT) growth and shortening is presented. In assembly (growth) state, one GTP-tubulin will attach to the growing tip of the protofilament (PF) firstly and then, after the hydrolysis of GTP in the penultimate PF unit, the curved PF segment is slightly straightened with one new PF unit lying into the MT cylinder surface. In disassembly (shortening) state, one tubulin unit will de- tach from the tip of PF, and then the GDP (or GDP+Pi) capped tip segment of PF will be further curved with one new tubulin unit out of the MT surface (the phos- phate is assumed to be released simultaneously). The PF can switch between the assembly and disassembly states with external force dependent rates stochastically. Each assembly or disassembly process contributes to one step of growth or shortening of MT with step size L =0.615 9 F=0 F=1 F=2 F=4 F=8 10 [GTP]=0.1 [GTP]=0.2 [GTP]=0.5 [GTP]=1 [GTP]=2 [GTP]=5 [GTP]=10 5 [Tubulin] (b) 500 400 300 200 100 ) n m i ( e m i t e f i l n a e M 0 0 (d) 103 102 101 100 ) n m i ( e m i t e f i l n a e M 0 5 Force (pN) 10 F=0 F=1 F=2 F=4 F=8 (a) 101 (c) 101 100 [Tubulin] 100 [Tubulin] ) 1 − s m n ( h t w o r g f o y t i c o e v n a e M l 20 0 −20 −40 −60 −80 −100 10−1 ) 1 − s m n ( h t g n e l h t w o r g n a e M x 104 F=0 F=1 F=2 F=4 F=8 4 3 2 1 0 −1 10−1 FIG. 9: GTP-tubulin concentration dependent properties of MT growth and shortening. (a) The mean growth veloc- ity V [see formulation (3)] increases with GTP-tubulin con- centration. The plots also indicate V increases with exter- nal force [see Fig. 5(c)]. (b) and (d) The mean lifetime of bead attachment to MT increases with GTP-tubulin concen- tration, but increases first and then decreases with external force. For high GTP-tubulin concentration, the critical force Fc6 under which the mean lifetime gets its maximum is al- most a constant (about 4 pN), see also Fig. 7(d). (c) The mean growth length of MT before bead detachment does not change monotonically with external force and GTP-tubulin concentration, so there exists critical force under which the maximum is obtained. But for high GTP-tubulin concentra- tion, mean growth length increases with [Tubulin]. (d) For any GTP-tubulin concentration, the mean lifetime does not change monotonically with external force. The optimal value of external force, under which the mean lifetime is maximum, increases with GTP-tubulin concentration, but is almost in- variable for large [Tubulin]. (c) (d) Both the mean growth length and mean lifetime do not change monotonically with external force, so there exists optimal values under which the corresponding maximum is reached (see Fc6, Fc7 Fig. 7). nm. This model can fit the recent experimental data measured by Akiyoshi et al [27] well. From this model, interesting properties of MT growth and shortening are found: Under large external force or high GTP-tubulin concentration, the MT is mainly in as- sembly state; The mean lifetime of bead attachment to MT and mean growth length during this period (in ex- periments, the external force is applied to MT through a bead attached to the growing tip of MT) increase firstly and then decrease with the external force, but roughly speaking, they all increase with the GTP-tubulin concen- tration; The growth speed of MT increases with GTP- F=0 F=1 F=2 F=4 F=8 5 [Tubulin] 10 g ) s / m n ( V d e e p S h t w o r G (b) 25 20 15 10 5 0 10−6 10−5 10−4 [Tubulin] F=0 F=1 F=2 F=4 F=8 F=0 F=1 F=2 F=4 F=8 10−3 F=0 F=1 F=2 F=4 F=8 1000 (a) 800 600 400 200 ) 2 1 ρ + ρ ( / ) 2 p + p ( 1 (c) 0 0 40 30 20 10 0 0 ) m µ ( g l h t g n e L h t w o r G ) 100 (d) m µ ( g* l h t g n e L h t w o r G 80 60 40 20 0 0 5 [Tubulin] 10 5 [Tubulin] 10 FIG. 10: The ratio of probability p1 + p2 that MT in as- sembly state to probability ρ1 + ρ2 that MT in disassembly state, and the mean growth length lg, l∗ g in each growth pe- riod all increase linearly with GTP-tubulin concentration (a) (c) (d) [see Tab. II for definitions]. The growth speed Vg of MT increases with [Tubulin] but tends to a limit constant for saturated concentration [see Eq. (5) for the formulation of Vg]. tubulin concentration but has an external force depen- dent limit. For the sake of experimental verification, al- together eleven critical forces are defined, including the force under which the mean lifetime or mean growth 10 length reach its maximum, the mean assembly speed is equal to the mean disassembly speed, the probabilities of MT in assembly and disassembly states are equal to each other, the detachment rates of bead during assembly and disassembly states are the same, the mean dwell times in assembly and disassembly states are the same, the mean growth velocity of MT is vanished, etc. Almost all of the above critical forces decrease with the GTP-tubulin concentration, since high GTP-tubulin concentration is favorable for MT growth and under low GTP-tubulin concentration, MT will shortens its length in average. Roughly speaking, GTP-tubulin and external force are helpful to MT assembly, but there exists optimal values external force for the mean lifetime of bead on MT and mean growth length of MT. Acknowledgments This study is funded by the Natural Science Founda- tion of Shanghai (under Grant No. 11ZR1403700). The author thanks Michael E. Fisher of IPST in University of Maryland for his initial introduction and inspiration of the present study, and is also very appreciated for the referees' critical comments and valuable suggestions, due to which many changes have been done. [1] F. Julicher and J. Prost, Phys. Rev. Lett. 75, 2618 rison, Nat. Struct. Mol. Biol. 12, 138 (2005). (1995). [2] M. J. Schnitzer and S. M. Block, Nature 388, 386 (1997). [3] R. D. Vale, Cell 112, 467 (2003). [4] M. Schliwa, Molecular Motors (Wiley-Vch, Weinheim, 2003). [5] A. O. Sperry, Molecular Motors: Methods and Protocols (Methods in Molecular Biology Vol 392) (Humana Press Inc., Totowa, New Jersey, 2007). [6] A. B. Kolomeisky and M. E. Fisher, Ann. Rev. Phys. Chem. 58, 675 (2007). [7] G. M. Cooper, The Cell: A Molecular Approach, 2nd Edn (Sinauer Associates, Inc., Sunderland, Mass., 2000). [8] J. Howard, Mechanics of Motor Proteins and the Cy- toskeleton (Sinauer Associates, Sunderland, MA, 2001). [9] J. Howard, Phys. Biol. 3, 54 (2006). [10] J. Howard and A. A. Hyman, J. Cell. Biol. 10, 569 (2009). [11] S. Westermann, A. Avila-Sakar, H.-W. Wang, H. Nieder- strasser, J. Wong, D. G. Drubin, E. Nogales, and G. Barnes, Mol Cell. 17, 277 (2005). [12] J. J. Miranda, P. D. Wulf, P. K. Sorger, and S. C. Har- [13] E. L. Grishchuk, M. I. Molodtsov, F. I. Ataullakhanov, and J. R. McIntosh, Nature 438, 384 (2005). [14] S. Westermann, H. W. Wang, A. Avila-Sakar, D. G. Dru- bin, E. Nogales, and G. Barnes, Nature 440, 565 (2006). [15] A. D. Franck, A. F. Powers, D. R. Gestaut, T. Gonen, T. N. Davis, and C. L. Asbury, Nat. Cell Biol. 9, 832 (2007). [16] J. R. McIntosh, E. L. Grishchuk, M. K. Morphew, A. K. Efremov, K. Zhudenkov, V. A. Volkov, I. M. Cheeseman, A. Desai, D. N. Mastronarde, and F. I. Ataullakhanov, Cell 135, 322 (2008). [17] A. F. Powers, A. D. Franck, D. R. Gestaut, J. Cooper, B. Gracyzk, R. R. Wei, L. Wordeman, T. N. Davis, and C. L. Asbury, Cell 136, 865 (2009). [18] Q. Gao, T. Courtheoux, Y. Gachet, S. Tournier, and X. Hea, Proc. Natl. Acad. Sci. USA 107, 13330 (2010). [19] T. L. Hill, Proc. Natl. Acad. Sci. USA 82, 4404 (1985). [20] M. I. Molodtsov, E. L. Grishchuk, A. K. Efremov, J. R. McIntosh, and F. I. Ataullakhanov, Proc. Natl. Acad. Sci. USA 102, 4353 (2005). [21] E. Salmon, Curr. Biol. 15, R299 (2005). [22] S. W. Grill, K. Kruse, and F. Julicher, Phys. Rev. Lett. 94, 108104 (2005). [23] A. Efremov, E. L. Grishchuk, J. R. McIntosh, and F. I. Ataullakhanov, Proc. Natl. Acad. Sci. USA 104, 19017 (2007). [24] J. W. Armond and M. S. Turner, Biophys. J. 98, 1598 (2010). [25] C. L. Asbury, J. F. Tien, and T. N. Davis, Trends Cell Biol. 21, 38 (2011). [26] M. E. Fisher and A. B. Kolomeisky, Proc. Natl. Acad. Sci. USA 98, 7748 (2001). [27] B. Akiyoshi, K. K. Sarangapani, A. F. Powers, C. R. Nelson, S. L. Reichow, H. Arellano-Santoyo, T. Gonen, J. A. Ranish, C. L. Asbury, and S. Biggins, Nature 468, 576 (2010). [28] D. Bray, Cell movements: from molecules to motility, 2nd Edn (Garland, New York, 2001). [29] A. B. Kolomeisky and M. E. Fisher, Biophys. J. 80, 149 (2001). [30] R. A. Walker, E. T. O'Brien, N. K. Pryer, M. E. Soboeiro, W. A. Voter, H. P. Erickson, and E. D. Salmon, The Journal of Cell Biology 107, 1437 (1988). [31] R. A. Walker, N. K. Pryer, and E. D. Salmon, The Jour- nal of Cell Biology 114, 73 (1991). [32] M. E. Janson, M. E. de Dood, and M. Dogterom, The Journal of Cell Biology 161, 1029 (2003). [33] B. Derrida, J. Stat. Phys. 31, 433 (1983). [34] Y. Zhang, Phys. Lett. A 373, 2629 (2009). [35] S. Redner, A Guide to First-Passage Processes (Cam- bridge University Press, 2001). [36] P. A. Pury and M. O. C´aceres, J. Phys. A: Math. Gen. 36, 2695 (2003). [37] A. B. Kolomeisky, E. B. Stukalin, and A. A. Popov, Phys. Rev. E 71, 031902 (2005). [38] Y. Zhang, Physica A 383, 3465 (2009). [39] G. I. Bell, Science 200, 618 (1978). [40] F. Verde, M. Dogterom, E. Stelzer, E. Karsenti, and S. Leibler, The Journal of Cell Biology 118, 1097 (1992). [41] M. Dogterom and B. Yurke, Science 278, 856 (1997). [42] K. S. B. Ajit P. Joglekar and E. D. Salmon, Curr. Opin. Cell. Biol. 22, 57 (2010). 1 and k3 are much larger than k0 [43] A. P. Joglekar and A. J. Hunt, Biophys. J. 83, 42 (2002). [44] From the parameter values listed in Tab. III, one can see that the rates k0 2 and k0 4, so under low external force and high free GTP-tubulin concentration, the model depicted in Fig. 2(a) is a good approximation of our generalized model depicted in 2(b). [45] To keep as less parameters as possible, in our two-state model, we assume that, the bead only can detach from MT from sub-states 1 and 1′. The reasons are as follows: in assembly state, the experimental data in [27, 41] [or see Fig. 3(b)] imply the growth speed of MT increases 11 with external force (Note, the definition of force direc- tion in [27] is different from that in [41]. In this study, the force direction definition is the same as in [27], i.e., the force is positive if it points to the MT growth direc- tion), so the corresponding force distribution factor δg [see Eq. (18)] should be positive since the growth speed Vg = k1k2L/(k1 + k2) [see Eq. (5)]. Consequently, the probability ¯p1 = k2/(k1 + k2) that MT in state 1 [see Eq. (12)] increases but the probability ¯p2 = k1/(k1 + k2) that MT in state 2 decreases with external force, i.e., as the increase of external force, the assembly MT would more like to stay in state 1. Meanwhile, from the exper- imental data in [27] one sees the detachment rate from assembly state increases with external force. Therefore, the more reasonable choice is to assume that the bead can only detach from state 1 but not state 2. Through similar discussion, one also can see that it is more rea- sonable to assume that, in disassembly state, the bead can only detach from state 1′. At the same time, the experimental data in [30] imply the catastrophe rate decreases with GTP-tubulin concentration [Tubulin] [or see Fig. 4(b)]. Since k1 = k0 1[Tubulin], the probability ¯p1 = k2/(k1 + k2) decreases with [Tubulin], but the prob- ability ¯p2 = k1/(k1 + k2) increases with [Tubulin]. This is why we assume the catastrophe takes place at state 1. 2, δg and k3, k0 4, δs by fitting formulations (5) and (6) to the experi- mental data of growth and shortening speeds respectively [see Fig. 3(b)], and then get k0 c , δc by fitting formulations (16) and (17) to the catastrophe and res- cue rates [see Fig. 3(c)], k0 d, δd are determined by fitting formulations (13) and (14) to the correspond- ing data plotted in Fig. 3(a). Finally all the parameters are slightly adjusted according to the experimental data about the mean lifetime of bead attachment to MT [see formulation (8) and Fig. 3(d)]. All the fitting are done by the nonlinear least square program lsqnonlin in Matlab. In each fitting, We randomly choose 1000 initial values of the parameters and adopt the parameter values which fit the experimental data best. r , δr and k0 a, δa and k0 [46] In our fitting, we firstly get the parameters k0 1, k0 [47] The parameter values listed in Tab. III do not fit well to the GTP-tubulin concentration dependent growth speed of MTs obtained in [30, 32], since the corresponding data in [30, 32] are much different from that in [27]. Without external force, but under similar GTP and tubulin con- centration, the growth speed of MT measured in [30] is about 43 nm/s, and about 20 nm/s in [32], but it is only about 5 nm/s in [27]. In this study, we get the param- eter values mainly based on the data measured in [27]. One reason is that, from our model, if the GTP-tubulin concentration is nonzero, the growth speed of MT will always positive [see formulation (5), Vg (cid:13) 0 if k1 (cid:13) 0]. However, this might not be true for the data in [30, 32]. So, it might be impossible to get a believable fitting pa- rameters for formulation (5) from data in [30, 32] since the data in [30, 32] cannot be described by a formulation like (5). In [41], the velocity-force data are measured un- der tubulin concentration 25µM. However, the zero force growth speed obtained there is about 20 nm/s, which is also much larger than that obtained in [27]. Conse- quently, the theoretical results based on the parameter values listed in Tab. III do not fit well to their data ei- ther. One can verify that the velocity-force data in [41] can be well described by formulation (5) but with param- eters k0 2 = 53.17 s−1 and δg = 4.13. The difference among these experimental data might due to the differences of experimental techniques, methods or materials. 1 = 2.99 s−1µM−1, k0 [48] If the simple model depicted in Fig. 2(a) is employed to 12 describe the dynamic properties of MT, then Fc2 = Fc11 and Fc3 = Fc9. The reason is as follows. At steady state, the probabilities p, ρ that MT in assembly and disassem- bly states are p = kr/(kc + kr) and ρ = kc/(kc + kr) respectively. So the mean growth velocity of MT is V = k1p − k2ρ = (k1kr − k2kc)/(kc + kr). Then the criti- cal force Fc2 satisfies k1(Fc2)kr(Fc2) = k2(Fc2)kc(Fc2). Meanwhile, l∗ s = Vs/Kr = Vs/kr = k2L/kr, so l∗ s (Fc11) is equivalent to k1(Fc11)kr(Fc11) = k2(Fc11)kc(Fc11) which means Fc2 = Fc11. At the same time, p = ρ is equivalent to kc = kr, so Fc3 = Fc9. But for our model as depicted in Fig. 2(b), Fc2 6= Fc11 and Fc3 6= Fc9 [see Figs. 8(b) and 8(c)]. g = Vg/Kc = Vg/kc = k1L/kc and l∗ g(Fc11) = l∗
1204.3518
3
1204
2012-05-29T15:58:18
Physical descriptions of the bacterial nucleoid at large scales, and their biological implications
[ "physics.bio-ph", "cond-mat.soft", "q-bio.GN" ]
Recent experimental and theoretical approaches have attempted to quantify the physical organization (compaction and geometry) of the bacterial chromosome with its complement of proteins (the nucleoid). The genomic DNA exists in a complex and dynamic protein-rich state, which is highly organised at various length scales. This has implications on modulating (when not enabling) the core biological processes of replication, transcription, segregation. We overview the progress in this area, driven in the last few years by new scientific ideas and new interdisciplinary experimental techniques, ranging from high space- and time-resolution microscopy to high-throughput genomics employing sequencing to map different aspects of the nucleoid-related interactome. The aim of this review is to present the wide spectrum of experimental and theoretical findings coherently, from a physics viewpoint. We also discuss some attempts of interpretation that unify different results, highlighting the role that statistical and soft condensed matter physics, and in particular classic and more modern tools from the theory of polymers, plays in describing this system of fundamental biological importance, and pointing to possible directions for future investigation.
physics.bio-ph
physics
Physical descriptions of the bacterial nucleoid at large scales, and their biological implications. Vincenzo G. Benza,1 Bruno Bassetti,2, 3 Kevin D. Dorfman,4 Vittore F. Scolari,5, 6 Krystyna Bromek,7 Pietro Cicuta,7 and Marco Cosentino Lagomarsino5, 6 1Dipartimento di Fisica e Matematica, Universit`a dell'Insubria, Como, Italy 2Universit`a degli Studi di Milano, Dip. Fisica, Via Celoria 16, 20133 Milano, Italy 3I.N.F.N. Milano, Via Celoria 16, 20133 Milano, Italy 4Department of Chemical Engineering and Materials Science, University of Minnesota - Twin Cities, 421 Washington Ave. SE, Minneapolis, Minnesota 55455, United States 5Genomic Physics Group, UMR 7238 CNRS "Microorganism Genomics" 6University Pierre et Marie Curie, 15 rue de l' ´Ecole de M´edecine Paris, France 7Cavendish Laboratory and Nanoscience Centre, University of Cambridge, Cambridge CB3 0HE, U. K. Recent experimental and theoretical approaches have attempted to quantify the physical organization (compaction and geometry) of the bacterial chromosome with its complement of proteins (the nucleoid). The genomic DNA exists in a complex and dynamic protein-rich state, which is highly organised at various length scales. This has implications for modulating (when not directly enabling) the core biolog- ical processes of replication, transcription, segregation. We overview the progress in this area, driven in the last few years by new scientific ideas and new inter- disciplinary experimental techniques, ranging from high space- and time-resolution microscopy to high-throughput genomics employing sequencing to map different as- pects of the nucleoid-related interactome. The aim of this review is to present the wide spectrum of experimental and theoretical findings coherently, from a physics viewpoint. In particular, we highlight the role that statistical and soft condensed matter physics play in describing this system of fundamental biological importance, specifically reviewing classic and more modern tools from the theory of polymers. We also discuss some attempts towards unifying interpretations of the current re- sults, pointing to possible directions for future investigation. I. INTRODUCTION The demarcation line between a "physical" and a "biological" system is rapidly becoming anachronistic, to the point that it possibly hinders research in both disciplines. This is par- ticularly true in the context of gene expression, which is a fundamental process in biology at the molecular level, common to all life on earth: the genetic code is read out ("transcribed") from DNA and written into RNA, which, in the case of messenger RNA, is then translated into proteins. In order for the right number of proteins to be produced in response to changes in environment, internal states, and stimuli, all cells are capable of tightly regulating this sequence of events (Alberts et al., 2008). The physico-chemical implementation of this gene regulation 2 1 0 2 y a M 9 2 ] h p - o i b . s c i s y h p [ 3 v 8 1 5 3 . 4 0 2 1 : v i X r a 2 process takes place primarily through specific DNA-binding interactions of transcription factors, which repress or promote transcription. It is becoming very clear that the genome's conformational properties as a polymer come into play in the processes involved in the regulatory fine-tuning of gene expression, in particular its topological, chemical, geometric and mechanical properties. These properties can influence the activity of the transcription factors and can play a role in the coordination of a large scale cellular response. Evidence for this level of regulation has been found in the different kingdoms of life. We focus here on the efforts to describe the genome's physical state in the case of bacteria, where it is (perhaps surprisingly) less explored than for higher life forms. Understanding this problem requires stringent, quantitative experiments with the standards of physics, together with up-to-date physical models and arguments. Since considerable knowl- edge has been developed in polymer science over the last 50 years, it is very tempting to try to apply this knowledge to understand the energy and time scales involved in maintaining the DNA at the same time compact and accessible inside a cell, and the role of its geometry, structure and compaction in gene regulation. This review discusses challenges that arise in the biological arena, in which mature exper- imental and theoretical tools from the physical sciences might now allow significant progress. The review is primarily aimed at our colleagues in the physical sciences: it should commu- nicate a feel for the main questions and the main challenges, and our understanding that a comprehensive physical approach is possible and necessary at this point. We will not explain the physical models in great technical detail, and we hope this work will also be of interest to biologists who could take the references given here as a starting point and a "compass" in order to evaluate different modelling approaches. Ultimately, we believe that optimal progress in this area will take place in collaboration, and this review might contribute to establishing a common ground and language. BOX 1: Main factors affecting nucleoid organization. Supercoiling by topoisomerases. These enzymes affect the winding of DNA, and play a role in the creation of the observed branched structure of plectonemic loops along the genome. Nucleoid associated proteins (NAPs). These proteins bind to DNA with different spe- cific modes, each responsible for a different aspect of organization, ranging from double-strand bridging to nucleoprotein filament formation. Confinement. Millimeters of genomic DNA are confined within the small cell volume of a bacterium. Molecular crowding. There is a high concentration of macromolecules present in the cytoplasm, This factor could affect nucleoid organization at different levels, for example creating a general effective self-attraction favoring collapse, and strong depletion attraction between large objects, such as ribosomes. Repli- cation and transcription. These are the non-equilibrium processes of DNA and mRNA production that continuously take place in dividing cells. The first affects the sheer amount of genome present and supercoiling and makes both highly nonsteady, the second can cre- ate uneven ribosome concentrations. Both contribute to non-thermal force and displacement fluctuations of the chromosome. 3 II. BACKGROUND Bacteria are single cell organisms of fundamental importance in nature and to mankind. They are "prokaryotes", in the etymological sense that they lack a nucleus as a compartment enclosed by a membrane (in contrast to the "eukaryotic" cells that make up animals and plants). Indeed, all the DNA, RNA, and proteins in the bacterial cell are always present together in the same single compartment. While the classical picture of bacteria (still implicitly adopted by many theoretical and experimental investigators) views them as little more than a bag (or "well-stirred reactor") of proteins and DNA, it is now clear that this tenet is flawed on many different levels. Despite their lack of membrane-bound organelles, bacteria have a high degree of intracellular spatial organization, related to most cellular processes. Perhaps the highest organized structure of the bacterial cell is the genome itself. In most bacteria, the chromosome is a single circular DNA molecule confined by the cell membrane. In E. coli (the best studied bacterial system), the chromosome consists of about 4.7 million base pairs (bp) and has a total length of 1.5 mm (Stavans & Oppenheim, 2006; Trun & Marko, 1998). Bacterial DNA is organized into a specific structure called the nucleoid, which is composed of DNA, RNA and proteins, and occupies a well-defined region of the cell (Sherratt, 2003). The nucleoid is organized by a set of nucleoid associated proteins, or "NAPs" (such as Dps and transcription factors Fis, H-NS, IHF, HU), which can modify the shape of the DNA both at local and global levels (Luijsterburg et al., 2006; Ohniwa et al., 2011). Since the linear size of the genomic DNA is orders of magnitude larger than the length of the cell, it must be packaged and organized in such a way that the resulting structure is compact, while still allowing the primary information processing, genome replication and gene expression (Thanbichler et al., 2005). In fact, recent evidence strongly suggests that changes in chromosome architecture can directly affect the accessibility and activity of the regulatory proteins at the local level as well as at larger scales. In addition, the genome can efficiently control gene expression by changing the way DNA-binding regulatory proteins can access their target sites via the chromosome architecture (Dillon & Dorman, 2010). Historically, the nucleoid has been visualized by transmission electron microscopy (TEM), phase-contrast microscopy and confocal scanning light microscopy; an overview of early microscopic visualization of the nucleoid can be found in (Robinow & Kellenberger, 1994). These studies showed that the nucleoid mostly occupies a separate subcellular region, without being bound by a membrane, and that thin DNA threads extrude from this region. The double-stranded genomic DNA is generally torsionally constrained in bacteria, typically in such a way that its linking number (the number of times each single strand of DNA winds around the other) is lower than in the relaxed configuration. In biological words this is referred to as "negatively supercoiled", and the specific difference in linking number is called superhe- lical density. The name is due to the fact that supercoiled DNA develops nonzero writhe, or "supercoils", i.e. it is wrapped around itself in the manner of a twisted telephone cord. This torsional constraint has important consequences for gene expression, as the mechanical stress carried by a negatively supercoiled configuration can locally weaken the interaction between the two strands. The resultant breaking of base-pair bonds between the two helices is required for the initial steps of transcription, as well as DNA replication and recombination. Since most DNA-binding proteins bind to DNA sensitively to the arrangement of the two strands, and in particular to their average distance, negative supercoiling usually also facilitates protein bind- ing (Dillon & Dorman, 2010). The level of supercoiling is tightly regulated by the cell, and it can be changed by the action of specific enzymes such as topoisomerases and gyrases. The average supercoiling is generally negative in bacterial cells (except for thermophilic bacteria (Confalonieri et al., 1993)). In the case of E. coli, the supercoiling superhelical density is maintained around the value σ = −0.025, where the supercoil density σ is defined as the relative change in linking number due to the winding (or unwinding) of the double helix, where negative values of σ correspond to an unwound helix (Stavans & Oppenheim, 2006). This 4 value is set by the constraining action of DNA-binding proteins and the combined activity of topoisomerases. Deviations larger than 20% to either side of σ are detrimental to cell growth. For example, overwinding causes the formation of DNA structures that impede transcription and replication, while excessive underwinding leads to poor chromosome segregation (Stavans & Oppenheim, 2006). The most important features linking nucleoid organization and cell physiology are summa- rized here, and discussed further in the review. See also BOX 1 on page 2 for a brief description of the main factors affecting nucleoid organization. (i) At large scales, it is seen from in vitro experiments that the nucleoid is composed of topologically unlinked dynamic domain structures; these are due to supercoiling forming plectonemes and toroids (Marko & Cocco, 2003), and stabilized by nucleoid-associated proteins, for example Fis. This combination of effects gives the chromosome the shape of a branched tree-like polymer visible from TEM (Kavenoff & Bowen, 1976; Postow et al., 2004; Skoko et al., 2006). Topological domains are thought to be packaged during replication, otherwise it appears that their boundaries are "fluid" and randomly dis- tributed (Postow et al., 2004; Thanbichler & Shapiro, 2006). (ii) Strong compaction is experimentally observed in vivo, and possibly arises from confine- ment within the cell boundaries, but also from various factors such as molecular crowd- ing (de Vries, 2010) and supercoiling (Stuger et al., 2002). The degree of compaction changes with the cell's growth conditions and in response to specific kinds of stress. The E. coli chromosome, with a linear size of 1.5mm, occupies a volume of 0.1-0.2 µm3 (the bare DNA volume is about a factor 20-30 smaller), which brings the need to study the folding geometry of the nucleoid (Stavans & Oppenheim, 2006); (iii) Supercoiling and nucleoid organization play an important role in gene expression (Dillon & Dorman, 2010). For example, during rapid growth, while several chromosome equiv- alents are present in the cell due to multiple ongoing replication cycles, an increased production of some nucleoid-associated proteins is observed, compacting the chromosome and probably giving rise to specific transcription patterns in a way that is not yet fully characterized. (iv) RNA polymerase, the nucleoid-bound enzyme responsible for gene transcription, is con- centrated into transcription foci or "factories," which can affect the nucleoid structure by bringing together distant loci (Grainger et al., 2005; Jin & Cabrera, 2006). (v) For cells that are not replicating the genome, the position of genetic loci along the chro- mosome is linearly correlated with their position in the cell (Breier & Cozzarelli, 2004; Viollier et al., 2004; Wiggins et al., 2010). The exact subcellular positioning of different loci varies in different bacteria (Toro & Shapiro, 2010). (vi) During replication, daughter chromosomes demix and segregate before cell division. In E. coli, the two arms of the chromosome are segregated in an organized <left-right- left-right> asymmetric fashion from the center of the cell (Nielsen et al., 2007; Toro & Shapiro, 2010; Wang et al., 2006). Replicated chromosomal loci are thought to be immediately recondensed, as they appear to preserve the linear arrangement while they are moved in opposite directions to assume their final position in the incipient daughter cell (Thanbichler et al., 2005; Thanbichler & Shapiro, 2006). In short, the nucleoid's physical organization plays a major role in the most important cellular processes, such as cell division, DNA replication, and gene expression. Explaining these links is a long-standing open problem in microbiology. Today, it can be revisited with new quantitative experiments, including both the "-omics" approaches and more sophisticated 5 and controlled experimental techniques allowing the analysis of nucleoids both in vitro and in vivo. This is paralleled by a renewed interest in the quantitative characterization of bacterial physiology (Scott & Hwa, 2011; Scott et al., 2010; Zaslaver et al., 2009), of which the nucleoid constitutes a fundamental part. Also note that the relative chromosomal positioning and the orientation of genes are subject to natural selection on evolutionary time-scales as shown by the comparison of the genetic maps of different bacteria (Ochman & Groisman, 1994; Rocha, 2008). Consequently, the field is blossoming with a new wave of studies, hypotheses and findings. While many new pieces of evidence are available, the challenge of building coherent pictures for physical nucleoid organization and its role in the different cell processes remains open. Our scope here is to present the main hypotheses and the experimental and modeling tools that have been put forward in order to understand the physical aspects of nucleoid organization, and give the interested reader an ordered account of the known facts. III. MEASUREMENTS We start by reviewing some of the salient experimental findings at the scale of whole- nucleoids, with a particular emphasis on the more recent results. The evidence that we will discuss emerges from a combination of experimental biology, biophysics, high throughput biol- ogy and bioinformatic approaches (see BOX 2 on page 6). Most of the studies reviewed here are based on E. coli or Caulobacter. BOX 2: Several available experimental techniques can probe the nucleoid at large scales. 6 Advanced microscopy together with cell-biological techniques yield information about struc- ture and dynamics of the nucleoid (Right). High-resolution tracking of tagged loci allows measurement of the local viscoelastic properties of the nucleoid. Static configurations of chromosomal loci in fixed cells allow determination of their spatial arrangement within nondi- viding cells and following cell division. Dynamic tracking at long time scales gives information on the chromosome's "choreography" followed over a cell cycle and on the macrodomain subcellular arrangement. Fusions of NAPs with fluorescent proteins also enable evaluation of their localization within the nucleoid. Nucleoids can be purified and manipulated outside of a cell in order to access more directly their biophysical properties (Left). This procedure implies release of confinement and crowding, and dilution of binding proteins. As a result, purified nucleoids are several times larger in radius than the size of a cell. Traditionally, purified nucleoids were imaged by electron microscopy, showing a ramified plectonemic structures. More recently, investigators have concentrated on characterizing their organization as polymers tracking of labelled loci, fluorescence correla- tion spectroscopy(FCS) and probing them mechanically (AFM). In addition, information on nucleoid organization can be obtained from high-throughput experimental datasets (Bottom). Transcriptomics (using sequencing or microarrays) can be used to probe transcriptional response to nucleoid perturbations, such as NAP deletions, changes in the average level of supercoiling, or local release of a plectonemic loop. Another important source of data is protein occupancy, for example by NAPs, and its correlation with gene expression. This information is obtained both by microarray (CHiP-chip) and by next-generation sequencing techniques (ChIP-seq). Recombination has been used to define macrodomains, as compartments within which recombination between chromosomal segments was more likely than recombination with segments laying outside of the compartment. Finally, Chromosome Conformation Capture (3C) techniques probe the spatial vicinity of pairs of chromosomal loci in the (average) cell. They can be combined with sequencing in order to produce high-throughput data sets (Hi-C). Experiments probing the nucleoid at large scalesFluorescently tagged loci and NAPs In VivoHigh-throughput molecular biology techniquesPurified nucleoidslysisLoci TrackingFCS / AFM AFMStructural unitsSubdiffusionFiber formationTranscriptomicsBinding of NAPsChIP-chip, ChIP-seq binding profilesResponse to nucleoid perturbationsDetection of macrodomain organizationtimemean-square displacementcellSpatial arrangement in single cellsHigh-resolution dynamic trackingShort-time subdiffusive behaviorLong-time trajectoriesLinear chromosome organizationSegregation choreographyMacrodomain arrangementNAP clustersOriLROriLROriLRDetection of spatially co-localized lociRandom recombination librariesChromosome Conformation Capture (3C)OriCTerLRrecombination barriersx3C crosslinkinggenomeOriC 7 Strikingly, the intracellular lo- a. Chromosome spatial arrangement and compartmentalization. calization of a given chromosomal locus in a cell is remarkably deterministic, as revealed by fluorescent tagging of chromosomal loci on E. coli and Caulobacter crescentus (Liu et al., 2010; Nielsen et al., 2006; Niki et al., 2000; Viollier et al., 2004; Wang et al., 2006; Wiggins et al., 2010). The series of chromosome segments is localized along the long axis of the cell in the same order as their positions along the chromosome map, with the interlocus distance typically linearly proportional to (arclength) genomic distance. The precise location of individual loci varies in the known bacteria, and might depend on DNA-membrane tethering interactions (Toro & Shapiro, 2010). A recent microscopy study on E. coli (Meile et al., 2011) considered the posi- tioning of the chromosome in the short-axis section of the cell. They found that the Ter region occupies the periphery of the nucleoid, at a larger distance from the longitudinal axis with re- spect to the rest of the chromosome. In newborn or non-replicating cells, the two chromosome arms are spatially arranged such that loci on the left arm of the chromosome lie in one half of the cell and loci on the right arm lie in the opposite half, with the replication origin between them. It is tempting to interpret the resulting sausage-shaped structure as a chromosomal "fiber." In a recent study (Wiggins et al., 2010), the cell-to-cell variability of loci positioning in non-replicating cells was used to estimate an internal elasticity, which, perhaps not surprisingly, appears to be much higher than expected from a naive estimate for a linear polymer. Even more recently (Umbarger et al., 2011), high-throughput Chromosome Conformation Capture (3C) techniques have been used in combination with live-cell fluorescent tagging of loci, in order to determine the global folding architecture of the Caulobacter crescentus swarmer cell genome. These data indicate that a chromosomal fiber exists also in this case, spanning the whole chromosomal ring. Additionally to the linear spatial arrangement of loci according to their chromosomal coordinate, loci of the left chromosomal arm tend to be very proximal to symmetric loci on the right arm. The resulting structure is a compressed ring-like fiber, which, the authors argue, typically takes an eight shape, free to roll around the long cell axis. They also find that the symmetry in the cross-chromosomal arm interactions is determined by the protein-dense attachment point to the cell membrane at the old pole of the cell, triggered by the binding of the ParB protein to its target parS binding sites. Moving the parS pole-anchoring site by 400 Kb along the chromosome (but not the replication origin) determines a sliding of the whole interaction structure, as in a tank crawler. This sliding is slightly asymmetric, suggesting the presence of supplementary attachment points between chromosomal arms or between the chromosome and the cell body. Another important recent discovery, consistent with the mentioned correlation between chro- mosome arms and cell halfs, is the existence of "macrodomains" (Esp´eli & Boccard, 2006; Moulin et al., 2005; Valens et al., 2004) often described as chromosomal isolated compart- ments. The first evidence in this direction (Valens et al., 2004) came from measurements of the recombination frequency between loci, see figure 1A. All else being equal, this should be proportional to the probability that the two chromosomal segments come into contact within the cell. For a well mixed polymer, the recombination frequency should be uniform. How- ever, experiments show a highly non uniform pattern, compatible with a compartmentalized structure with clear boundaries. Four macrodomains of a few hundred Kb in size have been identified, corresponding to regions surrounding the replication origin and terminus, and to two symmetric regions at the edges of the Ter macrodomain, see box and figure 1C,D,E. The remaining "non-structured regions" appear to have different physical properties. Subsequent studies have confirmed the presence of macrodomains and measured their dynamics using flu- orescently labelled loci (Esp´eli & Boccard, 2006; Esp´eli et al., 2008; Lesterlin et al., 2005), see figure 1B,D. While the molecular mechanisms responsible for this level of organization are not yet clear, the same authors also found that the Ter macrodomain appears to be condensed by a single DNA-binding protein (MatP, figure 2) with a small set of specific binding sites (Mercier et al., 2008). Other proteins with macrodomain-specific DNA-binding properties have recently been identified (Cho et al., 2011; Dame et al., 2011; S´anchez-Romero et al., 2010; Tonthat et al., 8 FIG. 1 Chromosome compartmentalization and spatial arrangement of genes. (A) An example of the data that was used for the genetic definition of Right and Ter macrodomains, reprinted by permission from Macmillan Publishers Ltd. from Valens et al. (2004), c(cid:13)2004. The plot shows the frequency of recombination events between the genetic position indicated by the arrow and other positions probed along the genome (The x-axis indicates the base pair coordinates of the chromosome). This should be uniform for a well-mixed polymer. On the contrary, experiments show a highly non-uniform pattern, compatible with a compartmentalized structure with clear boundaries. Dashed vertical lines indicate the delimitation of macrodomains. Colored bars show the extent of the defined macrodomains. (B) Effective diffusion constants obtained from loci tracking experiments, reprinted by permission from John Wiley and Sons Ltd. Esp´eli et al. (2008), c(cid:13)2008 . The plot shows clear differences between the behaviour of loci placed within macrodomains and in non structured regions (NSR1-4,NSL-2). (C) After Valens et al. (2004) (reprinted by permission from Macmillan Publishers Ltd. c(cid:13)2004), graphical representation of the macrodomains and their boundaries within the E coli genome. Structured macrodomains are indicated as colored arcs, black dashed arcs indicate non-structured regions. (D) From Esp´eli et al. (2008) (reprinted by permission from John Wiley and Sons Ltd. c(cid:13)2008), the genetic insulation of the macrodomains correlates with spatial insulation in subcellular territories. The positions of 10 foci were superimposed according to the barycenter of their trajectory during 30 intervals of 10 s at the home position. Foci from tags at OriC, NSR-3, Right-2 and Ter-6 were plotted. (E) Schematic of macrodomains' localization within the cell. Ter is localized at the periphery of the nucleoid, towards the cell membrane Meile et al. (2011) (reprinted by permission from Biomed Central, c(cid:13)2011). 2011), and appear to be conserved in related bacteria. In view of the results from this thread of work, macrodomains might be seen as a process of microphase separation triggered by specific protein binding. Interestingly, as mentioned above macrodomain-like regions also emerge from independent large-scale genomic data (Berger et al., 2010; Mathelier & Carbone, 2010; Scolari et al., 2011). During replication, the chromosomes segregate following a well-defined "choreography," which has been the subject of multiple studies (Berlatzky et al., 2008; Jun & Wright, 2010; Toro & Shapiro, 2010). While segregation is not the main focus here, it is useful to discuss it briefly, as the existing approaches to this problem (both experimental and theoretical) are intimately linked with chromosome organization and will be mentioned in the following. Specifically, spa- tial reorganization of the segregating chromosome arms appears to preserve qualitatively the relationship between loci distance along the chromosome and in the cell. Moreover, reorga- nization ensures that the two replication forks remain in opposite halves of the cell during replication and that the relative orientation of the two reorganized nucleoids in pre-division D% recombinants LeftTer17'29'Ori1'13'26'47'62'81'oriC1 Mbp2 Mbp3 Mbp4 Mbp0E. coliTerLeftRightLeftTerOriRightNSNS1101001000τ (sec)D = MSD/4τ00,00010,00060,00050,00040,00030,0002BCA025 5075 025 5075 E 9 FIG. 2 The protein MatP organizes the Ter macrodomain by specific binding. The circular chromo- some of E. coli divided into 4 macrodomains and 2 non structured zones (see figure 1). (A) From (Mercier et al., 2008), ChIP-chip assays for MatP binding show a specific affinity of this protein with Ter macrodomains, indicating that the macrodomain is condensed by MatP, and suggesting that other macrodomains might be condensed by dedicated proteins with a small set of specific binding sites. (B) From (Mercier et al., 2008), the localization bias of MatP binding sites is conserved among en- terobacteria, (S. typhimurium LT2 with genome size 4683 Kb, E. carotovora SCRI1043 - 5064 Kb), Vibrio (V. cholerae Cl 2 - 961 kbp ), and Pasteurella (Y. pestis KIM 4 - 600 kbp) species. (C) MatP controls DNA compaction in the Ter macrodomain. The histogram represents the proportion of cells with given interfocal distances between foci of two Ter MD markers in WT and matP deletion mutant cells. (D) Fluorescence microscopy images from the two experiments (scale bars 2µm). These results indicate that the effect of MatP on foci colocalization is associated with compaction. (A-D) reprinted from Mercier et al. (2008), c(cid:13)2008, with permission from Elsevier. cells is not random. Quite interestingly, the spatial separation of sister chromosomes is not a continuous process, but has been observed to proceed through "snaps" (Esp´eli et al., 2008; Joshi et al., 2011), suggesting the existence of energetic or entropic barriers for separation, possibly overcome by active processes. There is debate on what is the main driver for chromosome segregation: Entropic repulsion forces due to strong confinement into a box of linear polymers have been proposed to explain this behavior at least in part (Jun & Wright, 2010); in experiments on replicating B. subtilis the chromosome compaction and spatial organization have been hypothesized to result from non- equilibrium dynamics (Berlatzky et al., 2008), rather than from an entropic repulsion process; in Caulobacter crescentus, a contribution from bidirectional extrusion of the newly synthesized DNA from the transcription complex has been postulated to contribute to chromosome seg- regation (Jensen et al., 2001; Toro & Shapiro, 2010). Experiments using inhibition of protein synthesis by chloramphenicol show that this produces nucleoids with a more rounded shape and induces fusion of separate nucleoid bodies (van Helvoort et al., 1996). Upon release from protein synthesis inhibition, the two nucleoids reoccupy the DNA-free cell independently of cell elongation (Helvoort et al., 1998). The control of the segregation mechanism by protein synthe- sis processes could be indirect: ongoing protein synthesis can affect nucleoid compactness and segregation at multiple levels, including the decrease in the amount of enzymes modifying the topology of the DNA or carrying out transcription and DNA replication. The idea that mem- brane protein synthesis activates nucleoid segregation directly has also been proposed (Norris, 1995) (in bacteria the presence of cotranscriptional translation creates a direct physical link between the genome and the membrane (Woldringh & Nanninga, 2006)) Also note that current arguments based on segregation by entropy and confinement might turn out to be inconsistent with interpreting macrodomains as the result of a microphase separation, since a variety of ACBD 10 interactions, and specifically protein-DNA binding processes could play an important role in defining the free energy of the nucleoid. b. Supercoil domains and nucleoid associated proteins. At smaller scales, the circular chromosome of E. coli is organized in plectonemic loops, or "supercoil domains." Those regions are separated by topological barriers formed by nucleoid associated proteins (NAPs) such as Fis and H-NS, see figure 3. These proteins bridge two strands of DNA by binding to both, and prevent the propagation of torsional energy. As a result, they also prevent the spreading of uncontrolled effects on gene expression in case of accidental DNA breaks or mechanical strain, caused for example by advancing replication forks (Postow et al., 2004; Skoko et al., 2006; Stavans & Oppenheim, 2006). Multiple NAPs have been identified, each with its specific binding properties (reviewed in (Luijsterburg et al., 2006)). Besides bridging double strands, they can change the local shape of DNA inducing bends or hinges and form nucleoprotein filaments. Additionally, NAPs often have multiple DNA binding modes which might be dependent on physiological factors. A good example of this behavior is H-NS. Its binding results in DNA -- H-NS -- DNA bridges, but also forms a rigid nucleoprotein filament which could act as zipper in vitro (Amit et al., 2003; Dame et al., 2006; Wiggins et al., 2009). It seems that the nucleoprotein filament formation may be important, as it has been found to be a structure shared by other NAPs including HU (van Noort et al., 2004) and StpA (Lim et al., 2011). Biologically, this could hypotetically provide a mechanism for environmental sensing by NAPs. It is thus possible that our current understanding of NAP binding, being based on the limited number of conditions tested, is incomplete and generalization of a DNA binding property to other solution conditions may be dangerous. FIG. 3 Branched plectonemic conformation of the genomic DNA molecule in E. coli and connection with nucleoid associated proteins. (A) From Postow et al. (2004) (Reprinted by permission from CSHL Press, c(cid:13)2004), electron micrograph of a purified E. coli chromosome. The branched plectonemic structure is visible. From image analysis, Postow and coworkers measured a roughly exponential distribution for the length of supercoiling loops, with the average around 10-15Kb. Note that the extraction and purification process might interfere with many of the properties of the conditions in which the nucleoid structure is naturally found. (B) Electron micrograph of Fis binding to plasmid DNA (from (Schneider et al., 2001)). This dimeric NAP stabilizes plectonemic branches of supercoiled DNA (schematized in the right panel). In particular, Schneider and coworkers measured a Fis to DNA ratio of 1 dimer per 325 bp DNA. (C) From (Schneider et al., 2001), electron micrograph of H- NS binding to plasmid DNA, elongated complexes (polymers) are formed (schematized in the right panel), presumably containing two DNA duplexes. Schneider and coworkers measured a ratio of H-NS to DNA of 1 molecule per 10 bp. (B-C) reprinted from Schneider et al. (2001) by permission of Oxford University Press. 11 The structure of supercoil domains was studied in E. coli by Postow and coworkers, through analysis of the supercoiling-sensitive transcription of more than 300 genes following relaxation by restriction enzymes in vivo, and by electron microscopy (Postow et al., 2004). They con- cluded that domain barriers may vary dynamically and/or across a population, but they follow an exponential length distribution. The average domain size is (cid:39) 10 − 15 Kb, implying the existence of about 200 − 400 domains (Postow et al., 2004; Stavans & Oppenheim, 2006). Branches of the same typical length are visible directly from electron micrographs of purified nucleoids (Kavenoff & Bowen, 1976; Postow et al., 2004). Thus, the genome topology may be visualized as a branched structure with supercoiled domains that are subject to modula- tion by nucleoid-associated proteins, and active processes such as DNA transcription and DNA replication. The factors responsible for establishing the boundaries of supercoiled domains and the deter- minants of domain size and number are still largely unknown. While H-NS and Fis, along with MukB, the analogue of the eukaryotic SMC (Structural Maintenance of Chromosomes) protein, could be involved in stabilizing plectonemic conformations, their precise roles and importance in this context have not yet been established in a definitive fashion, either in vitro or in vivo (Dillon & Dorman, 2010; Grainger et al., 2006; Maurer et al., 2009; Skoko et al., 2006). FIG. 4 Examples of high-throughput data an bioinformatic analyses concerning nucleoid organization (A) Left panel: ChIP-seq binding profile of H-NS, plotted using data from Kahramanoglou et al. (2011); right panel: sketch representing the extended protein occupancy domains detected by Vora et al. (2009). These are polymer-like nucleoprotein complexes, presumably related to nucleoid organi- zation. (B) Chromosomal sectors emerge from the local correlation (plotted here) between expression levels for wild type in log-phase growth and a codon bias index, related to translation pressure (from Mathelier & Carbone (2010), reprinted by permission from Macmillan Publishers Ltd: Molecular Systems Biology, c(cid:13)2010). These sectors correlate well with macrodomain organization. (C) Top panel: effective transcriptional regulatory network (red links) and areas of influence of Fis and H-NS (colored circles) obtained from transcriptomics experiments combining NAP mutants and perturba- tions in supercoiling background (from Marr et al. (2008), Reprinted by permission from BioMed Central, c(cid:13)2008). Bottom panel: clusters of transcriptional response to nucleoid perturbations (left, data from Marr et al. (2008) and similar esperiments) and NAP binding (right) correlate with the organization of macrodomains (outer arcs) and chromosomal segments (inner arcs) of the genome (from Scolari et al. (2011), reproduced by permission of The Royal Society of Chemistry). By contrast, a consistent amount of information on the binding of NAPs in different condi- tions and its effects on the cell state is available from high-thoughput experimental techniques (figure 4.) Nucleoid associated proteins can modulate the nucleoid conformation structure in response to changes in environmental conditions (Luijsterburg et al., 2006). This can re- BOccupancyRNAAC4Kb383000038350003840000384500015 kbAnnotationsRead countsBindingregions 12 sult in large-scale changes in gene expression (Dillon & Dorman, 2010). The local mechani- cal action of NAPs on DNA is often well-characterized by single molecule experiments (Lui- jsterburg et al., 2006), which also lead to the observation of chromatin-like nucleoprotein "fibers" (Kim et al., 2004). Large-scale NAP binding data in specific growth conditions was obtained from high-throughput experiments involving microarrays (CHiP-chip) or sequencing (CHiP- seq) (Grainger et al., 2006, 2007, 2008; Kahramanoglou et al., 2011; Oshima et al., 2006; Wade et al., 2007). Furthermore, transcriptomics studies profiled the changes in gene expression upon different nucleoid perturbations, such as NAP deletion and/or altered supercoiling (Berger et al., 2010; Blot et al., 2006; Bradley et al., 2007; Marr et al., 2008). Many of these data sets show linear regions of dense binding that often correspond to macrodomain boundaries, and associate with global or NAP-dependent transcriptional response and its correlation with codon bias (Mathelier & Carbone, 2010; Scolari et al., 2011). The physical origin of this preference in binding and the gene expression changes at the boundaries of macrodomains is not precisely clear. One possibility is that macrodomain bound- aries might be co-localized by NAP structures. A recent interesting experimental study (Vora et al., 2009) looked at protein occupancy along the genome regardless of protein identity. This work uncovered extended polymer-like domains rich in bound proteins (including NAPs) with an average length of 1.6Kb, associated with transcriptionally silent or transcriptionally enhanced regions (and also with high intrinsic DNA curvature.) When cells enter the stationary phase a radical, global condensation of the nucleoid oc- curs. It is believed that this is a mechanism via which the cell can protect its DNA in harsh conditions (Kim et al., 2004). AFM studies have shown that the structure of the DNA dif- fers at the supercoiling level (Kim et al., 2004) and that action of Dps and CbpA, the NAPs that replace Fis in this growth phase, is quite different. The Dps and CbpA proteins pro- duce compact aggregates (which can protect DNA from degradation by nucleases) rather than binding to distributed sites as Fis does (Cosgriff et al., 2010). Interestingly, the action of Fis counters Dps-induced compaction through a transcriptional response affecting the expression of topoisomerase and gyrase (Ohniwa et al., 2006). c. Compaction by molecular crowding, specific proteins, transcription factories, and confinement. A different question concerns identifying the main factors contributing to nucleoid compaction and organization. The main candidates are macromolecular crowding, electrostatic self-attraction, supercoiling and nucleoid proteins. Macromolecular crowding, or the high concentration of macromolecules present in the cy- toplasm, is generally believed to be an important determinant on the basis of theoretical ar- guments (de Vries, 2010; Odijk, 1998), which predict a possible phase separation mechanism between nucleoid and cytoplasm. Note that the generic term "macromolecular crowding" might include depletion interactions (see below), together with a number of additional effects of en- tropic and energetic nature. DNA condensation by crowders can be observed in vitro under very controlled and well- understood conditions, for example by experiments using dextran or PEG, demonstrating that naked DNA can be directly condensed by these crowders (Estevez-Torres & Baigl, 2011; Huang et al., 2007; Xu & Muller, 2012; Zhang et al., 2009) Similar (but less controlled) behavior, is shown by purified nucleoids (Zimmerman, 2004). At the same time, experimental studies on isolated nucleoids obtained from mutants lacking various NAPs (Zimmerman, 2006a) suggest that the effects of crowding on compaction are substantial and independent of the NAP com- posite background. It has also been suggested that the action of NAPs could be aimed at antagonizing compaction rather than compacting the nucleoid (Zimmerman, 2006b). However, this must be a complicated, combined effect involving forces of different nature. In the absence of crowding and confinement, it is obvious that the radius of gyration of the genome would be smaller if it were organized, e.g. in a branched structure of plectonemic loops stabilized by 13 DNA-bridging NAPs (Postow et al., 2004; Trun & Marko, 1998). As a particular case of crowding effect, it has been proposed that the (attractive) depletion interactions, well known in colloid science, might play an important role in chromosome or- ganization (Marenduzzo et al., 2006a) This force is due to reduction in total solvent excluded volume upon formation of a molecular complex. Depletion interactions are consequential when large molecular assemblies are formed in presence of smaller particles. In fast-growth condi- tions, genome-bound RNA polymerase is localized into a few transcriptionally active foci or "transcription factories," (Cabrera et al., 2009; Jin & Cabrera, 2006), analogous to the eukary- otic case (Marenduzzo et al., 2006a,b). Depletion interactions were suggested to explain the formation of these macromolecular assemblies (Marenduzzo et al., 2006b). Interestingly, the formation of these foci has been associated with the presence of the NAP protein HU (Berger et al., 2010). We add that a similar argument might hold for the local compaction of the sur- roundings of OriC, rich in ribosomal RNA producing regions, in a macrodomain-microphase. In other words, ribosome-rich ribosomal RNA transcripts, attached to the genome through RNA polymerase, could help compact the Ori region. If this should be the case, the compaction properties of this region would change with the number of ribosomes being synthesized, i.e. with growth rate and translation efficiency (Scott & Hwa, 2011) (One can also speculate that, for the same reason, newly-replicated DNA could be sequentially aggregated during replica- tion.) On one hand, experiments show that transcription of ribosomal RNA operons (which are generally located in the Ori macrodomain) is related to the compaction of nucleoids ob- served upon inhibition of translation (Cabrera et al., 2009). On the other hand, these processes must be intersected with the (non entropic) binding of NAPs, given the evidence connecting specific NAPs to the compaction of the Ter macrodomain (MatP) and indirectly to the effect of the HU protein on organization of the Ori macrodomain. A very recent study (Wang et al., 2011a) systematically addressed the chromosomal local- ization and role in spatial organization of the nucleoid of five major NAPs (HU, Fis, IHF, Stpa, H-NS) using fluorescent protein fusions and super-resolution fluorescence microscopy. In the growth conditions tested all the proteins showed scattered distributions in the nucleoid, except for H-NS, which seemed to form (on average) two distinct foci per chromosome copy, bringing together different (even distant) H-NS targets, see figure 5. Thus, H-NS should be added to the list of NAPs with a compacting action on the nucleoid associated with the formation of specific foci. The long-range interactions between H-NS binding targets were validated by 3C, and show no apparent coherence with the macrodomain structure; loci pairs that are near to H-NS targets but are not targets show no 3C signal and wider distributions of subcellular distances (evaluated with microscopy). In general, DNA associated with many NAPs has a much larger surface, which should enhance the depletion interactions. For instance, DNA coated with an H-NS nucleoprotein filament will have a diameter of about 20 nm instead of the 2 nm of naked DNA. H-NS nucleoprotein filament formation could strongly enhance the depletion attractions with respect to other NAPs not forming filaments, and filaments formed at remote locations on the contour of the chromosomal DNA could find each other in a crowded environment. This may explain the results from super-resolution imaging mentioned above (Wang et al., 2011a). We note again that this is one of the counter-intuitive aspects of molecular crowding, since one might think that it would be easier to encounter another filament if there were no "obsta- cles". The confusion can come from thinking about diffusion, which is hindered in a crowded environment due to the higher effective viscosity, versus thermodynamics, where the depletion interactions increase the probability of observing two nucleofilaments in contact at equilibrium. Adopting a simple Kramers-type model for the kinetics, it is not obvious whether the rate of association would go down due to the prefactor or go up due to the lower free energy minimum. Implicit in the entire discussion implicating crowding in aggregation in this cellular context is the idea that the distribution of nucleoprotein filaments is at near-equilibrium even though the cell is dividing rapidly. Finally, depending on the degree of autonomous compaction, the confinement exerted by 14 FIG. 5 Effect of H-NS on chromosome compaction. From Wang et al. (2011a), Reprinted with permission from AAAS. H-NS was shown by Wang and coworkers to form a small set of foci in the cell, which bring together distant H-NS targets. Together with MatP (figure 2), this evidence proves a role of drives of energetic origin in chromosome compaction. (A) Super-resolution fluorescence imaging of fluorescent protein fusion of H-NS showing compact H-NS clusters in the nucleoid. The E. coli cells are shown in the bright-field image (left). The z coordinate of each localization is color-coded (top bar). In comparison, a conventional fluorescence image of the same cells is shown (right). (B) Scattered distribution of HU in the nucleoid. (Left) Bright-field image; (right) super-resolution image. Similar diffuse distributions were observed for Fis, IHF, and StpA. (C) Effect of deleting H-NS on the subcellular distribution of H-NS target genes (hdeA, hchA) versus non-targets (lacZ). The 2D histograms of the relative hdeA, hchA, and lacZ locus positions normalized to the cell dimensions are shown, the genes were fluorescently labelled. The first quartile of the cell plots the (color-coded) probabilities of finding the fluorescent locus in the particular position, the rest of the cell is filled with mirror images to aid the eye (grid size 100 to 200 nm). In each case, 2000 to 5000 gene locus positions were analyzed. H-NS deletion has little effect on lacZ distribution, despite this locus is only a few Kb apart from the closest H-NS binding site. the cell wall might play a relevant role in nucleoid organization and segregation (Jun & Wright, 2010). This is expected to be particularly significant in fast-growth conditions, where the genome needs to be highly accessible for transcription (and thus will not be condensed) and the amount of genome per cell is higher due to overlapping replication rounds (Nielsen et al., 2007). To summarize, the degree of condensation and the geometry of the nucleoid are strongly dependent on the growth phase and growth rate of the bacteria. Such changes may be modeled by means of equilibrium states, slowly evolving in accordance to "external" control parameters (e.g. the concentration of the various NAPs). However, it seems likely that nonequilibrium processes are also important, and relevant aspects of the problem might be lost in attempting to describe the nucleoid condensation process purely in terms of an approach to thermodynamic equilibrium. For example, on the other side of the spectrum in terms of biological complexity, the recently found scale-invariant structure of the human genome (Lieberman-Aiden et al., 2009; Mirny, 2011) was suggested not to be the result of an equilibrium state, but similar in nature to the so-called "crumpled globule," or "fractal globule," (Grosberg et al., 1993) since loci that are near along the genome arclength coordinate are also physically proximal in three- dimensional space. This proximity contrasts with what happens in an (equilibrium) collapsed CAB 15 polymer, where the linear structure is completely mixed in the globule. Let us quote a few experimental findings, some of which are very recent, supporting the role of nonequilibrium processes, specifically for bacteria. The initial conditions, i.e. the choreog- raphy of DNA replication, appear to play a central role in defining the final structure of the nucleoid (Daube et al., 2010). In artificial E. coli strains with two distant origins instead of just one (Wang et al., 2011b), the two origins initiate replication synchronously at the expected separate positions of the genetic loci associated with them. Replication forks move indepen- dently, indicating that replication does not occur in a single replication factory and that the replication machinery is recruited to origins rather than vice versa. Most importantly, in these experiments progression of replication plays a major role in determining the space-time pattern of locus segregation. The large scale structure of the B. subtilis nucleoid (Berlatzky et al., 2008) has been observed at various stages of the replication process. The newborn portions of the chain are compacted and sequentially conveyed towards the poles, resulting in an ordered, spiraling structure. A strong correlation between space coordinate and genomic coordinate is preserved, similar to the linear behavior observed in E. coli (Wiggins et al., 2010). A chore- ography of this sort is found also in Caulobacter (Jensen et al., 2001). Finally, the large scale spiral structure of the nucleoid of Bdellovibrio bacteriovorous (Butan et al., 2011) also sug- gests a metastable steady state, sustained by cooperative motion and/or energy exchanges. It seems difficult to disregard the deterministic replication-segregation dynamics in describing such phenomenology (Breier & Cozzarelli, 2004; Toro & Shapiro, 2010). Tracking studies of fluorescently labeled chromosomal d. Viscoelasticity and structural units. loci have evaluated in vivo dynamic properties of the nucleoid with fairly high time reso- lution, measuring for example the mean-square displacement (MSD) of the loci or the time autocorrelation function (Esp´eli et al., 2008; Meile et al., 2011; Weber et al., 2010a). In general, these measurements give information on the local relaxation time scales of the nu- cleoid and its viscoelastic behavior, see figure 6. On one hand, for large time-scales (Esp´eli et al., 2008), loci mobility correlates well with macrodomain structure. In particular, the MSD saturates at the spatial scale of the macrodomain size. On the other hand, especially for smaller time scales, the mean square displacement of a locus is seen to follow a power law: M SD ≡< ((cid:126)x(s, t + ∆t) − (cid:126)x(s, t))2 >≈ c · (∆t)α (where s is its arclength genomic coordinate) and the exponent α seems to be universally close to 0.4, independent of s (Weber et al., 2010a). Perhaps surprisingly, an extra-chromosomal RK-2 plasmid showed the same behavior, while smaller RNA particles had a higher subdiffusive exponent. It must be mentioned that the localization of RK-2 plasmids appears to be highly regulated (Derman et al., 2008; Kolatka et al., 2008). The underlying viscoelasticity of the bacterial cytoplasm surrounding the nu- cleoid is still poorly understood. Some characterisation has been approached in vivo via FRAP measurements of diffusing GFP (Konopka et al., 2006). The observed anomalous diffusion has been modeled using phenomenological approaches. There are a variety of dynamical models exhibiting anomalous diffusion, such as Langevin equations with time dependent viscosity (also equivalent to fractional Langevin equations), continuous time random walks, and random walks over a fractal object (Burov & Barkai, 2008; Condamin et al., 2008; He et al., 2008). Weber et al. (Weber et al., 2010a,b) compared their data with the results obtained on the basis of the first two approaches, and the fractional Langevin equation gave a more satisfactory agreement. The anomalous diffusion exponent per se does not give quantitative information on the ge- ometry of the nucleoid. Physical measurements, possible on purified nucleoids (Cunha et al., 2001, 2005; Romantsov et al., 2007), can enrich the scenario. In particular, Romantsov and coworkers (Romantsov et al., 2007) have obtained, using fluorescence correlation spectroscopy, the time-dependent coarse-grained density distribution of purified nucleoids. They conclude that the polymer appears to be composed of a set of "structural units" defined by a measur- 16 FIG. 6 Optical techniques in nucleoid dynamic visualization. (A) From Meile et al. (2011) (reprinted by permission from Biomed Central, c(cid:13)2011), using phase contrast for cell visualization, membrane staining (FM 4-64), DNA staining (DAPI), and YFP-ParB for foci dynamic measurement, it is pos- sible to produce "informed maps" of the bacteria by overlaying different micrographs. (B) Movies can be used to to follow foci dynamics. The plot, from Esp´eli et al. (2008) ( reprinted by per- mission from John Wiley and Sons Ltd. c(cid:13)2008), shows a typical trajectory of a fluorescent focus, from which dynamical properties of the nucleoid can be inferred. (C) From Esp´eli et al. (2008), reprinted by permission from John Wiley and Sons Ltd. c(cid:13)2008. The trajectories of fluorescent dots define territories in which genetic loci are placed within the cell. (D) From Weber et al. (2010a) (http://prl.aps.org/abstract/PRL/v104/i23/e238102), reprinted, c(cid:13)2010 by the American Physical Society. A measurement of the ensemble-averaged MSD for live and fixed cells using fluorescent loci. The authors showed that the dynamical exponent of the foci is universal (and close to 0.4). The ob- served anomalous diffusion has been explained phenomenologically as a consequence of the viscoelastic behaviour of the nucleoid. We hypothesize that this could be a consequence of a fractal organization of the nucleoid similar to a "fractal globule", see Discussion. (E, F) Measurements on purified nucleoids, from Romantsov et al. (2007), reprinted with permission from Elsevier, c(cid:13)2007. Using Fluorescence Correlation Spectroscopy, nucleoid volume and amplitude of FCS correlation functions have been measured at varying supercoiling (induced by different concentrations of chloroquine drug). The plots show that the nucleoid reach a maximum volume and minimum FCS correlation at full supercoil- ing relaxation. The FCS correlation signal can be used to deduce the size of structural units of the polymer. able correlation length. The measurements were performed at varying degrees of supercoiling, induced by different concentrations of the chloroquine drug. The size of the structural units was found to vary from 50-100Kb in high (positive or negative) supercoiling to 3Kb at zero supercoiling. The diameter of the purified nucleoid varied from 2.5µm in high supercoiling to 3.5µm in low supercoiling. The authors also estimated the typical diameter of the structural units from the diffusion constant (obtained from the decay of the fluorescence autocorrelation function) and Stokes-Einstein's relation. Perhaps surprisingly, the resulting size of structural units was near 70-80 nm regardless of supercoiling. Thus the emerging picture for the uncon- fined genome is that of a string of ∼ 100 highly dense "beads", each containing ∼ 100 Kuhn lengths (effective independently jointed elementary polymer segments) of DNA each. These values apply in the presence of supercoiling but in the absence of crowding and confinement effects, as for purified nucleoids most of the cytoplasmic (and probably a considerable part of the DNA-binding) proteins are probably diluted away. 17 IV. MODELS We will now review a few modeling approaches put forward in recent years, and point to some more classic work in polymer physics that we believe could be relevant in this context. As the reader will have observed, the wealth of existing experimental results is appealing on one hand, but, on the other hand, does not offer any clear grasp on a small set of relevant ingre- dients necessary for building coherent physical descriptions of the nucleoid. Consequently, the approaches adopted in the literature are highly diverse and heterogenous in terms of premises, methods, ingredients, and points of view. The notable efforts to understand entropic aspects of chromosome segregation (reviewed in (Jun & Wright, 2010)) have revived the studies on entropic forces of single and multiple confined polymers dating back to Edwards and DeGennes (DeGennes, 1979; Edwards & Freed, 1969; Jun & Mulder, 2006; Jung & Ha, 2010). As we have pointed out above, the role of bound proteins on nucleoid entropy is typically disregarded in these arguments. A comprehensive review of the literature on confined DNA in different contexts is provided in Marenduzzo et al. (2010). Considering this approach, there remains the open issue that the observed segregation and compaction times might be too short to be compatible with an entropic process. For linear polymers this time has been evaluated by Molecular Dynamics simulations to scale like the square of the number of monomers, N 2, which, for sufficiently large N , will be smaller than the O(N 3) chain diffusion time (Arnold & Jun, 2007). However, it is not straightforward to use these scaling relations for empirically relevant estimates. Moreover, these estimates will depend on the model used to represent the structure of the nucleoid and its correlation with the replication process, linking this problem to other unanswered questions. Overall, it seems likely that entropy is only part of the story, and active / nonequilibrium processes of different kinds might play a role in chromosome segregation. We will now turn our attention to work concerning nucleoid organization and cellular ar- rangement. Buenemann and Lenz have attempted to understand the linear arrangement of chromosomal loci in terms of a purely geometrical model (Buenemann & Lenz, 2010), where a linear polymer in the form of a string of blobs, is confined within a cylinder and locked at one or at a few loci. This constrained geometry obviously provides an ordering mechanism, as long as the blobs are large enough with respect to the cylinder's diameter. This model makes predictions on the spatial arrangement of the chromosome in mutants of C. crescentus (Viollier et al., 2004) and on the cell-cycle dependent ordering in E. coli. A very recent simulation study (Fritsche et al., 2011) explains the linear ordering observed in E. coli as a product of confinement and entropic repulsion of a string of linearly arranged chromosomal loops. In order to show this, they represent the chromosome as a confined circular self-avoiding chain under the constraint that consecutive loops, identical in size, are distributed along the arclength co- ordinate, while the Ter region does not contain such loops. Their simulations show both linear ordering along the cell axis and Ter region occupying the outer periphery of the nucleoid (Meile et al., 2011), as properties of the equilibrium states. Intriguingly, the same study suggests that this linear-loop ordering could be a consequence of the transcription network organization. To support this point, they simulate a polymer where transcription factor-target pairs are coupled by attractive harmonic interactions, and show that the linear ordering is recovered. It is well- known that transcription factor-target distances have a statistical tendency to be short along the chromosome (Warren & ten Wolde, 2004). Vettorel and coworkers (Vettorel et al., 2009) performed an abstract study motivated by the possible nature of the compartmentalization and the structural units of a generic (eukaryotic or prokaryotic) chromosome forming a crumpled or fractal globule mentioned in section III (figure 7). Specifically, they explore metastable collapsed states of polymers, where the total size scales as N 1/3 as in an equilibrium compacted globule, but a much higher degree of com- partmentalization is present. At odds with the intrinsic disorder of the equilibrium globule, in a fractal globule, for any pair of loci separated by a chain length s, the distance R(s) has 18 the scaling behavior R(s) ≈ s1/3 (Mirny, 2011). In other words, both the equilibrium globule (i.e. the equilibrium collapsed structure emerging from polymer self-attraction or unfavorable entropy of mixing) and the fractal globule have mass fractal dimension Df = 3, but a generic volume (and in particular a blob in the DeGennes sense) inside the equilibrium globule can include non-sequential segments, while in the fractal globule it contains a single sequential seg- ment (Grosberg et al., 1993). This structure can be understood as resulting from a process where condensation sequentially involves larger and larger scales in s, so that the genomic proximity is preserved in a scale invariant fashion. As a consequence, distant portions of the chromosome will occupy different compartments within the globule. In order to obtain a fractal globule in a simulation, a constraint preventing entanglement must exist (Vettorel et al., 2009). To obtain this condition, Vettorel and coworkers considered a semi-dilute or concentrated solution of mutually disentangled (unconcatenated) rings. Also note that because of the constraint preventing entanglement during collapse, a fractal globule is generally larger in size than an equilibrium globule of equal chain length (while for both the total size scales as R ≈ N 1/3). The fractal globule configuration has proven to be relevant for the description of genomic DNA organization in non-dividing eukaryotic (human) cells (Lieberman- Aiden et al., 2009; Mirny, 2011). In particular, it has been found that the human chromosomes display this structure from ∆s ≈ 500Kb to ∆s ≈ 7Mb. This range of sizes might be relevant to bacteria that have genomes spanning a few Mb (also because the lower cutoff might be related to the fiber organization of chromatin, which in bacteria is different). Thus, as the authors speculate, the fractal globule description might be useful for the nucleoid as well. We will further discuss this point in the light of the available data in section V. We have already mentioned above the theoretical work on the role of macromolecular crowd- ing on compaction (de Vries, 2010; Odijk, 1998), and of depletion interactions on loop formation (and possibly on macrodomain organization (Marenduzzo et al., 2006b)). A recent simulation study (Junier et al., 2010) has concentrated on the chromosome-shaping role of transcription factories. Considering a self-avoiding worm-like chain with a fixed hard-core repulsion radius, and short-ranged bridging protein complexes, they show that this system can take "micro- structured" collapsed globule configurations, where bridging complexes cluster and regions of high and low densities of interacting sites coexist in a microphase separated thermodynamic state. More abstract analytical studies could provide a useful context for understanding the role of NAP binding (Diamant & Andelman, 2000; Kantor & Kardar, 1996). Finally, some attention has been devoted recently to the role played by the branched plec- tonemic structure of the nucleoid (Odijk & Ubbink, 1998; Ubbink & Odijk, 1999), which is believed to have relevant implications for transcription (Dillon & Dorman, 2010; Postow et al., 2004). Provided the correct questions are formulated, this topic has the advantage of being placed in a strong framework developed in the past 30 years, building on the classical calcula- tions of Zimm and Stockmayer (Zimm & Stockmayer, 1949) for considering arbitrarily ramified ghost chains as gaussian networks, and obtaining their equilibrium properties (Farago & Kan- tor, 2000; Graessley, 1980; Sommer & Blumen, 1995). For example, some recent work has focused on the induction of loops involving multiple polymer segments (Sumedha & Weigt, 2008). In this framework, reliable estimates for the self-avoiding case can be obtained with Flory-like arguments, taking into account the fact that branched polymers have higher internal topological complexity, which makes the repulsive interactions stronger than between linear or circular chains. The same approach is also possibly relevant for studying dynamic aspects of loci mobility (Dolgushev & Blumen, 2009; Jasch et al., 2003), which can be compared with simulations and more phenomenological models (Weber et al., 2010b). Finally, a number of very refined results based on renormalization group / field theoretical methods have been ob- tained in a more abstract context using the "randomly branched polymers" model (or "lattice animals"), typically defined as the ensemble of all the clusters of connected sites (monomers) on a regular lattice. While this ensemble is probably too general, it is possible that these results have implications for questions related to the structure of the nucleoid. For example, recent 19 FIG. 7 Main universal features of the "fractal globule" model. (A) From Mirny (2011) (with kind permission from Springer Science+Business Media), root-mean squared end-to-end distance R(s) and probability of a contact P (s) as a function of the genomic distance s between the ends of a subchain of N = 32000 monomers, from simulations of a fractal globule and a collapsed (equilibrium) globule. Notice the different scalings in the two situations. All random collapsed configurations of a polymer in two dimensions behave as a fractal globule, as topological entanglement is forbidden. This does not hold in three dimensions, as a large number of knots can be generated. In a kinetically constrained situation where knots are not present, this property is restored, and the resulting metastable config- uration has the property that segments that are close in arclength distance, are also close in space. It has been argued that this configuration might be relevant for eukaryotic chromatin within a range of length and time scales. (B) From Lieberman-Aiden et al. (2009) (reprinted with permission from AAAS), contact probability as a function of genomic distance averaged across the genome (blue) of an eukaryotic cell obtained by the Hi-C technique. It shows a power law scaling between 500 kb and 7 Mb (shaded region) with a slope of 1.08 (fit shown in cyan). This analysis indicates that the eukaryotic DNA might be organized similarly to a fractal globule. work based on Langevin dynamics (Janssen & Stenull, 2011) analyzes their collapsed regime, obtaining a fractal dimension Df = 2.52, intermediate between the swollen chain (Df = 2) and the fully compacted globular state (Df = 3). AB 20 V. DISCUSSION. HYPOTHESES AND PARADOXES CONCERNING NUCLEOID GEOMETRY AND DYNAMICS. We would like to discuss here some speculations on the possible links between the experi- mental and theoretical results discussed above. Let us start with an analogy to the geometrical organization of eukaryotic chromatin, where different geometrical features are observed at different scales, ranging from the known fiber organization up to the arrangement in preferred "chromosome territories" within the nucleus. In this case, experimental techniques such as Chromosome Conformation Capture (3C) and its high-throughput variants (Lieberman-Aiden et al., 2009), or FISH (fluorescence in situ hybridisation, where fluorescent tags are attached to pairs of DNA loci) allow, for instance, to measure the mean absolute distance, R, and the contact probability, Pc, of two genomic loci at arclength distance s. For bacteria, these techniques entail a number of specific difficulties, but the data are starting to be available, as already mentioned (Umbarger et al., 2011). However, the study by Umbarger and coworkers focuses on the large-scale nucleoid 3D architecture, rather than on more detailed properties of the interaction map. Apart from saturation at large genomic distances, R(s) is typically found to increase with s (Mateos-Langerak et al., 2009) with an approximate power-law behavior, R(s) ∼ sν. Here, ν is a scaling exponent, interpretable as 1/Df , which empirically varies with scale of observation and cell type in eukaryotes. If a stretch of length s of a polymer spans a region of size R, in D dimensions, it can occupy a volume V ∼ RD ∼ sνD. Thus, heuristically, one expects that the probability of one end of the stretch meeting the other scales as Pc ∼ 1/V ∼ s−Dν. Experimentally, for loci on the same chromosome, Pc(s) decreases as a power-law Pc(s) ∼ s−1, for a set of length scales in the approximate interval 0.5-7Mb for s (Lieberman-Aiden et al., 2009), which provides evidence for a fractal globule-like organization, ν = 1/3. This is confirmed by FISH data for R(s) on a smaller range of genomic lengths (Mateos-Langerak et al., 2009). However, experimental data on R(s) for chromatin are complex, as the ν exponent is cell- type specific and varies with genomic length, reflecting different degrees and modes of chromatin compaction. At short genomic distances ν is found in the range 0.2−0.6 at short distances, and R(s) reaches a plateau (i.e., ν ∼ 0) at order 10 Mb genomic distances, because of chromosome territories (Barbieri et al., 2011; Mateos-Langerak et al., 2009; Shopland et al., 2006). For the E. coli nucleoid, the only available quantitative data (Wiggins et al., 2010) indicate that R(s) might scale like s for non-replicating chromosomes at scales of 0.3-2 Mb. In principle, a link between nucleoid geometry and the measured nucleoid local dynamics is expected. The anomalous diffusion of chromosomal loci in E .coli has been modeled in terms of fractional Langevin equations (Weber et al., 2010b); such an approach correctly reproduces the temporal behavior of the loci, but disregards the geometry of the structure containing them. Very likely, for any polymer, this structure has a fractal character, and one would like to understand how this influences the motion of the loci. A minimal model would consider a relaxation equation over the fractal. For a purely self-avoiding polymer one has Df ≈ 1.7. In the case of a self-attracting polymer where attraction is screened by self-avoidance, i.e. at the θ point (DeGennes, 1979; Grosberg & Khokhlov, 1994), it is reasonable to assume a mass fractal dimension Df = 2, as for a ghost chain. When the tendency towards compaction increases, one expects that Df will increase accordingly. In order to illustrate this point with an example, we can consider protein structures. Data from the Protein Data Bank (Berman et al., 2000) for 200 proteins with a number of amino acids ranging from N ≈ 100 to N ≈ 10000 give values of Df from 2.3 up to 2.6 (Enright & Leitner, 2005). On a larger scale, high resolution X-ray spectroscopy has resolved the ribosome structure at the atomic level. Such data indicate (Lee, 2006) that the heavier 50S unit is fully compacted, with Df = 3, while the lighter 30S unit has Df = 2.8; it has been argued that the sparser structure of 30S is compatible with a dynamic geometry, as required in the translation process. Folded proteins are generally described as an harmonic network, by the 21 so-called Gaussian Network Model (Reuveni et al., 2010), and we can try to apply a similar reasoning to the nucleoid. Close to the fully collapsed regime of a very long polymer such as a bacterial chromosome, where one expects Df ≈ 3, the harmonic approximation strictly does not apply, because hard-core repulsion acts against chain compression. The total energy can then be written as an harmonic "Rouse" term, describing waves propagating along the chain, plus a self-attraction term and a self-avoidance term. The relaxation dynamics in such conditions (neglecting hydrodynamic interactions) has been studied within a continuum model approach by Pitard and Orland (Pitard & Orland, 1998). They find that the relaxation time τ of the globule scales with the polymer length N as τ ≈ N 5/3. Taking into account that the globule size R scales as R ≈ N 1/3 (i.e. it fills space), this implies that R2 ≈ (τ 3/5)2/3 ≈ (τ )2/5 (since N ≈ τ 3/5). This scaling appears to coincide with the experimental value for the anomalous diffusion exponent α = 0.4 measured by Weber and coworkers (Weber et al., 2010a), previously quoted in the text. In other words, the Rouse subdiffusive dynamics of a collapsing (and thus off-equilibrium) globule follows the same scaling law as the observed local dynamics of nucleoid loci within a range of time scales. While this might be simply a coincidence, it leads us to speculate that the measured dynamic exponent for the mean-square displacement might be the consequence of a fractal-globule-like nature of the nucleoid, at least within a range of length scales. This argument can be recast in more generic terms. Rouse polymer relaxation dynamics was originally explored by De Gennes (DeGennes, 1976), who obtained the scaling relation z = 2+Df , where z ≡ 2/α is the so-called dynamical exponent. At the θ point (DeGennes, 1979; Grosberg & Khokhlov, 1994) where one has Df = 2, the DeGennes's relation gives α = 1/2, the Rouse result for non-interacting chains. As the polymer dimension Df increases, a smaller value of α is to be expected; in the compacted configuration, where Df = 3, the relation gives z = 5, which is the result reported above. DeGennes's work is based on scaling arguments, but is confirmed by field-theoretical methods (Wiese, 1998) for two-body interactions. To our knowledge, in the collapsed regime the relation has been proved only at the level of mean field (Pitard & Orland, 1998) by modeling, in the spirit of a virial expansion, the effective interaction with an attractive two-body term and a repulsive three-body term. In conclusion, if the nucleoid behaves like a fractal globule (i.e. an off-equilibrium polymer collapsing because of self attraction, but where entanglement is prevented by topological con- straints), or more in general if it has fractal dimension Df = 3, from mean field theory one expects the subdiffusion exponent α = 0.4. Conversely, if the DeGennes relation is valid, the experimental result α ≈ 0.4, observed in E. coli and in large plasmids (Weber et al., 2010a), implies that the nucleoid fractal dimension could be Df ≈ 3. To our knowledge the available experimental result that comes closer to a direct measure- ment of Df deals with the mean square displacement of fluorescently tagged replisomes (Reyes- Lamothe et al., 2008). In the approximation of constant replication fork velocity along the mother DNA, replication time in this experiment and genomic arclength distance have equal scaling. Hence, neglecting the global movements due to chromosomal segregation, the repli- some's anomalous diffusion exponent αR measures the effective fractal dimension Df,R of the replicating DNA. Specifically, the scaling M SD(t) ∼ tαR implies αR ∼ 2/Df,R. Obviously, this measured Df,R in principle contains errors due to fluctuations of a stationary background as well as large scale effects associated with coherent restructuring of the nucleoid. The latter pro- cesses are relevant for segregation, but in the initial phases of replication one can assume that they can be disregarded, as we are in the presence of a "weak perturbation" of the stationary (non-replicating) structure. In such a case, the time fluctuations of the fork velocity and the steady state anomalous diffusion of genetic loci would be the main corrections to be taken into account in order to estimate Df from the measured Df,R. Quite interestingly the experimen- tal estimates of the exponent αR from (Reyes-Lamothe et al., 2008) are αR ∼ 0.66 and 0.58 for experiments with 30s and 5min time lapses respectively. If these numbers were confirmed by further measurements and accurate data analysis, they would support the hypothesis that 22 Df ∼ 3 for a range of chromosomal scales, independently on the reasoning presented above, based on the DeGennes scaling relation z = 2 + Df . Clearly, for the nucleoid one expects a structure with a range of fractal dimensions in dif- ferent scale regimes, as suggested by the case of chromatin and the ribosome. The existence of a range of fractal dimensions is also supported by the microscopy results for fluorescent pairs of loci reporting a linear correlation between loci distance in the cell and along the chromo- some (Wiggins et al., 2010). The linear correlation could result from any sequentially ordered segregation process generating a uniform mass density. For example, one might consider a pe- riodic winding of the chromosome (K´ep`es & Valliant, 2003; Mathelier & Carbone, 2010), that could be produced by a segregation choreography of the type observed in B. subtilis as well as in E. coli. However, this needs to be reconciled with the observed subdiffusion. As for the case of chromatin, a hierarchical structure is very logical, since some chromosomal functions, such as transcription, replication and DNA repair, require a certain degree of plasticity, and are not compatible with full compaction at all scales and at all times. Thus, the linear correlation of loci subcellular position and genomic distance can be consistent with Df ≈ 3, but further work is needed to determine the range of spatial scales where these properties apply. The main objection against the fractal-globule as a long-lasting transient state supported by topological constraints is the ubiquitous presence of topoisomerases, DNA enzymes able to cut and paste strands and thus easily resolve these constraints. An interesting theoretical study has focused on the entanglement of tethered rings (Marko, 2009); it is argued that entanglements would "condense", i.e. aggregate in space, in physically relevant situations, which, in presence of enzymes, would facilitate further the resolution of topological constraints. This kind of objection holds for both the eukaryotic and the prokaryotic case. It cannot be excluded that the non-equilibrium constraints leading to the fractal structure observed in Hi-C experiments are caused by something else, or more in general, and more plausibly, that a different physical process than simple topological constraints leads to the observed phenomenology. However, the generic reasoning presented above for connecting Df and α might be robust with respect to these considerations. Other approaches to chromatin organization aim at reproducing the interlocus distance R(s) and the distribution of interacting loci Pc(s) with alternative polymer models, such as a collapsing self-avoiding walk in a solution of organizing proteins which can bind and act as discrete self-attraction points, representing organizing proteins (Nicodemi & Prisco, 2009). The spirit of this kind of study is to go beyond a dominant role of entropy, and take more seriously the "energetic part" of the free energy, and in particular the organizing proteins. The consequences of this hypothesis are explored by analyzing the resulting equilibrium structures for the polymer. Nicodemi and coworkers have recently found that such a model polymer could be close the θ point for empirically relevant protein concentrations, and small variation of the concentration of binding proteins around this state could recapitulate a considerable part of the observed phenomenology of nuclear eukaryotic DNA (Barbieri et al., 2011). Finally, as mentioned above, it is also worthwhile to consider whether equilibrium statistical mechanics is even the proper starting point to understand the structure of the bacteria nucleoid. Cells expend a considerable amount of energy maintaining steady-state, non-equilibrium envi- ronments. A classic example is the membrane potential, whose existence requires an elaborate mechanism for pumping protons through the membrane. Given that the genomic information is arguably the most important part of the cell, containing both the instructions for the current cell and the inheritable information for the next generation, it seems unlikely that bacteria have evolved such that the structure of the nucleoid is resigned to equilibrium. VI. CONCLUSIONS. To conclude, we briefly review some of the main features of the partial and emerging picture of the nucleoid, from the physics viewpoint. All these problems still need to be understood in 23 a quantitative framework for bacterial physiology (Scott & Hwa, 2011), and in particular for varying growth rates (and subcellular compositions) and during adaptation to different growth conditions (Muskhelishvili et al., 2010). A first problem that can be isolated is the explanation of its compaction/condensation prop- erties. Likely mechanisms that can influence (positively or negatively) nucleoid condensation (and more than one can be at play) include (i) supercoiling, bending and looping, in interplay with binding of NAPs, which can cause punctual or polymer-like links building aggregation foci (H-NS, MatP) and stabilize a ramified plectonemic loop structure (Fis), (ii) consequences of molecular crowding, in the form of both phase separation and depletion interactions, and (iii) (nonequilibrium) segregation after replication, which could be induced by different physical processes. A second, related problem is the geometry of the nucleoid, which requires understanding how the subunits are arranged in the cell at different scales and times. It is likely that the organization principles in a given cell state (determined e.g. by growth rate and growth phase) are different at different scales. At the micron scale, the experiments seem to converge towards a linearly-arranged sausage-shaped structure, sometimes wrapped by the Ter region, and the main outstanding questions seem to relate to the physical mechanisms behind segregation and its choreography. Below this scale, the existence of macrodomains and transcription foci still elude a physical explanation, which could be microphase separation stabilized by short-ranged attractions of chemical (organizing proteins such as MatP) or of entropic (ribosome-induced depletion interactions) origin. At an even smaller scale, an organization in blobs or fibers seems to be equally elusive, despite the existence of numerous pieces of evidence for different aspects of NAP binding and plectonemic loop formation and stabilization. Finally, it is important to point out that within the layered information given here there lies more than one unresolved question. For example, if macrodomains are microphases struc- tured by protein binding, then certainly these proteins must play an important role in the configurational entropy of the nucleoid, which is not considered in the arguments concerning entropy-driven chromosome segregation. Also, if the genome is compacted (at least in a range of scales) in a fractal or conventional globule configuration by attractive interactions of en- tropic or energetic origin, this will greatly affect its entropy, and thus its mechanical properties, loci dynamics and the interactions between segregating chromosomes. Equally important, the supercoiling-independent size of structural units measured for purified nucleoids (whose size varies with supercoiling) appears challenging for theoretical explanations. While we are cer- tainly far from a coherent and consistent physical description of the nucleoid, there is a clear abundance of existing data and many ongoing experiments merging quantitative biophysics and high-throughput molecular biology. These emerging results, together with the fragmented but partially successful modeling approaches, make us believe that we might be on the verge of resolving at least some of the existing issues regarding the physics of the bacterial nucleoid. Acknowledgments We are very grateful to Bianca Sclavi for discussions, feedback, and help with the revision of this manuscript, and to Mario Nicodemi, Andrea Parmeggiani, Eric Siggia, Georgi Muskhe- lishvili, Ivan Junier, Zhicheng Long, Avelino Javer, Matteo Osella, Matthew Grant, and Eileen Nugent for useful discussions. We also thank Christine Hardy for kindly allowing us to reprint Fig. 3A. This work was supported by the International Human Frontier Science Program Or- ganization, grant RGY0069/2009-C. 24 References Alberts, B., Bray, D., Lewis, J., Raff, M., Roberts, K., & Watson, J.D. 2008. Molecular Biology of the Cell. 5th edn. Garland. Amit, Roee, Oppenheim, Amos B., & Stavans, Joel. 2003. Increased bending rigidity of single DNA molecules by H-NS, a temperature and osmolarity sensor. Biophys J, 84(4), 2467 -- 2473. Arnold, Axel, & Jun, Suckjoon. 2007. Time scale of entropic segregation of flexible polymers in implications for chromosome segregation in filamentous bacteria. Phys Rev E Stat confinement: Nonlin Soft Matter Phys, 76(3 Pt 1), 031901. Barbieri, M, Chotalia, M, Lavitas, L-M, Fraser, J, Dostie, J., Pombo, A, & Nicodemi, M. 2011. Complexity of Chromatin Folding: the Strings and Binders Switch Model. submitted. Berger, Michael, Farcas, Anca, Geertz, Marcel, Zhelyazkova, Petya, Brix, Klaudia, Travers, Andrew, & Muskhelishvili, Georgi. 2010. Coordination of genomic structure and transcription by the main bacterial nucleoid-associated protein HU. EMBO Rep, 11(1), 59 -- 64. Berlatzky, Idit Anna, Rouvinski, Alex, & Ben-Yehuda, Sigal. 2008. Spatial organization of a replicating bacterial chromosome. Proc Natl Acad Sci U S A, 105(37), 14136 -- 14140. Berman, H. M., Westbrook, J., Feng, Z., Gilliland, G., Bhat, T. N., Weissig, H., Shindyalov, I. N., & Bourne, P. E. 2000. The Protein Data Bank. Nucleic Acids Res, 28(1), 235 -- 242. Blot, N., Mavathur, R., Geertz, M., Travers, A., & Muskhelishvili, G. 2006. Homeostatic regulation of supercoiling sensitivity coordinates transcription of the bacterial genome. Embo Rep., 7(7), 710 -- 715. Bradley, M. D., Beach, M. B., de Koning, A. P. J., Pratt, T. S., & Osuna, R. 2007. Effects of Fis on Escherichia coli gene expression during different growth stages. Microbiology, 153, 2922 -- 2940. Breier, Adam M, & Cozzarelli, Nicholas R. 2004. Linear ordering and dynamic segregation of the bacterial chromosome. Proc Natl Acad Sci U S A, 101(25), 9175 -- 9176. Buenemann, Mathias, & Lenz, Peter. 2010. A geometrical model for DNA organization in bacteria. PLoS One, 5(11), e13806. Burov, S., & Barkai, E. 2008. Fractional Langevin equation: overdamped, underdamped, and critical behaviors. Phys Rev E Stat Nonlin Soft Matter Phys, 78(3 Pt 1), 031112. Butan, Carmen, Hartnell, Lisa M, Fenton, Andrew K, Bliss, Donald, Sockett, R. Elizabeth, Subra- maniam, Sriram, & Milne, Jacqueline L S. 2011. Spiral architecture of the nucleoid in Bdellovibrio bacteriovorus. J Bacteriol, 193(6), 1341 -- 1350. Cabrera, Julio E, Cagliero, Cedric, Quan, Selwyn, Squires, Catherine L, & Jin, Ding Jun. 2009. Active transcription of rRNA operons condenses the nucleoid in Escherichia coli: examining the effect of transcription on nucleoid structure in the absence of transertion. J Bacteriol, 191(13), 4180 -- 4185. Cho, Hongbaek, McManus, Heather R, Dove, Simon L, & Bernhardt, Thomas G. 2011. Nucleoid occlusion factor SlmA is a DNA-activated FtsZ polymerization antagonist. Proc Natl Acad Sci U S A, 108(9), 3773 -- 3778. Condamin, S., Tejedor, V., Voituriez, R., Bnichou, O., & Klafter, J. 2008. Probing microscopic origins of confined subdiffusion by first-passage observables. Proc Natl Acad Sci U S A, 105(15), 5675 -- 5680. Confalonieri, F., Elie, C., Nadal, M., de La Tour, C., Forterre, P., & Duguet, M. 1993. Reverse gyrase: a helicase-like domain and a type I topoisomerase in the same polypeptide. Proc Natl Acad Sci U S A, 90(10), 4753 -- 4757. Cosgriff, Sarah, Chintakayala, Kiran, Chim, Ya Tsz A, Chen, Xinyong, Allen, Stephanie, Lovering, Andrew L, & Grainger, David C. 2010. Dimerization and DNA-dependent aggregation of the Es- cherichia coli nucleoid protein and chaperone CbpA. Mol Microbiol, 77(5), 1289 -- 1300. Cunha, S., Odijk, T., Suleymanoglu, E., & Woldringh, C. L. 2001. Isolation of the Escherichia coli nucleoid. Biochimie, 83(2), 149 -- 54. Cunha, S., Woldringh, C. L., & Odijk, T. 2005. Restricted diffusion of DNA segments within the isolated Escherichia coli nucleoid. J Struct Biol, 150(2), 226 -- 32. Dame, Remus T., Noom, Maarten C., & Wuite, Gijs J L. 2006. Bacterial chromatin organization by 25 H-NS protein unravelled using dual DNA manipulation. Nature, 444(7117), 387 -- 390. Dame, Remus T, Kalmykowa, Olga J, & Grainger, David C. 2011. Chromosomal macrodomains and associated proteins: implications for DNA organization and replication in gram negative bacteria. PLoS Genet, 7(6), e1002123. Daube, Shirley S, Bracha, Dan, Buxboim, Amnon, & Bar-Ziv, Roy H. 2010. Compartmentalization by directional gene expression. Proc Natl Acad Sci U S A, 107(7), 2836 -- 2841. de Vries, Renko. 2010. DNA condensation in bacteria: Interplay between macromolecular crowding and nucleoid proteins. Biochimie, 92(12), 1715 -- 1721. DeGennes, P. G. 1979. Scaling concepts in Polymer Physics. Cornell University Press. DeGennes, P.G. 1976. Dynamics of entangled polymer solutions I. The Rouse model. Macromolecules, 9, 587. Derman, Alan I, Lim-Fong, Grace, & Pogliano, Joe. 2008. Intracellular mobility of plasmid DNA is limited by the ParA family of partitioning systems. Mol Microbiol, 67(5), 935 -- 946. Diamant, H., & Andelman, D. 2000. Binding of molecules to DNA and other semiflexible polymers. Phys Rev E, 61(6 Pt B), 6740 -- 9. Dillon, Shane C, & Dorman, Charles J. 2010. Bacterial nucleoid-associated proteins, nucleoid structure and gene expression. Nat Rev Microbiol, 8(3), 185 -- 195. Dolgushev, Maxim, & Blumen, Alexander. 2009. Dynamics of semiflexible treelike polymeric networks. J Chem Phys, 131(4), 044905. Edwards, S F, & Freed, K F. 1969. The entropy of a confined polymer. I. Journal of Physics A: General Physics, 2(2), 145. Enright, Matthew B, & Leitner, David M. 2005. Mass fractal dimension and the compactness of proteins. Phys Rev E Stat Nonlin Soft Matter Phys, 71(1 Pt 1), 011912. Esp´eli, O, & Boccard, F. 2006. Organization of the Escherichia coli chromosome into macrodomains and its possible functional implications. J Struct Biol, 156(2), 304 -- 10. Esp´eli, O, Mercier, R, & Boccard, F. 2008. DNA dynamics vary according to macrodomain topography in the E. coli chromosome. Mol Microbiol, 68(6), 1418 -- 27. Estevez-Torres, Andre, & Baigl, Damien. 2011. DNA compaction: fundamentals and applications. Soft Matter, 7, 6746 -- 6756. Farago, Oded, & Kantor, Yacov. 2000. Elasticity of Gaussian and nearly Gaussian phantom networks. Phys. Rev. E, 62(5), 6094 -- . Fritsche, Miriam, Li, Songling, Heermann, Dieter W, & Wiggins, Paul A. 2011. A model for Escherichia coli chromosome packaging supports transcription factor-induced DNA domain formation. Nucleic Acids Res, Oct. Graessley, William W. 1980. Linear Viscoelasticity in Gaussian Networks. Macromolecules, 13(2), 372 -- 376. Grainger, D. C., Hurd, D., Harrison, M., Holdstock, J., & Busby, S. J. 2005. Studies of the distribution of Escherichia coli cAMP-receptor protein and RNA polymerase along the E. coli chromosome. Proc Natl Acad Sci U S A, 102(49), 17693 -- 8. Grainger, D. C., Hurd, D., Goldberg, M. D., & Busby, S. J. W. 2006. Association of nucleoid proteins with coding and non-coding segments of the Escherichia coli genome. Nucleic Acids Research, 34(16), 4642 -- 4652. Grainger, D. C., Aiba, H., Hurd, D., Browning, D. F., & Busby, S. J. 2007. Transcription factor distribution in Escherichia coli: studies with FNR protein. Nucleic Acids Res, 35(1), 269 -- 78. Grainger, D. C., Goldberg, M. D., Lee, D. J., & Busby, S. J. 2008. Selective repression by Fis and H-NS at the Escherichia coli dps promoter. Mol Microbiol, 68(6), 1366 -- 77. Grosberg, Rabin, Havlin, & Neer. 1993. Crumpled Globule Model of the Three-Dimensional Structure of DNA. Europhys. Lett., 23 (5), 373 -- 378. Grosberg, A., & Khokhlov, A. 1994. Statistical Physics of Macromolecules. AIP, NY. He, Y., Burov, S., Metzler, R., & Barkai, E. 2008. Random time-scale invariant diffusion and transport 26 coefficients. Phys Rev Lett, 101(5), 058101. Helvoort, J. M. Van, Huls, P. G., Vischer, N. O., & Woldringh, C. L. 1998. Fused nucleoids resegregate faster than cell elongation in Escherichia coli pbpB(Ts) filaments after release from chloramphenicol inhibition. Microbiology, 144 ( Pt 5)(May), 1309 -- 1317. Huang, Wei-Hua, Zinchenko, Anatoly A., Pawlak, Cyril, Chen, Yong, & Baigl, Damien. 2007. Dynamic conformational behavior and molecular interaction discrimination of DNA/binder complexes by single-chain stretching in a microdevice. Chembiochem, 8(15), 1771 -- 1774. Janssen, Hans-Karl, & Stenull, Olaf. 2011. Collapse transition of randomly branched polymers: Renor- malized field theory. Phys Rev E Stat Nonlin Soft Matter Phys, 83(5 Pt 1), 051126. Jasch, F., von Ferber, Ch, & Blumen, A. 2003. Dynamics of randomly branched polymers: con- figuration averages and solvable models. Phys Rev E Stat Nonlin Soft Matter Phys, 68(5 Pt 1), 051106. Jensen, R. B., Wang, S. C., & Shapiro, L. 2001. A moving DNA replication factory in Caulobacter crescentus. EMBO J, 20(17), 4952 -- 4963. Jin, D. J., & Cabrera, J. E. 2006. Coupling the distribution of RNA polymerase to global gene regulation and the dynamic structure of the bacterial nucleoid in Escherichia coli. J Struct Biol, 156(2), 284 -- 91. Joshi, Mohan C, Bourniquel, Aude, Fisher, Jay, Ho, Brian T, Magnan, David, Kleckner, Nancy, & Bates, David. 2011. Escherichia coli sister chromosome separation includes an abrupt global transition with concomitant release of late-splitting intersister snaps. Proc Natl Acad Sci U S A, Jan. Jun, S, & Mulder, B. 2006. Entropy-driven spatial organization of highly confined polymers: lessons for the bacterial chromosome. Proc Natl Acad Sci U S A, 103(33), 12388 -- 93. Jun, Suckjoon, & Wright, Andrew. 2010. Entropy as the driver of chromosome segregation. Nat Rev Micro, 8(8), 600 -- 607. Jung, Y., & Ha, B-Y. 2010. Overlapping two self-avoiding polymers in a closed cylindrical pore: Implications for chromosome segregation in a bacterial cell. Phys Rev E, 82 (5),, 51926 -- 51931. Junier, Ivan, Martin, Olivier, & K´ep`es, Fran¸cois. 2010. Spatial and Topological Organization of DNA Chains Induced by Gene Co-localization. PLoS Comput Biol, 6(2), e1000678. Kahramanoglou, Christina, Seshasayee, Aswin S N., Prieto, Ana I., Ibberson, David, Schmidt, Sabine, Zimmermann, Jurgen, Benes, Vladimir, Fraser, Gillian M., & Luscombe, Nicholas M. 2011. Direct and indirect effects of H-NS and Fis on global gene expression control in Escherichia coli. Nucleic Acids Res, 39(6), 2073 -- 2091. Kantor, & Kardar. 1996. Conformations of randomly linked polymers. Phys Rev E Stat Phys Plasmas Fluids Relat Interdiscip Topics, 54(5), 5263 -- 5267. Kavenoff, R., & Bowen, B. C. 1976. Electron microscopy of membrane-free folded chromosomes from Escherichia coli. Chromosoma, 59(2), 89 -- 101. K´ep`es, F., & Valliant, C. 2003. Transcription-based solenoidal model of chromosome. ComPlexUs, 1, 171 -- 180. Kim, J., Yoshimura, S. H., Hizume, K., Ohniwa, R. L., Ishihama, A., & Takeyasu, K. 2004. Funda- mental structural units of the Escherichia coli nucleoid revealed by atomic force microscopy. Nucleic Acids Res, 32(6), 1982 -- 92. Kolatka, Katarzyna, Witosinska, Monika, Pierechod, Marcin, & Konieczny, Igor. 2008. Bacterial partitioning proteins affect the subcellular location of broad-host-range plasmid RK2. Microbiology, 154(Pt 9), 2847 -- 2856. Konopka, Michael C, Shkel, Irina A, Cayley, Scott, Record, M. Thomas, & Weisshaar, James C. 2006. Crowding and confinement effects on protein diffusion in vivo. J Bacteriol, 188(17), 6115 -- 6123. Lee, Chang-Yong. 2006. Mass fractal dimension of the ribosome and implication of its dynamic characteristics. Phys Rev E Stat Nonlin Soft Matter Phys, 73(4 Pt 1), 042901. Lesterlin, C, Mercier, R, Boccard, F, Barre, FX, & Cornet, F. 2005. Roles for replichores and 27 macrodomains in segregation of the Escherichia coli chromosome. EMBO Rep, 6(6), 557 -- 62. Lieberman-Aiden, Erez, van Berkum, Nynke L, Williams, Louise, Imakaev, Maxim, Ragoczy, To- bias, Telling, Agnes, Amit, Ido, Lajoie, Bryan R, Sabo, Peter J, Dorschner, Michael O, Sandstrom, Richard, Bernstein, Bradley, Bender, M. A., Groudine, Mark, Gnirke, Andreas, Stamatoyannopou- los, John, Mirny, Leonid A, Lander, Eric S, & Dekker, Job. 2009. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science, 326(5950), 289 -- 293. Lim, Ci Ji, Whang, Yixun R., Kenney, Linda J., & Yan, Jie. 2011. Gene silencing H-NS paralogue StpA forms a rigid protein filament along DNA that blocks DNA accessibility. Nucleic Acids Res, Dec. Liu, Xun, Wang, Xindan, Reyes-Lamothe, Rodrigo, & Sherratt, David. 2010. Replication-directed sister chromosome alignment in Escherichia coli. Mol Microbiol, 75(5), 1090 -- 1097. Luijsterburg, M. S., Noom, M. C., Wuite, G. J., & Dame, R. T. 2006. The architectural role of nucleoid-associated proteins in the organization of bacterial chromatin: a molecular perspective. J Struct Biol, 156(2), 262 -- 72. Marenduzzo, D., Micheletti, C., & Orlandini, E. 2010. Biopolymer organization upon confinement. J Phys Condens Matter, 22(28), 283102. Marenduzzo, Davide, Finan, Kieran, & Cook, Peter R. 2006a. The depletion attraction: an underap- preciated force driving cellular organization. J Cell Biol, 175(5), 681 -- 686. Marenduzzo, Davide, Micheletti, Cristian, & Cook, Peter R. 2006b. Entropy-driven genome organi- zation. Biophys J, 90(10), 3712 -- 3721. Marko, J. F., & Cocco, S. 2003. The micromechanics of DNA. Phys. world, March, 37 -- 41. Marko, John F. 2009. Linking topology of tethered polymer rings with applications to chromosome segregation and estimation of the knotting length. Phys Rev E Stat Nonlin Soft Matter Phys, 79(5 Pt 1), 051905. Marr, C, Geertz, M, Hutt, MT, & Muskhelishvili, G. 2008. Dissecting the logical types of network control in gene expression profiles. BMC Syst Biol, 2, 18. Mateos-Langerak, Julio, Bohn, Manfred, de Leeuw, Wim, Giromus, Osdilly, Manders, Erik M M, Verschure, Pernette J, Indemans, Mireille H G, Gierman, Hinco J, Heermann, Dieter W, van Driel, Roel, & Goetze, Sandra. 2009. Spatially confined folding of chromatin in the interphase nucleus. Proc Natl Acad Sci U S A, 106(10), 3812 -- 3817. Mathelier, Anthony, & Carbone, Alessandra. 2010. Chromosomal periodicity and positional networks of genes in Escherichia coli. Mol Syst Biol, 6(May), 366. Maurer, Sebastian, Fritz, Jørgen, & Muskhelishvili, Georgi. 2009. A Systematic In Vitro Study of Nucleoprotein Complexes Formed by Bacterial Nucleoid-Associated Proteins Revealing Novel Types of DNA Organization. Journal of Molecular Biology, 387(5), 1261 -- 1276. Meile, Jean-Christophe, Mercier, Romain, Stouf, Mathieu, Pages, Carine, Bouet, Jean-Yves, & Cornet, Franois. 2011. The terminal region of the E. coli chromosome localises at the periphery of the nucleoid. BMC Microbiol, 11(1), 28. Mercier, Romain, Petit, Marie-Agnes, Schbath, Sophie, Robin, Stephane, Karoui, Meriem El, Boccard, Frederic, & Esp´eli, Olivier. 2008. The MatP/matS site-specific system organizes the terminus region of the E. coli chromosome into a macrodomain. Cell, 135(3), 475 -- 485. Mirny, Leonid A. 2011. The fractal globule as a model of chromatin architecture in the cell. Chromo- some Res, 19(1), 37 -- 51. Moulin, L, Rahmouni, AR, & Boccard, F. 2005. Topological insulators inhibit diffusion of transcription-induced positive supercoils in the chromosome of Escherichia coli. Mol Microbiol, 55(2), 601 -- 10. Muskhelishvili, Georgi, Sobetzko, Patrick, Geertz, Marcel, & Berger, Michael. 2010. General organ- isational principles of the transcriptional regulation system: a tree or a circle? Mol Biosyst, 6(4), 662 -- 676. 28 Nicodemi, Mario, & Prisco, Antonella. 2009. Thermodynamic pathways to genome spatial organization in the cell nucleus. Biophys J, 96(6), 2168 -- 2177. Nielsen, H. J., Ottesen, J. R., Youngren, B., Austin, S. J., & Hansen, F. G. 2006. The Escherichia coli chromosome is organized with the left and right chromosome arms in separate cell halves. Mol Microbiol, 62(2), 331 -- 8. Nielsen, H. J., Youngren, B., Hansen, F. G., & Austin, S. 2007. Dynamics of Escherichia coli chromo- some segregation during multifork replication. J Bacteriol, 189(23), 8660 -- 6. Niki, H., Yamaichi, Y., & Hiraga, S. 2000. Dynamic organization of chromosomal DNA in Escherichia coli. Genes Dev, 14(2), 212 -- 23. Norris, V. 1995. Hypothesis: chromosome separation in Escherichia coli involves autocatalytic gene expression, transertion and membrane-domain formation. Mol Microbiol, 16(6), 1051 -- 1057. Ochman, H., & Groisman, E. A. 1994. The origin and evolution of species differences in Escherichia coli and Salmonella typhimurium. EXS, 69, 479 -- 493. Odijk, T. 1998. Osmotic compaction of supercoiled DNA into a bacterial nucleoid. Biophys Chem, 73(1-2), 23 -- 9. Odijk, Theo, & Ubbink, Job. 1998. Dimensions of a plectonemic DNA supercoil under fairly general perturbations. Physica A: Statistical and Theoretical Physics, 252(1-2), 61 -- 66. Ohniwa, R. L., Morikawa, K., Kim, J., Ohta, T., Ishihama, A., Wada, C., & Takeyasu, K. 2006. Dynamic state of DNA topology is essential for genome condensation in bacteria. EMBO J, 25(23), 5591 -- 602. Ohniwa, Ryosuke L, Ushijima, Yuri, Saito, Shinji, & Morikawa, Kazuya. 2011. Proteomic analyses of nucleoid-associated proteins in Escherichia coli, Pseudomonas aeruginosa, Bacillus subtilis, and Staphylococcus aureus. PLoS One, 6(4), e19172. Oshima, Taku, Ishikawa, Shu, Kurokawa, Ken, Aiba, Hirofumi, & Ogasawara, Naotake. 2006. Es- cherichia coli histone-like protein H-NS preferentially binds to horizontally acquired DNA in asso- ciation with RNA polymerase. DNA Res, 13(4), 141 -- 153. Pitard, E., & Orland, H. 1998. Dynamics of the swelling or collapse of a homopolymer. Europhys. Lett., 41(4), 467 -- 472. Postow, L, Hardy, CD, Arsuaga, J, & Cozzarelli, NR. 2004. Topological domain structure of the Escherichia coli chromosome. Genes Dev, 18(14), 1766 -- 79. Reuveni, Shlomi, Granek, Rony, & Klafter, Joseph. 2010. Anomalies in the vibrational dynamics of proteins are a consequence of fractal-like structure. Proc Natl Acad Sci U S A, 107(31), 13696 -- 13700. Reyes-Lamothe, R, Possoz, C, Danilova, O, & Sherratt, DJ. 2008. Independent positioning and action of Escherichia coli replisomes in live cells. Cell, 133(1), 90 -- 102. Robinow, C., & Kellenberger, E. 1994. The bacterial nucleoid revisited. Microbiol Rev, 58(2), 211 -- 232. Rocha, Eduardo P C. 2008. The organization of the bacterial genome. Annu Rev Genet, 42, 211 -- 233. Romantsov, T., Fishov, I., & Krichevsky, O. 2007. Internal structure and dynamics of isolated Es- cherichia coli nucleoids assessed by fluorescence correlation spectroscopy. Biophys J, 92(8), 2875 -- 84. S´anchez-Romero, Mara Antonia, Busby, Stephen J W, Dyer, Nigel P, Ott, Sascha, Millard, Andrew D, & Grainger, David C. 2010. Dynamic distribution of SeqA protein across the chromosome of Es- cherichia coli K-12. MBio, 1(1). Schneider, R., Lurz, R., Lder, G., Tolksdorf, C., Travers, A., & Muskhelishvili, G. 2001. An archi- tectural role of the Escherichia coli chromatin protein FIS in organising DNA. Nucleic Acids Res, 29(24), 5107 -- 5114. Scolari, Vittore F, Bassetti, Bruno, Sclavi, Bianca, & Lagomarsino, Marco Cosentino. 2011. Gene clusters reflecting macrodomain structure respond to nucleoid perturbations. Mol Biosyst, 7(3), 878 -- 888. Scott, Matthew, & Hwa, Terence. 2011. Bacterial growth laws and their applications. Curr Opin Biotechnol, 22(4), 559 -- 565. Scott, Matthew, Gunderson, Carl W, Mateescu, Eduard M, Zhang, Zhongge, & Hwa, Terence. 2010. 29 Interdependence of cell growth and gene expression: origins and consequences. Science, 330(6007), 1099 -- 1102. Sherratt, David J. 2003. Bacterial Chromosome Dynamics. Science, 301(5634), 780 -- 785. Shopland, Lindsay S, Lynch, Christopher R, Peterson, Kevin A, Thornton, Kathleen, Kepper, Nick, von Hase, Johann, Stein, Stefan, Vincent, Sarah, Molloy, Kelly R, Kreth, Gregor, Cremer, Christoph, Bult, Carol J, & O'Brien, Timothy P. 2006. Folding and organization of a contiguous chromosome region according to the gene distribution pattern in primary genomic sequence. J Cell Biol, 174(1), 27 -- 38. Skoko, Dunja, Yoo, Daniel, Bai, Hua, Schnurr, Bernhard, Yan, Jie, McLeod, Sarah M., Marko, John F., & Johnson, Reid C. 2006. Mechanism of Chromosome Compaction and Looping by the Escherichia coli Nucleoid Protein Fis. Journal of Molecular Biology, 364(4), 777 -- 798. Sommer, J U, & Blumen, A. 1995. On the statistics of generalized Gaussian structures: collapse and random external fields. Journal of Physics A: Mathematical and General, 28(23), 6669. Stavans, Joel, & Oppenheim, Amos. 2006. DNA-protein interactions and bacterial chromosome archi- tecture. Physical Biology, 3(4), R1 -- . Stuger, Rogier, Woldringh, Conrad L., van der Weijden, Coen C., Vischer, Norbert O. E., Bakker, Bar- bara M., van Spanning, Rob J.M., Snoep, Jacky L., & Weterhoff, Hans V. 2002. DNA Supercoiling by Gyrase is Linked to Nucleoid Compaction. Molecular Biology Reports, 29(1), 79 -- 82. Sumedha, & Weigt, Martin. 2008. A thermodynamic model for the agglomeration of DNA-looping proteins. Journal of Statistical Mechanics: Theory and Experiment, 2008(11), P11005. Thanbichler, M., Wang, S. C., & Shapiro, L. 2005. The bacterial nucleoid: a highly organized and dynamic structure. J Cell Biochem, 96(3), 506 -- 521. Thanbichler, Martin, & Shapiro, Lucy. 2006. Chromosome organization and segregation in bacteria. J Struct Biol, 156(2), 292 -- 303. Tonthat, Nam Ky, Arold, Stefan T, Pickering, Brian F, Dyke, Michael W Van, Liang, Shoudan, Lu, Yue, Beuria, Tushar K, Margolin, William, & Schumacher, Maria A. 2011. Molecular mechanism by which the nucleoid occlusion factor, SlmA, keeps cytokinesis in check. EMBO J, 30(1), 154 -- 164. Toro, Esteban, & Shapiro, Lucy. 2010. Bacterial chromosome organization and segregation. Cold Spring Harb Perspect Biol, 2(2), a000349. Trun, N. J., & Marko, J. F. 1998. Architecture of a bacterial chromosome (review). American Society for Microbiology News, 64, 276. Ubbink, J., & Odijk, T. 1999. Electrostatic-undulatory theory of plectonemically supercoiled DNA. Biophysical journal, 76(5), 2502 -- 2519. Umbarger, Mark A., Toro, Esteban, Wright, Matthew A., Porreca, Gregory J., Ba, Davide, Hong, Sun-Hae, Fero, Michael J., Zhu, Lihua J., Marti-Renom, Marc A., McAdams, Harley H., Shapiro, Lucy, Dekker, Job, & Church, George M. 2011. The three-dimensional architecture of a bacterial genome and its alteration by genetic perturbation. Mol Cell, 44(2), 252 -- 264. Valens, M, Penaud, S, Rossignol, M, Cornet, F, & Boccard, F. 2004. Macrodomain organization of the Escherichia coli chromosome. EMBO J, 23(21), 4330 -- 41. van Helvoort, J. M., Kool, J., & Woldringh, C. L. 1996. Chloramphenicol causes fusion of separated nucleoids in Escherichia coli K-12 cells and filaments. J Bacteriol, 178(14), 4289 -- 4293. van Noort, J., Verbrugge, S., Goosen, N., Dekker, C., & Dame, R. T. 2004. Dual architectural roles of HU: formation of flexible hinges and rigid filaments. Proc Natl Acad Sci U S A, 101(18), 6969 -- 74. Vettorel, Thomas, Grosberg, Alexander Y, & Kremer, Kurt. 2009. Statistics of polymer rings in the melt: a numerical simulation study. Phys Biol, 6(2), 025013. Viollier, Patrick H, Thanbichler, Martin, McGrath, Patrick T, West, Lisandra, Meewan, Maliwan, McAdams, Harley H, & Shapiro, Lucy. 2004. Rapid and sequential movement of individual chro- mosomal loci to specific subcellular locations during bacterial DNA replication. Proc Natl Acad Sci U S A, 101(25), 9257 -- 9262. Vora, T., Hottes, A. K., & Tavazoie, S. 2009. Protein occupancy landscape of a bacterial genome. 30 Molecular Cell, 35, 247 -- 253. Wade, J. T., Struhl, K., Busby, S. J., & Grainger, D. C. 2007. Genomic analysis of protein-DNA interactions in bacteria: insights into transcription and chromosome organization. Mol Microbiol, 65(1), 21 -- 6. Wang, Wenqin, Li, Gene-Wei, Chen, Chongyi, Xie, X. Sunney, & Zhuang, Xiaowei. 2011a. Chromo- some organization by a nucleoid-associated protein in live bacteria. Science, 333(6048), 1445 -- 1449. Wang, X., Liu, X., Possoz, C., & Sherratt, D. J. 2006. The two Escherichia coli chromosome arms locate to separate cell halves. Genes Dev, 20(13), 1727 -- 31. Wang, Xindan, Lesterlin, Christian, Reyes-Lamothe, Rodrigo, Ball, Graeme, & Sherratt, David J. 2011b. Replication and segregation of an Escherichia coli chromosome with two replication origins. Proc Natl Acad Sci U S A, 108(26), E243 -- E250. Warren, P. B., & ten Wolde, P. R. 2004. Statistical analysis of the spatial distribution of operons in the transcriptional regulation network of Escherichia coli. J Mol Biol, 342(5), 1379 -- 1390. Weber, Stephanie C., Spakowitz, Andrew J., & Theriot, Julie A. 2010a. Bacterial Chromosomal Loci Move Subdiffusively through a Viscoelastic Cytoplasm. Phys. Rev. Lett., 104(23), 238102. Weber, Stephanie C, Theriot, Julie A, & Spakowitz, Andrew J. 2010b. Subdiffusive motion of a polymer composed of subdiffusive monomers. Phys Rev E Stat Nonlin Soft Matter Phys, 82(1 Pt 1), 011913. Wiese, K.J. 1998. Dynamics of Selfavoiding tethered Membranes I Model A Dynamics (Rouse Model). Eur. Phys. J., B1, 269 -- 272. Wiggins, Paul A., Dame, Remus Th, Noom, Maarten C., & Wuite, Gijs J L. 2009. Protein-mediated molecular bridging: a key mechanism in biopolymer organization. Biophys J, 97(7), 1997 -- 2003. Wiggins, Paul A., Cheveralls, Keith C., Martin, Joshua S., Lintner, Robert, & Kondev, Jan´e. 2010. Strong intranucleoid interactions organize the Escherichia coli chromosome into a nucleoid filament. Proceedings of the National Academy of Sciences, 107(11), 4991 -- 4995. Woldringh, C. L., & Nanninga, N. 2006. Structural and physical aspects of bacterial chromosome segregation. J Struct Biol, 156(2), 273 -- 83. Xu, Weilin, & Muller, Susan J. 2012. Polymer-monovalent salt-induced DNA compaction studied via single-molecule microfluidic trapping. Lab Chip, 12(3), 647 -- 651. Zaslaver, Alon, Kaplan, Shai, Bren, Anat, Jinich, Adrian, Mayo, Avi, Dekel, Erez, Alon, Uri, & Invariant distribution of promoter activities in Escherichia coli. PLoS Itzkovitz, Shalev. 2009. Comput Biol, 5(10), e1000545. Zhang, Ce, Shao, Pei Ge, van Kan, Jeroen A., & van der Maarel, Johan R C. 2009. Macromolecular crowding induced elongation and compaction of single DNA molecules confined in a nanochannel. Proc Natl Acad Sci U S A, 106(39), 16651 -- 16656. Zimm, Bruno H., & Stockmayer, Walter H. 1949. The Dimensions of Chain Molecules Containing Branches and Rings. The Journal of Chemical Physics, 17(12), 1301+. Zimmerman, S. B. 2004. Studies on the compaction of isolated nucleoids from Escherichia coli. J Struct Biol, 147(2), 146 -- 58. Zimmerman, S. B. 2006a. Cooperative transitions of isolated Escherichia coli nucleoids: implications for the nucleoid as a cellular phase. J Struct Biol, 153(2), 160 -- 75. Zimmerman, S. B. 2006b. Shape and compaction of Escherichia coli nucleoids. J Struct Biol, 156(2), 255 -- 61.
1510.00062
1
1510
2015-09-30T22:41:11
The avian tectorial membrane: Why is it tapered?
[ "physics.bio-ph" ]
While the mammalian- and the avian inner ears have well defined tonotopic organizations as well as hair cells specialized for motile and sensing roles, the structural organization of the avian ear is different from its mammalian cochlear counterpart. Presumably this difference stems from the difference in the way motile hair cells function. Short hair cells, whose role is considered analogous to mammalian outer hair cells, presumably depends on their hair bundles, and not motility of their cell body, in providing the motile elements of the cochlear amplifier. This report focuses on the role of the avian tectorial membrane, specifically by addressing the question, "Why is the avian tectorial membrane tapered from the neural to the abneural direction?"
physics.bio-ph
physics
TM short hc BP BM tall hc FIGURE 1. The anatomy of the avian inner ear. TM: the tectorial membrane, BM: the basilar membrane, BP: the basilar papilla. After O. Gleich and G. A. Manley [8]. The avian tectorial membrane: Why is it tapered? Kuni H Iwasa∗,† and Anthony J Ricci∗,∗∗ ∗Otolaryngology & Head and Neck Surgery, Stanford University, Stanford, CA 93405, USA ∗∗Molecular & Cellular Physiology, Stanford University, Stanford, CA 93405, USA †NIDCD, NIH, Bethesda, MD 20892, USA Abstract. While the mammalian- and the avian inner ears have well defined tonotopic organizations as well as hair cells specialized for motile and sensing roles, the structural organization of the avian ear is different from its mammalian cochlear counterpart. Presumably this difference stems from the difference in the way motile hair cells function. Short hair cells, whose role is considered analogous to mammalian outer hair cells, presumably depends on their hair bundles, and not motility of their cell body, in providing the motile elements of the cochlear amplifier. This report focuses on the role of the avian tectorial membrane, specifically by addressing the question, “Why is the avian tectorial membrane tapered from the neural to the abneural direction?” INTRODUCTION The avian and mammalian ears have similarities and differences. The innervation pattern of avian tall hair cells resembles that of mammalian inner hair cells in having primarily afferent innervation, suggesting their role as the primary sensor. In contrast, the innervation pattern of short hair cells in the avian ear is exclusively efferent and resembles that of outer hair cells, suggesting the role of motor element in the cochlear amplifier [3, 7, 11]. Since bending of hair bundles (including tilting motion of the apical surface of hair cells [1]) is the only motile response in short hair cells, such a physiological role requires that the hair bundles of short hair cells exert force between the tectorial membrane and the basilar papilla. There are marked differences. One such difference is the absence of pillar cells and the tunnel of Corti in the avian ear. In addition, the presence of electrical tuning [4] and absence of somatic motility in avian hair cells make them apart. Yet another is the morphology of the tectorial membrane (TM). Is it true that the difference in the motile element of the cochlear amplifier is the reason for the structural difference? Here this issue is examined by focusing upon the morphology of the tectorial membrane. The stiffness of the avian TM would be low compared with the mammalian one, being devoid of collagen [17]. It shows perforations reminiscent of Swiss cheese in electron micrographs due to fixation instead of continuous low density mass as observed in unfixed preparations [19]. Since collagen is largely responsible for the mechanical anisotropy of the mammalian TM [16], its absence would make the material properties of the avian TM very different from the mammalian one, in which stiffness gradient and the significance of the resonance frequency has been documented [5, 6, 12, 14, 18]. The avian TM is tapered toward the abneural side (Fig. 1). The steepness of tapering increases toward the basal end with higher best frequencies [13]. In contrast to the flat and featureless surface of the mammalian tectorial membrane [15], the avian TM has an elaborate structure on the side facing the basilar papilla. A thin veil-like structure descends to the basilar papilla, surrounding each hair cell [9]. A dome like recess above each hair cell and each hair bundle makes contact with the tectorial membrane at the dome [9]. What is the functional significance of these structural features? Like outer hair cells in the mammalian ear, short hair cells are located nearly above the middle of the basilar membrane, the movement of which in the up-down direction [10] would effectively transmit bending stress to the short hair cells’ hair bundles. In contrast, tall hair cells, which functions as the sensor, are located away from the center of the basilar membrane, at a location seemingly unsuitable to sense the motion of the basilar membrane directly. Thus it is likely short hair cells are involved in amplification by a hair bundle active process [2, 22, 23], transmitting the mechanical vibration of the basilar membrane to the hair bundles of tall hair cells. The problem is how that can happen? To address this question, it requires to identify the vibrational mode, which involves the basilar membrane, the basilar papilla, and the TM. Since reported measurements are limited to the basilar membrane with Mössbauer technique [10], we need to assume how other elements could move. Since not much else is known for certain, the models considered herein are largely based on morphological observations and rather speculative. Since hair cell bundles are stimulated by shear, it would be logical to assume that shear motion between the basilar papilla and the TM stimulates the hair bundles of tall hair cells, similar to the mammalian ear. However, there are clear differences. The orientation of hair bundles in the avian ear is not as uniform as in mammalian cochlea. Although the orientation is relatively uniform in the basal region, it forms domain structure in a more apical region [8, 13]. In the following, we examine the mode of motion that is consistent with the bundle alignment in the basal area, then consider a possible mechanism for domain formation in more apical region. SLIDING MODE OF MOTION M m A B FIGURE 2. Two possible modes of motion of the avian inner ear. Two squares represent two parts of the tectorial membrane. The smaller one on top of short hair cells and the larger one on top of tall hair cells. Electron micrographs show that hair bundles are oriented primarily in the direction orthogonal to the longitudinal axis of the basilar papilla, particularly in the basal higher frequency region. This orientation is consistent with the shear between the tectorial membrane and the basilar papilla, analogous to the mammalian ear (Fig. 2 A). However, there are appreciable differences between the two. The avian ear does not have pillar cells or a tunnel of Corti. Thus it may not have well-defined pivoting points for creating shear from the up-down motion of the basilar membrane. In addition, the tectorial membrane (TM) is tapered, making it thicker on top of tall hair cells. The TM is attached to a wall, where it is thickest, nearer to tall hair cells. How can such a morphology be consistent with the assumption that short hair cells serve as the cochlear amplifier? To explain this puzzle, Charles Steele suggested that the TM should have resonance near the best frequency [21]. Indeed, resonance would make it work. However, a question still remains. Why should it tapered toward short hair cells? The avian TM works despite its shape? Is there an advantage of this morphology? 10 K X0 f0 8 6 4 2 0 0.0 0.2 0.4 0.6 0.8 1.0 m/M FIGURE 3. Force amplitude KX0/ f0, normalized to that of the driving external force, applied to the mass M at w = w 0, is plotted against the mass ratio m/M using the reduced drag coefficient h = h /(Mw 0) as a parameter, which ranges from 0.1 to 0.5 incremented by 0.1. The peak is at small values of mass ratio m/M for the small value of h and systematically shifts to larger values as h increases. Since so many factors remain uncharacterized, it would be advantageous to examine the simplest possible model to obtain a conceptual understanding of the issue rather than more complex ones, which would require many more details. For this reason, two-mass models are used here for describing the TM (Fig. 2). One has a smaller mass m, located at the top of short hair cells and another with a larger mass M, located at the top of tall hair cells. The small mass m is connected to the larger mass M with a spring with stiffness k and the larger mass M is connected with the wall by another spring with stiffness K. Let X the displacement of the bigger mass from its equilibrium position, and that of smaller one be x. Let h 2 be viscous drag on mass M and h 1 be negative drag on mass m to mimic the amplifying role of short hair cells (Fig. 2 A). The set of equations is given by, (cid:18)m d2 dt2 − h 1 d2 (cid:18)M dt2 + h 2 dt(cid:19) x − k(X − x) = f0 exp[iw t] + K(cid:19)X − k(X − x) = 0, d d dt (1) (2) where f0 exp[iw t] is a periodic external force of angular frequency w applied to mass m, which is located just above the center of the basilar membrane. Here it is assumed that up-down motion of the basilar membrane results in this external force. Following Steele [21], let w 0 be the resonance frequency of the tectorial membrane, i.e. w 2 0 = K/M = k/m and seek a steady state solution. Under this assumption, the force applied to the mass M is plotted for a special case of h 1 = h 2 = h (Fig. 3). For a given value of M, this force is larger for smaller values of the reduced drag h where the ratio m/M is smaller. Since the reduced drag is defined by h = h /(Mw 0), this parameter is smaller for higher frequency w 0. Thus the plot shows that the force is maximal for smaller ratio m/M at larger frequencies. This result is consistent with the experimental finding that the sharpness of tapering of the TM increases from the apical to the basal end [13, 20]. One can argue that the force relevant to the transduction of tall hair cells would be better related to drag h dX/dt rather than the force KX0. That does not change the optimal mass ratio because drag is represented by h KX0, maximizing at the same mass ratio. The condition for resonance is not very strict because those peaks exist for 1 ≤ w /w 0 ≤ 1.4. JUMPING MODE OF MOTION Domain structure While hair bundles in the basal area of the basilar papilla are aligned perpendicular to the longitudinal axis of the papilla, those in more apical area are aligned in a patchwork fashion, particularly in the middle between neural and abneural ends. How can this pattern be explained? Suppose the motion of the tectorial membrane relative to the basilar papilla is perpendicular to the surface of the basilar papilla (Fig. 2 B), how do hair bundles bend? Each hair bundles is attached to the tectorial membrane at the edge of a dome-like recess [9]. The fibrous veils, which descend from the tectorial membrane and reach the basilar papilla, surround each hair cell at its margins. Some parts of the seemingly extendable veils form triangles with hair 1.8 1.6 K pY0 1.4 f 0 1.2 1.0 0.8 0.0 0.2 0.4 0.6 0.8 1.0 m/ M FIGURE 4. Force amplitude KpY0/ f0, normalized to that of the driving external force, is plotted against the mass ratio m/M, assuming h 1 = −h 2 = h , Kp = kp, w 2 0 = Kp/M, and w /w 0 = 1. The reduced drag coefficient h = h /(Mw 0) is a parameter, which ranges from 0.3 to 0.7 incremented by 0.1. The peak is at small values of mass ratio m/M for small values of h . The peak systematically shifts to larger values as h increases. X0/(X0 + x0) = 0.1 bundles and parts of the basilar papilla [9], which is much denser and therefore expected to be much stiffer. This structure could couple elastic elongation and shortening of the veils, resulting in a bending of the hair bundles. Since bending of a hair bundle results in local displacement of the TM, bending of the neighboring hair bundles in the same direction is energetically favorable if those hair bundles are oriented in the same direction. For this reason, up-down motion of the TM favors domain formation in hair bundle orientations. The size of domains is limited to even out those local deformations of TM to make the average motion in the up-down direction. Equations of motion In this mode of motion, we assume for simplicity, that two masses do not interact directly, but only indirectly being mediated by the basilar papilla, which pivots at the supporting point (Fig. 2 B). Let X0 the distance of the large mass M from the pivoting point, and x0 that of the smaller mass m. Notice that x0 > X0 and for that reason, the torque is more effectively produced by the force applied by the mass m. The equations of motion would be given by, d dt d2 (cid:18)M dt2 + h 1 d2 (cid:18)m dt2 + h 2 + Kp(cid:19)Y − KpX0q = 0, + kp(cid:19) y − kpx0q = 0, d dt (cid:18)I d2 dt2 + h 3 d dt + kpx0 + KpX0 + k (cid:19)q − KpY − kpy = f0 exp[iw t], (3) (4) (5) where I is the moment of inertia of the basilar papilla, q the angle of rotation of the basilar papilla around its pivoting point and h j’s (j=1,2,3) are either positive or negative, depending on whether they have damping or amplifying effect. The quantity y is the perpendicular displacement of the mass m, and Y that of the mass M. Here kp and Kp represent elasticity due to hair bundles and veils. The quantity k represents the elastic elements not associated with hair bundles. Since this is a three body problem, which may not have a periodic solution in general, we examine a special case. Since we are interesting in the mass ratio m/M, we regard the basilar papilla as mediating energy transfer between the smaller mass m and the bigger mass M, letting I = h 3 = k = 0. The longer lever arm of the hair bundles of short hair cells have is advantageous for short hair cells to function as an amplifier. Under this condition, the force applied to the mass M potted against the mass ratio m/M has a maximum for a given h , similar to the sliding mode (Fig. 4). We assumed that short hair cells provide negative damping whereas tall hair cells do not. This requires that active force generation at short hair cells is greater than in tall hair cells. In addition, the source of viscous drag is important. Up-down motion of the tectorial membrane with respect to the basilar papilla requires volume changes in the gap between the tectorial membrane and the basilar papilla, even though changes are small. Since the soft and possibly porous structure of the tectorial membrane may not allow fluid movement at auditory frequencies, the fluid in the gap is likely go through veils surrounds each hair cells, creating an extra viscous drag. In general, such drag would reduce the energy efficiency. However, it could be beneficial in this particular situation if it makes the drag of tall hair cells net positive and that of short hair cell net negative. Dominance of modes The modes of motion described above are not mutually exclusive and are likely coexist. The observation that basal hair bundles are well aligned than more apical bundles and that basal hair bundles have flatter appearance than more apical ones [8, 13] could indicate a relative significance of the sliding mode at higher frequencies. It is likely that in more apical area the jumping mode, which is consistent with domain formation, is dominant. The tapering of the TM is a feature optimizing the sensitivity of tall hair cells in the presence of active motion of the hair bundles of short hair cells. DISCUSSION One could argue that studying the modes of motion in the avian inner ear could be premature: First of all, very few physiological data are available. The only data on the vibration of the basilar membrane was published in 1987 by Gummer at al. using the Mössbauer effect [10], showing traveling wave and that the amplitude is somewhat smaller compared with the mammalian counterpart. The mechanical properties of the basilar papilla nor of the TM have not been measured. Moreover, studying the modes of motion may not be easy. Even in the mammalian cochlea, which has been studied much more extensively, the details are becoming clearer only recently. Nonetheless the avian ear would be an interesting subject for a couple of reasons. One such reason is that the avian basilar papilla has some resemblance to the mammalian ear, such as tonotopic organization and differentiated roles of two types of hair cells, and yet it has much simpler morphology, such as the absence of pillar cells and the tunnel of Corti besides different TM morphologies, possibly suggesting simpler modes of motion. In addition, its cochlear amplifier must depend solely on an active process in hair bundles, which may be based on fast adaptation [2, 22, 23] and possibly, in part, avian prestin [1]. Also the electrical properties of chick hair cells [4] are quite different from mammalian hair cells in having an intrinsic electrical tuning mechanism that likely in part shapes the mechanical components, whether they are the hair bundle or prestin [1]. For this reason, the mechanical motion in the avian ear does not have to be as sharply tuned as in the mammalian one. It is possible that the relatively simple structure of the avian ear could possibly be advantageous for identifying the modes of vibration. Assuming a local TM resonance [21] is attractive in view of its role in the mammalian cochlea [6, 12, 14, 18]. For the jumping mode of motion, it would be reasonable because hair bundles, which have systematic gradient [8, 13], would make a significant contribution to the stiffness. How realistic is the assumption of resonance for the sliding mode of motion? Could Young’s modulus, density, and the dimension of the TM be compatible with the assumption of resonance from the basal end to the apical end? Can a gradient of the mechanical properties of the TM exist? These issues are perhaps experimentally testable. The difference between short and tall hair cells is another requirement for our models models presented here. Specifically, our models require that the net effect of tall hair cells should be drag and that of short hair cells should be negative drag. This requires substantial difference in the magnitude of active force production in the two hair cell types. Indeed, some difference in bending responses, both voltage-induced and not, in these hair cell types has been observed [1]. However, it remains unclear whether or not the differences are large enough to satisfy our requirement. There are other possible modes of motion that are not considered here. The degree of simplification of the present models in describing the TM is obvious. In addition, there are other factors to consider. For example, does the basilar membrane move together with the basilar papilla as a solid body or can they possibly have a difference in the phase as well as amplitude? Photomicrographs show an appreciable space between the basilar papilla with embedded hair cells and the basilar membrane [13]. In addition, these two largely parallel structures appear to have separate anchoring points on the neural side. This feature could possibly allow an external force to drive basilar papilla more than the tectorial membrane. ACKNOWLEDGMENTS Thanks to Drs. Christine Köppl and Goeff Manley of Oldenburg, Germany and Charles Steele of Stanford for discussion. REFERENCES 1. Beurg M, Tan X, Fettiplace R (2013) A prestin motor in chicken auditory hair cells: active force generation in a nonmammalian species. Neuron 79:69–81 2. Choe Y, Magnasco MO, Hudspeth AJ (1998) A model for amplification of hair-bundle motion by cyclical binding of Ca2+ to mechanoelectrical-transduction channel. Proc Natl Acad Sci USA 95:15321–15326 Fischer FP (1992) Quantitative analysis of the innervation of the chicken basilar papilla. Hear Res 61:167–178 3. 4. Fuchs PA, Nagai T, Evans MG (1988) Electrical tuning in hair cells isolated from the chick cochlea. J Neurosci 8:2460–2467 5. Gavara N, Chadwick RS (2012) Determination of the elastic moduli of thin samples and adherent cells using conical atomic 6. Ghaffari R, Aranyosi AJ, Freeman DM (2007) Longitudinally propagating traveling waves of the mammalian tectorial force microscope tips. Nat Nanotechnol 7:733–736 membrane. Proc Natl Acad Sci U S A 104:16510–16515 7. Gleich O (1989) Auditory primary afferents in the starling: correlation of function and morphology. Hear Res 37:255–267 8. Gleich O, Manley GA (2000) The hearing organ of birds and crocodilia. In: Dooling RJ, Fay RR, N PA (eds) Comparative Hearing: Birds and Reptiles, New York: Springer, pp. 70–138 9. Goodyear R, Holley M, Richardson G (1994) Visualisation of domains in the avian tectorial and otolithic membranes with 10. Gummer AW, Smolders JW, Klinke R (1987) Basilar membrane motion in the pigeon measured with the Mössbauer technique. monoclonal antibodies. Hear Res 80:93–104 Hear Res 29:63–92 11. Hirokawa N (1978) The ultrastructure of the basilar papilla of the chick. J Comp Neurol 181:361–374 12. Jones GP, Lukashkina VA, Russell IJ, Elliott SJ, Lukashkin AN (2013) Frequency-dependent properties of the tectorial membrane facilitate energy transmission and amplification in the cochlea. Biophys J 104:1357–1366 13. Köppl C, Gleich O, Schwabedissen G, Siegl E, Manley GA (1998) Fine structure of the basilar papilla of the emu: implications for the evolution of avian hair-cell types. Hear Res 126:99–112 14. Lamb JS, Chadwick RS (2011) Dual traveling waves in an inner ear model with two degrees of freedom. Phys Rev Lett 15. Lim DJ (1986) Functional structure of the organ of corti: a review. Hear Res 22:117–146 16. Masaki K, Gu JW, Ghaffari R, Chan G, Smith RJH, Freeman DM, Aranyosi AJ (2009) Col11a2 deletion reveals the molecular basis for tectorial membrane mechanical anisotropy. Biophys J 96:4717–4724 17. Pickles JO, Brix J, Manley GA (1990) Influence of collagenase on tip links in hair cells of the chick basilar papilla. Hear Res 107:088101 50:139–143 18. Richter CP, Emadi G, Getnick G, Quesnel A, Dallos P (2007) Tectorial membrane stiffness gradients. Biophys J 93:2265–2276 19. Runhaar G (1989) The surface morphology of the avian tectorial membrane. Hear Res 37:179–187 20. Smith CA, Konishi M, Schuff N (1985) Structure of the barn owl’s (tyto alba) inner ear. Hear Res 17:237–247 21. Steele CR (1997) Three-dimensional mechanical modeling of the cochlea. In: Lewis ER, Long GR, Lyon RF, Narins PM, Steele CR, Hecht-Poinar E (eds) Diversity in Auditory Mechanics, Singapore: World Scientific, pp. 455–461 22. Sul B, Iwasa KH (2009) Amplifying effect of a release mechanism for fast adaptation in the hair bundle. J Acoust Soc Am 23. Tinevez JY, Jülicher F, Martin P (2007) Unifying the various incarnations of active hair-bundle motility by the vertebrate hair 126:4–6 cell. Biophys J 93:4053–4067
1810.09919
1
1810
2018-10-23T15:40:10
Measuring Bilayer Surface Energy and Curvature in Asymmetric Droplet Interface Bilayers
[ "physics.bio-ph" ]
For the past decade, droplet interface bilayers (DIBs) have had an increased prevalence in biomolecular and biophysical literature. However, much of the underlying physics of these platforms are poorly characterized. To further the understanding of these structures, a study of lipid membrane tension on DIB membranes is measured by analysing the equilibrium shape of asymmetric DIBs. To this end, the morphology of DIBs is explored for the first time using confocal laser scanning fluorescence microscopy (CLSM). The experimental results confirm that, in accordance with theory, the bilayer interface of a volume asymmetric DIB is curved toward the smaller droplet and a lipid asymmetric DIB is curved toward the droplet with the higher monolayer surface tension. Moreover, the DIB shape can be exploited to measure complex bilayer surface energies. In this study, the bilayer surface energy of DIBs composed of lipid mixtures of phosphatidylgylcerol (PG) and phosphatidylcholine (PC) are shown to increase linearly with PG concentrations up to 25%. The assumption that DIB bilayer area can be geometrically approximated as a spherical cap base is also tested, and it is discovered that the bilayer curvature is negligible for most practical symmetric or asymmetric DIB systems with respect to bilayer area.
physics.bio-ph
physics
Measuring Bilayer Surface Energy and Curvature in Asymmetric Droplet Interface Bilayers Nathan E. Barlow,a,b Halim Kusumaatmaja,c Ali Salehi-Reyhani,a,b,e Nick Brooks,a,b Laura M. C. Barter,a,b Anthony J. Flemming,d Oscar Cesa,b,e,* a Department of Chemistry, Imperial College London, Exhibition Road, London, SW7 2AZ, UK. *[email protected] b Institute of Chemical Biology, Imperial College London, Exhibition Road, London, SW7 2AZ, UK c Department of Physics, University of Durham, South Road, DH1 3LE, UK d Syngenta, Jealott's Hill International Research Centre, Bracknell, Berkshire, RG42 6EY, UK e FABRICELL, Imperial College London, London SW7 2AZ Abstract: For the past decade, droplet interface bilayers (DIBs) have had an increased prevalence in biomolecular and biophysical literature. However, much of the underlying physics of these platforms are poorly characterized. To further the understanding of these structures, a study of lipid membrane tension on DIB membranes is measured by analysing the equilibrium shape of asymmetric DIBs. To this end, the morphology of DIBs is explored for the first time using confocal laser scanning fluorescence microscopy (CLSM). The experimental results confirm that, in accordance with theory, the bilayer interface of a volume asymmetric DIB is curved toward the smaller droplet and a lipid asymmetric DIB is curved toward the droplet with the higher monolayer surface tension. Moreover, the DIB shape can be exploited to measure complex bilayer surface energies. In this study, the bilayer surface energy of DIBs composed of lipid mixtures of phosphatidylgylcerol (PG) and phosphatidylcholine (PC) are shown to increase linearly with PG concentrations up to 25%. The assumption that DIB bilayer area can be geometrically approximated as a spherical cap base is also tested, and it is discovered that the bilayer curvature is negligible for most practical symmetric or asymmetric DIB systems with respect to bilayer area. Keywords: Droplet Interface Bilayers, Surface Energy, Membrane Asymmetry. 1. Introduction Droplet interface bilayers (DIBs)(1) (2) have typically been used to measure membrane bilayer characteristics such as permeability, membrane protein interactions or electrical behaviour. Additionally, interesting DIB morphological behaviour has been studied such as bilayer area modulation by mechanical oscillation,(3) membrane capacitance,(4) or evaporation from the aqueous phase.(5) On a practical level, DIBs have been shown to be particularly useful as they allow for the production of asymmetric membranes,(6) (7) where understanding membrane asymmetry is of high value as it is known to offset transmembrane potential,(8) (9) affect membrane bending rigidity,(10) control membrane protein conformation,(11) as well as membrane permeability.(12) (13) (14) Surface energy 𝛾 in bio-membranes is important to quantify as it is known to affect cellular functions such as membrane fusion, ion binding,(15) and integral protein activity.(16) However, measuring surface tension in lipid DIB membranes is challenging, and currently, the only accepted measurement method is made via direct visualization of the surface morphology using bright field microscopy, which along with known monolayer surface tensions, can be used to infer bilayer tension. This technique, established by many groups (17) (18) (19) (8) (20) (21) outputs a bilayer surface tension on the order of ~1 𝑚𝑁 𝑚−1 for DIBs made with lipids such as 1,2-diphytanoyl-sn-glycero-3-phosphatidylcholine (DPhPC). For a frame of reference, note that according to Kwok and Evans, the lysis tension for lecithin vesicles was found to be on the order of 3 to 4 𝑚𝑁 𝑚−1.(22) Notably, this high surface tension value (close to known rupture tensions) deviates from that of the vesicular analogue membrane tension, which is often assumed to be negligible.(23) For example, from optical techniques (laser tweezer traps), membrane tethers have been measured to have a surface tension of 3 × 10−3 𝑚𝑁 𝑚−1.(24) Vesicle fluctuation analysis can also be used to estimate vesicle membrane tension as low as of 10−3 𝑚𝑁 𝑚−1.(25) The surface tension of neutrophils have been calculated to be 0.03 𝑚𝑁 𝑚−1,(26) measured with micropipette aspiration,(27) (28) or the micropipette interfacial area-expansion method.(29) The lipid 1,2-dioleoyl-sn-glycero-3-phospho-(1'-rac-glycerol) (DOPG) was chosen as it is documented that the uncharged PC lipids reduce the surface tension of pulmonary surfactants that contain a large amount of the charged PG lipid.(30) (31) Certainly, as DIB membranes are high energy systems relative to their vesicular counterparts, measuring membrane tension in DIBs is unfortunately limited by stability. Furthermore, as DIB membrane oscillation cannot be captured optically, and as micropipette aspiration of DIBs would not affect any change in surface tension, it appears that morphological measurements are the only practical option to calculating surface tension. However, though it has been shown that symmetric DIB bilayer surface energies can be estimated using shape information from bright field images, bright field microscopy lacks the ability to capture precise information about membrane curvature due to lipid asymmetry, which can significantly affect the surface energy calculation. In this study, for the first time using confocal scanning fluorescence microscopy (CSLM) it is shown that, in DIBs composed of droplets of different volumes, there exists curvature in asymmetric bilayers of lipids with differing surface properties. Confocal microscopy was required as it provided a higher resolution image for the bilayer shape which isn't obscured by extraneous light from above and below the midplane of the DIB. This shape information can be applied to the calculation of membrane tension in accordance with a force balance i.e. Neumann's triangle(32) (the sine rule). Additionally, a free energy model is applied that describes the curvature behaviour with respect to lipid asymmetry and droplet volume difference. 2. Materials and Procedures 2.1 Lipid preparation and (DOPG) The lipids 1,2-diphytanoyl-sn-glycero-3-phosphatidylcholine (DPhPC), 1,2-dioleoyl-sn-glycero-3- phospho-(1'-rac-glycerol) 1-Oleoyl-2-[12-[(7-nitro-2-1,3-benzoxadiazol-4- yl)amino]dodecanoyl]-sn-Glycero-3-Phosphocholine (NDB PC) were purchased from Avanti Polar Lipids. Samples were prepared with 10 mg of solid lipid mixtures suspended in chloroform. The suspension was evaporated where a film was deposited on the vial surface. The film was desiccated for 30 minutes and re-suspended in a 0.25 𝑀 phosphate buffer solution at 7.4 pH. The samples were freeze- thaw cycled in liquid nitrogen and in a water bath at 60C, repeated 5 times each. The frozen samples were stored at -20C until used. Before use, the samples were thawed and diluted to 5 𝑚𝑔 𝑚𝐿−1 and extruded 11 times through 100 𝑛𝑚 Avanti PC membrane filters. For confocal microscopy, the fluorescent lipid NBD-PC was similarly deposited on a vial surface where it was suspended in the previously extruded lipid solutions to a molar concentration of 0.1%. It was assumed that the low concentration of NBD-PC does not appreciably affect the surface properties of the lipid monolayer or bilayer. 2.2 DIB formation DIBs were formed by pipetting lipid-in aqueous emulsions into acrylic wells filled with hexadecane. Acrylic wells are chosen for DIB manifolds as the droplet wettability was reduced and has a refractive index of 1.49 (33) which was not dissimilar to the supplier reported value for hexadecane at 1.43. DIBs were formed at 5 𝑚𝑔 𝑚𝐿−1 lipid concentration. The dynamics of monolayer formation have already been established,(34) which show that lipid-in DIBs require a short incubation period on the order of minutes as single droplets in hexadecane before they are pushed together with a needle to form interfaces. There is also a period on the order of minutes where the DIBs 'zip-up' to form a bilayer at equilibrium, curvature measurements are taken at this equilibrium state. Note it is assumed that negligible amounts of residual oil may be trapped in the bilayer, as previous experiments have shown that this DIB system can accommodate the mechanosensitive membrane protein MscL and retain functionality.(35) 2.3 Confocal Microscopy A Leica TCS SP5 confocal fluorescent microscope was used with a 10x objective set with an 84.5 𝜇𝑚 pinhole (1 airy unit). The field of view was set to 775x775 𝜇𝑚 (512x512 pixels) and the samples were acquired at a frequency of 400 𝐻𝑧 with 8 line averages. The excitation was achieved with three wavelengths of 458, 476 and 488 𝑛𝑚 and absorbance was set at between 510 and 550 𝑛𝑚. The images used to fit the model were acquired in the midplane of the droplets. During data collection, focal planes slightly above and below were viewed to confirm that the image was indeed acquired from the midplane. 2.4 Pendant Drop Measurements and Drop Shape Analysis (DSA) It has been shown by Lee et al that ionic screening of PC/PG vesicles is required to allow the lipids to coat an air/water monolayer surface.(36) In order to confirm that the lipids have absorbed on the monolayer DSA measurements can also be performed on the lipid solutions in hexadecane. Surface energies of mixtures of lipids were calculated with a pendant tensiometer (Krüss) by drop shape analysis (DSA). The lipids used for making DIBs were formed into aqueous droplets and were immersed in hexadecane from a flat needle 0.52 𝑚𝑚 in diameter. The Worthington number 𝑊𝑜(37) 𝑊𝑜 = ∆𝜌𝑔𝑉𝑑 𝜋𝛾𝐷𝑛 (2.1) is a dimensionless number which measures the ratio of gravitational to surface forces and is an analogue of the well-known Bond number 𝐵𝑜 = ∆𝜌𝑔𝐿 𝛾 in bubble systems, where L is the characteristic length.(38) It is often used to estimate the accuracy of the DSA technique, where a measurement is considered accurate at around 𝑊𝑜~1 and inaccurate for 𝑊𝑜 << 1.(39) Thus the volume of the droplet must be maximized in order to attain accurate surface energy measurements. For this system, the density difference between water and hexadecane is ∆𝜌 = 230 𝑘𝑔 𝑚−3, acceleration of gravity 𝑔 = 9.8 𝑚 𝑠−2, droplet volume 𝑉𝑑 = 0.1 − 0.5[× 10−9 ]𝑚3, needle diameter 𝐷𝑛 = 5.2 × 10−4 𝑚 and surface tension 𝛾 is on the order of 10−3 𝐽 𝑚−2.(39) Due to the low adhesion energy of DPhPC and DOPG monolayers, pendent drop measurements become troublesome as the gravitational potential energy of large droplets overwhelms the pendant droplet adhesion and falls off the needle before equilibrium is reached. This limits the possible range of experimental values of the 𝑊𝑜 to between 0.26 to 0.99. 3. Results and Discussion 3.1 Model Equation and Geometry It is shown that there may exist a bend in the bilayer between the droplets that form a DIB.(18) Under the assumption that the DIB retains axial symmetry, as demonstrated in Figure 3.1, the bilayer bend of radius 𝑟𝑏 can be modelled as a section of a spherical cap of height ℎ𝑏 and the droplets themselves can be modelled as intersecting spheres of radius 𝑟1 and 𝑟2 truncated at height ℎ1 and ℎ2 with spherical cap base radius 𝑎. Note that there is an important distinction between the effective bilayer curvature ( 1 𝑟𝑏 ) in a DIB and the lipid spontaneous curvature 𝑐0.(40) The curvature in the DIB is a non-local description of the droplet macrostructure. In this work, the lipids DOPG and DPhPC are used as they form stable bilayers(41) with differing surface energies. Though both DPhPC(42) and DOPG(43) have negative spontaneous curvature, planar and positive curvature can occur in DIB membranes. Figure 3.1: Image of (a) cartoon depicting asymmetric DIB with one droplet of radius 𝑟1 formed from DPhPC vesicles and one droplet formed from DPhPC doped with DOPG lipids of radius 𝑟2. A diagram (b) of an asymmetric DIB which exhibits curvature in the bilayer with surface energy 𝛾𝑏 of radius 𝑟𝑏 and with angle relative to the x-axis, 𝛩𝑏, which balances the surface energies 𝛾1 and 𝛾2with contact angles 𝛩1and 𝛩2. ℎ𝑏 is the spherical cap height of the bilayer. By setting the bilayer concavity toward droplet 2 (see Figure 3.1), due to a surface tension force balance(32) the equations 𝛾1 cos 𝛩1 + 𝛾2 cos 𝛩2 = 𝛾𝑏 cos 𝛩𝑏 𝛾1 sin 𝛩1 = 𝛾2 sin 𝛩2 + 𝛾𝑏 sin 𝛩𝑏 (3.1) (3.2) must hold for a given set of bilayer and droplet monolayer surface energies 𝛾𝑏, 𝛾1 and 𝛾2. This is equivalent to the analysis done several authors.(17) (18) (19) (8) (20) (21) The bilayer and droplet contact angles Θ𝑖 (where the index 𝑖 is the set [𝑏, 1,2]) are defined in Fig. 3.1(b). In practice, as droplet radius and position are relatively easy to measure and can be used to measure contact angle Θ1, Θ2, and Θ, the force balance of (3.1) and (3.2) can expressed as the equations 𝛾𝑏 = 𝛾2 ( sin 𝛩2 cot 𝛩1 + cos 𝛩2 cos 𝜃𝑏 − sin 𝜃𝑏 cot 𝜃1 ) 𝛾𝑏 = 𝛾1 ( sin 𝛩1 cot 𝛩2 + cos 𝛩1 sin 𝛩𝑏 cot 𝛩2 + cos 𝛩𝑏 ) (3.3) (3.4) The usefulness of the form in (3.3-4) becomes apparent if the DIB geometry is known along with single surface energy value 𝛾1or 𝛾2, in which case the bilayer surface energy 𝛾𝑏 can then be calculated. Thus, based on this a single experimental value of surface energy, both bilayer and monolayer surface energies can be calculated using geometric information from CLSM DIB images. The error propagation analysis of this equation is provided in the ESI. In section 3.5 we will theoretically consider how DIB asymmetry affect the bilayer area and curvature. To write down a set of equations which are uniquely solvable, we will assume that the volume of the droplets are known and conserved. Using simple geometry, 𝑟𝑖 = ( ), and the standard equation for 𝑎 sin 𝛩𝑖 the volume of a spherical cap, the volume of droplets 1 and 2 are (( 𝑎 sin 𝛩1 𝑎 (( sin 𝛩2 3 ) 3 ) 𝑉1 = 𝑉2 = 𝜋 3 𝜋 3 (2 + 3 cos 𝛩1 − cos3 𝛩1) + ( (2 + 3 cos 𝛩2 − cos3 𝛩2) − ( sin 𝛩𝑏 sin 𝛩𝑏 𝑎 𝑎 3 ) 3 ) (2 − 3 cos 𝛩𝑏 + cos3 𝛩𝑏)) (2 − 3 cos 𝛩𝑏 + cos3 𝛩𝑏)) (3.5) (3.6) Now that four equations and four variables remain, namely: equations (3.1), (3.2), (3.5), and (3.6) with variables 𝜃1, 𝜃2, 𝜃𝑏, and 𝑎, the system of equations can be solved. However, as there is no simple analytic solution, this system must be solved using numerical techniques. 3.2 Symmetric Lipid DIB Confocal Imaging Result As an experimental control, symmetric lipid DIBs were formed as shown in Figure 3.2. Here the monolayer surface energy of a pure DPhPC monolayer between water and hexadecane is taken as 1.18 𝑚𝑁 𝑚−1.(17) (34). A DIB made up of pure DPhPC with closely matching volumes that vary by less than 1% is shown under CLSM to exhibit no bilayer curvature. To verify that there is no appreciable bilayer curvature, the image is processed with standard techniques using the MATLAB image processing toolbox. All the original data is processed with a Gaussian filter to smooth the edges on the interface peaks and the MATLAB 'fminsearch' function was used to attempt to fit the interface shape to the equation of a circle and to a line. Unsurprisingly, the solver could not fit the interface to the equation of a circle, but could fit to a straight line with a root mean square error (RMSE) of 0.15 depicted as a red line in Figure 3.2a. The droplets positions and radii are measured using the MATLAB function "imfindcircles" which employs the Hough(44) transform. The dimensions of the symmetric DIB is Figure 3.2a was found to be 𝑟1 = 433 𝜇𝑚, 𝑟2 = 437 𝜇𝑚, 𝑟𝑏 = 𝑖𝑛𝑓, and 𝑎 = 221 𝜇𝑚. From equations (3.3) and (3.4), with the input value of 𝛾1 = 𝛾2 = 1.18 𝑚𝑁 𝑚−1, the bilayer surface energy was calculated to be 𝛾𝑏 = 2.04 𝑚𝑁 𝑚−1 with an error of 0.121 𝑚𝑁 𝑚−1 (see ESI for error propagation analysis), matching previously reported surface energy results from Taylor.(17) In contrast, a non-similar volume DIB (i.e. a volume ratio of 0.37) is shown to exhibit a circular curve in the bilayer which bends toward the smaller droplet, shown in Figure 3.2b. To calculate the bilayer curvature, the image is processed again on MATLAB using the image processing. Within the region of interest (ROI), the maximum intensity peak values are obtained along the vertical axis. These peak values are fit to the equation of a circle using the MATLAB function "fminsearch" to minimize the root mean squared error (RMSE) of the distance from a peak point to the fit circle. The droplet dimensions are also measured using the MATLAB function "imfindcircles". From this the ratio of the bilayer radius of curvature to the smaller droplet radius of curvature in the figure is measured to be 7.21 with a RMSE of 0.12 depicted as a red line. Based on the measured, normalized geometry of 𝑟1 = 397 𝜇𝑚, 𝑟2 = 535 𝜇𝑚, 𝑟𝑏 = 2859 𝜇𝑚, and 𝑎 = 270 𝜇𝑚, the surface energy for the bilayer is calculated to be 𝛾𝑏 = 1.93 𝑚𝑁 𝑚−1 with an error of 0.107 𝑚𝑁 𝑚−1. Here actual bilayer surface energy measurement is within error of the previously reported measurement, 2.04 𝑚𝑁 𝑚−1.(17) Figure 3.2: Filtered CLSM image (a) of a volume symmetric DIB formed from a single lipid type. Filtered image (b) of a volume asymmetric DIB (volume ratio of 0.37) that exhibits bilayer interface curvature. Both droplets consist of DPhPC with dilute (0.1% molar fraction) NDB-PC dye formed in hexadecane. The bilayer concavity faces the smaller droplet with higher Laplace pressure. Scale bars are 100 𝜇m. 3.3 Asymmetric Lipid DIB Confocal Imaging Result The results of the CLSM experiment on asymmetric DIBs is provided in Figure 3.3. Three DIBs of varying degrees of monolayer asymmetry, from lowest to highest, are shown to exhibit bilayer curvature, where the interface curvature is measured with a MATLAB script where the derivative of the fluorescence intensity plot is used to find the image edge threshold which is fit to the equation of a circle by minimizing the RMSE, see ESI for more details. The geometry furthermore can be used to calculate the bilayer and the monolayer surface energy. Note that in the following cases the pure DPhPC lipid droplet surface energy is assumed to remain 𝛾2 = 1.18 𝑚𝑁 𝑚−1. Figure 3.3(a) shows a bilayer curvature to droplet curvature ratio of 4.96 and a spherical cap base radius to droplet radius ratio of 0.44 at a RMSE of 0.31. Note that in Figure 3.3(a) the monolayer in the dark (leftmost) droplet 1 is composed of 6% DOPG from total lipid content, which is left dark to enhance the contrast in the bilayer threshold. The geometric measurements of the DIB are 𝑟1 = 431 𝜇𝑚, 𝑟2 = 431 𝜇𝑚, 𝑟𝑏 = 2138 𝜇𝑚, and 𝑎 = 194 𝜇𝑚, where a brightfield image of the dark droplet is used to measure the dimensions of the dark droplet. The increased surface energy for droplet 1 and the bilayer is calculated to be 𝛾1 = 1.70 𝑚𝑁 𝑚−1, and 𝛾𝑏 = 2.58 𝑚𝑁 𝑚−1 with an error of 0.149 𝑚𝑁 𝑚−1. Further increasing the DIB asymmetry shown in Figure 3.3(b) confirms that the bilayer radius of curvature ratio deceases to 3.34 with a spherical cap base radius to droplet radius ratio of 0.49 at a RMSE of 0.33. The asymmetric DIB is composed of 12% DOPG from total lipid content in the left droplet with dimensions measured to be slightly volume asymmetric, 𝑟1 = 452 𝜇𝑚, 𝑟2 = 428 𝜇𝑚, 𝑟𝑏 = 1420 𝜇𝑚, and 𝑎 = 225 𝜇𝑚. Similarly, the surface energy for droplet 1 and the bilayer is calculated to be 𝛾1 = 1.92 𝑚𝑁 𝑚−1, and 𝛾𝑏 = 2.68 𝑚𝑁 𝑚−1 with an error of 0.169 𝑚𝑁 𝑚−1. The third and highest stable asymmetric DIB formed in Figure 3.3(c) is composed of 25% DOPG from total lipid content. The bilayer radius of curvature ratio is measured at 2.23 and spherical cap base radius to droplet radius ratio of 0.58 with a RMSE of 0.516. The DIB dimensions are calculated to be 𝑟1 = 334 𝜇𝑚, 𝑟2 = 331 𝜇𝑚, 𝑟𝑏 = 738 𝜇𝑚, and 𝑎 = 185 𝜇𝑚, where the surface energy for droplet 1 and the to be 𝛾1 = 2.70 𝑚𝑁 𝑚−1, and 𝛾𝑏 = 3.33 𝑚𝑁 𝑚−1 with an error of bilayer 0.195 𝑚𝑁 𝑚−1. is calculated Figure 3.3: The filtered CLSM images of matching volume droplet DIBs in hexadecane with mismatched surface energies exhibits curvature in the bilayer towards the higher surface tension droplet. The dark droplet contains the DOPG and light droplet contains the DPhPC with NDBPC. The amount of DOPG in images (a), (b) and (c) are respectively 6%, 12% and 25% from total lipid content. Scale bars are 100 𝜇m. 3.4 Pendant Drop Measurements (DSA) Results To compare with the above DIB method of measuring surface energy, the results from the DSA measurements are provided in Figure 3.4. The measurement was taken for droplets that could attain equilibrium without falling from the flat syringe needle. Note that the Worthington number (𝑊𝑜) is close to 1 for most measurements. The results below can be used to verify the DIB morphology method for surface tension. Percent DOPG 𝜸𝟏, 𝑚𝑁 𝑚−1 STDEV 50.0 25.0 3.13 2.65 1.56 1.35 𝑾𝒐 0.26 0.71 12.5 6.3 0.0 2.02 1.74 1.50* 1.02 0.91 0.82 0.67 0.99 0.70 Figure 3.4: Table of surface energy measurements for a droplet of DPhPC with a given DOPG% that forms a monolayer between water and hexadecane. The error and Worthington number are provided for reference. *Note that the literature value for pure DPhPC is 1.18 𝑚𝑁 𝑚−1.(17) (34). The DSA results show good agreement with the DIB method as shown in Figure 3.5. This verifies the technique developed by several authors with respect to symmetric DIBs.(17) (18) (19) (8) (20) (21) However, here we have shown that with the use of CLSM, we can capture bilayer curvature data to be used in calculating asymmetric bilayer surface energy. This is useful as with significantly low surface energies (lower than 5 𝑚𝑁 𝑚−1) it is often difficult to obtain shape measurements from the standard DSA method as the droplets tend to fall from the needle.(34) Given that they are stable and stationary, by the DIB method, asymmetric bilayer surface energies can be calculated. We further note that the discrepancy in the measurement of pure DPhPC comes chiefly from the fact that DSA measurements become more difficult with low surface tensions, in this case 𝛾 lower than 1.5 𝑚𝑁 𝑚−1. Figure 3.5: Bilayer and monolayer surface energies obtained from DSA and DIB methods as a function of monolayer asymmetry in DIBs with a droplet composed of pure DPhPC and a droplet with a mixture of DOPG in DPhPC. For the DIB method, the droplet 2 surface energy is assumed to be 𝛾2 = 1.18 𝑚𝑁 𝑚−1 for pure DPhPC. The linear fits for the bilayer and monolayer surface energies have a Pearson's R-squared value of 0.94 and 0.98 respectively with a sample size is 𝑛 = 3. 3.5 Droplet Morphology Model Result The free energy model described in Section 3.1 was applied to investigate asymmetric and symmetric DIB morphology with the given system surface energies 𝛾1, 𝛾2, and 𝛾𝑏 by equation (3.1), (3.2), (3.5) and (3.6). A simple way to analyse the system is to view the interface diameter 𝑎 normalized by the droplet radius 𝑟𝑚. This is useful as it can be generalized and scaled for different droplet systems driven by surface energy minimization. By this assessment, the symmetric model is 𝑎 𝑟𝑚 = √(1 − ( 𝛾𝑏 𝛾𝑚 2 ) 1 2 ) which is shown by Figure 3.6, where the DIB monolayer surface energies 𝛾1 and 𝛾2 of droplets 1 and 2 are the reference values, i.e.: 𝛾𝑏 is in the form of . As the bilayer surface energy is decreased from 𝛾𝑏 𝛾𝑚 contact surface area between droplets. Here one can see that the ratio of spherical cap base radius 𝑎 to droplet radius 𝑟1 and 𝑟2 increase drastically following the arrow and gradually decreases until = 0 = 0) the DIB will start to 'zip up'. This 'zip up' process is defined as an increase in = 2 𝑜𝑟 ( 𝛾𝑏 𝛾𝑚 𝑎 𝑟𝑚 𝛾𝑏 𝛾𝑚 𝑜𝑟 ( 𝑎 𝑟𝑚 = 1). This is an unsurprising result as it has already been shown by (3.1) that the contact angle 𝛩𝑚 changes as cos 𝛩𝑚 = 𝛾𝑏 2𝛾𝑚 . This 'zip up' process occurs mainly up to the point where the bilayer surface energy matches that of the monolayers, or 𝛾𝑚 = 𝛾𝑏 𝑜𝑟 ( perturbation in bilayer surface energy will have a diminished effect on bilayer radius 𝑎. Note that as surface energy is finite, the DIB can only 'zip up' completely if 𝛾𝑏 = 0. = 1). After this point any small 𝛾𝑚 𝛾𝑏 There is no simple analytical solution to the asymmetric case. However, it can be solved using numerical techniques. A MATLAB script was employed to solve for the variables 𝑟1, 𝑟2, 𝑟𝑏 and 𝑎. The script employs the 'fmincon' function, which runs an 'interior-point' algorithm, to solve for the minimization of the free energy functional(21) 𝑓 of surface energy 𝛾 and surface area 𝐴, 𝑓 = 𝛾1𝑑𝐴1 + 𝛾2𝑑𝐴2 + 𝛾𝑏𝑑𝐴𝑏 (3.7) under the constraint that 𝑉1 and 𝑉2 (from equation 3.7 and 3.8) are constant. This script was used to solve for the ratio of the drop radii 𝑟1, 𝑟2 and the bilayer radius 𝑟𝑏. Figure 3.7 shows that, for asymmetric DIBs, the membrane radius 𝑟𝑏 will decrease with increasing asymmetry in monolayer until it matches the spherical cap base radius, or 𝑟𝑏 = 𝑎. For simplicity, here we surface tension 𝛾1 𝛾2 have used 𝛾2 = 𝛾𝑏 as the reference tension value. Note that a DIB with a bilayer of infinite radius 𝑟𝑏 → ∞ (zero mean curvature) exists at the symmetric limit. The model result also shows that for ~1.2 , small changes in asymmetry affect the bilayer asymmetric DIBs greater than the range of 𝛾1 𝛾2 radius of curvature significantly. Figure 3.6: Symmetric DIB model result for the ratio between the spherical cap base radius 𝑎 𝑎𝑛𝑑 𝑡ℎ𝑒 𝑑𝑟𝑜𝑝𝑙𝑒𝑡 𝑟𝑎𝑑𝑖𝑖 𝑟1 and 𝑟2 as a function of bilayer to monolayer tension 𝛾𝑏/𝛾𝑚. The droplets 'zip up' drastically with increasing monolayer surface energy up to the point the bilayer and monolayer energies match, after which the effect is less dramatic until the droplets 'zip up' completely and the droplet radius matches that of the bilayer radius. Note that the lipids in the DIB cartoons are for shape reference and not drawn to scale. Figure 3.7: Asymmetric DIB model result for the ratio of the droplet radii 𝑟1𝑎𝑛𝑑 𝑟2 to bilayer radius 𝑟𝑏 as a function of monolayer tension ratio 𝛾1/𝛾2. Here 𝛾2 = 𝛾𝑏. Due to the force balance, the bilayer deviates from the spherical cap base radius 𝑎, and following the arrow, it will continue to curve towards the higher tension droplet until the bilayer radius matches that of the droplet radius 𝑟1 = 𝑟𝑏. Due to a mass balance, the droplet diameter of 2 will increase above that of droplet 1. Note that the lipids are not drawn to scale in the DIB cartoon. Often bilayer area is approximated by the spherical cap base radius (or the linear distance between the intersecting circles) 𝐴𝑎𝑝𝑝𝑟𝑜𝑥 = 𝜋𝑎2. The definitions of bilayer area considering curvature (spherical cap area) is given as 𝐴𝑏 = 2𝜋𝑟𝑏ℎ𝑏. Therefore, the percent area deviation from the linear approximation can be defined as ∆𝐴, ∆𝐴 = ( 2𝜋𝑟𝑏ℎ𝑏 𝜋𝑎2 − 1) × 100% (3.8) As shown in Figure 3.8 for volume symmetric DIBs, increasing the monolayer asymmetry elicits only a modest deviation in surface area. However, if the droplet volume asymmetry is modified the area deviation can be magnified. Note that typically DIB droplets are roughly the same size, and high surface energy asymmetry does not appear to be stable experimentally. By applying this model, the area correction of the DIB bilayers found experimentally via CLSM can be determined. For the volume asymmetric droplet (Figure 3.2b), by equation 3.10, the area deviation ∆𝐴 is found to be 0.22%. Additionally, the area deviation for the lipid asymmetric DIBs is found to be 0.21, 0.63, and 1.6% for the 6, 12 and 25% DOPG in DPhPC respectively. This shows that, at least for the range of DIB asymmetry explored in this study, the linear approximation of area is a reasonable estimate. Indeed, according to Figure 3.8, even relatively high monolayer asymmetry manifests as a deviation of less than 5% for volume symmetric DIBs. Figure 3.8: Model result of the percent area deviation of DIBs with volume and lipid asymmetry given that the bilayer surface tension is set to the average of the two monolayer surface tensions. The effect of monolayer surface energy is exacerbated by increasing droplet volume differences. 3.6 Model and System Limitations 3.6.1 Practical Limitations of the Method The valid range of 𝛾1 and 𝛾2 for the model is limited by DIB stability as experimentally stable DIBs are only formed below a surface energy ratio of 2.5. Above this level the droplets are disposed to coalesce into one larger droplet. This can be explained by the fact that emulsion or DIB stability depends on a) the osmotic and Laplace pressure of the droplets and b) on the pressure balance across the membrane.(45) Inescapably, the difference in pressure between connecting droplets may lead to inherent instability for large droplet volume ratios which limits the practicality for extreme droplet volume ratios. 3.6.2 Limitations in Scalability The use of DIBs for measuring surface tension is limited in size to micro scale droplets. This is the case as other thermodynamic factors come into play on smaller length scales such as line tension, which becomes non-negligible once the droplet reaches length scales below 100 𝑛𝑚.(46) Additionally, it is important to note that attaining thermodynamic equilibrium can be somewhat troublesome for DIBs as they continually lose water mass due to evaporation, this is a particular problem for DIBs with diameters on a micron length scale.(5) The effect of evaporation on DIBs was characterized recently by Venkatesan et al,(47) and this effect is mitigated by using droplets on the order of 300 micron in diameter covered by a thick layer of oil, however the effect of gravity on droplet shape prevents the use of much larger DIBs without adding another level of complexity to the model.(48) Indeed, the model system is limited in scalability by the Bond number (see Section 2.4). For this study the density difference between water and hexadecane is ∆𝜌 = 230 𝑘𝑔 𝑚−3, acceleration of gravity 𝑔 = 9.8 𝑚 𝑠−2, and surface tension 𝛾 is on the order of 10−3 𝐽 𝑚−2. If the characteristic length is taken as droplet radius, then 𝐿 is on the order of 2.3 − 4.0 [× 10−4]𝑚, this implies a Bond number of approximately 0.1 to 0.4. A reasonable upper limit for DIB applications is a Bond number less than 1 (a droplet radius of 660 micrometres for the system at hand), as values greater than 1 imply a decreased effect of surface tension relative to gravity and will result in non-spherical droplets. Note that this model does not account for non-spherical droplets. 3.6.3 Model limitations The model is limited to static equilibria and cannot be used to probe the absolute surface tension of the = 𝑓(𝛩1, 𝛩2, 𝛩𝑏), can DIB membrane out of equilibrium, though the relative surface forces, such as 𝛾𝑏 𝛾2 be calculated from equations (3.3) and (3.4). In a similar vein, the model is only valid for systems that are under tension. More specifically, this model would not be particularly useful to measure the tension of adhering vesicles, as the mechanical tension is not necessarily known as the bodies can be deflated and the energetics can be affected by the expansion modulus.(49) 4. Conclusions For the first time it has been shown that asymmetric DIBs form curved surface in the bilayer due to a surface energy balance. It is analogous to the effect of volume differences, but here the surface energy asymmetry controls this behaviour. As shown by Taylor et al for symmetric DIBs(17), our study shows that the curvature effect in asymmetric DIBs can be employed as an alternative method of measuring interfacial tension of complex, asymmetric lipid monolayers or bilayers through the application of CLSM. The obtained interfacial tension values are in good agreement with droplet shape analysis results. Furthermore, the results obviate the negligible effect of area deviation with respect to DIB asymmetry; though the effect of curvature strongly affects the surface tension calculation, even the most asymmetric system in this experiment (with a surface energy ratio of approximately 2.2) corresponds to a deviation of only 1.6%. Thus, with DIB platforms, the bilayer interfacial area measurement is only affected by the extreme cases of high surface tension asymmetry and extreme volume mismatch, an important validation of an assumption made in many published DIB applications. A linear relationship between bilayer surface energy with respect to DOPG and DPhPC mixtures is shown up to 25% DOPG. However, this linear relationship is not necessarily the case for all lipid mixtures. For example, significant non-linearity and hysteresis in dynamic interfacial tension measurements as a function of the mole fraction of cholesterol in lecithin lipids has been observed.(50) The for phosphatidylcholine- phosphatidylethanolamine and sphingomyelin-ceremide mixtures, this implies a non-linear relationship for interfacial tension with respect to lipid concentrations.(51) Thus, the asymmetric DIB morphology method could be used to probe this non-linear surface tension behaviour by measuring the surface morphology as a function of lipid content and asymmetry. lipid-lipid complex has been formation of a shown There is a wide range of possibilities for future work measuring surface tension and curvature affects in DIBs, giant unilameller vesicles (GUVs)(52) or even cells. It has already been shown that curvature exists between adhering cells as observed in the biologically mediated cell-cell contact between C. elegans embryos(53) and between adhering vesicles.(42) Investigations of lipid flip-flop in bio- membranes,(54) to a marginal degree of success, have been performed using sum frequency vibrational spectroscopy,(55) indirectly with ceremide induced trans-bilayer movement in vesicles,(54) small angle neutron scattering,(56) and by molecular simulation.(57) Asymmetric DIBs or adhering vesicles offer an alternative measurement technique for the rate of lipid flip-flop, by directly measuring the decrease of interfacial curvature as the lipids flip from one droplet or vesicle to another. Here the challenge lies in distinguishing the rate of flip-flop from the rate of lateral lipid diffusion(58) between the monolayer and bilayer, as well as lipid uptake into the bilayer.(5, 50, 59) However, a recent publication has demonstrated a promising technique for determining bilayer flip-flop on DIB membranes via parallel capacitance based measurements on an integrated microfluidic device; in this study it was successfully shown that surface bound peptides (alamethicin) facilitate the movement of lipids between leaflets.(9) The present model is valid for stationary surfaces at equilibrium. It would also be interesting to extend the model to dynamic behaviour of micro-DIBs where the droplets change shape and the bilayer may even buckle.(60) The bilayer buckling indicates that the effective bilayer surface tension 𝛾𝑏 had dropped to zero.(5) Understanding interfacial physical chemistry is paramount to the development of DIBs as a tool for biological discovery, which is crucial for burgeoning fields such as synthetic biology and biotechnology. Competing Interests The authors declare no competing financial interests. Authors' Contributions N.E.B. performed experiment and model analysis, wrote the main text of the document. H.K. edited the text, discussed the model and data analysis, A.S-R. edited text, built the experimental setup and discussed the results, O.C., A.J.F, L.M.C.B and N.B. edited text and discussed results. Funding The research leading to these results has received funding from the European Union Seventh Framework Programme (FP7/2007-2013) under grant agreement no 607466. This work was supported by an Imperial College Fellowship awarded to A.S-R. This research was funded by and BBSRC EPSRC grants: EP/J017566/1, EP/L015498/1, EP/J021199/1 and EP/K503733/1. References Funakoshi K, Suzuki H, Takeuchi S. Lipid bilayer formation by contacting monolayers in a Hagan Bayley BC, Andrew Heron, Matthew A. Holden, William Hwang, Ruhma, Syeda JT, and 1. Mark Wallace. Droplet Interface Bilayers. Molecular Biosystems. 2008;4(12):1191-208. 2. microfluidic device for membrane protein analysis. Analytical chemistry. 2006;78(24):8169-74. 3. Najem JS, Freeman EC, Yasmann A, Sukharev S, Leo DJ. Mechanics of Droplet Interface Bilayer "Unzipping" Defines the Bandwidth for the Mechanotransduction Response of Reconstituted MscL. Advanced Materials Interfaces. 2016. 4. control of droplet interface bilayer area. Langmuir. 2011;27(23):14335-42. 5. morphologies of microscale droplet interface bilayers. Soft Matter. 2014;10(15):2530-8. Gross LC, Heron AJ, Baca SC, Wallace MI. Determining membrane capacitance by dynamic Mruetusatorn P, Boreyko JB, Venkatesan GA, Sarles SA, Hayes DG, Collier CP. Dynamic Taylor G, Nguyen M-A, Koner S, Freeman E, Collier CP, Sarles SA. Electrophysiological Elani Y, Purushothaman S, Booth PJ, Seddon JM, Brooks NJ, Law RV, et al. Measurements of Krylov AV, Pohl P, Zeidel ML, Hill WG. Water permeability of asymmetric planar lipid bilayers: Milianta PJ, Muzzio M, Denver J, Cawley G, Lee S. Water Permeability across Symmetric and Hwang WL, Chen M, Cronin B, Holden MA, Bayley H. Asymmetric Droplet Interface Bilayers. Nguyen M-A, Sarles SA, editors. Microfluidic Generation, Encapsulation and Characterization Traïkia M, Warschawski DE, Lambert O, Rigaud J-L, Devaux PF. Asymmetrical Membranes Freeman EC, Najem JS, Sukharev S, Philen MK, Leo DJ. The mechanoelectrical response of 6. Journal of the American Chemical Society. 2008;130(18):5878-9. 7. of Asymmetric Droplet Interface Bilayers. ASME 2016 Conference on Smart Materials, Adaptive Structures and Intelligent Systems; 2016: American Society of Mechanical Engineers. 8. droplet interface bilayer membranes. Soft Matter. 2016;12(12):3021-31. 9. interrogation of asymmetric droplet interface bilayers reveals surface-bound alamethicin induces lipid flip-flop. Biochimica et Biophysica Acta (BBA)-Biomembranes. 2018. 10. the effect of membrane asymmetry on the mechanical properties of lipid bilayers. Chemical Communications. 2015;51(32):6976-9. 11. and Surface Tension. Biophysical Journal. 2002;83(3):1443-54. 12. Biological Membranes to Protons, Solutes, and Gases. The Journal of General Physiology. 1999;114(3):405-14. 13. leaflets of different composition offer independent and additive resistances to permeation. J Gen Physiol. 2001;118(4):333-40. 14. Asymmetric Droplet Interface Bilayers: Interaction of Cholesterol Sulfate with DPhPC. Langmuir. 2015;31(44):12187-96. 15. Ohki S, Ohshima H. Divalent cation-induced phosphatidic acid membrane fusion. Effect of ion binding and membrane surface tension. Biochimica et Biophysica Acta (BBA)-Biomembranes. 1985;812(1):147-54. 16. Biophysica Acta (BBA)-Biomembranes. 2004;1666(1):62-87. 17. Taylor GJ, Venkatesan GA, Collier CP, Sarles SA. Direct in situ measurement of specific capacitance, monolayer tension, and bilayer tension in a droplet interface bilayer. Soft Matter. 2015;11(38):7592-605. 18. Droplet Lipid Bilayers. Langmuir : the ACS journal of surfaces and colloids. 2012;28(19):7442-51. 19. microdroplets coated with lipid monolayers. Soft Matter. 2013;9(25):5891-7. 20. bilayers. ASME 2015 Conference on Smart Materials, Adaptive Structures and Intelligent Systems; 2015: American Society of Mechanical Engineers. 21. environments. Nat Nano. 2011;6(12):803-8. 22. 1981;35(3):637-52. 23. 1996;71(3):1348-9. 24. extracted from neuronal growth cones. Biophysical journal. 1996;70(1):358-69. 25. undulation forces between giant vesicles and solid substrates. Physical Review E. 1995;51(5):4526. 26. 2000;33(1):15-22. Yanagisawa M, Yoshida T-a, Furuta M, Nakata S, Tokita M. Adhesive force between paired Villar G, Heron AJ, Bayley H. Formation of droplet networks that function in aqueous Kwok R, Evans E. Thermoelasticity of large lecithin bilayer vesicles. Biophysical journal. Jähnig F. What is the surface tension of a lipid bilayer membrane? Biophysical Journal. Rädler JO, Feder TJ, Strey HH, Sackmann E. Fluctuation analysis of tension-controlled Hochmuth RM. Micropipette aspiration of living cells. Journal of biomechanics. Hochmuth F, Shao J-Y, Dai J, Sheetz MP. Deformation and flow of membrane into tethers Hill WG, Rivers RL, Zeidel ML. Role of Leaflet Asymmetry in the Permeability of Model Lee AG. How lipids affect the activities of integral membrane proteins. Biochimica et Dixit SS, Pincus A, Guo B, Faris GW. Droplet Shape Analysis and Permeability Studies in Kancharala AK, Freeman E, Philen MK, editors. Energy harvesting from droplet interface Klenz U, Saleem M, Meyer M, Galla H-J. Influence of lipid saturation grade and headgroup Venkatesan GA, Lee J, Barati Farimani A, Heiranian M, Collier CP, Aluru NR, et al. Adsorption Barriga HMG, Booth P, Haylock S, Bazin R, Templer RH, Ces O. Droplet interface bilayer Guevorkian K, Colbert M-J, Durth M, Dufour S, Brochard-Wyart F. Aspiration of biological Evans E, Rawicz W, Smith BA. Back to the future: mechanics and thermodynamics of lipid Kinoshita K, Parra E, Needham D. New Sensitive Micro-Measurements of Dynamic Surface Lee S, Kim DH, Needham D. Equilibrium and Dynamic Interfacial Tension Measurements at Rowlinson JS, Widom B. Molecular Theory of Capillarity: Dover Publications; 1982. Kasarova SN, Sultanova NG, Ivanov CD, Nikolov ID. Analysis of the dispersion of optical 27. viscoelastic drops. Physical review letters. 2010;104(21):218101. 28. biomembranes. Faraday Discussions. 2013;161:591-611. 29. Tension and Diffusion Coefficients: Validated and Tested for the Adsorption of 1-Octanol at a Microscopic Air-Water Interface and its Dissolution into Water. Journal of Colloid and Interface Science. 2017. 30. charge: a refined lung surfactant adsorption model. Biophysical journal. 2008;95(2):699-709. 31. Schürch S, Goerke J, Clements JA. Direct determination of surface tension in the lung. Proceedings of the National Academy of Sciences. 1976;73(12):4698-702. 32. 33. plastic materials. Optical Materials. 2007;29(11):1481-90. 34. kinetics dictate monolayer self-assembly for both lipid-in and lipid-out approaches to droplet interface bilayer formation. Langmuir. 2015. 35. reconstitution and activity measurement of the mechanosensitive channel of large conductance from Escherichia coli. Journal of the Royal Society Interface. 2014;11(98). 36. Microscopic Interfaces Using a Micropipet Technique. 2. Dynamics of Phospholipid Monolayer Formation and Equilibrium Tensions at the Water−Air Interface. Langmuir. 2001;17(18):5544-50. 37. Worthington AM. On pendent drops. Proceedings of the Royal Society of London. 1881;32(212-215):362-77. Yang L, Kapur N, Wang Y, Fiesser F, Bierbrauer F, Wilson MC, et al. Drop-on-demand 38. satellite-free drop formation for precision fluid delivery. Chemical Engineering Science. 2018. 39. Berry JD, Neeson MJ, Dagastine RR, Chan DYC, Tabor RF. Measurement of surface and interfacial tension using pendant drop tensiometry. Journal of Colloid and Interface Science. 2015;454:226-37. 40. 2015;128(6):1065-70. 41. Langmuir. 2015;31(1):350-7. 42. negative spontaneous curvature lipids induced by phase separation. The European physical journal E, Soft matter. 2008;25(4):403-13. 43. spontaneous curvature of dioleoylphosphatidylglycerol. Chemistry and physics of lipids. 2008;154(1):64-7. 44. image processing. 1988;44(1):87-116. 45. Film Strength. Journal of Colloid and Interface Science. 2002;250(2):444-50. 46. domains. Langmuir. 2006;22(26):11041-59. 47. monolayer compression improves droplet interface bilayer formation using unsaturated lipids. Biomicrofluidics. 2018;12(2):024101. Jiao J, Rhodes DG, Burgess DJ. Multiple Emulsion Stability: Pressure Balance and Interfacial Yasmann A, Sukharev S. Properties of Diphytanoyl Phospholipids at the Air -- Water Interface. Sakuma Y, Imai M, Yanagisawa M, Komura S. Adhesion of binary giant vesicles containing Blecua P, Lipowsky R, Kierfeld J. Line tension effects for liquid droplets on circular surface Venkatesan GA, Taylor GJ, Basham CM, Brady NG, Collier CP, Sarles SA. Evaporation-induced Alley SH, Ces O, Barahona M, Templer RH. X-ray diffraction measurement of the monolayer Illingworth J, Kittler J. A survey of the Hough transform. Computer vision, graphics, and McMahon HT, Boucrot E. Membrane curvature at a glance. Journal of Cell Science. Petelska AD, Figaszewski ZA. Interfacial tension of the two-component bilayer lipid Petelska AD. Interfacial tension of bilayer lipid membranes. Central European Journal of Elani Y, Gee A, Law RV, Ces O. Engineering multi-compartment vesicle networks. Chem Sci. Contreras FX, Sánchez-Magraner L, Alonso A, Goñi FM. Transbilayer (flip-flop) lipid motion Ren H, Xu S, Wu S-T. Effects of gravity on the shape of liquid droplets. Optics Fujita M, Onami S. Cell-to-cell heterogeneity in cortical tension specifies curvature of contact 48. Communications. 2010;283(17):3255-8. 49. Bolognesi G, Friddin MS, Salehi-Reyhani A, Barlow NE, Brooks NJ, Ces O, et al. Sculpting and fusing biomimetic vesicle networks using optical tweezers. Nature communications. 2018;9(1):1882. 50. membrane modelling of cell membrane. Bioelectrochemistry and Bioenergetics. 1998;46(2):199-204. 51. Chemistry. 2011;10(1):16-26. 52. 2013;4(8):3332-8. 53. surfaces in Caenorhabditis elegans embryos. PloS one. 2012;7(1):e30224. 54. and lipid scrambling in membranes. FEBS letters. 2010;584(9):1779-86. 55. frequency vibrational spectroscopy. Biophysical Journal. 2005;89(4):2522-32. 56. Phospholipids in Vesicles: Kinetic Analysis with Time-Resolved Small-Angle Neutron Scattering. The Journal of Physical Chemistry B. 2009;113(19):6745-8. 57. flip-flops in a model membrane. The Journal of Chemical Physics. 2014;140(6):064901. 58. bilayer membranes. FEBS letters. 1984;174(2):199-207. 59. biomedical engineering. 1995;23(3):287-98. 60. Droplet Interface Bilayers. Physical review letters. 2018;120(23):238001. Guiselin B, Law JO, Chakrabarti B, Kusumaatmaja H. Dynamic Morphologies and Stability of Liu J, Conboy JC. 1,2-diacyl-phosphatidylcholine flip-flop measured directly by sum- Nakano M, Fukuda M, Kudo T, Matsuzaki N, Azuma T, Sekine K, et al. Flip-Flop of Vaz WLC, Goodsaid-Zalduondo F, Jacobson K. Lateral diffusion of lipids and proteins in Needham D, Zhelev DV. Lysolipid exchange with lipid vesicle membranes. Annals of Arai N, Akimoto T, Yamamoto E, Yasui M, Yasuoka K. Poisson property of the occurrence of
1901.00263
1
1901
2019-01-02T04:46:12
Base-pair mismatch can destabilize small DNA loops through cooperative kinking
[ "physics.bio-ph", "q-bio.BM" ]
Base pair mismatch can relieve mechanical stress in highly strained DNA molecules, but how it affects their kinetic stability is not known. Using single-molecule Fluorescence Resonance Energy Transfer (FRET), we measured the lifetimes of tightly bent DNA loops with and without base pair mismatch. Surprisingly, for loops captured by stackable sticky ends, the mismatch decreased the loop lifetime despite reducing the overall bending stress, and the decrease was largest when the mismatch was placed at the DNA midpoint. These findings show that base pair mismatch transfers bending stress to the opposite side of the loop through an allosteric mechanism known as cooperative kinking. Based on this mechanism, we present a three-state model that explains the apparent dichotomy between thermodynamic and kinetic stability of DNA loops.
physics.bio-ph
physics
Base-pair mismatch can destabilize small DNA loops through cooperative kinking Jiyoun Jeong and Harold D. Kim∗ School of Physics, Georgia Institute of Technology, 837 State Street, Atlanta, GA 30332-0430, USA (Dated: January 3, 2019) Base pair mismatch can relieve mechanical stress in highly strained DNA molecules, but how it affects their kinetic stability is not known. Using single-molecule Fluorescence Resonance Energy Transfer (FRET), we measured the lifetimes of tightly bent DNA loops with and without base pair mismatch. Surprisingly, for loops captured by stackable sticky ends, the mismatch decreased the loop lifetime despite reducing the overall bending stress, and the decrease was largest when the mismatch was placed at the DNA midpoint. These findings show that base pair mismatch transfers bending stress to the opposite side of the loop through an allosteric mechanism known as cooperative kinking. Based on this mechanism, we present a three-state model that explains the apparent dichotomy between thermodynamic and kinetic stability of DNA loops. Cellular DNA is constantly exposed to the possibil- ity of mispairing (i.e. non-complementary base pairing) [1]. Most commonly, mismatched base pairs result from base misincorporation during gene replication [2] and heteroduplex formation between slightly different DNA sequences during homologous recombination [3]. They can also arise from exposure to DNA damaging agents that modify nucleobases [4, 5]. Due to less favorable base pairing and stacking [6], mismatched base pairs can in- crease local flexibility of double-stranded DNA [7 -- 9], and consequently the capture rate of tightly bent loops [10]. For example, 1 to 3 bp-mismatch near the center of a short DNA fragment (<150 bp) was shown to increase the rate of DNA loop formation by one to two orders of magnitude [11, 12]. The kinetics of loop formation or cap- ture is intuitively understood by a one-dimensional free energy curve with the end-to-end distance as a single reaction coordinate (Figure 1(a)). Base pair mismatch would reduce the mechanical work required to bring two distant DNA sites to proximity, more so for a shorter end-to-end distance. Therefore, the base pair mismatch would lower the transition state relative to the unlooped state (dotted line, Figure 1(a)). Base pair mismatch is also expected to affect the break- age or release rate of small DNA loops that are captured by protein complexes [13] or by sticky ends of the DNA itself [14]. Looped DNA segments on the order of one persistence length are subject to a high level of mechan- ical stress; therefore, the free energy of the looped state is significantly lowered in the presence of the mismatch. According to the free energy diagram in Figure 1(a), the transition state, being at a slightly longer end-to-end dis- tance by ∆x‡, would be lowered to a lesser degree (Fig- ure 1(a)). Therefore, the one-dimensional model predicts that the rate of loop release would decrease in the pres- ence of base pair mismatch. Such prediction of mismatch-dependence seems plausi- ble considering the success of the model in predicting the length dependence of loop capture and release rates. In the length regime where the free energy of loop formation is dominated by bending energy, increasing DNA length effectively reduces the tilt in the free energy curve be- cause states at shorter end-to-end distances receive more stress relief, similar to the dotted line in Figure 1(a). This change predicts that loop capture and release rates mea- sured at different DNA lengths would be anti-correlated; loops associated with higher mechanical stress are cap- tured more slowly and released more quickly. This pre- diction has been confirmed for both DNA loops captured by Lac repressor [15] and DNA loops captured by sticky FIG. 1. (a) One-dimensional free energy landscape for DNA loop capture and release. The two minimum free energy states correspond to the looped and unlooped states. The transi- tion state (vertical line) is separated from the looped state by a small distance ∆x‡, which is equal to the capture ra- dius. The base pair mismatch is expected to increasingly un- tilt the solid curve toward shorter end-to-end distances, which results in the dotted curve. (b) Typical FRET trajectories of a DNA molecule undergoing loop capture (left) and loop release (right). The DNA molecule labeled with Cy3 (green) and Cy5 (red) is in the low FRET state when unlooped, and in the high FRET state when looped. A sudden increase or decrease in NaCl concentration at the 20-second time point (marked by a vertical dotted line) triggers the transition. arXiv:1901.00263v1 [physics.bio-ph] 2 Jan 2019 ∆x‡ No mismatch With mismatch End-to-end distance High-salt buffer Low-salt buffer ∆tunloop ∆tloop 0 5 10 Time (min) 15 20 0 1 3 2 4 Time (min) 5 6 (a) Free energy (b) 1 0 FRET efficiency 2 FIG. 2. (a) Schematic of a hairband loop captured by sticky ends. The schematic on top shows base-paired overhangs, Cy3 (green circle), Cy5 (red circle), and the biotin linker (black circle). In this geometry, the overhangs on opposite strands form a duplex that can stack at both nicks of the loop. Different positions of base pair mismatch tested in our experiments are marked on the linear form at the bottom. Only the bases on the overhangs are shown. (b) Loop capture time of the hairband molecules (108 bp) as a function of the central mismatch size (circles). Data with an off-center 3-bp-mismatch are also shown as triangles. The upright and flipped triangles represent the loop capture times for base pair mismatches placed at 20 and 10 bp away from the center of the molecule, respectively. Error bars, the standard errors of the mean, are smaller than the size of the symbols. (c) Hairband loop lifetime (loop release time) as a function of the central mismatch size. Error bars represent the standard errors of the mean. (d) Probability density of spontaneous kink positions along the coarse-grained minicircle (105 bp) with (red) and without a pre-existing flexible defect (black), which is placed at position 1. (e) Bending angle calculated from the minimum-energy conformation of a DNA minicircle (105 bp) with a defect. Top and bottom figures show bending angles at the defect and the site opposite to the defect, respectively, as a function of the defect stiffness relative to an intact base pair. The minimum-energy conformations of the two extreme cases of the defect stiffness (0 and 100%) are also shown along the curves with the defect position marked by X. (f ) Hairband loop lifetime as a function of the mismatch position (3-bp in size). For comparison, the horizontal dotted line shows the loop lifetime without the mismatch. Error bars represent the standard errors of the mean. ends [16, 17]. While increasing DNA length evens out the bending stress over the entire DNA molecule, the base pair mismatch tends to localize sharp bending. There- fore, the effect of base pair mismatch might be quite dif- ferent from that of increasing DNA length. before unlooping is defined as the loop release time or loop lifetime (τloop). All DNA molecules used in this study were shorter than 150-bp, the length regime where the free energy of loop formation is dominated by bend- ing energy. In this Letter, we investigated how base pair mismatch affects the stability of small DNA loops. As a model sys- tem for DNA loop capture and release, we used short double-stranded DNA molecules with sticky ends. To monitor loop capture and loop release events, we used the single-molecule FRET assay as previously published [14]. Briefly, DNA molecules labeled with Cy3 and Cy5 near their sticky ends were immobilized to a NeutrAvidin- coated glass surface through a biotin linker, and loop cap- ture or release was triggered by exchange of buffers with different NaCl concentrations (see Supplemental Mate- rial for more details). The first transition times in the FRET signals (∆t) of ∼150 individual DNA molecules were collected. The mean of ∆t spent in the unlooped state before looping is defined as the loop capture time (τunloop), and the mean of ∆t spent in the looped state We first tried the loop capture geometry used in DNA cyclization, which we term as the "hairband loop" (Fig- ure 2(a)). In this geometry, the complementary over- hangs protrude from different strands so that the sticky ends can anneal in trans and stack upon each other. In a previous study, we showed that this end stacking, or equivalently nick closing, substantially increases the hair- band loop stability [18]. Using the single-molecule FRET assay, we measured the hairband loop capture times with and without base pair mismatch in the center. As shown in Figure 2(b), hairband loop capture took less time in the presence of the mismatch as expected. The loop cap- ture time further decreased with increasing mismatch size (circles, Figure 2(b)). The base pair mismatch in the cen- ter position led to the largest decrease in the loop cap- ture time, and the decrease dropped as the mismatch was 0 1 3 Mismatch size (bp) 5 No mismatch 0 10 20 Distance from center (bp) (c) 100 80 60 40 20 0 loop(sec) (f) 100 80 60 40 20 0 loop(sec) 100 Central mismatch 20bp off-center 10bp off-center 0 3 Mismatch size (bp) 1 Defect Opposite to defect 20 40 60 80 Relative defect stiffness (%) 103 102 101 100 120 90 60 30 0360 0 unloop(sec) (e) Bending angle (degrees) (a) Hairband loop Center Cy3 Cy5 Biotin (b) 104 Sticky ends 20bp 10bp Center No defect Defect at 1 25 50 75 Kink location (bp) 105 0 1 0.05 0.04 0.03 0.02 0.01 (Nkink=1000) Probability distribution (d) placed further from the center (triangles, Figure 2(b)). These observations confirm previous findings that mis- matched base pairs reduce the energy barrier for loop formation by increasing DNA bendability [8, 11, 19, 20], and this barrier reduction is most effective when the mis- match is in the center [21]. Next, we measured the hairband loop release times or loop lifetimes (τloop) with and without the mismatch in the center. Since a mismatch could relieve the bending stress of the hairband loop, we thought that the loop life- time would become longer. To our surprise, we observed the exact opposite effect where the central mismatch de- creased the hairband loop lifetime (Figure 2(c)). Increas- ing the size of the mismatch from 1 bp to 3 bp led to a further decrease in the lifetime. This effect seemed to plateau past the mismatch size of 3 bp (Figure 2(c)). This result suggests that the mismatch-containing hairband loop is more kinetically unstable than the mismatch-free loop, which seems paradoxical through the lens of the one-dimensional model presented in Figure 1(a). We thus considered the possibility that the transition state depends on other reaction coordinates besides the end-to-end distance, such as the closing angles at the loop junction. Since base stacking at the nick(s) in the hairband loop is a key determinant of decyclization ki- netics [18], we asked whether the central mismatch could destabilize the hairband loop by allosterically inducing nick opening. To investigate such allosteric coupling, we calculated the curvature profile of a kinkable semi- flexible loop [22] containing a defect with zero rigidity from a Monte Carlo simulation (see Supplemental Mate- rial for details). As shown in Figure 2(d), a kink with a sharp bending angle appeared most frequently at the furthest end of the loop from the defect. We also calcu- lated the minimum energy conformation of a semiflexible loop while varying the rigidity of the defect and found that the bending angles of furthest points were highly correlated (Figure 2(e)). This loop-mediated correlation of sharp bending angles between most distant sites is termed cooperative kinking [23], and has been observed in torsionally strained DNA minicircles by cryo-electron microscopy and molecular dynamics simulations [23 -- 25]. We hypothesized that the enhanced flexibility of the central mismatch destabilizes the hairband loop prevent- ing nicks(s) on the opposite side from closing. This hy- pothesis provides a few testable predictions. First, if the mismatch were displaced from the midpoint of the DNA, the degree of destabilization would be dampened. In agreement with this prediction, we observed a longer loop lifetime when the mismatch was placed at a quarterpoint instead of the center (Figure 2(f)). Second, the cooper- ative kinking hypothesis requires nicks that can buckle under the bending stress, and therefore the mismatch- induced destabilization would be eliminated in a loop capture geometry free of end-stacking. We thus tested a different loop geometry referred to as the "hairpin loop", where the complementary overhangs protrude from the same strand (Figure 3(a)). In this geometry, the sticky 3 ends anneal in cis and cannot stack upon each other. Us- ing these new DNA constructs with a central mismatch of various sizes, we repeated loop capture and release ex- periments. Similar to hairband loop capture, the hairpin capture time decreased with the size of base pair mis- match (Figure 3(b)). However, in sharp contrast to the hairband loop, the hairpin loop lifetime increased with mismatch size (Figure 3(c)). The effect of the base pair mismatch on the hairpin loop stability is therefore con- sistent with the prediction of the one-dimensional model. Overall, the lifetimes of hairpin loops were shorter than those of hairband loops, which is consistent with easier rupture of DNA duplex in an unzipping geometry than in a shearing geometry [26 -- 28]. These results lend strong support to the idea that cooperative kinking governs the kinetic stability of a mismatch-containing hairband loop. The mismatch-dependence of the hairband loop release kinetics reveals the limitations of the one-dimensional two-state model (Figure 1(a)) and invites us to consider additional states and alternative reaction paths along an- other dimension. Here, we present two different paths (k(0) and k(m)) that are likely to be the dominant ones for mismatch-free and mismatch-containing DNA (Fig- ure 4(a)). Each path goes through three different states: unlooped, unstacked, and stacked. The loop capture rate is much greater in the presence of a central mismatch due to its enhanced flexibility (k(m) 1 ). The reverse rate is expected to be slower with the mismatch (k(m) 2 < k(0) 2 ) because of the weaker loop tension. Mismatch-free DNA 1 (cid:29) k(0) FIG. 3. (a) Schematic of a hairpin loop. The schematic shows the FRET pair (green and red circles), the biotin linker (black circle), and base-paired overhangs. In this geometry, the over- hangs on the same strand form a duplex like a zipper. (b) Loop capture time of the hairpin (105 bp) molecules as a func- tion of the central mismatch size. Error bars are omitted due to their small sizes. (c) Hairpin loop lifetime as a function of the central mismatch size. Error bars represent the standard errors of the mean. 0 1 3 1 0 5 Mismatch size (bp) 3 104 103 102 101 100 30 unloop(sec) (b) (c) 20 10 loop(sec) 0 (a) Hairpin loop Sticky ends 4 3 (cid:28) k(0) may occur. In comparison, DNA with a mismatch in the center can be sharply bent at a much lower energy cost, and therefore, the most dominant path toward the looped state will resemble a tweezers-like motion. As a result of this motion, the sticky ends anneal at a sharp angle, and the hairband loop with the mismatch faces a higher energy barrier for end-stacking (nick closing) than without (k(m) 3 ). The mismatch not only sup- presses end-stacking, but also promotes end-unstacking (nick opening) through cooperative kinking, which im- plies k(m) 4 . Hence, the apparent release rate of the hairband loop (kunloop) becomes faster with the mis- match than without because the looped state with the mismatch is heavily biased towards the unstacked state. In comparison, for the hairpin loop that cannot proceed to the stacked state, the three-state model is reduced to the two-state model, and the loop release rate is slower with the mismatch (k(m) 4 (cid:29) k(0) 2 < k(0) 2 ). The two paths boxed in Figure 4(a) represent the two most extreme paths in terms of kinetics, the top path for the slowest hairband loop capture and release, and the bottom for the fastest. In reality, there exists a contin- uum of paths going through the three states with inter- mediate rates, and the flexibility profile of DNA deter- mines the relative weights at which individual paths are taken. Therefore, any changes to the flexibility profile of DNA would lead to correlated changes in the hair- band loop capture and release rates. To test this idea, we measured hairband loop capture and release times of 16 unrelated sequences, all of the same length. Although limited in sample size, we observed a significant degree of correlation between the two times (Pearson correlation = 0.74, Figure 4(b)). This result suggests that cooperative kinking is a general mechanism that governs the kinetics of hairband loop capture and release. In conclusion, we demonstrate that base pair mis- match can constrain the geometry and interactions for DNA loop capture through cooperative kinking, and the close coupling between hairband loop geometry and end- stacking can give rise to correlated changes between loop capture and release times ("easy come, easy go"). We propose a three-state model that correctly describes the effect of mismatched base pairs on the apparent kinetics of loop capture and release. We expect the effect of mis- matched base pairs on protein-mediated DNA loops to be more complex because of the diversity in loop capture geometry [29]. Beyond passively captured DNA loops, it would be interesting to investigate whether base pair mismatches can also influence the kinetics of DNA loop extrusion [30, 31] through cooperative kinking. FIG. 4. (a) The three-state model for hairband loop closure and release. The three states from left to right are unlooped, unstacked, and stacked states. The looped state is a mixed state between the unstacked and stacked states. Therefore, the apparent loop capture rate (kloop) is equal to k1, but the apparent loop release rate (kunloop) depends on k2, k3, and k4. For the hairpin loop, k3 = 0, and therefore, kunloop is equal to k2. Two representative paths for central mismatch size 0 and m are highlighted with arc-like (top) and tweezers- like (bottom) motions, respectively. The vertical dotted lines imply the continuum of paths running parallel to the two extreme ones shown. (b) Correlation between loop capture and release times of 16 unrelated hairband DNA molecules of the same size (94bp). The loop capture and release times were measured in equilibrium (i.e. no buffer-exchange) at slightly elevated temperature of 34 ◦C with [NaCl] = 700mM. undergoes small bending fluctuations uniformly through- out its contour, and therefore, follows an arc-like trajec- tory toward the looped state where end-stacking (nick closing) and end-unstacking (nick opening) transitions ∗ Corresponding author. Molecular Mutagenesis 58, 235 (2017). Email: [email protected] [2] T. A. Kunkel and D. A. Erie, Annual Review of Genetics [1] N. Chatterjee and G. C. Walker, Environmental and 49, 291 (2015). No mismatch (0) (0) k3 k1 (0) (0) k4 k2 (m) (m) k3 k1 (m) (m) k4 k2 With mismatch State Pearson corr. = 0.74 100 200 unloop (sec) 300 400 (a) Path (b) 50 40 30 20 10 0 0 loop(sec) 5 [3] K.-C. Tham, R. Kanaar, and J. H. G. Lebbink, DNA [18] J. Jeong and H. D. Kim, bioRxiv (2018), Repair 38, 75 (2016). https://doi.org/10.1101/503490. [4] J. Cadet and J. R. Wagner, Cold Spring Harbor Perspec- tives in Biology 5, a012559 (2013). [5] A. Granzhan, N. Kotera, and M.-P. Teulade-Fichou, [19] K. A. Schallhorn, K. O. Freedman, J. M. Moore, J. Lin, and P. C. Ke, Applied Physics Letters 87, 033901 (2005). [20] C. Yuan, E. Rhoades, X. W. Lou, and L. A. Archer, Chemical Society Reviews 43, 3630 (2014). Nucleic Acids Research 34, 4554 (2006). [6] J. SantaLucia and D. Hicks, Annual Review of Biophysics [21] P. Ranjith, P. S. Kumar, and G. I. Menon, Physical and Biomolecular Structure 33, 415 (2004). review letters 94, 138102 (2005). [7] G. Rossetti, P. D. Dans, I. Gomez-Pinto, I. Ivani, C. Gon- zalez, and M. Orozco, Nucleic Acids Research 43, 4309 (2015). [8] M. Sharma, A. V. Predeus, S. Mukherjee, and M. Feig, The Journal of Physical Chemistry B 117, 6194 (2013). [9] S. Chakraborty, P. J. Steinbach, D. Paul, H. Mu, S. Broyde, J.-H. Min, and A. Ansari, Nucleic Acids Re- search 46, 1240 (2017). [10] A. Dittmore, S. Brahmachari, Y. Takagi, J. F. Marko, and K. C. Neuman, Physical Review Letters 119 (2017). [11] J. D. Kahn, E. Yun, and D. M. Crothers, Nature 368, 163 (1994). [12] R. Vafabakhsh and T. Ha, Science 337, 1097 (2012). [13] C. Tardin, Biochimie 142, 80 (2017). [14] J. Jeong, T. T. Le, and H. D. Kim, Methods 105, 34 (2016). [15] Y.-J. Chen, S. Johnson, P. Mulligan, A. J. Spakowitz, and R. Phillips, Proceedings of the National Academy of Sciences 111, 17396 (2014). [16] T. T. Le and H. D. Kim, Biophysical journal 104, 2068 (2013). [17] T. T. Le and H. D. Kim, Nucleic Acids Research 42, 10786 (2014). [22] A. Vologodskii and M. D. Frank-Kamenetskii, Nucleic Acids Research 41, 6785 (2013). [23] T. A. Lionberger, D. Demurtas, G. Witz, J. Dorier, T. Lillian, E. Meyhfer, and A. Stasiak, Nucleic Acids Research 39, 9820 (2011). [24] R. N. Irobalieva, J. M. Fogg, D. J. Catanese Jr, T. Sut- thibutpong, M. Chen, A. K. Barker, S. J. Ludtke, S. A. Harris, M. F. Schmid, W. Chiu, et al., Nature communi- cations 6, 8440 (2015). [25] T. Sutthibutpong, C. Matek, C. Benham, G. G. Slade, A. Noy, C. Laughton, J. P. K. Doye, A. A. Louis, and S. A. Harris, Nucleic acids research 44, 9121 (2016). [26] M. Mosayebi, A. A. Louis, J. P. K. Doye, and T. E. Ouldridge, ACS Nano 9, 11993 (2015). [27] J. Zhang, Y. Yan, S. Samai, and D. S. Ginger, The Journal of Physical Chemistry B 120, 10706 (2016). [28] S. R. Tee and Z. Wang, ACS Omega 3, 292 (2018). [29] A. R. Haeusler, K. A. Goodson, T. D. Lillian, X. Wang, S. Goyal, N. C. Perkins, and J. D. Kahn, Nucleic acids research 40, 4432 (2012). [30] M. Ganji, I. A. Shaltiel, S. Bisht, E. Kim, A. Kalichava, C. H. Haering, and C. Dekker, Science 360, 102 (2018). [31] J. F. Marko, P. De Los Rios, A. Barducci, and S. Gruber, bioRxiv (2018), https://doi.org/10.1101/325373. Supplemental Material to "Base-pair mismatch can destabilize small DNA loops through cooperative kinking" Jiyoun Jeong and Harold D. Kim∗ School of Physics, Georgia Institute of Technology, 837 State Street, Atlanta, GA 30332-0430, USA (Dated: January 3, 2019) PCA oxygen scavenging system [4] for 10 minutes. We then injected a high salt (1 M [NaCl]) imaging buffer into the flow-cell to promote sticky ends to capture the loop configuration. Decyclization measurements were done similarly, except that the NaCl concentration was changed from 2 M to 75 mM. The immobilized molecules were excited by a 532-nm laser continuously through an objective-type TIR microscope from the beginning of the buffer exchange. The time trajectories of FRET signals (Figure 1(b) of the main text) from the molecules were recorded by an EMCCD camera (DU-897ECS0-# BV, Andor) at a rate of 100 ms per frame for the mismatch- free molecules and 50 ms per frame for the molecules with a mismatch. C. Minicircle simulations The Monte Carlo simulation of a minicircle was im- plemented as previously described [2, 5]. A set of 105 connected nodes was used to create a coarse-grained rep- resentation of a DNA minicircle of 105 bp. The bending energy at each node was described by the kinkable worm- like chain model [5] with the parameters of b = 0.3 and h = 12 following the same notation used in Ref. [6]. We performed the simulation with and without a flexible de- fect of zero bending energy placed at a fixed location. For the case of no flexible spot, we first initialized the sim- ulation without allowing the kink formation. Once the kink-free simulation was equilibrated, we allowed sponta- neous kinks to appear. To construct the probability den- sity of kink positions, we ran the simulation and stop at the first appearance of a kink. We then recorded the po- sition of this kink and equilibrated back to the kink-free state. This procedure was repeated until we collected a distribution of 1000 kink positions. The same procedure was repeated in the presence of the hyperflexible spot to predict the effect of a flexible spot on the probability distribution of kink. I. MATERIALS AND METHODS A. Preparation of DNA molecules A 105-bp-long DNA molecule was extracted from yeast genomic DNA by polymerase chain reaction (PCR) to serve as a control DNA template without any struc- tural defect. To probe the effect of a permanent de- fect, we planned to introduce a DNA mismatch to the control molecule by mixing it with its mutant followed by a strand-exchange reaction. To do so, we addition- ally prepared a set of mutated DNA molecules that dif- fer from the control only in a certain location in which we put a mutation of size equal to 1bp, 3bp, or 5bp. To make such molecules, first, the mutated templates of the control DNA were synthesized from Eurofins Ge- nomics (EXTREMer oligos) and duplexed via PCR. Each of the duplexed products was then incorporated into a pJET1.2\blunt vector (ThermoFisher) and cloned into DH5α Escherichia coli cells. Finally, the cloned frag- ments of DNA were extracted via colony PCR from the cells and were sequenced to ensure the correct mutation was made at the desired location. To modify these molecules to carry a FRET pair (i.e. Cy3 and Cy5), biotin, and single-stranded sticky ends, we followed our standard preparation protocol [1], which involves a series of PCR and strand exchange reactions that can be found in elsewhere. For introducing a DNA mismatch in the final construct, we mixed the Cy3-labled control molecule with one of the Cy5-labeled mutated molecules with a ratio of 4:1 in the strand-exchange re- action. The final DNA construct generated by this protocol carries a 50 protruding sticky end on each end and makes a hairband loop upon end-annealing as shown in Figure 2(a) of the main text. We also made hairpin loops by having sticky ends on the same DNA strand (Figure 3(a) of the main text). A complete list of all DNA sequences can be found in Tables S1 and S2 below. B. single-molecule FRET looing and unlooping assay We followed our previous single-molecule FRET assay that employs the sudden salt-exchange protocol [2, 3]. For cyclization, DNA molecules were deposited on a pas- sivated surface of a flow-cell and were incubated at a low salt (10 mM [NaCl) imaging buffer containing the PCD- arXiv:1901.00263v1 [physics.bio-ph] 2 Jan 2019 2 No mismatch 1bp-mismatch (central) 3bp-mismatch (central) 5bp-mismatch (central) 3bp-mismatch (10 bp off-center) 3bp-mismatch (20 bp off-center) Supplementary Table S1: DNA sequences of hairband molecules. 50−TGAATTTACGTGCCAGCAACAGA[T]AGCCGCGATCGCCATGGCAACGAGGTCG CACACGCCCCACACCCAGACCTCCCTGCGAGCGGGCATGGGTACAATCATTCGAG CTCGTTGTAG-30 30−CACGGTCGTTGTCTATCGGCGCTAGCGGTACCGTTGCTCCAGCGTGTGCGGG GTGTGGGTCTGGAGGGACGCTCGCCCGTACCCATGTTAGTAAGCTCGAGCAAC ATCACTTAAATG-50 50−TGAATTTACGTGCCAGCAACAGA[T]AGCCGCGATCGCCATGGCAACGAGGTCG CACACGCCCCAGACCCAGACCTCCCTGCGAGCGGGCATGGGTACAATCATTCGAG CTCGTTGTAG-30 30−CACGGTCGTTGTCTATCGGCGCTAGCGGTACCGTTGCTCCAGCGTGTGCGGG GTCTGGGTCTGGAGGGACGCTCGCCCGTACCCATGTTAGTAAGCTCGAGCAAC ATCACTTAAATG-50 50−TGAATTTACGTGCCAGCAACAGA[T]AGCCGCGATCGCCATGGCAACGAGGTCG CACACGCCCCGGGCCCAGACCTCCCTGCGAGCGGGCATGGGTACAATCATTCGAG CTCGTTGTAG-30 30−CACGGTCGTTGTCTATCGGCGCTAGCGGTACCGTTGCTCCAGCGTGTGCGGG GCCCGGGTCTGGAGGGACGCTCGCCCGTACCCATGTTAGTAAGCTCGAGCAAC ATCACTTAAATG-50 50−TGAATTTACGTGCCAGCAACAGA[T]AGCCGCGATCGCCATGGCAACGAGGTCG CACACGCCCGCGCGCCAGACCTCCCTGCGAGCGGGCATGGGTACAATCATTCGAG CTCGTTGTAG-30 30−CACGGTCGTTGTCTATCGGCGCTAGCGGTACCGTTGCTCCAGCGTGTGCGGG CGCGCGGTCTGGAGGGACGCTCGCCCGTACCCATGTTAGTAAGCTCGAGCAAC ATCACTTAAATG-50 50−TGAATTTACGTGCCAGCAACAGA[T]AGCCGCGATCGCCATGGCAACGAGGTCG TGGACGCCCCACACCCAGACCTCCCTGCGAGCGGGCATGGGTACAATCATTCGAG CTCGTTGTAG-30 30−CACGGTCGTTGTCTATCGGCGCTAGCGGTACCGTTGCTCCAGCACCTGCGGG GTGTGGGTCTGGAGGGACGCTCGCCCGTACCCATGTTAGTAAGCTCGAGCAAC ATCACTTAAATG-50 50−TGAATTTACGTGCCAGCAACAGA[T]AGCCGCGATCGCCATGGCGGTGAGGTCG CACACGCCCCACACCCAGACCTCCCTGCGAGCGGGCATGGGTACAATCATTCGAG CTCGTTGTAG-30 30−CACGGTCGTTGTCTATCGGCGCTAGCGGTACCGCCACTCCAGCGTGTGCGGG GTGTGGGTCTGGAGGGACGCTCGCCCGTACCCATGTTAGTAAGCTCGAGCAAC ATCACTTAAATG-50 No mismatch 1bp-mismatch (central) 3bp-mismatch (central) Supplementary Table S2: DNA sequences of hairpin molecules. 50−TGAATTTACG(CT)GTGCCAGCAACAGA[T]AGCCACATCGCCATGGCAACGAGG TCGCACACGCCCCACACCCAGACCTCCCTGCGAGCGGGCATGGGTTGCATGTCAG CTATGGATCCATTCGTAAATTCA-30 30−CACGGTCGTTGTCTATCGGTGTAGCGGTACCGTTGCTCCAGCGTGTGCGGGG TGTGGGTCTGGAGGGACGCTCGCCCGTACCCAACGTACAGT(CG)ATACCTAGGT- 50[Cy3] 50−TGAATTTACG(CT)GTGCCAGCAACAGA[T]AGCCACATCGCCATGGCAACGAGG TCGCACACGCCCCAGACCCAGACCTCCCTGCGAGCGGGCATGGGTTGCATGTCAG CTATGGATCCATTCGTAAATTCA-30 30−CACGGTCGTTGTCTATCGGTGTAGCGGTACCGTTGCTCCAGCGTGTGCGGGG TCTGGGTCTGGAGGGACGCTCGCCCGTACCCAACGTACAGT(CG)ATACCTAGGT- 50[Cy3] 50−TGAATTTACG(CT)GTGCCAGCAACAGA[T]AGCCACATCGCCATGGCAACGAGG TCGCACACGCCCCGGGCCCAGACCTCCCTGCGAGCGGGCATGGGTTGCATGTCAG CTATGGATCCATTCGTAAATTCA-30 30−CACGGTCGTTGTCTATCGGTGTAGCGGTACCGTTGCTCCAGCGTGTGCGGGG CCCGGGTCTGGAGGGACGCTCGCCCGTACCCAACGTACAGT(CG)ATACCTAGGT- 50[Cy3] Continued on next page 3 5bp-mismatch (central) 50−TGAATTTACG(CT)GTGCCAGCAACAGA[T]AGCCACATCGCCATGGCAACGAGG TCGCACACGCCCGCGCGCCAGACCTCCCTGCGAGCGGGCATGGGTTGCATGTCAG CTATGGATCCATTCGTAAATTCA-30 30−CACGGTCGTTGTCTATCGGTGTAGCGGTACCGTTGCTCCAGCGTGTGCGGGC GCGCGGTCTGGAGGGACGCTCGCCCGTACCCAACGTACAGT(CG)ATACCTAGGT- 50[Cy3] Both top (50 to 30) and bottom (30 to 50) sequences are shown. The underlined sequences represent sticky ends. A Cy5 fluorophore is internally attached at the thymine base colored in red. A Cy3 fluorophore is either at the green thymine base or the 50 end of the bottom strand. A biotin molecule is linked to the thymine base shown as [T]. Hairpin molecules includes a 2-nt gap (indicated by sequences in parentheses) near each end of the top strand before sticky ends. ∗ Corresponding author. Email: [email protected] [1] T. T. Le and H. D. Kim, Journal of Visualized Experiments (2014), 10.3791/51667. [2] T. T. Le and H. D. Kim, Nucleic Acids Research 42, 10786 (2014). [3] J. Jeong and H. D. Kim, bioRxiv (2018), https://doi.org/10.1101/503490. [4] C. E. Aitken, R. A. Marshall, and J. D. Puglisi, Biophysical Journal 94, 1826 (2008). [5] X. Zheng and A. Vologodskii, Biophysical journal 96, 1341 (2009). [6] A. Vologodskii and M. D. Frank-Kamenetskii, Nucleic Acids Research 41, 6785 (2013).
1605.04978
2
1605
2016-08-09T19:41:17
Effects of RNA branching on the electrostatic stabilization of viruses
[ "physics.bio-ph", "cond-mat.soft", "q-bio.BM" ]
Many single-stranded (ss) RNA viruses self assemble from capsid protein subunits and the nucleic acid to form an infectious virion. It is believed that the electrostatic interactions between the negatively charged RNA and the positively charged viral capsid proteins drive the encapsidation, although there is growing evidence that the sequence of the viral RNA also plays a role in packaging. In particular the sequence will determine the possible secondary structures that the ssRNA will take in solution. In this work, we use a mean field theory to investigate how the secondary structure of the RNA combined with electrostatic interactions affects the efficiency of assembly and stability of the assembled virions. We show that the secondary structure of RNA may result in negative osmotic pressures while a linear polymer causes positive osmotic pressures for the same conditions. This may suggest that the branched structure makes the RNA more effectively packaged and the virion more stable.
physics.bio-ph
physics
Effects of RNA branching on the electrostatic stabilization of viruses Gonca Erdemci-Tandogan,1 Jef Wagner,1 Paul van der Schoot,2, 3 Rudolf Podgornik,4, 5, 6 and Roya Zandi1 1Department of Physics and Astronomy, University of California, Riverside, California 92521, USA 2Group Theory of Polymers and Soft Matter, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands 3 Institute for Theoretical Physics, Utrecht University, Leuvenlaan 4, 3584 CE Utrecht, The Netherlands 4Department of Physics, University of Massachusetts, Amherst, MA 01003 5Department of Theoretical Physics, J. Stefan Institute, SI-1000 Ljubljana, Slovenia 6Department of Physics, University of Ljubljana, SI-1000 Ljubljana, Slovenia Many single-stranded (ss) RNA viruses self assemble from capsid protein subunits and the nucleic acid to form an infectious virion. It is believed that the electrostatic interactions between the negatively charged RNA and the positively charged viral capsid proteins drive the encapsidation, although there is growing evidence that the sequence of the viral RNA also plays a role in packaging. In particular the sequence will determine the possible secondary structures that the ssRNA will take in solution. In this work, we use a mean field theory to investigate how the secondary structure of the RNA combined with electrostatic interactions affects the efficiency of assembly and stability of the assembled virions. We show that the secondary structure of RNA may result in negative osmotic pressures while a linear polymer causes positive osmotic pressures for the same conditions. This may suggest that the branched structure makes the RNA more effectively packaged and the virion more stable. I. INTRODUCTION Many single-stranded (ss) RNA viruses package their genome concurrently with the self-assembly of the whole capsid in such a way, that small protein subunits spon- taneously assemble around the nucleic acid to built a complete protein shell (capsid) [1]. In the prevail- ing paradigm this assembly is predominantly driven by generic, nucleotide sequence independent, electrostatic interactions [2] between the negative charges on the RNA phosphate backbone and the positive charges on the virus capsid proteins (CP) [3 -- 8]. Recent experiments have in- deed abundantly verified the importance of the "charge- matching hypothesis", based on the preponderance of electrostatic interactions between the capsid proteins and the RNA for proper genome packaging [9]. However, besides the importance of electrostatics, packaging experiments suggest that there must exist a correlation between the specific details of the nucleic acid structure and the efficient virus assembly [10 -- 13]. In a beautifully designed experiment Comas-Garcia et al. [10] have set the viral RNA1 of Brome Mosaic Virus (BMV) and the RNA of Cowpea Chlorotic Mottle Virus (CCMV) to compete against each other for capsid proteins be- longing to CCMV exclusively. Although both RNAs are of similar length, BMV RNA was shown to outcompete the CCMV RNA, therefore suggesting that electrostat- ics alone is not enough for efficient genome encapsidation and that further structural details of RNA, apart from its generic charge, could play a role in the genome encap- sidation [10, 14]. Even further away from the presumed non-specificity of the genome - CP interactions are indications, from both in vitro and in vivo studies, that the capsid self- assembly is achieved via a directed capsid assembly me- diated by the highly specific, non-electrostatic interac- tions between sections of RNA and capsid proteins; these sections of RNA are thought to contain packaging sig- nals and are repeated along the genome according to the symmetry of the capsid [15]. Contrary to the generic electrostatic charge matching, the essence of the pack- aging signal hypothesis is thus that the viral genomes have local secondary or tertiary structures with high CP affinity, serving as heterogeneous nucleation sites for the formation of capsids [16, 17]. Quite interestingly, in a recent experiment on Satellite Tobacco Mosaic Virus (STMV), Sivanandam et al. find that reducing the num- ber of charges on the N-terminal section of capsid pro- teins through mutations results into the encapsidation of shorter RNAs than the wild type ones. However, unex- pectedly a single mutation in one specific location along the N-terminal completely stops the self-assembly [13]. Investigating the nature of how and which structural de- tails of RNA could be important for virus assembly is thus urgently required to ascertain on which point along the axis of "charge-matching" to "packaging signals" hy- potheses the viruses actually drive and regulate their as- sembly. Viral RNAs are found to be compact and highly branched [18] due to the base-pairing between the nu- cleotides, engendering compactification and folding of the molecule. Indeed, it appears that the compactness of the ssRNA wild-type viral genomes is one of the principal characteristics of their nucleotide sequence, setting them distinctly apart from randomized sequences [11, 19], and that the physical compactness of the viral genome can be regarded as a primary factor among evolutionary con- straints [20]. While theoretical arguments suggest that the details of the RNA structure are important for its efficient pack- aging in the small volume of the virus capsid [13, 21 -- 25], it remains overall poorly understood how the RNA sequence chemical composition together with its length affect the compactification and the packaging efficiency. Based on simple scaling arguments, it has been shown that genome secondary structures, or more specifically branching, lower the free energy of RNA encapsidation [21, 22]. As far as the length of RNA is concerned, there is a clear correlation with the number of positive charges on the virus coat proteins, structurally due to their ex- tended N-tails, for many ssRNA viruses [22, 23, 26 -- 28]. This correlation ratio is ∼ 1.6 for many wild type viruses [27], implying that the number of negative charges on the RNA is in fact larger than the number of positive charges on the protein motifs, making these viruses overcharged. Furthermore, when virus coat proteins encapsidate a linear polymer, e.g., poly(styrene sulfonate) (PSS), two different results are obtained: both highly overcharged (correlation ratio ∼ 9 [29]) and undercharged (correla- tion ratio between 0.45 and 0.6 [30]) virus-like parti- cles (VLP). The overcharging phenomenon has been dis- cussed in many theoretical papers with different conclu- sions dependening mostly on the details of the model under consideration [26 -- 33]. What one would hope for is that the important characteristics of the RNA genome packaging would robustly depend on some well defined characteristics of the genome, a hypothesis recently pro- posed in our work [24], where we showed that the sec- ondary structure of RNA, as quantified by its branchi- ness, coupled to electrostatic interactions enhances the genome encapsidation capacity and could robustly ex- plain the overcharging actually observed in virions. While understanding the detailed role of electrostat- ics and structure of RNA on the self-assembly is the fo- cus of what follows, we also aim additionally to under- stand what controls the virions or VLP stability or what the main factors are that enhance this stability before the disassembly of the capsid. Viruses seem to release their genome during the disassembly [34], which would imply that the genome not just leaves, but is in fact ac- tively pushed from the capsid - a scenario that has been shown as specifically valid for bacteriophages, where the repulsive DNA-DNA interactions act like a coiled osmotic spring ejecting the genome. The corresponding osmotic pressure is in fact quite large and positive, surpassing even 50 atm, and stemming mostly from the combina- tion of electrostatic and hydration interactions that are dominant in the range of DNA densities relevant for bac- teriophage packing [2]. Contrary to DNA in bacteriophages, the osmotic pres- sure in ssRNA viruses is not easy to measure directly and in the absence of experiments one thus has to rely on the- oretical estimates. There have been several theoretical studies that investigate the osmotic pressure of ssRNA viruses [28, 31, 35 -- 37]. Siber and Podgornik showed that the filled ssRNA virions exhibit a small residual negative osmotic pressure, which depends strongly on the amount of capsid charges and can be turned positive with rela- 2 tively higher capsid charge [28]. In addition, Javidpour et al. studied the effects of multivalent ions, which can fundamentally change the nature of electrostatic interac- tions [38], on the osmotic pressure and the stability of the virus like empty shells, showing that the multivalent ions can turn a positive electrostatic osmotic pressure into a negative one [36]. Furthermore, recent all atom molecular dynamics simulations showed that the osmotic pressure inside an empty Poliovirus capsid is negative, suggesting that the mechanism might be connected with excess charges on the capsid that prevent the solution ion to exchange with the capsid [37], a scenario at odds with what we know about the permeability of capsids. While there have thus been several lines of investigation regarding the nature and specifically the sign of the cap- sid osmotic pressure, there exist no studies taking into account the role of the secondary structure of RNA in the osmotic pressure of ssRNA viruses or virus like par- ticles, another aspect that we elucidate further below. In this paper, we extend our previous analysis and in- vestigate how the secondary structure of the RNA affects the osmotic pressure of ssRNA viruses and what are the repercussions for stability of the virions. We show that the secondary structure of RNA may indeed result in negative osmotic pressures at conditions where a linear polymer would exhibit positive osmotic pressures. This may suggest that having a branched structure makes not only RNA more effectively packaged but also makes a virion more stable. The paper is organized as follows. In the next section, we introduce the model and the funda- mentals of the theory together with the basic quantities that we will calculate. In Sec. III, we present the results for osmotic pressure as well as the effect of RNA branch- ing on the free energy minimum, defining the optimum length of RNA, the optimum number of branched points and the optimum charge ratios of the system, together with the corresponding ion concentration and RNA den- sity profiles. Section IV discusses effects of different mod- els, boundary conditions and different parameterizations that might correspond to different types of viruses. Fi- nally, we summarize our findings. In the appendix, we derive in detail the model free energy of the encapsida- tion. II. MODEL To elucidate the role of genome in the assembly of spherical RNA viruses, we model RNA as a generic, neg- atively charged, flexible branched polyelectrolyte that in- teracts with positive charges residing on the inner surface of the capsid. More specifically, we consider only the case of annealed branched polymers because the strength of RNA base-pairing is relatively weak and may easily be affected by the interaction with the positive inner surface charges of the shell during encapsidation. For simplicity, we model the capsid as a thin sphere and assume that the charges are not localized but smeared out uniformly on the inner surface of the sphere. We note that while a thin shell is a good approximation for the capsid of some viruses like Dengue and yellow fever [39], the capsid pro- teins of some other viruses contain N-terminal tails which are highly positively charged and point into the capsid cavity in a brush-like fashion [26]. (cid:90) The mean-field free energy functional of a polyelec- trolyte chain confined within a charged shell in a univa- lent salt solution, under the ground state approximation, can be written as (cid:104) a2 6 ∇Ψ(r)2 + W(cid:2)Ψ(r)(cid:3) (cid:105) ∇Φ(r)2 − 2µ cosh(cid:2)βeΦ(r)(cid:3) + βτ Φ(r)Ψ2(r) (cid:90) (cid:105) βF = − β2e2 (cid:104) d3r 8πλB + d2r βσ Φ(r) . (1) Here β denotes the inverse of the thermal energy kBT , a the statistical step (Kuhn) length of the polymer, τ the linear charge density of the polymer, σ the surface charge density of the shell, Ψ(r) the monomer density field at position r, and Φ(r) the mean electrostatic poten- tial. The parameter µ is the fugacity of the monovalent salt ions corresponding to the concentration of salt ions in the bulk. λB = e2β/4π0, is the Bjerrum length, a measure of the dielectric constant () of the solvent and is about 0.7 nm for water at room temperature. The first term of Eq. (1) is the entropic cost of non- uniform polymer density and the last two lines of Eq. (1) correspond to the electrostatic interactions between the polymer, the shell and the salt ions on the level of the Poisson-Boltzmann theory [28]. The standard form of this free energy can be found in references [28, 40]. For completeness we also provide a step by step derivation of Eq. (1) for a linear polymer in the appendix. The self-interaction term W [Ψ] in Eq. (1) is associated with the self repulsion of the polyelectrolyte and the en- ergy of an annealed branched polymer [41 -- 44], W [Ψ] = 1 2 υΨ4 − 1√ a3 (feΨ + a3 6 fbΨ3), (2) where υ is the excluded volume term and fe and fb are the fugacities of the end- and branch-points of the annealed polymer, respectively. A detailed derivation of Eq. 2 is given in [45]. In this model, the stem-loop or hair-pin configurations of RNA are counted as the end points. The number of end- and branch-points Ne and Nb of the polymer are related to the fugacities fe and fb in a standard way by Ne = −βfe Nb = −βfb and (3) . ∂F ∂fe ∂F ∂fb We have two additional constraints in the problem. First, the total number of monomers inside the capsid is fixed [46], (cid:90) 3 a constraint that we enforce by introducing a Lagrange multiplier, E, when minimizing the free energy. Second, the number of the end points depends on the number of branched points so that Ne = Nb + 2, (5) since we consider only a single polymer with no closed loops. Thus, fe is not a free parameter. For our calcu- lations, we change fb and find fe through Eqs. (3) and (5). The polymer is linear if fb = 0, and the number of branched points increases with fb. By varying the free energy functional with respect to fields Ψ(r) and Φ(r), we obtain a coupled set of non-linear differential equations coupling the monomer density with the electrostatic potential in the interior of the capsid, and the usual Poisson-Boltzmann equation for the exte- rior of the capsid. The monomer density field in fact satisfies the modified Edwards equation a2 6 ∇2Ψ(r) = −EΨ(r) + βτ Φin(r)Ψ(r) + 1 2 ∂W ∂Ψ , (6) (8) while the electrostatic potential satisfies the modified Poisson-Boltzmann equation in the interior of the cap- sids ∇2Φin(r) = 1 Dβe sinh(cid:2)βeΦin(r)(cid:3) − Dµβe2 Ψ2(r), (7) and the standard Poisson-Boltzmann equation in the ex- terior 2λ2 λ2 τ Dβe sinh(cid:2)βeΦout(r)(cid:3), ∇2Φout(r) = 1 λ2 √ where λD = 1/ 8πλBµ is the Debye screening length. The boundary condition (BC) for the electrostatic poten- tial is obtained by minimizing the free energy, n·∇Φin − n·∇Φout = 4πλBσ/βe2 assuming the surface charge den- sity σ is fixed. The concentration of the polymer out- side of the capsid is assumed to be zero. The BC for the inside monomer density field Ψ is of Neumann type (n·∇Ψs = 0), that can be obtained from the energy minimization [46]. However, due to the short-ranged self- repulsions of the polymer, Dirichlet type BC (Ψs = 0) might be preferable so that the polymer density goes to zero on the surface of the capsid. In our calculations we use both types of BCs and find that our conclusions do not depend on their detailed nature so that our conclu- sions are robust. We start with the Neumann BC but discuss the impact of the Dirichlet BC later in Sec. IV. Using Eq. (1), we can also obtain the osmotic pressure due to the genome encapsidation, i.e., the force exerted on the virus capsid by the genome per unit surface area, defined as (cid:18) ∂F ∂V (cid:12)(cid:12)(cid:12)(cid:12)Qc,N (cid:19) (cid:12)(cid:12)(cid:12)(cid:12)Qc,N =0 − ∂F ∂V P (N ) = − , (9) N = d3r Ψ2(r), (4) where V is the volume of the capsid and we subtracted the part of the osmotic pressure for the empty capsid. In the calculation of the pressure, we keep the total number of monomers N and the total number of charges on the capsid Qc = 4πb2σ constant with b the radius of the capsid. III. RESULTS We numerically solve the nonlinear coupled differential equations, Eqs. (6), (7), (8), subject to the constraints given in Eqs. (4) and (5) to obtain the fields Ψ and Φ and the parameter fe. Electrostatic potential and poly- mer concentration profiles as a function of r, the dis- tance from the center of the shell, are shown in Fig. 1(a) and (b), respectively for 10mM (solid and dashed lines) and for 100mM (dotted and dotted-dashed lines) salt concentrations for a linear polymer with fb = 0 (solid and dotted lines) and a branched polymer with fb = 3.0 (dashed and dotted-dashed lines). The total number of monomers enclosed in the shell is N = 1000 for both profiles shown in the figure. Independent of the amount of salt and degree of branching, the polymer concentra- tion is always larger right next to the surface due to the electrostatic attraction between the polymer and capsid, but it is higher for the branched polymers than the linear one (Fig. 1 (b)). Note that in all cases the genome pro- files remain nearly constant inside the shell but increase noticeably in the vicinity of the capsid wall. In addition, we investigated the distribution of branch and end points inside the capsid for 10mM and for 100mM salt concentrations. Figure 1(c) illustrates the concentration of endpoints Ce(r) = 1√ a3 feΨ(r) (solid line for 10 mM and dotted line for 100 mM ) and branch √ a3 6 fbΨ3(r) (dashed lines for 10 mM and points Cb(r) = dotted-dashed lines for 100 mM ), obtained from Eq. (3). As shown in Fig. 1(c), the number of branch points in- creases in the vicinity of the capsid wall at both salt concentrations; however, it increases even more at the lower salt concentration indicating more segments inter- act with the wall. The end points, on the other hand, mainly distributed over the interior of the shell. Figure 1(d) shows the fractions of end points Ce/C (solid lines for 10 mM and dotted line for 100 mM ) and fraction of branch points Cb/C (dashed lines for 10 mM and dotted- dashed lines for 100 mM ) as a function of r. Once the fields Ψ and Φ are obtained, we insert them into Eq. (1) to calculate the free energy of chain-capsid complex, F . To obtain the encapsidation free energy, F , we need to calculate the free energy of a chain free in solution and that of a positively charged capsid and then subtract them both from the chain-capsid complex free energy, F given in Eq. (1). The capsid self-energy (F (N = 0)) due to the elec- trostatic interactions is calculated through Eqs. (7) and (8) in the limit as N → 0, and should be explicitly sub- tracted from the encapsidation free energy. The focus of this paper is on the solution conditions in which the capsid proteins can spontaneously self-assemble in the 4 FIG. 1: For N = 1000 and two different salt concentrations µ corresponding to 10 mM (solid and dashed lines) and 100 mM (dotted and dotted-dashed lines), (a) Electrostatic po- tential profile for a linear polymer with fb = 0 (solid and dot- ted lines) and branched polymer with fb = 3.0 (dashed and dotted-dashed lines) (b) Concentration profile corresponding to two different degree of branching for a linear polymer with fb = 0 (solid and dotted lines) and for a branched polymer with fb = 3.0 (dashed and dotted-dashed lines). (c) Concen- tration profile of endpoints (solid and dotted lines) and branch points (dashed and dotted-dashed lines) for a branched poly- mer with fb = 3.0. (d) Fraction of endpoints (solid and dotted lines) and branch points (dashed and dotted-dashed lines) for a branched polymer with fb = 3.0. Other parameters are υ = 0.5 nm3, τ = −1 e, σ = 0.4 e/nm2, b = 12 nm, a = 1 nm and T = 300 K. absence of genome as seen in different kind of experi- ments [6, 47]. Note that the free energy associated with a free chain (both linear and branched) is negligible un- der the experimental conditions [22, 28, 31]. To avoid the problem of proper free energy rescaling, we furthermore calculate the osmotic pressure of RNA trapped inside the capsid and investigate the impact of its secondary struc- ture on the stability of capsid. Through the calculation of osmotic pressure, we have been able to confirm all our conclusions obtained through the free energy calculation. In order to get the osmotic pressure, we first calcu- late the free energy of the system as a function of the monomer number N for both linear and branched chains and then insert it in Eq. (9). A plot of the osmotic pres- sure P vs. the monomer number N is given in Fig. 2(a) for both linear and branched polymers at two different salt concentrations. The solid and dotted lines correspond to linear polymers with fb = 0 and dashed and dotted- dashed lines to branched polymers with fb = 3.0. The salt concentrations are 10 mM (solid and dashed lines) and 100 mM (dotted and dotted-dashed lines). As is clear from the figure, the osmotic pressure goes through a minimum and this minimum is displaced towards longer chains as we increase the degree of branching, i.e., more monomers can be encapsidated with increasing fb. For example, the minimum of pressure is at N ≈ 523 for a linear polymer fb = 0, and increases to N ≈ 851 for a (a)(b)(c)(d)8.09.010.011.012.0-0.05-0.04-0.03-0.02-0.01r(nm)Φ(V)8.09.010.011.012.0-0.015-0.0050.0050.0156.07.08.09.010.011.012.00.00.20.40.60.81.0r(nm)(CeC,CbC)9.09.510.010.511.011.512.00.00.20.40.60.81.0r(nm)C(nm-3)9.010.011.012.00.00.20.40.60.88.09.010.011.012.00.00.10.20.30.4r(nm)(Ce,Cb)(nm-3) 5 (a) Osmotic pressure as a function of monomer FIG. 2: numbers for a linear polymer with fb = 0 (solid and dot- ted lines) and a branched polymer with fb = 3 (dashed and dotted-dashed lines). Solid and dashed lines correspond to the salt concentration, µ = 10 mM and dotted and dotted- dashed lines represent the salt concentration, µ = 100 mM . (b) Osmotic pressure for N = 1200 as a function of fugacity of branch points, fb, at 10 mM (dotted lines) and 100 mM (dotted-dashed lines) salt concentrations. Other parameters are υ = 0.5 nm3, τ = −1 e, σ = 0.4 e/nm2, b = 12 nm, a = 1 nm and T = 300 K. branched polymer with fb = 3 at 100 mM salt. At 10 mM salt, the minimum of the free energy is at N ≈ 628 for fb = 0 and at N ≈ 719 for fb = 3. Figure 2 (b) shows the osmotic pressure in terms of the degree of branching fb for 10 mM (dotted lines) and 100 mM (dotted-dashed lines) salt concentrations with N = 1200. When fb = 0 (linear polymer), the osmotic pressure is positive but changes the sign as fb increases regardless of the salt concentration. The figure shows that the pressure becomes more negative as the degree of branching increases indicating that the secondary struc- ture of the genome makes the virus more stable. To further investigate the role of branching on the as- sembly of viral shells, we study the impact of branch- ing on the minimum free energy, the optimal number of monomers, the optimal number of branched points, and the ratio of the chain charge to the capsid charge. A plot of the encapsidation optimum free energy Fmin vs. the FIG. 3: For 10 mM (dotted lines) and 100 mM (dotted- dashed lines) salt concentrations, (a) Optimum free energy (units of kBT ) (b) Optimum number of monomers (c) Ratio of number of branched points to the number of monomers at the minima (d) Ratio of number of polymer charges to the capsid charges at the minima as a function of fugacity of branch points, fb. Other parameters are υ = 0.5 nm3, τ = −1 e, σ = 0.4 e/nm2, b = 12 nm, a = 1 nm and T = 300 K. branching fugacity fb is given in Fig. 3(a) at two differ- ent salt concentrations. For branched polymers, the free energy becomes deeper, indicating that compared to the linear polymers, the branchiness confers more stability to the capsid at both salt concentrations. This effect could explain why some RNAs are encapsidated more efficiently than others, or indeed linear polyelectrolytes. Note that the effect of branching is more apparent at high salt con- centrations. Expectedly, for low salt concentrations, elec- trostatics overwhelms all the other interactions and the impact of branching becomes less pronounced; neverthe- less, the minimum moves towards the longer chains for branched polymers compared to linear ones. Figure 3(b) shows the optimal number of encapsidated monomers associated with the minimum of free energy as a function of fb. As illustrated in the figure, more monomers are packaged as the degree of branching in- creases. For example, at 100 mM for a linear polymer, fb = 0, the optimum number of monomers is N ≈ 534 and it increases to N ≈ 1211 for a branched polymer with fb = 3.0. At 10 mM salt, the optimum monomer number for linear polymer is N ≈ 638 and for branched one is Nmin ≈ 773 , with fb = 3.0. Figure 3(c) is a plot of the ratio of number of branched points to the opti- mal number of monomers vs. the branching fugacity. As expected, the ratio increases for higher fb values. The fact that longer, branched chains can be more 0500100015002000-3-2-10123NP(cid:1)105Pa(cid:2)(a)(b)0.00.51.01.52.02.53.0-2.0-1.5-1.0-0.50.00.5fbP(cid:1)105Pa(cid:2)(b)(c)(d)0.00.51.01.52.02.53.0-1400-1200-1000-800-600-400-200fbFmin(kBT)(a)0.00.51.01.52.02.53.060080010001200fbNmin0.00.51.01.52.02.53.00.000.050.100.150.200.25fbNbNmin0.00.51.01.52.02.53.0-1.6-1.4-1.2-1.0-0.8fbQpQc easily encapsidated by capsid proteins could straightfor- wardly explain one of the reasons why viruses are over- charged. The total charge of the virion is Q = Qp + Qc = τ N + 4πb2σ where the first term corresponds to the genome charge and the second one to that of the cap- sid. Figure 3(d) shows the charge ratio of the genome to the capsid vs. the fugacity of branched points for two different salt concentrations at the minima of the free en- ergy for υ = 0.5 nm3, τ = −1 e, σ = 0.4 e/nm2, b = 12 nm, a = 1 nm and T = 300 K. The virion becomes over- charged for the values of fb > 2 at 10 mM and fb > 1 at 100 mM . IV. DISCUSSION AND SUMMARY We have investigated the role of RNA sequence speci- ficity, as it transpires through the RNA branchiness in the electrostatic encapsidation of RNA viruses. Specifically, we addressed in detail the dependence of the free energy and the osmotic pressure of a confined self-interacting RNA constrained within a spherical, charged capsid. The sequence specificity was modeled through an annealed distribution of RNA end- and branch-points, and the electrostatics was addressed within a mean-field Poisson- Boltzmann framework, allowing us to study explicitly the impact of branching and genome-capsid electrostatic interaction on the optimal length of the encapsidated genome. While the details of our model can be subject to criticism and RNA sequence specificity could enter on other more detailed levels of description, we do believe that the coupling between RNA self-interaction and cap- sid electrostatics represents a robust mechanism of en- capsidation and virion stabilization. To confirm that the results derived within our model of RNA branching, corresponding to a simple description of the RNA secondary structure, are indeed robust we also propose an alternative self-interacting linear chain model of RNA based on the assumption that RNA can be de- scribed as a linear polymer, i.e., possesses no branch- points and only two end-points, but self-interacts with short-ranged attractive interactions describing the self- pairing of RNA segments [40]. As for the rest, we as- sume again that the capsid wall can be modeled as a thin, charged spherical shell with uniform surface charge density. The free energy corresponding to this model is again given by Eq. (1), except that the polymer chain is now linear, implying that fe, fb −→ 0, and the self interaction term W [Ψ] thus changes to W [Ψ] = (v − a3βsw)Ψ4 + 1 2 1 6 uΨ6, (10) (11) with s the average fraction of self-interacting chain seg- ments, i.e., base-pairs, and w is the corresponding short- range binding energy. Note that we included the next, 6 Ψ6 term in the virial expansion in Eq. 11, with u > 0 in order to stabilize the free energy since (v − a3βsw) can in general become negative. Variation of the free en- ergy yields the same Euler-Lagrange equations as given in Eqs. (6), (7), (8) subject to the constraint, Eq. (4). The results of this calculation are presented in Fig. 4 that illustrates the encapsidation free energy as a func- tion of the number of monomers, N . As illustrated in the figure, the positions of the free energy minima move towards longer polymers (larger N ) and the depth of the minima increase with increasing s, the average fraction of bound segments. At 10 mM salt, Fig. 4 shows that the minimum of the encapsidation free energy is located at N = 632 for s = 0 and at N = 740 for s = 0.04. The effect is again more pronounced at 100 mM salt in which the location of the minimum moves from N = 524 for s = 0 to N = 903 for s = 0.04. w is chosen 1 kBT and u = 0.5 nm6 in our calculations. It thus seems that this rather different model, though presenting the same salient features of the system, yields the same qualitative behavior as discussed above for branched polymers. This substantiates our claim that the coupling between RNA self-interaction and capsid electrostatics represents a robust mechanism of encapsi- dation and virion stabilization. In addition to investigating the different ways of mod- eling the secondary structures of RNA, we also studied the impact of different boundary conditions on the en- capsidation free energy and osmotic pressure. While all the results presented above correspond to the Neumann BC, n.∇Ψs = 0, we found that our conclusions do not depend on the type of BCs in that we obtained quali- tatively the same results for the Dirichlet BC, Ψs = 0. Although the Dirichlet BC changes the polymer density profile (see the inset of Fig. 5), the behavior of the free energy and the osmotic pressure remains qualitatively remarkably unaffected in that the minimum of the free energy does get deeper and moves towards longer chains as branching increases. As is clear from Fig. 5, at 100 mM salt the minimum of the free energy at N ≈ 401 for a linear polymer with fb = 0, is displaced to N ≈ 1103 for a branched polymer with fb = 8.5 when the Neumann BC is replaced by the Dirichlet BC for the polymer den- sity field. Furthermore, for the Dirichlet BC at 10 mM salt, the free energy minimum is displaced from N = 599 for fb = 0 to N = 735 for fb = 8.5. Note that the value of fb used for Dirichlet is chosen such that the ratio of num- ber of branch points to the number of total monomers is almost the same as those for Neumann case. We also calculated the osmotic pressure for Dirich- let BC using both branched and self-interacting linear chains. Consistent with the free energy results, we found that as the degree of branching or the average fraction of self-interacting chain segments increases, the osmotic pressure as a function of N becomes more negative and its minimum moves towards longer chains. Further, we examined the impact on the free energy of the capsid surface charge density (0.3 ≤ σ ≤ 0.9), 7 polymer charge density (−2.0 ≤ τ ≤ −0.5) and Kuhn length (0.5 ≤ a ≤ 2.0). For both Dirichlet and Neumann BCs, we found that the optimal number of encapsidated monomers for linear chains is always such that number of charges on the polymer is less than those on the capsid, i.e., the virus-like particles (VLP) are undercharged. In contrast, we found that the optimal length of the encapsi- dated branched polymers is larger than that of the linear polymers for all cases examined, resulting in overcharg- ing of VLPs in many cases. We emphasize that while our findings are consistent with previous mean-field PE theories in that the VLPs with a linear polymer is under- charged [28], our results for linear polymers differ from recent numerical simulations [23] and the scaling theories [22] on the assembly of viral particles. While the over- charging for linear polymers, observed in Ref. [22] is due to the charges on the N-terminals and in Ref. [23] could be due to the solution conditions or the protein charge distribution, it is found that the branched structure of the polymer enhances overcharging, consistent with our studies. It is difficult to determine the topology of large single- stranded viral RNAs in solution, but recent experiments indicate that the secondary structure does play an im- portant role in the efficient packaging of RNA [10, 14]. The secondary structures can be predicted using a num- ber of softwares, such as RNAsubopt (a program in the Vienna RNA package [48]), RNAfold (another program in the Vienna RNA package [48]) and mfold [49]. All these software tools, that are progressively unreliable for longer chains, estimate the free energy changes according to the base-pairing and the loop closure of ssRNA and the secondary structure of RNA results from base-pairing of G, U, C and A nucleotides. RNAfold and mfold calcu- late the possible sets of base-pairing corresponding to the minimum free energy, while RNAsubopt has an option to generate Boltzmann weighted secondary structures which can be used to calculate a meaningful ensemble average of any quantity. This software was successfully used [11, 20] to calculate the maximum ladder distance (MLD) and we applied RNAsubopt to calculate the thermally averaged number of branch points for RNA1 of BMV and CCMV to shed light on the experiments noted in the introduction on the competition between RNA1 of CCMV and BMV. We generated the ensemble of secondary structures us- ing the RNA1 sequences of both BMV and CCMV ob- tained form the National Center for Biotechnology Infor- mation Genome Database [50], and then calculated the thermally averaged number of branched points of RNA1 of BMV and CCMV. We found that RNA1 of BMV has 65 branched points vs. 60.5 branched points of RNA1 of CCMV [51]. These numbers confirm the experimental results of Comas-Garcia et al. [10] that RNA1 of BMV would be preferentially packaged over RNA1 of CCMV. We note that although these programs were designed for the short RNAs, many important results have been ex- tracted through finding the ensemble average of the de- sired quantities for viral genomes of length 2500− 10000 FIG. 4: Encapsidation free energy (units of kBT ) as a func- tion of monomer number for a self-interacting linear chain model with s=0 (solid and dotted) and s=0.04 (dashed and dotted-dashed lines) at two different values of µ, correspond- ing to salt concentrations 10 mM (solid and dashed lines) and 100 mM (dotted and dotted-dashed lines). The arrow indi- cates the monomer number at which the full virus particle is neutral (Qp = Qc). Inset shows the position of the minimum Nmin vs. the average fraction of self-paired bases, s, for 100 mM salt concentration. Other parameters are υ = 0.5 nm3, w = 1 kBT , u = 0.5 nm6, τ = −1 e, σ = 0.4 e/nm2, b = 12 nm, a = 1 nm and T = 300 K. FIG. 5: Encapsidation free energy (units of kBT ) vs monomer numbers for a linear chain with fb = 0 (solid and dotted lines) and a branched chain with fb = 8.5 (dashed and dotted-dashed lines) at two different sal concentrations µ, 10 mM (solid and dashed lines) and 100 mM (dotted and dotted- dashed lines) with the Dirichlet BC. The arrow indicates the monomer number at which the full virus particle is neutral (Qp = Qc). Other parameters take the values υ = 0.05 nm3, τ = −1 e, σ = 0.4 e/nm2, b = 12 nm, a = 0.5 nm and T = 300 K. Inset shows the concentration profile for N=1000 with two different branching fugacities, fb = 0 (linear chain) for the dotted line, and fb = 8.5 (branched chain) for the dotted-dashed lines. ◆◆◆◆0200400600800100012001400-1500-1000-5000500100015002000NF(kBT)0.000.04500700950Nmins◆◆◆◆0200400600800100012001400-1500-1000-5000500100015002000NF(kBT)0.04.08.012.00.00.350.65C(nm-3)r(nm) 8 (cid:90) υ 2 (cid:90) N 0 dr ρ2 m(r)+ dsV (r(s)) drdr(cid:48) ρc(r)υc(r − r(cid:48))ρc(r(cid:48)). (13) ds r2(s)+ (cid:90) (cid:90) where βH = 3 2a2 (cid:90) N 0 + β 2 The first term in Eq. (13) describes the ideal entropy of the chain, the second corresponds to the short range steric repulsions between monomers and the third term is an external potential acting on the chain. The last term corresponds to the electrostatic interactions between the charges of monomers and ions. In Eq. (13) υc is the Coulomb interaction υc = 1 4π0 1 r − r(cid:48) , (14) and ρc is the charge density operator given by (cid:90) N N +(cid:88) 0 ρc(r) = τ ds δ(r − r(s)) N−(cid:88) + e δ(r − r+ i ) − e δ(r − r− i ) + ρ0(r). (15) i i Here, τ is the uniform monomer charge density along the polyelectrolyte and ρ0(r) is the charge density of the in- ner wall capsid in this system. To calculate the following integral in the partition function (cid:82) (cid:82) drdr(cid:48) ρc(r)υc(r−r(cid:48)) ρc(r(cid:48)), (16) we introduce a local charge density ρc(r) and its auxiliary field φ(r) using the following identity 1 = D[ρc(r)]δ(ρc(r) − ρc(r)) D[ρc(r)]D[φ(r)]eiβ(cid:82) dr(ρc(r)− ρc(r))φ(r) = (17) (cid:90) (cid:90) (cid:90) nucleotides [11, 20]. The theoretical models presented in this paper clearly indicate the important role of the secondary structure of RNA on the assembly of ssRNA viruses. The sec- ondary structure can be indeed invoked to explain the overcharging observed in RNA viruses, while it promotes the efficiency of RNA packaging by increasing the com- pactness of RNA in order to better fit into a small capsid. As shown above, the secondary structure of RNA clearly effects the osmotic pressure of the capsid; regardless of the details of the model as well as calculational details such as the form of the BCs, we obtain consistently nega- tive osmotic pressures resulting from the presence of the negatively charged chain. The osmotic pressure becomes more negative for a branched polymer compared to the linear one. Non-specific electrostatic interactions have emerged as the driving force for virus assembly through both the ex- perimental as well as the theoretical studies [9, 14, 24, 27, 28]. In our two simple models we generalized the imple- mentation of electrostatic interactions by coupling it to RNA topology. While this is an important step in real- ism of the modeling, the present level of description still cannot include the specific interactions (or packaging sig- nals) into a complete picture of virus assembly. Further investigations on both specific and non-specific interac- tions could help understanding the structure of viruses and take steps on the development of antiviral drugs. This work was supported by the National Science Foundation through Grant No. DMR-13-10687 (R.Z.). R. P. acknowledges the financial support of the Agency for research and development of Slovenia (ARRS) under Grants No. P1-0055 and J1-7435. The authors would like to thank the Aspen Center for Physics where some of the work was discussed during the Physics and Mathematics of Viral Assembly workshop. V. APPENDIX Derivation of the free energy We consider RNA as a single polyelectrolyte in a good solvent in the presence of salt ions. There are N monomers of the polyelectrolyte chain, N + positive and N− negative salt ions in the solvent. The microscopic de- grees of freedom are the position of the monomers (r(s)) and positive (r+ i ) ions . The partition function can be written as path integral over all configu- rations i ) and negative (r− (cid:90) Acknowledgments Zsalt = D[r+ i ]D[r− i ]e− β 2 where the second line is the Fourier transform of the delta function. The auxiliary field φ(r) will turn out to be the electrostatic potential. We then replace the density operator ρc by the corresponding fluctuating density field ρc [52]. Multiplying Eq. 16 by Eq. 17 and using Eqs. 14 and 15 and the Hubbard-Stratonovich transformation, we find (cid:90) D[φ(r)] ( (cid:82) dr(∇φ(r))2 (cid:90) (cid:90) dre−iβeφ(r))N + dreiβeφ(r))N− 0 dsφ(r(s))e−iβ(cid:82) drρ0(r)φ(r). e−iβτ(cid:82) N ( Zsalt = e− β0 2 (18) Z = Dr(s)Dr+ i Dr− i e−βH (12) We use the same procedure as above to obtain the con- tribution of excluded volume interaction to the partition (cid:90) dreiβeφ(r))N− = exp − dr ∇Ψ0(r)2 + V (r)Ψ0(r)2 D[r(s)]e−βH1[r(s)] ≈ e−N E0 = e −N min{ <Ψ0HΨ0> <Ψ0Ψ0> } (cid:18) (cid:90) (cid:26) a2 6 + iβτ φ(r)Ψ0(r)2 + iυ ψ(r)Ψ0(r)2 − λ(Ψ0(r)2 − N V ) (cid:27)(cid:19) (25) function, 2 υ(cid:82) dr ρ2 (cid:90) e− 1 m(r) = D[ψ(r)]e− 1 2 υ(cid:82) drψ2(r)e−iυ(cid:82) N 0 ds ψ(r(s)), (19) with ψ the auxiliary field representing the monomer den- sity field. Plugging Eqs. 18 and 19 into Eq. 12, we find the partition function (cid:90) (cid:90) (cid:90) Z[N +, N−] = D[r(s)]D[φ(r)]D[ψ(r)] ( dre−iβeφ(r))N + (cid:82) N 0 ds r2(s)−(cid:82) N 0 dsφ(r(s))−iβ(cid:82) drρ0(r)φ(r) (cid:82) dr(∇φ(r))2−iβτ(cid:82) N 2 υ(cid:82) drψ2(r)−iυ(cid:82) N e− 1 0 ds ψ(r(s)). e− β0 0 dsV (r(s)) − 3 2a2 e ( 2 (20) We now switch to the grand-canonical ensemble modify- ing only the terms associated with the salt ions Ξ[µ] = µN ++N− N +!N−! Z[N +, N−], (21) with µ the fugacity (density) of the monovalent salt ions related to the concentration of salt ions in the bulk. In- serting Eq. 20 into Eq. 21, the grand canonical partition function can be written as Ξ = D[φ(r)]D[ψ(r)]e−βH1[φ(r),ψ(r)] (cid:90) ∞(cid:88) N± (cid:90) D[r(s)]e−βH2[r(s)] (22) with the effective free energies (cid:90) N (cid:16) 3 ds βH1[r(s)] = 0 and βH2[φ(r), ψ(r)] = (cid:90) 2a2 r2(s) + V (r(s)) + iβτ φ(r(s)) (23) + iυ ψ(r(s)) (cid:17) (cid:16) β0 (∇φ(r))2 + iβρ0(r)φ(r) 2 dr − 2µ cos(βeφ(r)) + υψ2(cid:17) 1 2 (24) 9 The polymer part of the partition function is similar to 6 ∇2 + the Feymann integral of the Hamiltonian H = − a2 U (r) with the potential U (r) = V (r) + iβτ φ(r) + iυ ψ(r) and imaginary time t → is [40]. We assume that the chain is very long (total number of monomers N → ∞) with a well defined energy gap such that the ground state approximation is valid. Thus, we have with Ψ0 the eigenfunction and E0 the eigenenergy of the ground state. The Lagrange multiplier λ is introduced to normalize the wave function. Plugging Eq. 25 into Eq. 22 and integrating out the ψ field, we find the grand canonical partition function as (cid:90) Ξ = D[Φ(r)]e−βF (26) with βF = (cid:90) (cid:26) a2 6 dr ∇Ψ0(r)2+V (r)Ψ0(r)2+βτ Φ(r)Ψ0(r)2 + 1 2 υΨ0(r)4 − λ(Ψ0(r)2 − N V ) − β0 2 ∇Φ(r)2 + βρ0(r)Φ(r) − 2µ cosh(βeΦ(r)) (27) (cid:27) where we introduce the transformation Φ → iφ with Φ being the mean electrostatic potential. Due to the ab- sence of an external potential, V (r) = 0 and the capsid charge density is ρ0(r) = σδ(z) with σ the surface charge density. This leads then to Eq. 1 considering the con- straint given in Eq. 4. Note that Eq. (27) is for a linear chain with f1 = 0 and f3 = 0. For branched polymers in the absence of electrostatic interactions, see Ref. [45]. [1] M. Rossmann and V. Rao, Viral molecular machines, vol. 726 (Springer Science & Business Media, 2011). [2] A. Siber, A. L. Bozic, and R. Podgornik, Phys. Chem. Chem. Phys. 14, 3746 (2012), 5905. [3] P. Ni, Z. Wang, X. Ma, N. C. Das, P. Sokol, W. Chiu, B. Dragnea, M. Hagan, and C. C. Kao, J. Mol. Biol. 419, 284 (2012). [4] P. van der Schoot and R. Zandi, Phys. Biol. 4, 296 (2007). 10 [5] A. Losdorfer Bozic, A. Siber, and R. Podgornik, J. Biol. (2008). Phys. 39, 215 (2013). [29] Y. F. Hu, R. Zandi, A. Anavitarte, C. M. Knobler, and [6] A. Zlotnick, R. Aldrich, J. M. Johnson, P. Ceres, and W. M. Gelbart, Biophys. J. 94, 1428 (2008). M. J. Young, Virology 277, 450 (2000). [7] H.-K. Lin, P. van der Schoot, and R. Zandi, Phys. Biol. 9, 066004 (2012). [8] R. Kusters, H.-K. Lin, R. Zandi, I. Tsvetkova, B. Drag- nea, and P. van der Schoot, J. Phys. Chem. B 119, 1869 (2015). [9] R. F. Garmann, M. Comas-Garcia, M. S. T. Koay, J. J. L. M. Cornelissen, C. M. Knobler, and W. M. Gelbart, J. Virol. 88, 10472 (2014). [10] M. Comas-Garcia, R. D. Cadena-Nava, A. L. N. Rao, C. M. Knobler, and W. M. Gelbart, J. Virol. 86, 12271 (2012). [11] A. M. Yoffe, P. Prinsen, A. Gopal, C. M. Knobler, W. M. [30] R. D. Cadena-Nava, Y. F. Hu, R. F. Garmann, B. Ng, A. N. Zelikin, C. M. Knobler, and W. M. Gelbart, J. Phys. Chem. B 115, 2386 (2011). [31] P. van der Schoot and R. Bruinsma, Phys. Rev. E 71, 061928 (2005). [32] C. L. Ting, J. Z. Wu, and Z. G. Wang, PNAS 108, 16986 (2011). [33] M. F. Hagan and R. Zandi, Curr. Opin. Virol. 18, 36 (2016). [34] W. H. Roos, I. L. Ivanovska, A. Evilevitch, and G. J. Wuite, Cell Mol Life Sci. 64, 1484 (2007). [35] R. F. Bruinsma, Euro. Phys. J. E 19, 303 (2006). [36] L. Javidpour, A. L. Boi, A. Naji, and R. Podgornik, J. Gelbart, and A. Ben-Shaul, PNAS 105, 16153 (2008). Chem. Phys. 139, 154709 (2013). [12] F. Li Tai, W. M. Gelbart, and A. Ben-Shaul, J. Chem. Phys. 135, 155105 (2011). [13] V. Sivanandam, D. Mathwes, R. Garmann, G. Erdemci- Tandogan, R. Zandi, and A. Rao, Scientific Reports 6, 26328 (2016). [37] Y. Andoh, N. Yoshii, A. Yamada, K. Fujimoto, H. Ko- jima, K. Mizutani, A. Nakagawa, A. Nomoto, and S. Okazaki, J. Chem. Phys. 141, 165101 (2014). [38] A. Naji, M. Kandu c, J. Forsman, and R. Podgornik, J. Chem. Phys. 139, 150901 (2013). [14] R. F. Garmann, M. Comas-Garcia, C. M. Knobler, and [39] A. L. Bozic, A. Siber, and R. Podgornik, J. Biol. Phys. W. M. Gelbart, Acc. Chem. Res. 49, 48 (2016). 38, 657 (2012). [15] P. Stockley, R. Twarock, S. Bakker, A. Barker, A. Boro- davka, E. Dykeman, R. Ford, A. Pearson, S. Phillips, and N. Ranson, J. Biol. Phys. 39, 277 (2013). [16] N. Patel, E. Dykeman, R. Coutts, G. Lomonossoff, D. Rowlands, S. Phillips, N. Ranson, R. Twarock, R. Tuma, and P. Stockley, Proc. Natl. Acad. Sci. 112, 2227 (2015). [40] I. Borukhov, D. Andelman, and H. Orland, Euro. Phys. J. B 5, 869 (1998). [41] T. C. Lubensky and J. Isaacson, Phys. Rev. A 20, 2130 (1979). [42] T. T. Nguyen and R. F. Bruinsma, Phys. Rev. Lett. 97, 108102 (2006). [43] S. I. Lee and T. T. Nguyen, Phys. Rev. Lett. 100, 198102 [17] R. Zandi, P. van der Schoot, D. Reguera, W. Kegel, and (2008). H. Reiss, Biophys. J. 90, 1939 (2006). [44] K. Elleuch, F. Lequeux, and P. Pfeuty, J. Phys. I France [18] R. F. Garmann, A. Gopal, S. S. Athavale, C. M. Knobler, W. M. Gelbart, and S. C. Harvey, RNA 21, 877 (2015). [19] G. Erdemci-Tandogan, J. Wagner, P. van der Schoot, and R. Zandi, J. Phys. Chem. B 120, 6298 (2016). [20] L. Tubiana, A. L. Bozic, C. Micheletti, and R. Podgornik, Biophysical J. 108, 194 (2015). [21] R. Zandi and P. van der Schoot, Biophys. J. 96, 9 (2009). [22] P. van der Schoot and R. Zandi, J. Biol. Phys. 39, 289 (2013). [23] J. D. Perlmutter, C. Qiao, and M. F. Hagan, eLife 2, e00632 (2013). [24] G. Erdemci-Tandogan, J. Wagner, P. van der Schoot, R. Podgornik, and R. Zandi, Phys. Rev. E 89, 032707 (2014). [25] S. W. Singaram, R. F. Garmann, C. M. Knobler, W. M. Gelbart, and A. Ben-Shaul, J. Phys. Chem. B. 119, 13991 (2015). [26] T. Hu, R. Zhang, and B. I. Shklovskii, Physica A 387, 3059 (2008). [27] V. A. Belyi and M. Muthukumar, PNAS 103, 17174 (2006). [28] A. Siber and R. Podgornik, Phys. Rev. E 78, 051915 5, 465 (1995). [45] J. Wagner, G. Erdemci-Tandogan, and R. Zandi, J. Phys.:Condens. Matter 27, 495101 (2015). [46] H. Ji and D. Hone, Macromolecules 21, 2600 (1988). [47] L. Lavelle, M. Gingery, M. Phillips, W. M. Gelbart, C. M. Knobler, R. D. Cadena-Nava, J. R. Vega-Acosta, L. A. Pinedo-Torres, and J. Ruiz-Garcia, J. Phys. Chem. B 113, 3813 (2009). [48] I. L. Hofacker, W. Fontana, P. F. Stadler, L. S. Bonho- effer, M. Tacker, and P. Schuster, Monatsh. Chem. 125, 167 (1994). [49] M. Zuker, Nucleic Acids Res. 31, 3406 (2003). [50] www.ncbi.nlm.nih.gov. [51] The standard deviation for the number of branch points in the ensemble of RNA secondary structures is around 2.5 for the RNA1 sequence of both BMV and CCMV. [52] A. G. Moreira and R. R. Netz, Electrostatic Effects in Soft Matter and Biophysics, Chapter: Field-Theoretic Approaches to Classical Charged Systems, vol. 46 of NATO Science Series (Springer Netherlands, 2001).
1211.0990
1
1211
2012-11-05T20:22:30
Ensemble-based characterization of unbound and bound states on protein energy landscape
[ "physics.bio-ph", "q-bio.BM" ]
Characterization of protein energy landscape and conformational ensembles is important for understanding mechanisms of protein folding and function. We studied ensembles of bound and unbound conformations of six proteins to explore their binding mechanisms and characterize the energy landscapes in implicit solvent. First, results show that bound and unbound spectra often significantly overlap. Moreover, the larger the overlap the smaller the RMSD between bound and unbound conformational ensembles. Second, the analysis of the unbound-to-bound changes points to conformational selection as the binding mechanism for four of the proteins. Third, the center of the unbound spectrum has a higher energy than the center of the corresponding bound spectrum of the dimeric and multimeric states for most of the proteins. This suggests that the unbound states often have larger entropy than the bound states considered outside of the complex. Fourth, the exhaustively long minimization, making small intra-rotamer adjustments, dramatically reduces the distance between the centers of the bound and unbound spectra as well as the spectra extent. It condenses unbound and bound energy levels into a thin layer at the bottom of the energy landscape with the energy spacing that varies between 0.8-4.6 and 3.5-10.5 kcal/mol for the unbound and bound states correspondingly. At the same time, the energy gap between the two lowest states in the full ensemble varies between 0.9 and 12.1 kcal/mol. Finally, the analysis of protein energy fluctuations showed that protein vibrations itself can excite the inter-state transitions, thus supporting the conformational selection theory.
physics.bio-ph
physics
Ensemble-based characterization of unbound and bound states on protein energy landscape Anatoly M. Ruvinsky1*, Tatsiana Kirys1,2, Alexander V. Tuzikov2, and Ilya A. Vakser1,3 1Center for Bioinformatics, The University of Kansas, Lawrence, Kansas 66047, USA 2United Institute of Informatics Problems, National Academy of Sciences, 220012 Minsk, Belarus 3Department of Molecular Biosciences, The University of Kansas, Lawrence, Kansas 66045, USA *To whom correspondence should be addressed: email: [email protected] 1 Abstract Characterization of protein energy landscape and conformational ensembles is important for understanding mechanisms of protein folding and function. We studied ensembles of bound and unbound conformations of six proteins to explore their binding mechanisms and characterize the energy landscapes in implicit solvent. First, results show that bound and unbound spectra often significantly overlap. Moreover, the larger the overlap the smaller the RMSD between bound and unbound conformational ensembles. Second, the analysis of the unbound-to-bound changes points to conformational selection as the binding mechanism for four of the proteins. Third, the center of the unbound spectrum has a higher energy than the center of the corresponding bound spectrum of the dimeric and multimeric states for most of the proteins. This suggests that the unbound states often have larger entropy than the bound states considered outside of the complex. Fourth, the exhaustively long minimization, making small intra-rotamer adjustments, dramatically reduces the distance between the centers of the bound and unbound spectra as well as the spectra extent. It condenses unbound and bound energy levels into a thin layer at the bottom of the energy landscape with the energy spacing that varies between 0.8-4.6 and 3.5-10.5 kcal/mol for the unbound and bound states correspondingly. At the same time, the energy gap between the two lowest states in the full ensemble varies between 0.9 and 12.1 kcal/mol. Finally, the analysis of protein energy fluctuations showed that protein vibrations itself can excite the inter-state transitions, thus supporting the conformational selection theory. 2 Introduction Relationships between protein energy landscape, structure, and function have been a subject of numerous studies resulted in the development of the funnel shape energy landscape theory (1-5). This theory has been further extended by the conformational selection paradigm to include the ensemble-based description of proteins and protein-protein interactions (5-8). The concept suggests that bound and bound-like conformations may co-exist in solution within a large ensemble of unbound conformations. By shifting equilibrium in the unbound ensemble towards the bound-like conformations, binding forces select a bound conformation corresponding to the free energy minimum. Recent studies focused on reconstruction of the native ensembles (9-12). An ensemble of ubiquitin structures reflecting dynamics up to the microsecond time scale was refined against residual dipolar couplings. All crystallographically determined bound conformations of ubiquitin were found within 0.8 Å root mean square deviation (RMSD) of the Cα atoms (11). Another RDC-optimized ensemble of ubiquitin consistent with the microsecond time scale dynamics was created by Monte Carlo sampling of the ‘‘Backrub’’ motions (12). Cold denaturation and protein encapsulation were combined with NMR to probe the ensemble (13). Single-molecule experiments corroborated the theory of multiple interconverting conformations and revealed their relation to the fluctuating catalytic reactivity (14). Room-temperature X-ray crystallography was able to detect such conformations in proline isomeraze (15). Best et al. (9) showed that an ensemble of highly homologous X-ray structures can also reproduce structural diversity in the native ensemble probed by NMR spectroscopy in solution. A protocol combining molecular dynamics (MD) simulations of an X-ray structure with information from the NMR relaxation experiments has been suggested for studying protein conformational ensembles in solution(16). In general MD simulations have been instrumental in mapping the conformational space (17-19). Alternative methods for generating conformational ensembles without solving explicit equations of motion have been actively developed (see Ref. (20) for a review). Large conformational ensembles are routinely used in protein structure prediction (21, 22) and studies of allosteric interactions (23, 24). Despite the significant progress achieved in generating protein ensembles, their energetic properties and relation to the unbound-to-bound conformational changes are not well understood (25). How to generate a bound-like structure from the unbound one is one of the main problems in structure prediction of protein complexes. Although MD simulations showed that the interface side chains – “anchor residues” - sample bound-like conformations (26, 27), criteria for selecting such conformations from the MD snapshots are yet to be determined. Current docking protocols are much more successful when bound conformations are used, but become less reliable in a common case when only unbound structures are known (28, 29). To advance the docking protocols, the relation between the energy landscape and conformational changes upon binding should be unraveled. Recent large-scale studies of conformational changes upon binding focused on the relationship between single bound and single unbound conformations (30-32). However, how well the change between two selected conformations characterizes transition between the unbound and bound states within conformational ensembles as well as the transformation of the (free) energy landscapes is still unclear. In this study, we investigate structural similarity between the ensembles of bound and unbound conformations for six proteins and characterize their energy landscapes in implicit solvent. We 3 consider impact of the energy minimization in the Generalized Born (GB) model on the distance between the ensembles of bound and unbound conformations and the ensembles’ spectral properties (the energy spacing, the spectrum gap between the lowest states, and the spectrum width). Our focus on the energy minimization in implicit solvent was motivated by a recent study (33) showing that ranking protein structures by minimized GB energies can distinguish the near- native structures from decoys better than ranking based on the energy minimization either in vacuum or explicit solvent. First, our study shows that although the shortly minimized GB energies of the bound and unbound ensembles often significantly overlap, the center of the unbound spectrum tends to have a higher energy than the centers of the bound spectrum of the dimeric and multimeric states. Moreover, the larger the overlap the smaller the RMSD between bound and unbound conformational ensembles. Second, the existence of the structurally different equipotential states in both ensembles suggests that unbound states have larger entropy than the bound states. The entropy-driven modeling of the unbound-to-bound conformational changes suggests a novel direction in advancing protein-protein docking algorithms, which, in fact, commonly neglect entropy effects. Third, the results show that the bound conformations of the RNase A interface pre-exist in the unbound ensemble, indicating conformational selection as the binding mechanism. Pancreatic trypsin inhibitor, ubiquitin, and lysozyme C also have high similarity between bound and unbound interfaces as well as small deviations that can be attributed to flash cooling (34, 35) or variations in the crystallization conditions. Fourth, the exhaustively long minimization by the Adopted Basis Newton-Raphson algorithm results in small mostly intra-rotamer adjustments that dramatically reduce the distance between the centers of the bound and unbound spectra as well as the spectra extent. It condenses unbound and bound energy levels into a thin layer at the bottom of the energy landscape. At the same time, the whole spectrum from the shortly minimized states to the bottom of the folding funnel can cover up to 40.3% of the lowest energy, indicating that the folded states may significantly differ in energy. The average energy spacing at the bottom of the energy landscape varies between 0.8-4.6 and 3.5-10.5 kcal/mol for the unbound and bound states correspondingly. The energy gap between the two lowest states varies between 0.9 and 12.1 kcal/mol. Finally, the results show that protein vibrations itself can stimulate the inter-state transitions, thus supporting the conformational selection theory. We suggest an approach for estimating the number of normal modes involved in conformational transition and show that, on average, 20 low-energy normal modes are needed to describe transition between two neighboring energy states. At the same time, transitions between the two lowest states may involve an order of magnitude larger number of the modes. Results and Discussion Bound and unbound energy bands Figure 1 shows minimized energy spectra of six proteins (see Methods) represented by the ensembles of their conformations determined by X-ray crystallography and NMR (Tables 1 and S1). Short and long minimizations (SM and LM) were applied to characterize the topography of the energy landscape in the Generalized Born model (see Methods). The energy minimizations caused small in-rotamer re-adjustments of the exposed side chains resulting in a typical RMSD≤0.7 Å between the minimized and the non-minimized structures, which did not substantially change neither the sizes of the bound and unbound ensembles nor the distance between them (Table 2). Nevertheless, these changes were enough to significantly condense both spectra of the unbound and bound proteins. Figure 1 shows that the span of the spectra and the 4 spacing between energy minima after LM are significantly smaller than that after SM. The ratio between the overall energy span (including the SM and LM bands) in the unbound ensemble and the lowest energy in the protein spectrum is 40.3% for ovomucoid, 13.7% for PTI, 26.9% for ubiquitin, 21.5% for RNase A, 22.8% for CheY and 24.2% for lysozyme C. Excluding the lowest ratio for PTI as an outlier, the ratio for other five proteins decreases 1.7 times, with a 2.5 times increase of the number of atoms from ovomucoid to lysozyme C. The ratio decrease is expected because the lowest energy is a function of the total number of protein residues, whereas the energy extent relates mainly to the surface residues that are able to change their conformations in solution. The outlying ratio for PTI may result from the insufficient size of its unbound ensemble, which is the smallest among the proteins in our set (Table 1). Ovomucoid has 51 residues, compared to 56 residues of PTI, but its unbound ensemble is 4.6 times larger than the ensemble of PTI. The ratio between the energy span of the LM band and the lowest energy is 7.0% for ovomucoid, 4.9% for PTI, 14.0% for ubiquitin, 5.5% for RNase A, 5.0% for CheY and 3.1% for lysozyme C (Fig. 2). Thus, LM reduces the width of the energy bands, condensing protein states into a thin layer at the bottom of the energy landscape. The energy distance (the ruggedness) between the centers of the SM and LM energy bands was calculated as an arithmetic mean of energies in a protein ensemble after short and long minimizations (see Methods). Interestingly, the energies of the unbound ensembles decrease more upon minimization than the energies of the bound ensembles (Fig. S1) despite the fact that both ensembles have equipotential energy levels (Fig. 1). On average, the unbound proteins lose 0.6 kcal/mol per heavy atom or 4.6 kcal/mol per residue. The bound structures from dimers and multimers lose less: 0.4 and 0.3 kcal/mol per heavy atom or 3.1 and 2.5 kcal/mol per residue accordingly. The centers of the unbound energies after SM were higher than the centers of the bound energies for all proteins, with the exception of ubiquitin. This suggests that the unbound-to-bound conformational changes guided by inter-molecular interactions often follow a path that decreases the internal energy of the binding proteins. Such mechanism increases the binding affinity. The energy decrease can be achieved by improving the interface packing upon binding. The prevalence of the disorder-to-order interface transitions over the reverse transitions corroborates this hypothesis (30). The two-sample t-test showed statistical significance of the difference between the SM-bands’ centers at the 5% level for all proteins, except ubiquitin (Table S2). LM resulted in a significant decrease of the distance between the centers and a cancellation of the statistical significance of the difference between the centers of the unbound energies and bound energies for dimeric states of lysozyme C and ovomucoid, and multimeric states of PTI and lysozyme C. Comparison of the centers of the bound spectra of the structures extracted from dimers and multimers (Table S3) shows that LM resulted in statistically significant difference between the corresponding centers for RNase and PTI. For these proteins, the center of the dimeric bound band of the dimer states is lower than the center of the multimeric bound band. protein states have equal energies !! ≈ !! , then the choice of the bound state is guided by the entropy contribution to the binding free energy !!,! = !"#!!,! (36), where !!,!  is the number of An overlap between the unbound and bound energies, shown in Figure 1, suggests that microstates associated with the unbound macrostates 1 and 2, and  ! is the gas constant. An conformational selection of a bound conformation may be guided by entropy. Indeed, if two 5 unbound macrostate contributes –(! − !") = −! + !"#$ to the binding free energy. Therefore, Thus, if !! ≈ !! , a less-populated conformation is the most effective binder, in agreement with a less-populated state increases the negative value of the binding free energy to a lesser extent. the concept of conformational selection (37). Lower entropy means that protein conformations that are candidates for its bound state reside in narrow energy basins formed by the intra- molecular interactions. We intend to verify this hypothesis in our future study. The entropy- driven modeling of the bound-like conformations may improve performance of protein-protein docking algorithms, which commonly neglect entropy effects. Energy spacing, fluctuations and conformational changes Figure 3 summarizes calculations of the energy spacing in the ensembles. The spacing in the unbound ensembles averaged over the protein set is 2.9 kcal/mol, which is two times less than the average spacing in the bound ensembles. On average, the dimeric and multimeric states are separated by 6.0 kcal/mol and 6.4 kcal/mol accordingly. Larger variations of the energy spacing in bound ensembles may be a result of a smaller size of these ensembles (Table 1). Considering the proteins separately, one can see that the energy spacing between the unbound states varies between 0.8 and 4.6 kcal/mol, which includes the Hyeon and Thirumalai’s estimate of 0- 3kcal/mol for the energy landscape roughness or the barrier (38) and 3.2-3.5kcal/mol barrier measured by single-molecule dynamic force spectroscopy for a complex of GTPase Ran and the nuclear transport receptor importin-b (39). Note that the average spacing of 2.9 kcal/mol is approximately equal to the maximum barrier found in the Hyeon and Thirumalai study (38) but less then the maximum barrier found by Nevo et al (39). The energy spacing between the bound states is larger and falls in the intervals of 3.5-10.5kcal/mol and 3.8-10.2kcal/mol for the dimeric and multimeric states correspondingly. The energy gap between the two lowest energy minima in the joint ensemble of the bound and unbound states is 0.9 kcal/mol for RNase A, 6.1 kcal/mol for PTI, 12.1 kcal/mol for CheY, 7.9 kcal/mol for ubiquitin, 6.5 kcal/mol for ovomucoid and 9.9 kcal/mol for Lysozyme C. The smallest energy gap was found for RNase A, which has the smallest interface and all-atom RMSD between bound and unbound states (Table 2). On the other hand, the largest gap in the CheY spectrum corresponds to the largest RMSD between its bound and unbound ensembles. ensemble can be calculated as !" = ∆!! = !" !! , where !! = !! + !! + !! (!)  is a To explore whether an inter-state transition in principle can result from protein vibrations, one specific heat capacity per molecule at constant volume, !! = !! = 3/2 are the contributions of can estimate a number of the normal modes that needs to be involved in the protein energy ! is the three translational and three rotational degrees of freedom, and !! ! = !!! fluctuation equal to the inter-state barrier. The molecular energy fluctuation in a canonical contribution of the 3!!" − 6 internal vibrations in a protein with !!" atoms. For low-energy ! modes ℏ!! ≪ !" and therefore !!! (!)   ≈ 1 and !! ! ≈ ! + !!! ! , where ! is a number of the low-energy modes. For high-energy modes  ℏ!! ≫ !" and !!! (!) exponentially goes to !!!!! transition between any two states separated by the !" barrier: ! = !" !" ! − 3. It gives 23 zero. One can find the low-bound estimate for the number of low-energy modes needed for a and 32 normal modes for the standard ambient temperature of 298.15K and the barrier of 3.0 and 3.5 kcal/mol correspondingly. Since functional modes are often found among the lowest 20-30 modes (40-44), we can suggest that the protein vibrations indeed can excite the inter-state 6 transitions related to protein function. Thus, an external stimulus (e.g., a ligand or a partner- protein) may not be needed for changing protein conformation, which supports the conformational selection paradigm. Interestingly, many more modes are needed for an energy fluctuation covering the distance between the two lowest states in all the proteins in our set, except RNase A, which shows the smallest distance between bound and unbound ensembles. This is supported by a study of conformational changes in myosin, calmodulin, NtrC, and hemoglobin (45), which showed that the first 20 modes contribute ≤50% of the conformational changes in these molecules. The first 30 modes of the [AChET]4–ColQ complex account for 75% of the conformational change in the tetramer (43). For a typical 1000 atoms protein having 2994 normal modes, a fluctuation of 12.1 kcal/mol is achieved when at least 13% of the normal modes get involved. It was shown that the protein energy fluctuation can increase up to 38 kcal/mol (46), which is more than enough for a transition over the largest gap/barrier considered in this study. Distance between unbound and bound conformational ensembles and binding mechanisms The size of the ensemble is controlled by ambient parameters (temperature, pH, salt concentration etc.) and dependent on protein sequence composition (Table 2). Protein-protein interactions can either select a group of bound-like conformations from the unbound ensemble or transform the whole ensemble into a new group of bound-like conformations. To find out which mechanism takes place, all-atom and interface RMSDs were calculated between all bound and unbound structures (Table 2). The smallest distance between the bound and unbound ensembles was found for RNase A and lysozyme C. RNase A shows the all-atom/interface RMSD of 0.7/0.3 Å. The ensembles of lysozyme C are separated by 0.8/0.7 Å of the all-atom/interface RMSD. Interestingly, the unbound ensemble of lysozyme C encompasses X-ray structures only, and RNase A has the second largest share (35%) of X-ray structures in its unbound ensemble among the proteins in our set. The low bound of the unbound-to-bound all-atom and interface RMSDs varies within 0.7-1.9 Å and 0.3-1.6 Å. The largest distance between the bound and unbound ensembles of CheY corresponds to the smallest overlap between their bound and unbound SM spectra (Figure 1), which disappears after LM. Thus, the majority of the SM and all LM bound conformations of CheY have lower energies than the unbound ones. It is likely that the entropy discussed above makes these lower-energy states unfavorable for the unbound ensemble. In addition to the RMSD analysis, we calculated the share of the unbound residues within 1 Å of their bound conformations for all pairs of the bound and unbound structures (Figures 4 and 5). This metric also showed the lowest similarity between the CheY ensembles at 0.39 level (39% of all the residues, Fig. 5). The ensembles of RNase A, lysozyme C, and PTI had the highest similarity at 0.9, 0.89 and 0.86 levels correspondingly. Comparison of the bound and unbound interfaces revealed 13 unbound structures of RNase A with all interface residues within 1Å from the bound conformations (Fig. 4). Note that 2 Å is a typical size of a rotamer (32). Thus, dimerization of RNase A with another protein can be completely described by the conformational selection mechanism (5-8). Contrary to that, forming a multimer involving RNase A invokes induced fit to expose its C-terminal (Fig. S2), which forms an interface β- strand that swaps with the N-terminal helix in the RNase trimer. None of the unbound structures has the exposed C-terminal. This further suggests that some proteins may employ various 7 binding mechanisms, from the induced fit, to “lock-and-key” and conformational selection, and their combination, depending on the binding partner (6, 7). Interestingly, flash cooling used to determine approximately 90% of macromolecular structures (34) results in a 0.2 - 0.8 Å backbone RMSD between the structures determined at cryogenic and room temperatures (35). It can also change the conformational distribution of up to one third of the protein side chains (47). Taking this into account, we can assume that crystallographic conditions may distort the structure by 1 Å RMSD of all atoms. The low bound of the all-atom RMSD between bound and unbound ensembles (Table 2) suggests that, in addition to RNase A, the conformational selection likely guides binding processes of pancreatic trypsin inhibitor, ubiquitin (11), and lysozyme C. Materials and Methods Generation of the Protein Set To compile a set of proteins with multiple bound and unbound conformations, a subset of protein complexes with small changes in the backbone upon binding (all-atom RMSD ≤ 2 Å) was selected from the non-redundant DOCKGROUND set 3.0 (48). The subset covers 71% of the DOCKGROUND set of 233 complexes. The subset was narrowed down to proteins that are monomers in the unbound state of the biological assembly. Their sequences were used to identify homologous proteins in PDB (sequence identity > 98% by BLAST (49)). The unbound protein structures with small ligands were excluded. All PDB entries found for each query-protein were put into three ensembles: unbound monomers, dimers and multimers. Only proteins with more than five unbound and bound structures were retained. Selected structures were analyzed for disordered residues and mutations. If some of the structures had a disordered terminal, it was deleted in all members of the ensemble. All fragments with ≤ 3 disordered residues at the interface and ≤ 5 at the non-interface were reconstructed by a program Profix from the Jackal package (http://wiki.c2b2.columbia.edu/honiglab_public/index.php/Software). Structures with disordered fragments longer than five residues were discarded. Point mutations were reversed by Profix. The resulting set consisted of six proteins (Tables 1 and S1) with multiple X-ray and NMR-derived bound and unbound conformational states and 100% sequence identity between the states. Minimization Protocol The MMTSB Tool Set (50) and the Generalized Born method that calculates Born radii by analytic volume integration (CHARMM: GBMV method 2) were used to minimize solvation free energy of the proteins (51, 52). The method was parameterized to accurately reproduce electrostatic solvation energies from standard Poisson theory. A non-polar contribution to the solvation free energy was calculated by the ASP model considering the exposed surface area (53). Each protein was subjected to 50 steps of the steepest descent minimization (short minimization; SM) followed by 104 steps of the Adopted Basis Newton-Raphson minimization (long minimization; LM). The CHARMM22 force field was used. The dielectric constant was set to 1 for protein and 80 for solvent. In minimizations, a switching function was used for truncating long-range interactions between 16Å and 18Å. Each bound protein was minimized within its complex to keep interface unchanged. The analysis showed that protein energy 8 changed ≤1.5% between 500 and 104 steps of LM. The average RMSD between all heavy atoms of the initial and minimized structures after short minimization was 0.1 Å. LM produced the average all-atom RMSD at 0.7 Å between the initial and the minimized structures. As can be seen from Table 2, LM did not change substantially the RMSD-based size of the conformational ensembles and the distance between the unbound and bound ensembles. lowest energy was calculated as the absolute value of 100% ∙   (Δ!! + Δ!! )/!! , where !! is the Characterization of the energy spectrum lowest protein energy in the joint ensemble of bound and unbound structures, and Δ!!,! are the The ratio of the spectrum width in the ensemble of the SM and LM unbound structures to the overlap in the SM and LM ensembles, then the ratio was calculated as 100% ∙   (!!"# − !!"# )/!! , where !!"# ,!"# are the lowest and the highest energies in the unbound spectrum. energy span in the unbound ensemble after the SM and LM correspondingly. If the energy spans The ruggedness of the energy landscape was calculated as !!! − !!" , where !!" , !!" are the !!!" − !!!!!" /(! − 1) = !!!! (!!!" − !!!" ) (! − 1), where {!!!" } is an ordered set of the LM energies, ! is the number of average energies in a protein ensemble after SM and LM accordingly. The energy spacing was calculated as the average distance between energy levels: structures in the LM ensemble. Acknowledgements This study was supported by grant R01GM074255 from the NIH. 9 1. 16. 17. 19. 6. 7. 4. 5. 8. 9. 2. 3. Frauenfelder H, Sligar SG, & Wolynes PG (1991) The energy landscapes and motions of proteins. Science 254(5038):1598-1603. Dill KA & Chan HS (1997) From Levinthal to pathways to funnels. Nat Struct Biol 4(1):10-19. Wolynes PG, Onuchic JN, & Thirumalai D (1995) Navigating the folding routes. Science 267(5204):1619-1620. Dill KA, Ozkan SB, Shell MS, & Weikl TR (2008) The protein folding problem. Annu Rev Biophys 37:289-316. Tsai CJ, Kumar S, Ma B, & Nussinov R (1999) Folding funnels, binding funnels, and protein function. Protein Sci 8(6):1181-1190. Boehr DD, Nussinov R, & Wright PE (2009) The role of dynamic conformational ensembles in biomolecular recognition. Nat Chem Biol 5(11):789-796. Csermely P, Palotai R, & Nussinov R (2010) Induced fit, conformational selection and independent dynamic segments: an extended view of binding events. Trends Biochem Sci 35(10):539-546. Tsai CJ, Ma B, & Nussinov R (1999) Folding and binding cascades: shifts in energy landscapes. Proc Natl Acad Sci U S A 96(18):9970-9972. Best RB, Lindorff-Larsen K, DePristo MA, & Vendruscolo M (2006) Relation between native ensembles and experimental structures of proteins. Proc Natl Acad Sci U S A 103(29):10901-10906. 10. Wlodarski T & Zagrovic B (2009) Conformational selection and induced fit mechanism underlie specificity in noncovalent interactions with ubiquitin. Proceedings of the National Academy of Sciences of the United States of America 106(46):19346-19351. Lange OF, et al. (2008) Recognition dynamics up to microseconds revealed from an RDC-derived ubiquitin ensemble in solution. Science 320(5882):1471-1475. Friedland GD, Lakomek NA, Griesinger C, Meiler J, & Kortemme T (2009) A correspondence between solution-state dynamics of an individual protein and the sequence and conformational diversity of its family. PLoS Computational Biology 5(5):e1000393. Babu CR, Hilser VJ, & Wand AJ (2004) Direct access to the cooperative substructure of proteins and the protein ensemble via cold denaturation. Nat Struct Mol Biol 11(4):352- 357. Yang H, et al. (2003) Protein conformational dynamics probed by single-molecule electron transfer. Science 302(5643):262-266. Fraser JS, et al. (2009) Hidden alternative structures of proline isomerase essential for catalysis. Nature 462(7273):669-673. Lindorff-Larsen K, Best RB, Depristo MA, Dobson CM, & Vendruscolo M (2005) Simultaneous determination of protein structure and dynamics. Nature 433(7022):128- 132. Karplus M & McCammon JA (2002) Molecular dynamics simulations of biomolecules. Nat Struct Biol 9(9):646-652. 18. Meinhold L, Smith JC, Kitao A, & Zewail AH (2007) Picosecond fluctuating protein energy landscape mapped by pressure temperature molecular dynamics simulation. Proc Natl Acad Sci U S A 104(44):17261-17265. Kohn JE, Afonine PV, Ruscio JZ, Adams PD, & Head-Gordon T (2010) Evidence of 15. 14. 11. 12. 13. 10 functional protein dynamics from X-ray crystallographic ensembles. PLoS Computational Biology 6(8). Jacobs DJ (2010) Ensemble-based methods for describing protein dynamics. Current Opinion in Pharmacology 10(6):760-769. Shortle D, Simons KT, & Baker D (1998) Clustering of low-energy conformations near the native structures of small proteins. Proc Natl Acad Sci U S A 95(19):11158-11162. Zhang Y & Skolnick J (2005) The protein structure prediction problem could be solved using the current PDB library. Proc Natl Acad Sci U S A 102(4):1029-1034. Kidd BA, Baker D, & Thomas WE (2009) Computation of conformational coupling in allosteric proteins. PLoS Comput Biol 5(8):e1000484. Hilser VJ, Dowdy D, Oas TG, & Freire E (1998) The structural distribution of cooperative interactions in proteins: analysis of the native state ensemble. Proc Natl Acad Sci U S A 95(17):9903-9908. Hegler JA, Weinkam P, & Wolynes PG (2008) The spectrum of biomolecular states and motions. HFSP J 2(6):307-313. Rajamani D, Thiel S, Vajda S, & Camacho CJ (2004) Anchor residues in protein-protein interactions. Proc Natl Acad Sci U S A 101(31):11287-11292. Smith GR, Sternberg MJ, & Bates PA (2005) The relationship between the flexibility of proteins and their conformational states on forming protein-protein complexes with an application to protein-protein docking. Journal of Molecular Biology 347(5):1077-1101. Norel R, Petrey D, Wolfson HJ, & Nussinov R (1999) Examination of shape complementarity in docking of unbound proteins. Proteins 36(3):307-317. Lensink MF & Wodak SJ (2010) Docking and scoring protein interactions: CAPRI 2009. Proteins 78(15):3073-3084. Ruvinsky AM, Kirys T, Tuzikov AV, & Vakser IA (2011) Side-chain conformational changes upon Protein-Protein Association. Journal of Molecular Biology 408(2):356-365. Guharoy M, Janin J, & Robert CH (2010) Side-chain rotamer transitions at protein- protein interfaces. Proteins 78(15):3219-3225. Kirys T, Ruvinsky AM, Tuzikov AV, & Vakser IA (2012) Rotamer libraries and probabilities of transition between rotamers for the side chains in protein-protein binding. Proteins 80(8):2089-2098. Chopra G, Summa CM, & Levitt M (2008) Solvent dramatically affects protein structure refinement. Proc Natl Acad Sci U S A 105(51):20239-20244. Garman E (2003) 'Cool' crystals: macromolecular cryocrystallography and radiation damage. Curr Opi Struct Biol 13(5):545-551. Juers DH & Matthews BW (2001) Reversible lattice repacking illustrates the temperature dependence of macromolecular interactions. J Mol Biol 311(4):851-862. Dill KA (1999) Polymer principles and protein folding. Protein science : a publication of the Protein Society 8(6):1166-1180. Boehr DD & Wright PE (2008) Biochemistry. How do proteins interact? Science 320(5882):1429-1430. Hyeon C & Thirumalai D (2003) Can energy landscape roughness of proteins and RNA be measured by using mechanical unfolding experiments? Proc Natl Acad Sci U S A 100(18):10249-10253. Nevo R, Brumfeld V, Kapon R, Hinterdorfer P, & Reich Z (2005) Direct measurement of protein energy landscape roughness. EMBO Rep 6(5):482-486. 11 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 43. 44. 45. 46. 42. 40. Tama F & Sanejouand YH (2001) Conformational change of proteins arising from normal mode calculations. Protein Eng 14(1):1-6. Dobbins SE, Lesk VI, & Sternberg MJE (2008) Insights into protein flexibility: The relationship between normal modes and conformational change upon protein-protein docking. Proc Natl Acad Sci U S A 105(30):10390-10395. Cui Q, Li G, Ma J, & Karplus M (2004) A normal mode analysis of structural plasticity in the biomolecular motor F(1)-ATPase. J Mol Biol 340(2):345-372. Zhang D & McCammon JA (2005) The association of tetrameric acetylcholinesterase with ColQ tail: a block normal mode analysis. PLoS Computational Biology 1(6):e62. Park J, Kahng B, Kamm RD, & Hwang W (2006) Atomistic simulation approach to a continuum description of self-assembled beta-sheet filaments. Biophys J 90(7):2510- 2524. Petrone P & Pande VS (2006) Can conformational change be described by only a few normal modes? Biophys J 90(5):1583-1593. Cooper A (1976) Thermodynamic Fluctuations in Protein Molecules. Proc Natl Acad Sci U S A 73(8):2740-2741. Fraser JS, et al. (2011) Accessing protein conformational ensembles using room- temperature X-ray crystallography. Proc Natl Acad Sci U S A 108(39):16247-16252. Gao Y, Douguet D, Tovchigrechko A, & Vakser IA (2007) DOCKGROUND system of databases for protein recognition studies: unbound structures for docking. Proteins 69(4):845-851. Altschul SF, Gish W, Miller W, Myers EW, & Lipman DJ (1990) Basic local alignment search tool. Journal of Molecular Biology 215(3):403-410. Feig M, Karanicolas J, & Brooks CL, 3rd (2004) MMTSB Tool Set: enhanced sampling and multiscale modeling methods for applications in structural biology. J Mol Graphics Model 22(5):377-395. Chocholousova J & Feig M (2006) Balancing an accurate representation of the molecular surface in generalized born formalisms with integrator stability in molecular dynamics simulations. J Comput Chem 27(6):719-729. Feig M, et al. (2004) Performance comparison of generalized born and Poisson methods in the calculation of electrostatic solvation energies for protein structures. J Comput Chem 25(2):265-284. 53. Wesson L & Eisenberg D (1992) Atomic Solvation Parameters Applied to Molecular- Dynamics of Proteins in Solution. Protein Sci 1(2):227-235. 47. 48. 49. 50. 51. 52. 12 41. Table 1. Ensembles of bound and unbound proteins. Protein Unbound structures Bound structures a Dimers Multimers RNase A Pancreatic trypsin inhibitor (PTI) Chemotaxis protein CheY Ubiquitin Ovomucoid Lysozyme C 49 27 73 394 124 45 32 18 6 8 6 19 3 29 24 aConsidered separately from the other subunit(s) in the dimers/multimers. 13 Table 2. Bound-to-bound and bound-to-unbound RMSDs. RMSD between bound structures, Å RMSD between bound and unbound structures, Å Protein PTI CheY RNase A min maxd min max min max min Ubiquitin max Ovomucoid min max Lysozyme C min max All atoms SMb LMc 0.3 0.3 7.2 7.2 0.3 0.1 2.4 2.3 0.6 0.6 1.1 1.2 0.4 0.1 1.7 1.5 0.7 0.6 1.2 1.2 0.4 0.2 1.9 1.8 Ia 0.3 7.1 0.1 2.3 0.6 1.2 0.3 1.6 0.6 1.2 0.2 1.8 Interface SM 0.1 11.8 0.0 2.7 0.5 1.7 0.1 2.5 0.5 1.6 0.1 3.2 I 0.1 11.6 0.0 2.7 0.5 1.7 0.4 2.4 0.5 1.6 0.1 3.0 LM 0.2 11.7 0.2 2.7 0.6 1.8 0.5 2.3 0.5 1.8 0.2 3.1 All atoms SM 0.8 7.4 1.0 2.5 1.9 2.9 1.0 4.1 1.1 2.2 0.8 1.9 I 0.7 7.3 1.0 2.5 1.9 2.9 1.0 4.1 1.2 2.2 0.8 1.9 LM 0.7 7.4 0.9 2.4 1.8 3.0 0.9 4.0 0.9 2.1 0.8 1.9 Interface SM 0.3 11.9 1.1 2.7 1.6 3.1 1.1 5.9 1.2 3.4 0.7 3.1 LM 0.4 11.8 1.0 2.8 1.3 2.9 1.0 6.1 1.1 3.1 0.7 3.3 I 0.3 11.7 1.1 2.7 1.6 3.1 1.1 5.9 1.3 3.5 0.7 3.1 aInitial (not minimized) protein structures. bStructures subjected to 50 steps of the steepest descent minimization (short minimization). cStructures subjected to 50 steps of the steepest descent minimization, followed by 104 steps of the Adopted Basis Newton-Raphson minimization (long minimization). dMinimum and maximum RMSDs. 14 Captions to the figures 1. Energy spectrum of the unbound and bound proteins. (A) PTI, (B) RNase A, (C) Lysozyme C, (D) CheY, (E) Ovomucoid and (F) Ubiquitin. SM and LM indicate the spectrum after the short minimization (50 steps of the steepest descent minimization) and long minimization (SM followed by 104 steps of the Adopted Basis Newton-Raphson minimization). U, D and M are unbound proteins (green), proteins crystallized as dimers (red), and multimers (blue). Open squares with error bars show the average energy (the band’s center) and the standard deviation. 2. The ratio of the ensemble width after long minimization to the lowest energy in the joint ensemble. The data is shown for unbound (¢) and bound ensembles, extracted from dimers (˜) and multimers (š). 3. Energy spacing at the bottom of the folding funnel. The figure shows the energy gap between the two lowest energy minima (p) in the joint ensemble of the bound and unbound states and the average distance between energy minima after the long minimization in the unbound (¢) and bound ensembles, extracted from dimers (˜) and multimers (š). 4. Similarity of bound and unbound interface conformations. (A) PTI, (B) RNase A, (C) Lysozyme C, (D) CheY, (E) Ovomucoid and (F) Ubiquitin. The similarity is calculated for each pair of bound and unbound structures as the share of the unbound interface residues within 1 Å RMSD from the bound interface residues. Bars and circles show the interface similarity between dimeric/multimeric and unbound conformations accordingly. The horizontal axis shows conformation in the bound ensembles. 5. Similarity of bound and unbound structures. (A) PTI, (B) RNase A, (C) Lysozyme C, (D) CheY, (E) Ovomucoid and (F) Ubiquitin. The similarity is calculated for each pair of bound and unbound structures as the share of the unbound residues within 1 Å RMSD from the bound ones. Bars and circles show the similarity between dimeric/multimeric and unbound conformations accordingly. The horizontal axis shows conformation in the bound ensembles. 15 Figure 1 16 Figure 2 17 Figure 3 18 Figure 4 19 Figure 5 20 SUPPORTING MATERIAL: Ensemble-based characterization of unbound and bound states on protein energy landscape A.M. Ruvinsky, T. Kirys, A.V. Tuzikov, and I.A. Vakser Table S1. PDB IDs of bound and unbound protein structures. Protein Unbound structures Bound structures a Dimers Multimers RNase 9rat_A,b 8rat_A, 7rat_A, 6rat_A, 5rat_A, 4rat_A, 3rat_A, 2rat_A, 2aas_A(N,32)c, 1rtb_A, 1rhb_A, 1rha_A, 1rbx_A, 1rat_A, 1kf8_A, 1kf7_A, 1kf5_A, 1fsa_A Pancreatic trypsin inhibitor (PTI) Chemotaxis protein CheY 4pti_A, 1pit_A(N,20), 1oa6_5(N,3), 1oa5_5(N,3) 1djm_A(N,27), 1cey_A(N,46) 1bzq_A,B,C,D, 2p45_A, 2p49_A,2p48_A,2p47 _A, 2p44_A,2p43_A, 2e33_B, 1dfj_E, 3ev6_A,B, 3ev5_A,B, 3ev4_A,B, 3ev3_A,B, 3ev2_A,B, 3ev1_A,B, 9rsa_A,B, 3jw1_A,B, 3euz_A,B, 2p4a_A,C 3tgk_I, 3tgj_I, 3tgi_I, 1f5r_I, 3tpi_I, 3fp8_I, 3fp6_I, 3btk_I, 2tpi_I, 2tgp_I, 2ptc_I, 2kai_I, 2ijo_I, 1tpa_I, 1fy8_I, 1bzx_I, 3gym_I,J 1ffg_A, 1kmi_Y, 1a0o_A,C,E,G 1r0r_I, 3sgb_I, 1ppf_I, 1cho_I, 1sgr_I, 1ds2_I 1p3q_V, 1s1q_B,D, 1yd8_U,V, 2c7m_B, 2fid_A, 1wrd_B 1ubq_A, 1ubi_A, 2nr2_A(N,144), 1xqq_A(N,128),2kn5_A(N,50), 2klg_A(N,20),2jzz_A(N,20), 1v81_A(N,10),1v80_A(N,10), 1d3z_A(N,10) 1tus_A(N,12), 1tur_A(N,12), 1omu_A(N,50), 1omt_A(N,50) 8lyz_A, 7lyz_A, 6lyz_A,6lyt_A,5lyz_A,5lyt_A,4lyz_A,4l yo_A,4lym_A,3lyz_A, 3lym_A,3exd_A,2zq4_A,2yvb_A,2lyz_ A,2lym_A,2hso_A,2hs9_A,2hs7_A,2ep e_A, 2cds_A, 2c8p_A, 2c8o_A,2aub_A, 2a6u_A, 1xek_A,1xej_A, 1xei_A, 1ved_A, 1vdt_A, 1vds_A,1vdq_A, 1uig_A, 1rfp_A, 1lzt_A, 1lza_A, 1lyz_A, 1lyo_A,1lsf_A,1lse_A, 1lsd_A,1lsc_A,1lsb_A,1lsa_A, 1jpo_A a Considered separately from the other subunit(s) in the dimers/multimers. b 9rat_A is an X-ray structure of a chain A from 9rat. c 2aas_A (N,32) are 32 NMR structures of a chain A from 2aas. 1uuz _D, 1zvy_B, 1zvh_L, 1zv5_L, 1xfp_L, 1sq2_L, 1rjc_B, 1ri8_B, 1jtt_L, 3g3a_B,D,F,H, 2znx_Z,Y, 2znw_Z,Y, 2i25_M,L Ubiquitin Ovomucoid Lysozyme C 1js0_A,B,C 2hex_A,B,C,D,E, 1mtn_H,D, 1cbw_I,D, 1bz5_A,B,C,D,E, 1bhc_A,B,C,D,E,F,G,H,I, J, 1b0c_A,B,C,D,E 3hfm_Y, 3d9a_C, 2yss_C, 2eks_C, 2eiz_C, 2dqj_Y, 2dqi_Y, 2dqh_Y, 2dqg_Y, 2dqe_Y, 2dqd_Y, 2dqc_Y, 1yqv_Y, 1xgu_C, 1xgt_C, 1xgr_C, 1xgq_C, 1xgp_C, 1vfb_C, 1ua6_Y, 1ndm_C, 1kir_C, 1kiq_C, 1kip_C 21 Table S2. Two-sample t-test for equal means of energies of the unbound and bound ensembles. P-values, % Dimeric ensemble vs Unbound ensemble Multimeric ensemble vs Unbound ensemble SMa 0.0001 0.00006 0.02 35 0.02 0.08 Protein RNase A CheY PTI Ubiquitin Ovomucoid Lysozyme C Table S3. Two-sample t-test for equal means of energies of the bound structures extracted from dimers and multimers. LMb 0.0002 8.8*10-41 0.13 42 41 70 SM 4 - 0.11 - - 0.4 LM 0.2 - 44 - - 58 P-values, % Multimeric ensemble vs Dimeric ensemble LM 0.1 2 96 Protein RNase A PTI Lysozyme C aShort minimization. bLong minimization. SM 13 73 47 22 Figure S1. Ruggedness of the energy landscape. The ruggedness was calculated as a difference between the average energies of a protein after short and long minimizations for each unbound (¢) and bound ensembles, extracted from dimers (˜) and multimers (š). The solid and dashed lines show linear fit to the data for the unbound proteins and the ones co-crystallized as dimers. The slopes of the solid and dash lines are 3.8 and 2.5 kcal/mol per residue. 23 Figure S2. Structure ensemble of RNase A. The unbound, dimeric and multimeric structures are in blue, red and gray. 24
1104.4102
1
1104
2011-04-20T19:07:57
SIR epidemics in monogamous populations with recombination
[ "physics.bio-ph", "cond-mat.stat-mech", "q-bio.PE" ]
We study the propagation of an SIR (susceptible-infectious-recovered) disease over an agent population which, at any instant, is fully divided into couples of agents. Couples are occasionally allowed to exchange their members. This process of couple recombination can compensate the instantaneous disconnection of the interaction pattern and thus allow for the propagation of the infection. We study the incidence of the disease as a function of its infectivity and of the recombination rate of couples, thus characterizing the interplay between the epidemic dynamics and the evolution of the population's interaction pattern.
physics.bio-ph
physics
Papers in Physics, vol. 3, art. 030001 (2011) Received: 12 August 2010, Accepted: 18 February 2011 Edited by: G. C. Barker Reviewed by: B. Blasius, ICBM, University of Oldenburg, Germany. Licence: Creative Commons Attribution 3.0 DOI: 10.4279/PIP.030001 www.papersinphysics.org ISSN 1852-4249 SIR epidemics in monogamous populations with recombination Dami´an H. Zanette1 ∗ We study the propagation of an SIR (susceptible -- infectious -- recovered) disease over an agent population which, at any instant, is fully divided into couples of agents. Couples are occasionally allowed to exchange their members. This process of couple recombination can compensate the instantaneous disconnection of the interaction pattern and thus allow for the propagation of the infection. We study the incidence of the disease as a function of its infectivity and of the recombination rate of couples, thus characterizing the interplay between the epidemic dynamics and the evolution of the population's interaction pattern. I. Introduction Models of disease propagation are widely used to provide a stylized picture of the basic mechanisms at work during epidemic outbreaks and infection spreading [1]. Within interdisciplinary physics, they have the additional interest of being closely re- lated to the mathematical representation of such di- verse phenomena as fire propagation, signal trans- mission in neuronal axons, and oscillatory chemical reactions [2]. Because this kind of model describes the joint dynamics of large populations of interact- ing active elements or agents, its most interesting outcome is the emergence of self-organization. The appearance of endemic states, with a stable finite portion of the population actively transmitting an infection, is a typical form of self-organization in epidemiological models [3]. Occurrence of self-organized collective behavior has, however, the sine qua non condition that in- formation about the individual state of agents must be exchanged between each other. In turn, this re- ∗E-mail: [email protected] 1 Consejo Nacional de Investigaciones Cient´ıficas y T´ecnicas, Centro At´omico Bariloche e Instituto Balseiro, 8400 Bariloche, R´ıo Negro, Argentina. quires the interaction pattern between agents not to be disconnected. Fulfilment of such requirement is usually assumed to be granted. However, it is not difficult to think of simple scenarios where it is not guaranteed. In the specific context of epi- demics, for instance, a sexually transmitted infec- tion never propagates in a population where sex- ual partnership is confined within stable couples or small groups [4]. In this paper, we consider an SIR (susceptible -- infectious -- recovered) epidemiological model [3] in a monogamous population where, at any instant, each agent has exactly one partner or neighbor [4, 5]. The population is thus divided into cou- ples, and is therefore highly disconnected. How- ever, couples can occasionally break up and their members can then be exchanged with those of other broken couples. As was recently demonstrated for SIS models [6, 7], this process of couple recombi- nation can compensate to a certain extent the in- stantaneous lack of connectivity of the population's interaction pattern, and possibly allow for the prop- agation of the otherwise confined disease. Our main aim here is to characterize this interplay between recombination and propagation for SIR epidemics. In the next section, we review the SIR model and its mean field dynamics. Analytical results are then 030001-1 Papers in Physics, vol. 3, art. 030001 (2011) / D. H. Zanette provided for recombining monogamous populations in the limits of zero and infinitely large recombina- tion rate, while the case of intermediate rates is studied numerically. Attention is focused on the disease incidence -- namely, the portion of the pop- ulation that has been infectious sometime during the epidemic process -- and its dependence on the disease infectivity and the recombination rates, as well as on the initial number of infectious agents. Our results are inscribed in the broader context of epidemics propagation on populations with evolv- ing interaction patterns [4, 5, 8 -- 11]. II. SIR dynamics and mean field de- scription In the SIR model, a disease propagates over a pop- ulation each of whose members can be, at any given time, in one of three epidemiological states: suscep- tible (S), infectious (I), or recovered (R). Suscep- tible agents become infectious by contagion from infectious neighbors, with probability λ per neigh- bor per time unit. Infectious agents, in turn, be- come recovered spontaneously, with probability γ per time unit. The disease process S → I → R ends there, since recovered agents cannot be in- fected again [3]. With a given initial fraction of S and I -- agents, the disease first propagates by contagion but later declines due to recovery. The population ends in an absorbing state where the infection has disap- peared, and each agent is either recovered or still susceptible. In this respect, SIR epidemics dif- fers from the SIS and SIRS models, where -- due to the cyclic nature of the disease, -- the infection can asymptotically reach an endemic state, with a constant fraction of infectious agents permanently present in the population. Another distinctive equilibrium property of SIR epidemics is that the final state depends on the initial condition. In other words, the SIR model possesses infinitely many equilibria parameterized by the initial states. In a mean field description, it is assumed that each agent is exposed to the average epidemiolog- ical state of the whole population. Calling x and y the respective fractions of S and I -- agents, the mean field evolution of the disease is governed by the equations x = −kλxy, y = kλxy − y, (1) where k is the average number of neighbors per agent. Since the population is assumed to re- main constant in size, the fraction of R -- agents is z = 1 − x − y. In the second equation of Eqs. (1), we have assigned the recovery frequency the value γ = 1, thus fixing the time unit equal to γ−1, the average duration of the infectious state. The contagion frequency λ is accordingly normalized: λ/γ → λ. This choice for γ will be maintained throughout the remaining of the paper. The solution to Eqs. (1) implies that, from an initial condition without R -- agents, the final frac- tion of S -- agents, x∗, is related to the initial fraction of I -- agents, y0, as [1] x∗ = 1 − (kλ)−1 log[(1 − y0)/x∗]. (2) Note that the final fraction of R -- agents, z ∗ = 1 − x∗, gives the total fraction of agents who have been infectious sometime during the epidemic process. Thus, z ∗ directly measures the total incidence of the disease. The incidence z ∗ as a function of the infectiv- ity kλ, obtained from Eq. (2) through the stan- dard Newton -- Raphson method for several values y0 of the initial fraction of I -- agents, is shown in the upper panel of Fig. 1. As expected, the dis- ease incidence grows both with the infectivity and with y0. Note that, on the one hand, this growth is smooth for finite positive y0. On the other hand, for y0 → 0 (but y0 6= 0) there is a transcritical bifurcation at kλ = 1. For lower infectivities, the disease is not able to propagate and, consequently, its incidence is identically equal to zero. For larger infectivities, even when the initial fraction of I -- agents is vanishingly small, the disease propagates and the incidence turns out to be positive. Finally, for y0 = 0 no agents are initially infectious, no in- fection spreads, and the incidence thus vanishes all over parameter space. III. Monogamous populations with couple recombination Suppose now that, at any given time, each agent in the population has exactly just one neighbor or, 030001-2 Papers in Physics, vol. 3, art. 030001 (2011) / D. H. Zanette e c n e d i c n i e c n e d i c n i 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0 0.8 0.6 0.4 0.2 0.0 0 0.8 0.6 0.4 = 0.2 y 0 y 0 0 mean field 1 2 3 0.8 0.6 0.4 0.2 = 0 y 0 static (r = 0) monogamous 1 infectivity 2 3 Figure 1: SIR epidemics incidence (measured by the final fraction of recovered agents z ∗) as a func- tion of the infectivity (measured by the product of the mean number of neighbors times the infec- tion probability per time unit per infected neigh- bor, kλ), for different initial fractions of infec- tious agents, y0. Upper panel: For the mean field equations (1). Lower panel: For a static (non- recombining) monogamous population, described by Eqs. (3) with r = 0. in other words, that the whole population is al- ways divided into couples. In reference to sexually transmitted diseases, this pattern of contacts be- tween agents defines a monogamous population [5]. If each couple is everlasting, so that neighbors do not change with time, the disease incidence should be heavily limited by the impossibility of propagat- ing too far from the initially infectious agents. At most, some of the initially susceptible agents with infectious neighbors will become themselves infec- tious, but spontaneous recovery will soon prevail and the disease will disappear. If, on the other hand, the population remains monogamous but neighbors are occasionally al- lowed to change, any I -- agent may transmit the disease several times before recovering. If such changes are frequent enough, the disease could per- haps reach an incidence similar to that predicted by the mean field description, Eq. (1) (for k = 1, i.e. with an average of one neighbor per agent). We model neighbor changes by a process of cou- ple recombination where, at each event, two cou- ples (i, j) and (m, n) are chosen at random and their partners are exchanged [6, 7]. The two pos- sible outcomes of recombination, either (i, m) and (j, n) or (i, n) and (j, m), occur with equal prob- ability. To quantify recombination, we define r as the probability per unit time that any given couple becomes involved in such an event. A suitable description of SIR epidemics in monogamous populations with recombination is achieved in terms of the fractions of couples of different kinds, mSS, mSI, mII, mIR, mRR, and mSR = 1 − mSI − mII − mIR − mRR. Evolution equations for these fractions are obtained by con- sidering the possible transitions between kinds of couples due to recombination and epidemic events [7]. For instance, partner exchange between two couples (S,S) and (I,R) which gives rise to (S,I) and (S,R), contributes positive terms to the time derivative of mSI and mSR, and negative terms to those of mSS and mIR, all of them proportional to the product mSSmIR. Meanwhile, for example, con- tagion can transform an (S,I) -- couple into an (I,I) -- couple, with negative and positive contributions to the variations of the respective fractions, both pro- portional to mSI. The equations resulting from these arguments read mSS = rASIR, mSI = rBSIR − (1 + λ)mSI, mII = rAIRS + λmSI − 2mII, mIR = rBIRS + 2mII − mIR, mRR = rARSI + mIR, mSR = rBRSI + mSI. (3) For brevity, we have here denoted the contribution of recombination by means of the symbols Aijh ≡ (mij +mih)2/4−mii(mjj +mjh+mhh), (4) 030001-3 fi Papers in Physics, vol. 3, art. 030001 (2011) / D. H. Zanette and Bijh ≡ (2mii + mih)(2mjj + mjh)/2 −mij(mij + mih + mjh + mhh)/2, (5) with i, j, h ∈ {S, I, R}. The remaining terms stand for the epidemic events. In terms of the couple fractions, the fractions of S, I and R -- agents are expressed as x = mSS + (mSI + mSR)/2, y = mII + (mSI + mIR)/2, z = mRR + (mSR + mIR)/2. (6) Assuming that the agents with different epidemi- ological states are initially distributed at random over the pattern of couples, the initial fraction of each kind of couple is mSS(0) = x2 0, mSI(0) = 2x0y0, mII(0) = y2 0, mIR(0) = 2y0z0, mRR(0) = z 2 0, and mSR(0) = 2x0z0, where x0, y0 and z0 are the initial fractions of each kind of agent. It is important to realize that the mean field -- like Eqs. (3) to (6) are exact for infinitely large popula- tions. In fact, first, pairs of couples are selected at random for recombination. Second, any epidemic event that changes the state of an agent modifies the kind of the corresponding couple, but does not affect any other couple. Therefore, no correlations are created by either process. In the limit without recombination, r = 0, the pattern of couples is static. Equations (3) become linear and can be analytically solved. For asymp- totically long times, the solution provides -- from the third of Eqs. (6) -- the disease incidence as a func- tion of the initial condition. If no R -- agents are present in the initial state, the incidence is z ∗ = (1 + λ)−1[1 + λ(2 − y0)]y0. (7) This is plotted in the lower panel of Fig. 1 as a function of the infectivity kλ ≡ λ, for various val- ues of the initial fraction of I -- agents, y0. When recombination is suppressed, as expected, the in- cidence is limited even for large infectivities, since disease propagation can only occur to susceptible agents initially connected to infectious neighbors. Comparison with the upper panel makes apparent substantial quantitative differences with the mean field description, especially for small initial frac- tions of I -- agents. Another situation that can be treated analyti- cally is the limit of infinitely frequent recombina- tion, r → ∞. In this limit, over a sufficiently short time interval, the epidemiological state of all agents is virtually "frozen" while the pattern of couples tests all possible combinations of agent pairs. Con- sequently, at each moment, the fraction of couples of each kind is completely determined by the in- stantaneous fraction of each kind of agent, namely, mSS = x2, mSI = 2xy, mII = y2, mIR = 2yz, mRR = z 2, mSR = 2xz. (8) These relations are, of course, the same as quoted above for uncorrelated initial conditions. Replacing Eqs. (8) into (3) we verify, first, that the operators Aijh and Bijh vanish identically. The remaining of the equations, corresponding to the contribution of epidemic events, become equivalent to the mean field equations (1). Therefore, if the distributions of couples and epidemiological states are initially uncorrelated, the evolution of the frac- tion of couples of each kind is exactly determined by the mean field description for the fraction each kind of agent, through the relations given in Eqs. (8). For intermediate values of the recombination rate, 0 < r < ∞, we expect to obtain incidence lev- els that interpolate between the results presented in the two panels of Fig. 1. However, these can- not be obtained analytically. We thus resort to the numerical solution of Eqs. (3). IV. Numerical results for recombin- ing couples We solve Eqs. (3) by means of a standard fourth- order Runge-Kutta algorithm. The initial condi- tions are as in the preceding section, representing no R -- agents and a fraction y0 of I -- agents. The dis- ease incidence z ∗ is estimated from the third equa- tion of Eqs. (6), using the long-time numerical so- lutions for mRR, mSR, and mIR. In the range of parameters considered here, numerical integration up to time t = 1000 was enough to get a satisfac- tory approach to asymptotic values. Figure 2 shows the incidence as a function of in- fectivity for three values of the initial fraction of I -- agents, y0 → 0, y0 = 0.2 and 0.6, and several values of the recombination rate r. Numerically, 030001-4 Papers in Physics, vol. 3, art. 030001 (2011) / D. H. Zanette e c n e d i c n i e c n e d i c n i e c n e d i c n i 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0 0.8 0.6 0.4 0.2 0.0 1.0 0 0.8 0.6 0.4 0.2 0.0 0 positive incidence m. f. 3 2 y t i v i t c e f n i 1 0 no incidence 4 6 2 8 recombination rate y0 0 1 10 10 5 3 r = 0 3 m. f. 2 10 3 1 r = 0 = 0.2 y0 = 0.6 y0 1 m. f. 2 3 10 1 r = 0 e c n e d i c n i 3 1 2 infectivity positive values of r give rise to incidences between those obtained for a static couple pattern (r = 0) and for the mean field description. Note that sub- stantial departure from the limit of static couples is only got for relatively large recombination rates, r > 1, when at least one recombination per cou- ple occurs in the typical time of recovery from the infection. Among these results, the most interesting situa- tion is that of a vanishingly small initial fraction of I -- agents, y0 → 0. Figure 3 shows, in this case, the epidemics incidence as a function of the recombi- nation rate for several fixed infectivities. We recall that, for y0 → 0, the mean field description predicts a transcritical bifurcation between zero and posi- tive incidence at a critical infectivity λ = 1, while in the absence of recombination the incidence is identically zero for all infectivities. Our numerical calculations show that, for sufficiently large values of r, the transition is still present, but the criti- cal point depends on the recombination rate. As r grows to infinity, the critical infectivity decreases approaching unity. 0.8 0.6 0.4 0.2 0.0 1 y0 0 3.0 2.5 2.0 l = 1.5 2 3 4 5 6 recombination rate Figure 3: SIR epidemics incidence as a function of the recombination rate r for a vanishingly small fraction of infectious agents, y0 → 0, and several infectivities λ. Straightforward linearization analysis of Eqs. (3) shows that the state of zero incidence becomes un- stable above the critical infectivity λc = r + 1 r − 1 . (9) Figure 2: SIR epidemics incidence as a function of the infectivity for three initial fractions of infec- tious agents, y0, and several recombination rates, r. Mean field (m. f.) results are also shown. The insert in the upper panel displays the boundary be- tween the phases of no incidence and positive inci- dence for y0 → 0, in the parameter plane of infec- tivity vs. recombination rate. the limit y0 → 0 has been represented by taking y0 = 10−9. Within the plot resolution, smaller val- ues of y0 give identical results. Mean field (m. f.) results are also shown. As expected from the an- alytical results presented in the preceding section, 030001-5 fi fi Papers in Physics, vol. 3, art. 030001 (2011) / D. H. Zanette This value is in excellent agreement with the nu- merical determination of the transition point. Note also that Eq. (9) predicts a divergent critical infec- tivity for a recombination rate r = 1. This implies that, for 0 ≤ r ≤ 1, the transition is absent and the disease has no incidence irrespectively of the infectivity level. For y0 → 0, thus, the recombina- tion rate must overcome the critical value rc = 1 to find positive incidence for sufficiently large infec- tivity. The critical line between zero and positive incidence in the parameter plane of infectivity vs. recombination rate, given by Eq. (9), is plotted in the insert of the upper panel of Fig. 2. V. Conclusions We have studied the dynamics of SIR epidemics in a population where, at any time, each agent forms a couple with exactly one neighbor, but neighbors are randomly exchanged at a fixed rate. As it had al- ready been shown for the SIS epidemiological model [6,7], this recombination of couples can, to some de- gree, compensate the high disconnection of the in- stantaneous interaction pattern, and thus allow for the propagation of the disease over a finite portion of the population. The interest of a separate study of SIR epidemics is based on its peculiar dynamical features: in contrast with SIS epidemics, it admits infinitely many absorbing equilibrium states. As a consequence, the disease incidence depends not only on the infectivity and the recombination rate, but also on the initial fraction of infectious agents in the population. Due to the random nature of recombination, mean field -- like arguments provide exact equations for the evolution of couples formed by agents in every possible epidemiological state. These equa- tions can be analytically studied in the limits of zero and infinitely large recombination rates. The latter case, in particular, coincides with the stan- dard mean field description of SIR epidemics. Numerical solutions for intermediate recombina- tion rates smoothly interpolate between the two limits, except when the initial fraction of infectious agents is vanishingly small. For this special situ- ation, if the recombination rate is below one re- combination event per couple per time unit (which equals the mean recovery time), the disease does not propagate and its incidence is thus equal to zero. Above that critical value, a transition ap- pears as the disease infectivity changes: for small infectivities the incidence is still zero, while it be- comes positive for large infectivities. The critical transition point shifts to lower infectivities as the recombination rate grows. It is worth mentioning that a similar transition between a state with no disease and an endemic state with a permanent infection level occurs in SIS epidemics with a vanishingly small fraction of in- fectious agents [6, 7]. For this latter model, how- ever, the transition is present for any positive re- combination rate. For SIR epidemics, on the other hand, the recombination rate must overcome a crit- ical value for the disease to spread, even at very large infectivities. While both the (monogamous) structure and the (recombination) dynamics of the interaction pat- tern considered here are too artificial to play a role in the description of real systems, they correspond to significant limits of more realistic situations. First, the monogamous population represents the highest possible lack of connectivity in the interac- tion pattern (if isolated agents are excluded). Sec- ond, random couple recombination preserves the instantaneous structure of interactions and does not introduce correlations between the individual epidemiological state of agents. As was already demonstrated for SIS epidemics and chaotic syn- chronization [7], they have the additional advan- tage of being analytically tractable to a large ex- tent. Therefore, this kind of assumption promises to become a useful tool in the study of dynamical processes on evolving networks. Acknowledgements - Financial support from is SECTyP -- UNCuyo and ANPCyT, Argentina, gratefully acknowledged. [1] R M Anderson, R M May, Infectious Diseases in Humans, Oxford University Press, Oxford (1991). [2] A S Mikhailov, Foundations of Synergetics I. Distributed active systems, Springer, Berlin (1990). 030001-6 Papers in Physics, vol. 3, art. 030001 (2011) / D. H. Zanette [3] J D Murray, Mathematical Biology, Springer, Berlin (2003). [4] K T D Eames, M J Keeling, Modeling dynamic and network heterogeneities in the spread of sexually transmitted diseases, Proc. Nat. Acad. Sci. 99, 13330 (2002). [8] T Gross, C J Dommar D'Lima, B Blasius, Epi- demic dynamics in an adaptive network, Phys. Rev. Lett. 96, 208 (2006). [9] T Gross, B Blasius, Adaptive coevolutionary networks: a review, J. R. Soc. Interface 5, 259 (2008). [5] K T D Eames, M J Keeling, Monogamous net- works and the spread of sexually transmitted diseases, Math. Biosc. 189, 115 (2004). [10] D H Zanette, S Risau -- Gusman, Infection spreading in a population with evolving con- tacts, J. Biol. Phys. 34, 135 (2008). [6] S Bouzat, D H Zanette, Sexually transmit- ted infections and the marriage problem, Eur. Phys. J B 70, 557 (2009). [11] S Risau -- Gusman, D H Zanette, Contact switching as a control strategy for epidemic outbreaks, J. Theor. Biol. 257, 52 (2009). [7] F Vazquez, D H Zanette, Epidemics and chaotic recombining monogamous populations, Physica D 239, 1922 (2010). synchronization in 030001-7
1704.08316
1
1704
2017-04-26T19:36:57
Magnetic Force Microscopy Characterization of Superparamagnetic Iron Oxide Nanoparticles (SPIONs)
[ "physics.bio-ph" ]
Superparamagnetic iron oxide nanoparticles (SPIONs), due to their controllable sizes, relatively long in vivo half-life and limited agglomeration, are ideal for biomedical applications such as magnetic labeling, hyperthermia cancer treatment, targeted drug delivery and for magnetic resonance imaging (MRI) as contrast enhancement agents. In order to understand how SPIONs interact with cells and cellular membranes it would be of interest to characterize individual SPIONs at the nanoscale in physiologically relevant conditions without labeling them. We demonstrate that Magnetic Force Microscopy (MFM) can be used to image SPIONs in air as well as in liquid. The magnetic properties of bare and SiO2 coated SPIONs are compared using MFM. We report that surface modification using (3-mercaptopropyl)-trimethoxysilane significantly improves adsorption and distribution of SPIONs on gold surfaces. To obtain proof of principle that SPIONS can be imaged with MFM inside the cell we imaged SPIONs buried in thin polymer films (polystyrene (PS) and poly methyl-methacrylate (PMMA)). This opens the possibility of visualizing SPIONs inside the cell without any labeling or modifications and present MFM as a potential magnetic analogue for fluorescence microscopy. The results of these studies may have a valuable impact for characterization and further development of biomedical applications of SPIONs and other magnetic nanoparticles.
physics.bio-ph
physics
31 Magnetic Force Microscopy Characterization of Superparamagnetic Iron Oxide Nanoparticles (SPIONs) Gustavo Cordova1, Simon Attwood2, Ravi Gaikwad2, Frank Gu3,4, Zoya Leonenko1,2,4, 1Department of Biology, University of Waterloo, Waterloo ON 2Department of Physics and Astronomy, University of Waterloo, Waterloo ON 3Department of Chemical Engineering, University of Waterloo, Waterloo ON 4Waterloo Institute for nanotechnology, University of Waterloo, Waterloo ON Corresponding author: E-mail: [email protected], Tel.: 519-888-4567 x38273 Received: Nov.7, 2013; Accepted: March 6, 2014; Published: March 15, 2014. Citation: Gustavo Cordova, Simon Attwood, Ravi Gaikwad, Frank Gu and Zoya Leonenko. Magnetic Force Microscopy Characterization of Superparamagnetic Iron Oxide Nanoparticles (SPIONs). Nano Biomed. Eng. 2014, 6(1), 31-39. DOI: 10.5101/nbe.v6i1.p31-39. Abstract Superparamagnetic iron oxide nanoparticles (SPIONs), due to their controllable sizes, relatively long in vivo half-life and limited agglomeration, are ideal for biomedical applications such as magnetic labeling, hyperthermia cancer treatment, targeted drug delivery and for magnetic resonance imaging (MRI) as contrast enhancement agents. In order to understand how SPIONs interact with cells and cellular membranes it would be of interest to characterize individual SPIONs at the nanoscale in physiologically relevant conditions without labeling them. We demonstrate that Magnetic Force Microscopy (MFM) can be used to image SPIONs in air as well as in liquid. The magnetic properties of bare and SiO2 coated SPIONs are compared using MFM. We report that surface modification using (3-mercaptopropyl)-trimethoxysilane significantly improves adsorption and distribution of SPIONs on gold surfaces. To obtain proof of principle that SPIONS can be imaged with MFM inside the cell we imaged SPIONs buried in thin polymer films (polystyrene (PS) and poly methyl-methacrylate (PMMA)). This opens the possibility of visualizing SPIONs inside the cell without any labeling or modifications and present MFM as a potential magnetic analogue for fluorescence microscopy. The results of these studies may have a valuable impact for characterization and further development of biomedical applications of SPIONs and other magnetic nanoparticles. Keywords: Magnetic force microscopy (MFM) imaging in liquid; superparamagnetic iron oxide nanoparticles (SPIONs); magnetic properties; surface modification for nanoparticle adsorption; silica coated SPIONs; MFM imaging of SPIONs in liquid and inside polymer films Introduction Biocompatible superparamagnetic iron oxide nanoparticles (SPIONs) have been widely used for biomedical applications such as tissue specific release of therapeutic agents, magnetic hyperthermia treatment for cancer patients, a wide range of cell separation techniques as well as contrast agents in MRI imaging. Inorganic coatings, such as aluminum, cadmium, gold and silica can be used to electrostatically stabilize http://www.nanobe.orgNano Biomed Eng2014; 6(1):31-39. doi: 10.5101/nbe.v6i1.p31-39.Research ArticleNano Biomed Eng 2014, Vol. 6, Issue 1 32 magnetic nanoparticles in a colloid [1]. These inorganic materials are most commonly used for post- synthesis modification of SPIONs which will adopt the core-shell nanoparticle structure. Recently, silica has received a great deal of attention for this purpose. Due to its biocompatibility, low cost, and allowance for covalent stabilization over a broad range of pHs, silica (ie. silicon dioxide) has become an ideal choice for post-synthesis modification of SPIONs to be used for biomedical applications [2, 3]. Silica is highly suitable for preserving the intrinsic magnetic properties of SPIONs by helping to prevent oxidation and aggregation of the SPION's magnetite core [4]. Although, in 2008 a study by Bumb et al. used SQUID magnetic analysis and showed that under low applied fields, higher magnetization values were observed for the silica SPION sample as compared to uncoated SPIONs, suggesting that silica separating the small particles may be leading to weak ferromagnetic ordering in the relatively large batches of nanoparticles that are required for SQUID analysis [4]. Today, SPIONs are popularly used in a large variety of therapeutic and diagnostic biomedical applications, both in vitro and in vivo. Most often SPIONs are used as magnetic resonance imaging (MRI) contrast enhancement agents. They are intravenously infused into the body to detect and characterize small lesions, tumours in organs, or to visualize the digestive tract [5]. Due to their high magnetization, SPIONs cause a critical decrease in the relaxation rate of water protons, and therefore are efficient MRI contrast agents. The enhanced contrast allows MRI to differentiate between different organs in the body as well as benign and malignant tissues [6]. Iron oxide nanoparticles are used most commonly for this purpose due to their low toxicity, chemical stability and biocompatibility. Although MRI is a powerful technique, its resolution is in the range of millimeters to micrometers and it does not give information about position of SPIONs at a single cell level. In order to develop better SPION- based contrast agents it is important to understand how SPIONs interact with the cell and cellular membrane at the nanoscale without labeling them. Currently, one of the most common methods for intracellular imaging of magnetic nanoparticles is fluorescence microscopy. A disadvantage of this technique is that nanoparticles must first be labeled or modified with fluorescent probes in order for the particles to be visualized, which may affect their interaction with the cellular membrane. Furthermore, the maximum resolution of this technique is limited to 300-500nm - half the wavelength of the light being used [7]. Relative to fluorescence microscopy, two-photon microscopy (TPM) offers improved resolution and has also been used to study cellular interactions with magnetic nanoparticles but still requires the particles to be labeled with a two-photon fluorescent dye [8]. Due to the relatively poor resolution and reliability of these techniques, a label-free in vitro detection method for magnetic nanoparticles, SPIONs especially, is of great interest. Magnetic force microscopy (MFM), because of its ability to localize and characterize magnetic nanoparticles at the nanoscale without labeling, offers great potential for this purpose. MFM has the capability to detect nanoscale magnetic domains as well as simultaneously obtain both atomic force microscopy phase and topography images. This technique has received limited attention as a potential tool for characterization of SPIONs, specifically in physiologically relevant conditions, and has only ever been used in liquid to image computer hard disks [9]. Most studies that have used MFM to characterize SPIONs, or other magnetic nanoparticles, have done so under ambient conditions (in air) [10, 11, 12, 13]. SPIONs, however, when applied in biomedicine, typically carry out their function in physiological conditions. Therefore, characterization of these particles should be undertaken in conditions similar and relevant to the physiological environment, i.e. in liquid. The potential of MFM is largely unexplored in this regard. In this study, we evaluate the applicability of MFM in air, as well as in liquid, to characterize bare and SiO2 coated SPIONs on mica. The magnetic properties of individual bare and SiO2 coated SPIONs are compared using MFM. For the first time we demonstrate that MFM imaging of SPIONs can be done in liquid. To mimic SPIONs buried inside the cell we imaged them inside thick polymer film (polystyrene (PS) and poly- methyl methacrylate (PMMA)). This will provide a platform for cellular studies on SPIONs without any labeling. Experimental Section SPION Synthesis Materials Iron (II) chloride (99%, Sigma), iron (III) chloride Nano Biomed Eng 2014, Vol. 6, Issue 1http://www.nanobe.org (99%, Sigma), tetramethylammonium hydroxide (TMAOH) (25% solution, Sigma), tetraethyl orthosilicate (TEOS) (99%, Sigma), ammonium hydroxide (28% solution, Sigma), and ethanol (99%, ACS) were purchased and used without further purification. (3-Mercaptopropyl) trimethoxysilane (MPTS) was purchased from Sigma (95%). SPION Synthesis Bare and SiO2 coated SPIONs were synthesized by a co-precipitation method as described previously [14]. Briefly, 2.5 mL of a mixed iron solution in deionized water (2 mol/L FeCl2 and 1 mol/L FeCl3) was added to a 0.7 mol/L tetramethylammonium hydroxide (TMAOH) solution under vigorous stirring, and the reaction was allowed to proceed open to the air at room temperature for 30 minutes while stirring. After 30 minutes, the black particles were separated from solution over a neodymium magnet, and washed at least thrice with an equivalent volume of pH 12 TMAOH solution (so as to maintain the equivalent particle concentration as immediately after the reaction) until the particles were no longer magnetically separable. This colloidal suspension was sonicated for 10 minutes (Branson Digital Sonifier 450, USA), and then 20 mL of the sonicated fluid was mixed with 20 mL pH 12 TMAOH and 160 mL ethanol. 7 mL tetraethylorthosilicate (TEOS) was then added to this suspension while stirring, and allowed to react at room temperature while stirring for approximately 18 hours. The SiO2 coated SPIONs were then magnetically recovered from solution, and washed thrice with ethanol and thrice with deionized water by magnetic decantation, and sonicated in deionized water for 10 minutes before further use. Deposition of SPIONs on mica substrate for imaging A SPION dilution of ~5.5 mg/mL was prepared using deionized water. A concentration of ~5.5 mg/mL of SPIONs was used because it was observed to provide a uniform distribution of SPIONs with a relatively small size distribution where individual particles could be observed. The SPION dilution was sonicated for 25 minutes. After this 10 µL of the SPION dilution was deposited on freshly cleaved mica (v-4 grade, SPI Supplies, PA, USA), covered with a petri dish and left to air dry for approximately six minutes. The sample was gently rinsed with four or five drops of deionized water and then immediately dried with a steady stream 33 of nitrogen for 3.5 minutes, placed in a sealed petri dish with nitrogen and left in a dessicator overnight. For polymer coated samples, a three percent (3%) solution of PMMA or 0.4% solution of polystyrene was used to spin coat the SPION samples with a coating of ~30 nm. Both PMMA and polystyrene dilutions were made in toluene. 40 µL aliquots of the respective polymer solutions were pipetted onto the already deposited SPION samples. A spin motor with an applied voltage of 1V for 15 seconds was used in order to spin coat both the PMMA and polystyrene. Substrate (mica) modification via 3-MPTS Circular mica substrates were freshly cleaved and stored in nitrogen. The mica substrates were then sputtered with a 2.5 nm layer of titanium using an electron beam evaporator. Subsequently, a 50 nm layer of gold was deposited onto the titanium layer without interruption of the vacuum ensuring that titanium dioxide did not form on the surface of the substrate. After sputtering, mica substrates were stored in nitrogen until needed. Sputtered gold wafers were immersed in a 40 mM solution of 3-MPTS in methanol for three hours. Substrates were then thoroughly rinsed with methanol and Millipore water. After immersion in methanol, substrates were then placed into an aqueous 0.01 M NaOH solution for one hour. Substrates were then immersed in a solution of SiO2 SPIONs (~6.9 mg/mL) for one hour. To finish, substrates were washed with Millipore water and dried with nitrogen gas. Samples were imaged immediately. MFM imaging All samples were imaged using a Nanowizard II atomic force microscope (JPK Instruments, Germany). MikroMasch NSC-18, cobalt/chromium coated magnetic cantilevers (MIKROMASCH CA, USA) were used to image the SPIONs. An external perpendicular magnetic field was applied to the sample during imaging in order to ensure magnetization of the SPIONs and improve the contrast of MFM phase images by placing a small permanent magnet directly underneath the mica during imaging [10]. All samples were imaged using AFM intermittent contact mode with MFM hover mode (or lift mode) at various distances from the substrate. Quantitative Analysis of Topography and MFM Images All particle analysis was conducted using Nano Biomed Eng 2014, Vol. 6, Issue 1http://www.nanobe.org 34 customized scripts written in Matlab. Raw images were flattened by subtracting a best fit plane (plane flatten). The median of each row was then found, and subtracted from each pixel in that row (line-by- line flatten). Next a course height (or magnetic signal if assessing the MFM image) threshold was used to highlight the features not associated with the flat substrate. The raw data was then re-loaded and a plane- flatten and line-by-line flatten were employed whilst excluding the previously highlighted regions. This procedure allowed the images to be correctly flattened which is extremely important for particle analysis. A sum of two Gaussians was then fitted to the histogram of height (or magnetic signal) data. For the topography images the position of the first peak represents the mica surface whereas for the MFM images this represents the background magnetic signal. Often a second height (or MFM) threshold was employed so as to carefully highlight the pixels associated with the particles. The maximum height for each particle was then measured relative to the mica surface from the topography image. The maximum magnetic signal shift was measured relative to the background signal from the MFM image. Thus our Matlab routine allows both the maximum height and corresponding magnetic signal to be assessed for each particle. Results and Discussion Bare vs. SiO2 coated SPIONs In order to compare the magnetic properties of bare and SiO2 SPIONs at the nanoscale, MFM imaging was A nm 7 0 C nm 22 0 B 2.0° 0 D 20 µm 20 µm 20 µm 12.0° 0 20 µm done on samples of each type of SPION deposited on mica as described in section 2.2. Panels A - D in Figure 1 show the AFM and MFM phase images of the bare and SiO2 coated SPIONs. AFM topography images (panels A and C) were analyzed to obtain the size distributions for both types of SPIONs. The distribution of particle size (diameter) is shown in Figure 1E and 1F and was determined using height distribution analysis. Both bare and silica coated SPIONs have non-symmetric right skewed size distributions with the SiO2 SPIONs having a broader distribution and a higher mean diameter than the bare SPIONs. The mean, median and mode diameter measured for bare and SiO2 coated SPIONs were 5.1 +/- 0.1 nm (standard error), 4.0 nm, 1.2 nm and 33 +/- 1 nm (standard error), 13 nm, 4 nm respectively. Although MFM detection of SPIONs has been reported, there is some doubt whether individual SPIONs can actually be distinguished by MFM because the magnetic field from SPIONs is proportional to the diameter of the particle and thus very small. In 2009, silica nanoparticles, with and without the presence of a magnetic core were compared using MFM [15]. When the magnetic core was absent, no MFM contrast was observed suggesting that only magnetic structures will cause measurable phase contrast. In Figure 1, MFM contrast is observed for SPIONs as small as ~3 nm. As a result, we can be confident that these are true MFM signals from the SPIONs being studied. d e s i l a m r o y c n e u q e r f 1.0 0.8 0.6 0.4 0.2 0N 1.0 0.8 0.6 0.4 0.2 0N 0 y c n e u q e r f d e s i l a m r o SiO2 coated spions Uncoated spions E F 25 50 Particle diameter (nm) 75 100 125 150 Fig. 1 Characterization of bare vs. SiO2 coated SPIONs. A-D. Tapping-mode AFM topography images (column 1) and MFM phase images (colum 2) of bare and SiO2 coated SPIONS (images A/B and C/D respectively). Both topography and MFM phase images obtained using a magnetic AFM probe in the presence of an externally applied perpendicular magnetic field in ambient conditions. MFM phase images obtained in lift mode at a scan height of 50 nm. Colour-scale and scale bars for both topography and phase images are shown in the bottom of each panel. E/F) Size distributions for both bare (E) and SiO2 coated (F) SPIONs. Nano Biomed Eng 2014, Vol. 6, Issue 1http://www.nanobe.org The contrast observed in the MFM images is caused by the interaction between MFM probe and the magnetic field from the SPIONs. These interactions cause a shift in the phase of the oscillating probe [16, 17]. The MFM images for both types of SPIONs shown in Figure 1, panels B and D demonstrate an effect called dipolar contrast, with half of the phase contrast being dark, and half being light for each individual magnetic structure. This dipolar contrast for magnetic nanoparticles is typically found only when external magnetic fields are applied perpendicularly to the measurement direction as is the case for this study [10, 15, 18, 19]. In order to directly compare the magnetic properties of individual bare and SiO2 coated SPIONs, the MFM phase-shift for both bare and SiO2 SPION, gathered from the MFM phase images (Figure 1, panels B and D), was plotted as a function of particle size (diameter) in Figure 2. A positive linear trend is observed for both types of SPIONs. This data suggests that the magnetic moment is proportional to the diameter (and therefore the volume) of the SPION. Computer simulations of MFM on SPIONs have also demonstrated that the phaseshift detected in MFM depends very strongly on the particle diameter [18]. A 2008 study by Bumb et al., noted that under low applied fields, higher magnetization values were observed for silica-coated 40 30 20 10 0 ) g e d ( t f i h s e s a h P −10 −20 −30 0 50 Bare spions in air Bare spions in air SiO2 coated spions in air SiO2 coated spions in air 100 Particle size (nm) 150 200 250 300 Fig. 2 MFM phase-shift vs. particle size for bare and SiO2 coated SPIONs. Phase-shift values obtained in MFM experiments on bare and SiO2 coated SPIONs described in Figure 4.2. Phase shift (measured in degrees) versus SPION size (nanometers) is shown. A positive linear trend is observed for both bare and SiO2 SPIONS. Phase shift values for bare SPIONs range from 0 to ~4 degrees and from 0 to ~20 degrees for SiO2 coated SPIONs. MFM analysis for this data set was done in ambient conditions. 35 SPION samples when compared to uncoated SPIONs; suggesting that silica separating the small particles may be leading to weak ferromagnetic ordering in the relatively large batches of nanoparticles that are required for SQUID analysis [4]. We found bare and SiO2 coated SPIONs behave identically when analyzed with MFM, demonstrating that the SiO2 coating has no effect on the magnetic properties of the SPIONs – contrary to large batch analysis using SQUID. A similar result was also observed by Neves et al. in 2010 which found that the response of MFM to magnetic nanoparticles is not affected by the presence of a silica coating [19]. MFM phase shift dependence on scan height To understand the limits to the detection of small SPIONs with MFM, we experimentally analyzed the magnetic force sensitivity of MFM by examining the relationship between MFM phase shift and distance between the probe and the sample in hover mode imaging (Figure 3). To mimic various biological media we embedded SPIONs under a 30 nm layer of PS on mica. From Figure 3, the phase contrast in the MFM images is observed to increase with distance between the probe and the sample surface (scan height). Figure 4A shows the plot of MFM phase-shift (degrees) versus particle size (nm) for each scan height of SiO2 SPIONs covered with PS as seen in Fig. 3. Best fit lines show positive linear trends for these data sets. This relationship is plotted in Figure 4B. The data shows that for this height range as scan height increases, MFM phase shift measured also increases reaching a plateau at approximately 300 nm. We observe the increase of phase-shift signal with the increase of SPIONs size (Fig. 4A) and therefore magnetic moment. Interestingly we observe the increase of magnetic signal when the distance between the probe and the sample increases from 50 to 200 nm and then it follows the plateau at 250 and 300 nm (Fig. 4B). Therefore we show that MFM method has enough sensitivity to detect SPIONs at small as well as at larger separations between the probe and the sample. MFM of SPIONs in liquid MFM imaging of SiO2 coated magnetic particles in air has been reported before [15, 19]. In this work, for the first time we present MFM images of SPIONs in liquid environment. This step served as a proof of Nano Biomed Eng 2014, Vol. 6, Issue 1http://www.nanobe.org 36 A 4° 0 D B E 10 μm C F Fig. 3 Consecutive MFM phase shift images taken at different scan heights for SiO2 coated SPIONS covered with PS. Scan heights are (A) 50 nm; (B) 100 nm; (C) 150 nm: (D) 200 nm; (E) 250 nm; (F) 300 nm. Images were taken using a magnetic probe in the presence of an externally applied perpendicular magnetic field in ambient conditions. Colour-scale and scale bars shown in (A) apply to all images. 50 hm height 100 nm height 150 nm height 200 nm height 250 nm height 300 nm height B 5 4 3 2 1 ) g e d ( t f i h s e s a h p n a e M 10 A ) g e d ( t f i h s e s a h P 8 6 4 2 0 25 50 75 100 Particle size (nm) 125 150 0 0 50 100 150 Lift height (nm) 200 250 300 350 Fig. 4 (A) MFM phase-shift (degrees) versus particle size (nm) is shown for each scan height of SiO2 SPIONs covered with PS shown in Figure 3. Best fit lines show positive linear trends for these data sets. (B) Mean phase shift (for each set of SPIONs of different sizes) versus MFM scan height values for the phase images shown in Figure 3. principle for using MFM in a liquid environment to image SPIONs in cells. In this experiment SPIONs were coated with PMMA according to the protocol outlined in Section 2.2. Coating the sample with PMMA temporarily secured the nanoparticles to the substrate and prevented the SPIONs from being removed from the substrate when they were exposed to water. Relatively small agglomerations of SPIONs covered with an approximately 30 nm layer of PMMA are shown in Figure 5. Considerably wider and more gradual structures are observed in when compared to the SPION structures seen in Figure 1. This comparison suggests that the SPIONs in Figure 5 are indeed coated with PMMA. In the MFM image, Figure 5 panel B, the same dipolar contrast that was observed in Figure 1 (panels B and D) can be seen. Fig. 5B shows clear contrast of SPIONs covered with PMMA and imaged in liquid. To the best of authors' knowledge, this is the first case of MFM imaging of SPIONs in liquid reported up to date. S P I O N s o n g o l d c o a t e d 3 - M P T S functionalized mica SPIONs are small nanoparticles; nevertheless we Nano Biomed Eng 2014, Vol. 6, Issue 1http://www.nanobe.org A B 37 nm 54 0 1 μm 7.5° 0 1 μm Fig. 5 MFM in liquid image of SiO2 coated SPIONs spin-coated with PMMA. Tapping-mode AFM topography image (A) and MFM phase image (B) of SiO2 coated SPIONs spin-coated with PMMA taken at a lift height of 50 nm. Images were taken using a magnetic AFM probe in the presence of an externally applied perpendicular magnetic field in liquid (water). A spin motor with an applied voltage of 1V for 15 seconds was used to spin coat a 3% solution of PMMA (in toluene) in order cover the SPION sample with ~30 nm of PS. Colour-scale and scale bars are shown in the bottom of each panel. had problems depositing them on mica surfaces as they are easily washed away without special modification of the surface. Therefore we used chemical means of adhering the SPIONs to the surface of the substrate. The most promising method attempted to date for securing the SiO2 coated SPIONs to the mica is the use of gold-coated mica substrates further modified with 3-mercaptopropyltrimethoxysilane (3-MPTS), an organosilane [20]. The key to this method is the formation of a covalent bond between the gold coated substrate and the SiO2 coated SPION via the 3-MPTS molecule, specifically via its thiol functional group. Formation of this covalent bond will provide stability in air and liquid environments. An illustration of this deposition method as well as the chemical structure of 3-MPTS is shown in Figure 6. This method is the most promising method attempted thus far for securing the SiO2 coated SPIONs to the mica. In the MFM phase image shown in Figure 7B even better MFM contrast from the SPIONs can A B CH3 O CH3 O OSi SH CH3 O Si O O Si O O Si O O Si O O S S S S Au Fig. 6 Deposition of SiO2 coated SPIONs via (3-MPTS). (A) Chemical structure of 3-MPTS with the red circle indicating the thiol functional group. (B) Illustration of SiO2 deposition with the blue region indicating the SiO2 coating containing the silanol functional group. be observed, as compared to Figure 1D. Individual SPIONs can now be distinguished from within small aggregates with high level of magnetic detail. Therefore we conclude that mica substrates coated with gold serve as better substrates for SPION deposition, especially for AFM and MFM imaging. The SPIONs in Figure 7 were observed even after the substrates were kept in aqueous solution for an hour followed by thorough rinsing. This demonstrates the formation of the covalent bond between the SiO2 coated SPIONs and the 3-MPTS and its ability to secure these SPIONs to a gold-modified substrate in the presence of liquid. From Figure 7, it can be seen that successful SPION deposition only occurred for SPIONs in a water solvent. TMAOH has been shown to be a strong etchant of SiO2 (~1 nm/min), which forms the coating around the SPIONs [5]. Since this etching process would have been occurring for the SPIONs in TMAOH, these particles were effectively uncoated bare SPIONs. As seen in Figure 6, this deposition process involving 3-MPTS is dependent upon the silanol functional groups present on the SiO2 coating of the SPIONs. Thus, with the SiO2 coating absent, deposition via 3-MPTS will not occur. Conclusions Despite great progress and a rapidly advancing field of research there have always been problems associated with magnetic nanoparticles and their applications; overcoming immunological reactions, avoiding toxic responses to intravenously injected particles, proper Nano Biomed Eng 2014, Vol. 6, Issue 1http://www.nanobe.org 38 A C B D 1 μm 1 μm 2 μm 2 μm Fig. 7 Comparson of SPIONs kept in water solvent (A, B) and TMAOH (C, D) on 3-MPTS functionalized gold coated mica. (A) and (C) are AFM topography images, (C) and (D) are MFM images. Images were taken using a magnetic AFM probe in the presence of an externally applied perpendicular magnetic field in ambient conditions. Scale bars are shown in the bottom left of each panel. Successful SPION deposition was only observed for the SPIONs kept in water solvent. clearance of particles, and the debate over whether or not to sacrifice more efficient uptake at the cost of negative side effects have been prominent issues within the field. Proper and relevant in vitro, and ultimately in vivo, characterization of the magnetic properties of these nanoparticles can now be added to this list. We demonstrated that MFM can be used to image SPIONs in air, in water and inside the polymer films. We report that surface modification with (3-mercaptopropyl)-trimethoxysilane significantly improves adsorption and distribution of SPIONs on gold surfaces. Our results show feasibility of using MFM for the detection of magnetic nanoparticles within cells without any labeling or modifications, thus demonstrating that MFM can be used as a potential magnetic analogue for fluorescence microscopy. These results may also add to further developments of SPIONs and their applications in biomedicine. Acknowledgments We acknowledge Timothy Leshuk for assisting in synthesizing the SPION particles and the University of Waterloo Institute for Nanotechnology for help with sample preparation. This work was funded by Natural Science and Engineering Council of Canada (NSERC). References [1] Choi KH, Lee SH, Kim YR. Magnetic behavior of Fe 3O 4nanostructure fabricated by template method. Journal of Magnetism and Magnetic Materials. 2007; 310(2): e861-e863. [2] Alcalá MD, Real C. Synthesis based on the wet impregnation method and characterization of iron and iron oxide-silica nanocomposites. Solid State Ionics. 2006; 177(9): 955-960. [3] Ma D, Guan J, Normandin F. Multifunctional nano- architecture for biomedical applications. Chemistry of Materials. 2006; 18(7): 1920-1927. [4] Bumb A, Brechbiel MW, Choyke PL, Fugger L, Eggeman A, Prabhakaran D, Hutchinson J, Dobson PJ. Synthesis and characterization of ultra-small superparamagnetic iron oxide nanoparticles thinly coated with silica. Nanotechnology. 2008; 19(33): 335601. [5] Laurent S, Forge D, Port M, Roch A, Robic C, Vander Elst L, Muller RN. Magnetic iron oxide nanoparticles: Synthesis, stabilization, vectorization, physicochemical characterizations, and biological applications. Chemical Reviews. 2007; 108(6): 2064-2110. Nano Biomed Eng 2014, Vol. 6, Issue 1http://www.nanobe.org [6] Goya GF, Grazu V, Ibarra MR. Magnetic Nanoparticles for Cancer Therapy. Current Nanoscience. 2008; 4(1): 1-16. [7] Bertorelle F, Wilhelm C, Roger J, Gazeau F, Ménager C, Cabuil V. Fluorescence-modified superparamagnetic nanoparticles: Intracellular uptake and use in cellular imaging. Langmuir. 2006; 22(12): 5385-5391. [8] Bergey EJ, Levy L, Wang XP, Krebs LJ, Lal M, Kim KS, Pakatchi S, Liebow C, Prasad PN. DC magnetic field induced magnetocytolysis of cancer cells targeted by LH-RH magnetic nanoparticles in vitro. Biomedical Microdevices. 2002; 4(4): 293-299. [9] Giles R, Cleveland JP, Manne S, Hansma PK, Drake B, Maivald P, Boles C, Gurley G, Elings V. Noncontact force microscopy in liquids. Applied Physics Letters. 1993; 63(5): 617-618. [10] Schreiber S, Savla M, Pelekhov DV, Iscru DF, Selcu C, Hammel PC, Agarwal G. Magnetic Force Microscopy of Superparamagnetic Nanoparticles. Small. 2008; 4(2): 270- 278. [11] Raşa M, Kuipers BWM,Philipse AP.Atomic force microscopy and magnetic force microscopy study of model colloids. Journal of Colloid and Interface Science. 2002; 250(2): 303-315. [12] Zhang Y, Yang M, Ozkan M, Ozkan CS. Magnetic Force Microscopy of Iron Oxide Nanoparticles and Their Cellular Uptake. American Institute of Chemical Engineers. 2009; 25(4): 923-928. [13] Shen H, Long D, Zhu L, Li X, Dong Y, Jia N, Zhou H, Xin X, Sun Y. Magnetic force microscopy analysis of apoptosis of HL-60 cells induced by complex of antisense oligonucleotides and magnetic nanoparticles. Biophysical Chemistry. 2006; 122(1): 1-4. [14] Alwi R, Telenkov S, Mandelis A, Leshuk T, Gu F, Oladepo S, Michaelian K. Silica-coated super paramagnetic iron oxide nanoparticles (SPION) as biocompatible contrast 39 agents in biomedical photoacoustics. Biomedical Optics Express. 2012; 3(10): 2500-2509. [15] Pacifico J, van Leeuwen YM, Spuch Calvar M, Sánchez Iglesias A, Rodríguez-Lorenzo L, Pérez-Juste J, Pastoriza Santos I, Liz Marzan LM. Field gradient imaging of nanoparticle systems: Analysis of geometry and surface coating effects. Nanotechnology. 2009; 20(9): 095708. [16] Hartmann U. Magnetic force microscopy.Annual Review of Materials Science. 1999; 29(1): 53-87. [17] Belliard L, Thiaville A, Lemerle S, Lagrange A, Ferre J, Miltat J. Investigation of the domain contrast in magnetic force microscopy. Journal of Applied Physics. 1997; 81(8): 3849-3851. [18] Mironov VL, Nikitushkin DS, Bins C, Shubin AB,Zhdan PA. Magnetic force microscope contrast simulation for low-coercive ferromagnetic and superparamagnetic nanoparticles in an external magnetic field. IEEE Transactions on Magnetics. 2007; 43(11): 3961-3963. [19] Neves CS, Quaresma P, Baptista PV, Carvalho PA, Araújo PJ, Pereira E, Eaton P. New insights into the use of magnetic force microscopy to discriminate between magnetic and nonmagnetic nanoparticles. Nanotechnology. 2010; 21(30): 305706. [20] Vakarelski IU, McNamee CE, Higashitani K. Deposition of silica nanoparticles on a gold surface via a self-assembled monolayer of (3-mercaptopropyl) trimethoxysilane. Colloids and Surfaces A: Physiochem. Eng. Aspects. 2007; 295(1): 16-20. Copyright© 2014 Gustavo Cordova Simon Attwood, Ravi Gaikwad, Frank Gu and Zoya Leonenko. 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 author and source are credited. Nano Biomed Eng 2014, Vol. 6, Issue 1http://www.nanobe.org
1307.0410
2
1307
2013-07-17T21:18:11
Elastic Properties and Line Tension of Self-Assembled Bilayer Membranes
[ "physics.bio-ph", "cond-mat.soft" ]
The elastic properties of a self-assembled bilayer membrane are studied using the self-consistent field theory, applied to a model system composed of flexible amphiphilic chains dissolved in hydrophilic polymeric solvents. Examining the free energy of bilayer membranes with different geometries allows us to calculate their bending modulus, Gaussian modulus, two fourth-order membrane moduli, and the line tension. The dependence of these parameters on the microscopic characteristics of the amphiphilic chain, characterized by the volume fraction of the hydrophilic component, is systematically studied. The theoretical predictions are compared with the results from a simple monolayer model, which approximates a bilayer membrane by two monolayers. Finally the region of validity of the linear elasticity theory is analyzed by examining the higher-order contributions.
physics.bio-ph
physics
Elastic Properties and Line Tension of Self-Assembled Bilayer Membranes The State Key Laboratory of Molecular Engineering of Polymers, Department of Macromolecular Science, Fudan University, Shanghai 200433, China Jianfeng Li Kyle A. Pastor and An-Chang Shi∗ Department of Physics and Astronomy, McMaster University, 1280 Main Street West, Hamilton, Ontario, Canada L8S 4M1 Friederike Schmid and Jiajia Zhou† Institut fur Physik, Johannes Gutenberg-Universitat Mainz, Staudingerweg 7, D-55099 Mainz, Germany The elastic properties of a self-assembled bilayer membrane are studied using the self-consistent field theory, applied to a model system composed of flexible amphiphilic chains dissolved in hy- drophilic polymeric solvents. Examining the free energy of bilayer membranes with different geome- tries allows us to calculate their bending modulus, Gaussian modulus, two fourth-order membrane moduli, and the line tension. The dependence of these parameters on the microscopic characteris- tics of the amphiphilic chain, characterized by the volume fraction of the hydrophilic component, is systematically studied. The theoretical predictions are compared with the results from a simple monolayer model, which approximates a bilayer membrane by two monolayers. Finally the region of validity of the linear elasticity theory is analyzed by examining the higher-order contributions. I. INTRODUCTION Amphiphilic molecules, such as lipids, surfactants and block copolymers, are composed of hydrophilic and hy- drophobic components. When dispersed in water, the amphiphilic molecules spontaneously self-assemble into a variety of structures, e.g., spherical and cylindrical mi- celles or bilayer membranes. The self-assembly is driven by the competing interactions between water and dif- ferent parts of the amphiphilic molecules. Specifically, in the case of bilayers the hydrophilic parts stay on the outside of the bilayer, while the hydrophobic blocks are hidden in the interior. At the mesoscopic scale, bilayer membranes exhibit several unique properties [1]: they can form closed membranes without edges, such as vesi- cles and cell walls; they are extremely flexible and highly deformable; and despite their flexibility, they keep their structural integrity even under strong deformations. This combination of stability and flexibility motivates a phe- nomenological description of bilayer membranes, in which the membranes are modelled as two-dimensional sur- faces. The properties of the membrane within this surface model are then described by a set of elastic parameters in- cluding the spontaneous curvature c0, the bending mod- ulus κM , the Gaussian modulus κG, and the line tension σ of an open membrane edge. One of the main objec- tives for experimentalists and theorists is to determine these elastic parameters and to understand their physi- cal origins. The elastic properties can be used to analyze and explain numerous phenomena associated with vesicle ∗[email protected][email protected] deformation, membrane fusion, and other relevant mem- brane activities. A number of techniques have been developed to mea- sure the membrane elastic properties. Particular atten- tion has been paid to the bending modulus κM . Exper- imentally, κM can be obtained either by monitoring the thermal fluctuations of a membrane [2 -- 8], or by directly measuring the force required to pull or deform a tether [9, 10]. Similar methods have been employed in simula- tion studies of the bending modulus of bilayers [11 -- 14]. Measuring the Gaussian modulus κG is more challenging, since its contribution to the free energy of closed mem- brane systems changes only if they undergo topological changes, due to the celebrated Gauss-Bonnet theorem [15 -- 17]. A topological change often involves unstable or metastable transient membrane states. Performing mea- surements in such states is a challenging task, both in experiments and simulations. In Ref. [18], the Gaussian modulus of phospholipid bilayers was measured by de- tailed observation of the phase transition from the lamel- lar phase to the inverted hexagonal phase. In Ref. [19], the difference between Gaussian moduli of two types of lipid bilayers was obtained. The Gaussian modulus has also been determined in simulations by quantifying the probability that a membrane patch closes up to form a vesicle [20]. Finally, a third important elastic parame- ter of membranes is the spontaneous curvature c0. For bilayer membranes made of two leaflets with identical composition, it is zero. We will focus on this special case in this work. The techniques described above are designed to de- termine just one specific elastic parameter at a time. An alternative strategy is to compare the free energy of membranes with different curvatures. By studying bilay- ers in different geometries (planes, cylinders, spheres), it is possible to obtain several elastic constants simulta- neously. This approach can be implemented quite nat- urally in theoretical studies, and there are also experi- mental studies based on the equilibrium between bilayer structures with different topologies [21]. In theoretical studies, the focus has been on relating the elastic prop- erties to the microscopic parameters of the amphiphiles. Based on the strong-segregation theory of grafted poly- mers, an analytic theory of the monolayer elastic con- stants was developed by Wang and Safran [22, 23]. This analytic theory was generalized to diblock copolymer bi- layers by Wang [24]. Similar properties for a bilayer composed of diblock copolymers were studied by Ohta and Nonomura [25], who investigated the dependence of the bending and Gaussian moduli on the architec- tural parameter of the diblock copolymers. Apart from these analytic theories, numerical calculations using the self-consistent field theory (SCFT) have been carried out to study the elastic properties of bilayers, using di- block copolymer/homopolymer blends to mimic the am- phiphile/solvent system. Laradji and Desai [26] employed SCFT to calculate the surface tension and bending mod- ulus from the power spectrum of capillary modes. Mat- sen [27] introduced a new SCFT method to evaluate the elastic properties of a diblock monolayer. In his work, monolayers with different surface curvature were stabi- lized by exerting a force on the interface. Muller and Gompper [28] extended the study to monolayers com- posed of pure diblock copolymers, mixtures of diblock copolymers, and triblock copolymers. Chang and Morse [29] used a pressure difference to stabilize a monolayer in curved geometries. Nunalee et al. [30] presented a spontaneous curvature map for monolayers in the param- eter space of the interaction parameter and the diblock architectural parameter. In Ref. [31], the spontaneous curvature, bending and Gaussian moduli of a monolayer were extracted using SCFT calculations and compared to those measured in experiments. However, despite its ac- curacy, the SCFT method has not yet been employed for determining the elastic properties of bilayer membranes except for the area compressibility [31]. We turn to discussing the line tension. Two types of line tensions can be associated with liposomes: The line tension of the phase boundary between two phases of lipids coexisting in the same membrane [32], and the edge free energy of an open lipid membrane. In the present work we focus on the latter. The line tension of an open edge is a key parameter for understanding the processes of disc-to-vesicle transformation, vesicle-pore formation and membrane fusion. These processes often involve membranes which are highly curved in compar- ison to their thickness; therefore the influence of large curvatures should be examined. Experimentally, measurements of line tensions rely on the creation of bilayer membranes with edges: One can either stabilize a bilayer membrane patch with open edges [33], or create pores on vesicle surfaces, or in planar mem- branes [34, 35]. The line tension has also been deter- 2 mined from computer simulations of various atomistic or coarse-grained models, both for mechanically stretched membranes [36 -- 40] and tensionless membranes [41 -- 44]. In comparison, theoretical studies of the line tension are scarce[45, 46]. In this paper, we report a unified systematic study of the elastic properties of self-assembled bilayer mem- branes. Specifically, we employ SCFT formulated in dif- ferent geometries to calculate the elastic moduli and line tension of a bilayer membrane. The bilayer membrane is described by a microscopic model of flexible amphiphilic chains (AB diblock copolymers) dissolved in hydrophilic solvent molecules (A homopolymers). Our SCFT calcu- lations are carried out for membranes with zero or nearly zero surface tension in various geometries. Open mem- branes in the form of disks and/or membranes with pores are stabilized using appropriate constraints. The size of these stable disks or pores can be varied, which enables us to evaluate the edge line tension. The spontaneous cur- vature, the bending and the Gaussian moduli can be ex- tracted similarly to Refs. [27, 28, 31] by fitting the SCFT free energies of membranes with different geometries to an appropriate energy expression for the continuum elas- tic model. In the elastic model, the bilayer membrane is repre- sented by a two-dimensional surface whose energy is de- scribed in terms of its elastic energy. If this tensionless bilayer is not highly curved, its free energy can be well represented by a linear elastic model (the Helfrich model [47, 48]) (1) F =Z (cid:2)2κM (M − c0)2 + κGG(cid:3) dA +Z σdL, where M = (c1 + c2)/2 and G = c1c2 are the local mean and Gaussian curvatures of the deformed bilayer (c1,2 are the two principal curvatures). The last term in Eq. (1) represents the edge energy of an open membrane, and it vanishes for closed membranes. The elastic constants of interest, c0, κM , κG and σ, are the spontaneous curva- ture, the bending modulus, the Gaussian modulus and the edge line tension, respectively. We emphasize that the Helfrich free energy corresponds to a linear elastic model, which includes only the lowest-order contribu- tions to the free energy from the curvatures. In the current work, we derive the elastic parameters from the SCFT free energy, focusing on the dependence of these quantities on the volume fraction of the hydrophilic blocks, fA. Furthermore, we carefully examine the in- fluence of large curvatures on the accuracy of the linear elasticity theory. We find that higher-order contributions to the membrane free energy become significant when the curvature is large. To analyze them, we use an extension of the linear elasticity theory that includes fourth-order contributions, F = Z h2κM (M − c0)2 + κGG + κ1M 4 +κ2M 2G + κ3G2idA +Z σdL, (2) where κ1, κ2 and κ3 are fourth-order curvature moduli. With the geometries employed in our calculations, we can determine two linear combinations of the three fourth- order moduli. Even though two such combinations do not suffice to calculate the three moduli explicitly, the values of these two constants clearly show that the higher-order terms are not negligible for highly-curved membranes. It should be noted that in writing Eq. (2) we have neglected fourth-order terms involving the derivative of the curva- tures, such as (∇M )2, ∇2M 2 and ∇2G. In addition, third-order terms are also not included in Eq. (2) due to the symmetry of the bilayers. For a bilayer membrane composed of two identical leaflets, odd-order powers of the curvature vanish in the bending energy expansion. The remainder of this paper is organized as follows: Section II describes the SCFT model of bilayer mem- branes and the geometric constraints used in the study. Our results on the elastic properties of the membranes are presented in section III, and compared with the pre- diction from the two-monolayer approximation when ap- propriate. Finally, section IV concludes with a brief sum- mary. II. MODEL AND METHOD In this section, we first review the theoretic frame- work for the calculation of the free energy of a bilayer membrane. The general theory of SCFT has been well documented in several review articles and monographs [49 -- 54], and we refer readers to them for details. We then describe the implementation of SCFT in different geometries, which we use in order to extract the elastic constants. A. Excess free energy of a bilayer membrane The molecular model for the bilayer membranes of interest consists of a binary mixture of AB diblock (amphiphilic) copolymers and A-type (hydrophilic) ho- mopolymers in a volume V . In the present model, the volume fraction of the hydrophilic blocks in the am- phiphile is denoted by fA. For simplicity, we assume that both the copolymers and the homopolymers have equal chain length characterized by a degree of polymerization N . Furthermore, we assume that the A/AB blend is in- compressible, and both monomers (A and B) have the same monomer density ρ0 (or the hardcore volume per monomer is ρ−1 0 ) and Kuhn length b. The interaction between the hydrophilic and hydrophobic monomers is described by the Flory-Huggins parameter χ. We formu- late the theory in the grand-canonical ensemble and use the chemical potential of the homopolymers as a refer- ence. The controlling parameter is the copolymer chem- ical potential µc, or its activity zc = exp(µc). Within the mean-field approximation, the grand free energy of a binary mixture has the form [55, 56]: 3 N F kBT ρ0 = Z drhχN φA(r)φB (r) − ωA(r)φA(r) − ωB(r)φB (r) − ξ(r)(1 − φA(r) − φB(r)) − ψδ(r − r1)(φA(r) − φB(r))i − zcQc − Qh, (3) where φα(r) and ωα(r) denote the local concentration and the mean field of the α-type monomers (α = A, B). The local pressure ξ(r) is a Lagrange multiplier intro- duced to enforce incompressibility. A second Lagrange multiplier, ψ, is used to implement constraints for stabi- lizing the bilayer in different geometries. A delta func- tion, δ(r−r1), is used to ensure that the ψ field only oper- ates on the interface at a prescribed position r1. The last two terms in Eq. (3) are the configurational entropies, re- lated to the single-chain partition functions for two types of polymer, Qc and Qh. For the copolymer, the partition function has the form Qc =R drqc(r, 1), where qc(r, s) is an end-integrated propagator, and s is a parameter that runs from 0 to 1 along the length of the polymer. The propagator satisfies the modified diffusion equation ∂ ∂s qc(r, s) = R2 g∇2qc(r, s) − ωα(r)qc(r, s), (4) where Rg = bpN/6 is the gyration radius of the copoly- mer chains. The mean field ωα(r) is a piece-wise function where α = A if 0 < s < fA and α = B if fA < s < 1. The initial condition is qc(r, 0) = 1. Since the copoly- mer has two distinct ends, a complementary propaga- tor q† It satisfies Eq. (4) with the right-hand side multiplied by −1, and the initial condi- tion q† c (r, 1) = 1. For the homopolymer, one propagator qh(r, s) is sufficient, and the single-chain partition func- c(r, s) is introduced. tion has the form Qh =R drqh(r, 1). In our calculations the modified diffusion equations are solved in real space using the Crank-Nicolson method [57]. The SCFT method employs a mean field approxima- tion to evaluate the free energy using a saddle-point tech- nique. This demands the functional derivatives of the expression (3) to be zero, δF δφα = δF δωα = δF δξ = δF δψ = 0. (5) Carrying out these functional derivatives leads to the fol- lowing mean-field equations: φA(r) = Z 1 0 ds qc(r, s)q† ds qh(r, s)qh(r, 1 − s) +zcZ fA φB(r) = zcZ 1 ωA(r) = χN φB(r) + ξ(r) − ψδ(r − r1), ωB(r) = χN φA(r) + ξ(r) + ψδ(r − r1), ds qc(r, s)q† c (r, s), c (r, s), 0 fA 1 = φA(r) + φB(r), φA(r1) = φB(r1). (6) (7) (8) (9) (10) (11) These equations can be solved by iteration. We are interested in the free energy of a system con- taining a bilayer membrane compared to that of a refer- ence system without a bilayer, i.e., a homogeneous blend of copolymers and homopolymers. The free energy for the homogeneous bulk phase can be computed analyti- cally, N Fbulk kBT ρ0V = (1 − φbulk)(cid:2) ln(1 − φbulk) − 1(cid:3) + φbulk(cid:2) ln φbulk − 1(cid:3) + χN (1 − fA)φbulk(cid:2)1 − φbulk + fAφbulk(cid:3) − µcφbulk, (12) where φbulk is the bulk copolymer concentration. In the grand canonical ensemble, the bulk copolymer concen- tration depends on the copolymer chemical potential µc via the relation: µc = ln φbulk − ln(1 − φbulk) + χN (1 − fA)(cid:2)1 − 2(1 − fA)φbulk(cid:3). (13) For a bilayer membrane, its excess free energy (F − Fbulk) is proportional to the area of the membrane. We can define an excess free energy density, F , as the free energy difference between the systems with and without the bilayer membrane, divided by the area, A, of the membrane, F = N (F − Fbulk) kBT ρ0A . (14) Another useful quantity characterizing the membrane is the copolymer excess per unit area, Ω = 1 AZ drhφc(r) − φbulki, (15) where φc(r) is the local concentration of the AB diblock copolymers. B. Geometrical constraints In order to extract information on the various elastic properties, we calculate the excess free energy of a bi- layer membrane in the following five geometries: (i) an 4 infinite planar bilayer membrane, (ii) a cylindrical bilayer membrane with a radius r, which is extended to infin- ity in the axial direction, (iii) a spherical bilayer with a radius r, (iv) an axially symmetric disk-like membrane patch with a radius R, and (v) a planar membrane with a circular pore of radius R. The first three geometries, which can be reduced to a one-dimensional problem by an appropriate coordinate transformation, are employed to extract the bending modulus and the Gaussian mod- ulus as well as higher-order curvature moduli. The last two geometries, which are two-dimensional systems due to their axial symmetry, are used to extract the line ten- sion of the membrane edge. To stabilize a bilayer membrane in these geometries, a constraint term, ψδ(r − r1)(φA − φB), has been included in the free energy expression (3). From a mathematical point of view, this term ensures that at the point speci- fied by r1, the concentrations of the hydrophilic and hy- drophobic monomers are equal, i.e., φA(r1) = φB(r1). In the cylindrical and spherical geometries, this constraint sets the curvature radius of the membrane, and in the disk/pore geometry, it defines the size of the disk/pore. It should be noted that in the cylindrical and spherical case, the constraint is applied to the outer monolayer only, and the inner monolayer is free. This allows the bi- layer to optimize its thickness. We found that constrain- ing the inner monolayer while removing the constraint from the outer monolayer lead to identical results. In order to ensure that the calculated membrane properties are not affected by finite size effects, we choose a cal- culation box large enough such that the monomer den- sities reach their bulk values at the boundaries. With these geometric constraints, we obtained the excess free energies for the five geometries, which are denoted F X , where X = 0, C, S, D, P for the planar, the cylindrical, the spherical, the disk and the pore geometry, respec- tively. Furthermore, we focus on tensionless membranes in this work. The chemical potential of the amphiphilic diblock copolymers, µc, is adjusted carefully such that the grand free energy of the bulk system and of a planar bilayer are identical, i.e., F 0 = 0. For a curved bilayer of finite thickness, the definition of the interface position or the membrane surface involves a certain degree of arbitrariness. To be consistent with our constraint method, we define the bilayer interface to be at the mid point of the two positions where A- and B- segment concentrations are equal. For a cylindrical mem- brane with radius r, the mean curvature is M = 1/(2r) and the Gaussian curvature vanishes. For a spherical membrane with radius r, the mean curvature and the Gaussian curvature are M = 1/r and G = 1/r2, re- spectively. With these curvatures, the modified Helfrich free energy (2) can be written in terms of the curvature c = 1/r in the cylindrical and spherical geometries, F C = −2κM c0c + F S = −4κM c0c + (2κM + κG)c2 + BSc4, c2 + BC c4, κM 2 (16) (17) where the higher order parameters BC = κ1/16 and BS = κ1 + κ2 + κ3 will be termed fourth-order cylindrical and spherical modulus, respectively. In the future stud- ies, one can use ellipsoidal geometry with revolutionary symmetry to extract these moduli separately, where the entire shape of the outer layer has to be fixed during the SCFT calculations. Following Ref. [31], we use the thick- ness, d = 4.3Rg, of a planar bilayer with equal fractions of hydrophilic and hydrophobic segments (fA = 0.50) as the unit of length, where the thickness in planar geometry is given by the copolymer excess (15), d = Ω. For the sur- face tension, a natural unit is the interfacial free energy per unit area between coexisting A and B homopolymers in the limit of large χN , γint = pχN/6 ρ0kBT b/√N . Using this convention, the modulus can be made dimen- sionless: κM = κM γintd2 , κG = κG γintd2 , (18) BC = κ1 16γintd4 , BS = κ1 + κ2 + κ3 γintd4 . (19) Similar to the procedures described above, the line ten- sion of an open edge can be calculated by examining the free energy of membranes as a function of the disk or pore sizes, F DA = 2πσR + const, F P A = 2πσR + const, (20) (21) where R is the diameter of the disk/pore. The defini- tion of the edge position is again somewhat arbitrary for a membrane of finite thickness. However, the resulting value for σ does not depend on the specific convention used for calculating R. Changing the convention will only affect the constant offset in the free energy expressions, Eqs. (20) and (21). Here, we define 2R as the distance between two diametrally opposed points in the bilayer mid-plane that satisfy the constraint φA(r) = φB(r). Us- ing the length unit defined above, we define a unit for the line tension, σ0 = kBT ρ0d2/N . The dimensionless line tension is then given by σ = σ σ0 . (22) III. RESULTS AND DISCUSSION In this section, we first present the SCFT results for the second-order elastic moduli (κM and κG) and discuss their relation to the microscopic parameter fA. We then examine highly curved bilayer membranes and demon- strate that the contribution to the free energy from the higher-order moduli can be significant. Finally we study the bilayer membrane in disk/pore geometries and com- pute the line tension by varying the structure size. Since we mainly focus on the effect of the hydrophilic volume 5 fraction, fA, on the bilayer properties, most of the re- sults are presented for a particular interaction strength between the A and B monomers. That is, we have fixed χN = 30 unless otherwise stated. For this intermediate segregation case, the chemical potential of the copolymer is around µ = 4.6kBT when a planar bilayer membrane becomes tensionless. It should be noted that the inter- action parameter is also important, but a change in the interaction parameter does not produce any qualitative changes to the fA-dependence of the elastic moduli. A. Linear Elasticity: Bending and Gaussian moduli From the SCFT free energy of cylindrical and spher- ical membranes at large radii or small curvatures, the second-order moduli, the bending modulus κM and the Gaussian modulus κG, can be calculated by fitting the free energy curve to the Helfrich model. The basic struc- tural properties of a self-assembled bilayer membrane are the spatial distribution of the hydrophilic and hydropho- bic monomers across it. A typical concentration profile for a bilayer membrane in the spherical geometry with a radius R = 7.7Rg and fA = 0.50 is shown in Fig. 1. Within the spherical bilayer membrane, since the area of the inner interface is smaller than the outer one, the hy- drophilic monomers (A-blocks) in the inner leaflet have to pack more closely and have a higher local concentration in comparison to the outer leaflet. At the same time, the hydrophilic monomers in the inner interface also have a wider distribution in the axial direction, which indicates that the hydrophilic chains in the same region are more stretched than the outer chains. The loss of conforma- tional entropy due to the local extension of hydrophilic chains is the main contribution to the bending energy of the membranes. The excess free energies for a tensionless bilayer with fA = 0.50 in the cylindrical and spherical geometries are shown in Fig. 2. The free energies are plotted as a function of the dimensionless curvature cd. In the in- set, we show the difference between the free energies ob- tained by different methods of applying the constraint. For highly curved membranes, the free energy with fixed inner monolayer is slightly higher than that with fixed outer monolayer. In general, one would expect that the free energy will depend on the specifics of how the con- straint is applied, with the exceptions of (meta)stable states and saddle points where the conjugated field ψ vanishes. However, for the tensionless bilayer membranes studied here, the deviation introduced by using different constraints is small, so we opt for the method of applying the constraint to the outer monolayer in this study. Ideally, one can fit the excess free energy as a polyno- mial function of the curvature up to fourth order, then extract the elastic properties from the fitting parameters using Eqs. (16) and (17). The zeroth-order term should be zero for a tensionless membrane, and indeed this is the case as shown in Fig. 2, where the excess free en- φ 1 0.8 0.6 0.4 0.2 0 φ cBφ cAφ hA 4 6 8 r/Rg 10 12 FIG. 1: The concentration profile for a bilayer membrane in the spherical geometry with a radius R = 7.7Rg and fA = 0.50. The concentration for the hydrophilic solvent (φhA), the hydrophilic (φcA) and hydrophobic (φcB) segments of the amphiphiles, are shown as dot-dash, dashed and solid lines, respectively. FC/γ FS/γ int int 0.08 0.06 10-4 0.04 10-5 0.02 0.00 t n i / γ F 0.0 0.5 1.0 0.0 0.2 0.4 0.6 0.8 1.0 cd FIG. 2: The excess free energy for a tensionless bilayer membrane with fA = 0.50 as a function of the dimensionless curvature cd in the cylindrical (square symbols) and spherical (sphere symbols) geometries. The curvature of the membrane is maintained by fixing the position of the outer monolayer. Alternatively, one can also fix the inner monolayer. The difference of the free energy obtained with these different implementations of the constraints is shown in the inset for the cylindrical (solid line) and spherical geometry (dashed line). ergies go to zero when the curvature goes to zero. The linear term in the free energy plot is associated with the spontaneous curvature of the membranes. For a bilayer membrane consisting of two identical leaflets, the spon- taneous curvature is zero by symmetry. This can be seen from Fig. 2 where the slope of the free energy curves also vanishes at the zero curvature. Furthermore, the symme- try of the membranes dictates that the higher odd-order terms also vanish. 6 The first non-zero coefficient in the curvature expan- sion of the membrane free energy is the second-order term, which is related to the bending and Gaussian mod- uli. Fig. 3 shows κM and κG, as a function of the hy- drophilic volume function fA. Beyond fA ≥ 1 −p2/χN (= 0.74 for χN = 30), bilayer membranes are unstable with respect to the homogeneous phase. In the stable regime below this value, the dependence on fA is quite different for the two quantities, κM and κG. The bend- ing modulus, κM , on the one hand, is not very sensi- tive to the amphiphilic architecture specified by fA. As one might expect, it exhibits a weak maximum close to fA = 0.50 (symmetric amphiphiles), which is however not exactly symmetric with respect to fA = 0.50 due to the different solubility of the amphiphilic molecules. Molecules with longer hydrophilic parts can be dissolved in the solvent more easily, resulting in a higher excess at the interface. The Gaussian modulus, κG, on the other hand, is a strongly varying monotonically decreas- ing function of the hydrophilic volume fraction, and its value changes from positive to negative, crossing zero at around fA = 0.41. The planar bilayer becomes unstable for positive values of κG. It is interesting to note that the ratio of the Gaussian modulus and the bending modulus, κG/κM , decreases from roughly 1 to -2 as fA is increased, changing its sign at about fA = 0.41. The effect of varying the chain interaction strength on the qualitative behavior of the second-order moduli is small, as shown in Fig 3 for different values of χN . The bending modulus κM increases with increasing in- teraction parameter χN . The overall dependence on the hydrophilic architecture fA is the same for different χN . Even more interestingly, the Gaussian modulus κG shows a similar relation. When the interaction strength in- creases, the magnitude of the modulus also increases, and when the interaction strength decreases, the magnitude of κG does the same. The zero crossing point for all inter- action strengths is the same at approximately fA = 0.41. The architectural dependence is not drastically affected as the Gaussian modulus κG is still a monotonically de- creasing function in all cases. This result shows that the choice of interaction parameter has no qualitative effect on the behaviour of the elastic moduli as a function of the amphiphile architecture. The calculated elastic properties of a bilayer membrane can be compared with those of its corresponding mono- layers. The monolayer properties have been well stud- ied with the SCFT by several authors [27 -- 31]. The bi- layer properties can be derived from monolayer param- eters by a two-monolayer approximation as discussed in Appendix A. In this simple model the bilayer membrane is depicted by two independent monolayers sticking to- gether without any specific interactions, and the elastic properties of the bilayer can be expressed in terms of the monolayer counterparts. For instance, for a symmetric (a) (a) ) 2 d n t i γ ( / M κ 2 0 1 (b) (b) ) 2 d n t i γ ( / G κ 2 0 1 χN=25 χN=30 χN=35 χN=25 χN=30 χN=35 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 6.0 4.0 2.0 0.0 -2.0 -4.0 -6.0 -8.0 0.30 0.40 0.50 fA 0.60 0.70 FIG. 3: The bending modulus κM and the Gaussian modulus κG of a bilayer membrane as a function of the hydrophilic fraction fA. The results for different interaction strength χN are shown in different symbols. We note that the reference energies, γint, increase with χN via γint ∝ √χN . This trivial dependence has been eliminated by plotting the dimensionless renormalized quantities. The dashed curves correspond to the results from the two-monolayer approximation, Eqs. (23) and (24), for χN = 30. bilayer [18, 58, 59] we have, κM = 2¯κM , (23) (24) κG = 2(cid:2)¯κG − 4¯κM ¯c0δ(cid:3). In the above expressions, over-barred quantities denote the properties of the monolayer, and δ is the distance between the midplane of the bilayer and the monolayer interface. The value of δ is a linearly decreasing function of fA (not shown here). We also calculated the monolayer properties with the SCFT and the results are similar to those reported in Ref. [31]. The spontaneous curvature of the monolayer increases in an approximately linear manner with the hydrophilic fraction fA, with ¯c0 = 0 for symmetric amphiphiles with fA = 0.50. In Fig. 3, the prediction from the monolayer properties are plotted as dashed curves for χN = 30.0. It is interesting to notice that the simple two-monolayer model is in good 7 qualitative agreement with the exact results from SCFT. The bending modulus of the bilayer is roughly twice that of the monolayer. In previous simulations and exper- iments, stacking monolayers are also found to reinforce the rigidity of the membrane [18, 60]. The prediction for the bending modulus from the two-monolayer approxi- mation must necessarily be symmetric about fA = 0.50, due to the fact that the bending modulus for monolayer is symmetric about fA = 0.50. The SCFT results are slightly different, with the maximum shifted to a smaller hydrophilic fraction. This can be attributed to the differ- ence between the two-monolayer approximation and the full bilayer calculation: In the latter case one considers two apposing (and possibly coupled) monolayers, while in the former case, the apposing monolayer is replaced by a hydrophobic homopolymer melt. In our case, where copolymers and homopolymers have equal length, one would expect the monolayer approximation to be better when the composition of copolymers approaches that of hydrophobic (B-)homopolymers (small fA). The Gaussian modulus for a bilayer involves not only the two curvature moduli of the monolayer, but also its spontaneous curvature, ¯c0. At ¯c0 = 0, the Gaussian mod- ulus is slightly negative. Due to the interplay of mono- layer stiffness and spontaneous curvature, it is further reduced at fA > 0.50, and shifted upwards for fA < 0.50. B. Nonlinear Elasticity: Higher-order Moduli In the above subsection, the SCFT results demonstrate that the Helfrich model gives an excellent description of the free energy of slightly curved bilayer membranes. The Helfrich expression essentially corresponds to a linear elasticity theory of membranes, where the higher-order contributions are ignored. It is therefore desirable to examine the region of validity of the linear elastic the- ory. Specifically, we expect that the higher-order terms in Eq. (2) can no longer be ignored in systems of highly curved membranes. Examining the relative contribu- tion of the higher-order terms should then allow us to establish regions of validity for the Helfrich model. In Ref. [14], Harmandaris and Deserno employed a series of mesoscopic simulations of a coarse-grained model to test the validity of the Helfrich model and found that the quadratic Helfrich energy remains valid for membranes with curvature radius comparable to the bilayer thick- ness. However, it should be noted that these authors em- ployed the Cooke model [61, 62], which replaces the lipid tail by two beads. This high level of coarse-graining may limit the ability of the model to describe highly curved membrane structures. In more refined models, one im- portant contribution to higher-order terms is the change of the packing of the lipid tails upon bending the bi- layer. In Ref. [63], Risselada et al. used the MARTINI model to study vesicles formed of lipid bilayers. The higher-order contribution to the Helfrich Hamiltonian is important for explaining the inverted domain sorting un- der uniaxial compression. The higher-order elastic con- stants are also shown to stabilize the intermediate states of open vesicles in the disk-to-vesicle transition [64]. In the SCFT calculations, the stretching of the amphiphilic chains is captured, which allows us to examine the effects of the highly curved membranes. 8 Quantitatively, the fourth-order term can contribute as much as 10% to the total free energy for highly curved membranes. This contribution should not be neglected. Due to the limited number of geometries employed in our calculations, we are only able to calculate the cylindrical and spherical fourth order moduli defined in Eq. (19). In Fig. 5, these two higher-order moduli are plotted as a function of the hydrophilic fraction fA. The general trend is that the higher-order contribution is positive for cylindrical geometry, but negative for spherical bilayers. (a) (a) 0.0252 0.0248 0.0244 0.0240 0.0236 0.0232 0.0670 (b) (b) 0.0660 0.0650 FC/(c2d2γ int) FS/(c2d2γ int) 0.2 0.0 -0.2 -0.4 -0.6 0.30 0.40 intd4) intd4) 102 BC/(γ 102 BS/(γ 0.50 fA 0.60 0.70 0.0640 0.0 0.2 0.4 0.6 0.8 1.0 (cd)2 FIG. 4: Rescaled free energy curves F/(cd)2 plotted as a function of squared curvature (cd)2 for (a) the cylindrical and (b) the spherical membranes. Symbols and parameters are the same as in Fig. 2. The effect of the higher-order terms on the membrane free energy can be demonstrated by plotting the rescaled free energy, F/(cd)2 , as a function of the effective cur- vature (cd)2, as shown in Fig. 4 for both cylindrical and spherical geometries. If the quadratic expression of the Helfrich model is strictly valid, the scaled free energy F/(cd)2 should remain constant in these plots. Contrary to this expectation, the SCFT results shown in Fig. 4 re- veal that the scaled free energy of the membrane changes as (cd)2 is increased. For small values of (cd)2, the scaled free energy is approximately a linear function of (cd)2. The slope of this linear function gives the quartic cor- rection to the free energy. For highly curved bilayers, contributions of terms with even higher order come into play, and the scaled free energy deviates from the straight line. From these observations we conclude that higher or- der terms should be included when describing the behav- ior of highly curved membranes within an elastic model. FIG. 5: Fourth-order moduli BC and BS as a function of the hydrophilic fraction fA. With our results for the second and fourth-order mod- uli, we can now determine a boundary separating the region where the fourth-order corrections are negligible from that where the fourth-order terms start to play an important role. We require that the relative energy dif- ference, ∆E = F C 4 − F C 2 F C 2 , (25) does not exceed a given threshold ∆E∗. Here F C 4 is the free energy for the cylindrical geometry including the fourth-order contributions to the free energy, and F C is the free energy of the cylindrical geometry assuming that the fourth-order moduli are negligible. A similar definition is used for the spherical geometry. Applying Eqs. (16) and (17), one obtains the following expressions for the critical curvatures for the cylindrical and spherical bilayers: 2 c∗ c∗ 2BC C = r κM ∆E∗ S = s (2κM + κG)∆E∗ BS (26) (27) . Using the values of the moduli as a function of hy- drophilic chain fraction (fA) as obtained in the previous sections, the boundary of validity for the linear elastic 0.70 0.65 0.60 0.55 A f 0.50 0.45 0.40 0.35 0.30 0.2 cyl sph 0.4 0.6 0.8 cd 1.0 1.2 FIG. 6: Curvature boundaries for both the cylindrical and spherical bilayer geometries. The results for the criteria ∆E∗ = 2% are shown by filled symbols and solid lines, and the results for ∆E∗ = 5% are shown by unfilled symbols and dashed lines. Regions where the second-order terms adequately describe the bending energy (with ∆E∗ = 2%) for both geometries are filled with crossed stripes. Unpatterned regions or regions filled with simple stripes correspond to curvatures where fourth-order corrections are needed to describe the bending energy for at least one geometry (with ∆E∗ = 2%). theory is determined for both the cylindrical and spher- ical geometries. The results for ∆E∗ = 2% and 5% are shown in Fig. 6. Let us first examine the criterion ∆E∗ = 2%, which corresponds in Fig. 6 to the filled symbols and solid lines. In the patterned regions the second-order description of the bilayer bending energy is valid up to the prescribed accuracy ∆E∗. In the unpatterned regions, this descrip- tion becomes less accurate and the fourth-order moduli must be included to accurately describe the membrane energy. Above a hydrophilic chain fraction of fA ≈ 0.58 the absolute boundary on the critical curvature is dom- inated by the spherical geometry. In this region, mem- branes in cylindrical geometry can be described by the second-order bending moduli for a larger span of curva- tures up to cd = 0.70 for a chain fraction of approxi- mately fA = 0.70, but for spherical geometry, the valid region is significantly reduced to about cd = 0.2 for fA = 0.70. This trend is reversed below fA ≈ 0.58 where the absolute boundary is set by the cylindrical geome- try. In this range of chain fractions, the spherical cor- rections are relatively less important, which implies that the second-order descriptions can be used up to curva- tures of cd ≈ 0.8 for a chain fraction length of fA = 0.30. The unpatterned regions corresponds to parameter sets in the (cd, fA)-plane where the fourth-order terms must be included in order to describe the bending energy of the bilayer up to the accuracy ∆E∗ = 2%, regardless of 9 membrane geometry. If the accuracy criterion is relaxed to ∆E∗ = 5%, we obtain the curves shown in Fig. 6 in unfilled symbols and dashed lines, and the valid region for the second-order description is extended to larger cur- vatures up to around cd ∼ 1. C. Line Tension of Membrane Edge The calculation for the line tension of the membrane edge must be performed in two dimensions. In our study, two different geometries, a circular disk and a circu- lar pore, are employed. The excess free energies of the tensionless disk-like membranes of different size are cal- culated using the SCFT, and the results are shown in Fig. 7(a). The free energy is a linear function of the ra- dius R. The slope of the curve is then used to obtain the line tension. A typical density profile of a disk-like membrane is shown in the inset of Fig. 7(a). The cross- section contour of the disk can be approximated by two semicircles being separated by two parallel lines. Closely examining this contour, we note that the thickness in the middle of the disk is slightly smaller than the diameter of the semicircle, and each semicircle is connected smoothly to the two parallel lines by two inverted arcs. In Fig. 7(b), we plot the excess free energy for a bilayer membrane with a pore of varying radius. Similar to the case of the disk structure, the free energy is a linear func- tion of the radius R. Similar curves had been obtained in Ref. [65], and in Ref. [39], where the free energy at very small radius was found to deviate from the linear depen- dence. A cross-section of the pore structure is shown in the inset of Fig. 7(b). The thickness near the edge is slightly larger than that away from the edge, similar to the disk geometry. The calculated line tension as a function of the hy- drophilic volume fraction is shown in Fig. 8. The line tensions obtained from the disk and pore geometries are almost identical, reflecting the fact that when the radius R is large in comparison to the bilayer thickness, there is no difference between the edge of a disk and the edge of a pore. As seen in Fig. 8, the line tension decreases from positive to negative in an approximately linear way as a function of the hydrophilic volume fraction. For most lipids, such as DOPC or DOPE, their hydrophilic fractions are less than 0.50, corresponding to a large pos- itive line tension. The results for the line tension imply that a closed membrane in the form of vesicles is the pre- ferred structure in such cases. On the other hand, the line tension of membranes formed from lipids with large hydrophilic heads can even become negative, implying that it is possible to use these big-headed lipids as edge stabilizers. This result is also supported by simulations [43] and experiments [35]. In Ref. [43], it is found that the line tension of the bilayer depends linearly on the lipid tail length, which corresponds to the hydrophobic frac- tion 1 − fA, and adding short-tail lipids (with large fA) can be used to decrease the line tension of a membrane 10 disk pore 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 -0.05 0 σ σ / 10.0 R/d 15.0 20.0 -0.10 0.30 0.40 0.50 fA 0.60 0.70 (a) (a)(a)(a)(a) 2.0 disk ) 3 d π 2 ( / A D F 1.5 1.0 0.5 0.0 0.0 5.0 (b) (b)(b)(b)(b) 2.0 pore ) 3 d π 2 ( / A P F 1.5 1.0 0.5 FIG. 8: The dimensionless line tension σ/σ0, obtained from the disk and pore geometries, versus hydrophilic fraction fA. The dashed line shows the prediction from the coupled monolayer approximation (28). as demonstrated in the previous section. Furthermore, the simple analysis in Appendix B is unable to consider the contribution due to the derivative terms because of the discontinuity at the junction of flat and curved mem- brane surface. Given all these possible corrections, the qualitative agreement between SCFT results and the ap- proximate theory (28) is remarkable and indicates that the monolayer model does capture the physics of the edge elastic properties. IV. SUMMARY To summarize, we have systematically investigated the elastic properties of a bilayer membrane, including the line tension of the edge, the bending modulus, the Gaus- sian modulus, and the fourth-order moduli, by apply- ing the self-consistent field theory to a molecular model where the membrane is composed of amphiphilic chains dissolved in hydrophilic solvent molecules. The free en- ergy of membranes subject to different geometric con- straints has been calculated. The elastic properties were extracted from the SCFT free energy of the membrane in five geometries: the planar, the cylindrical, the spheri- cal, the disk, and the pore geometry. In particular, SCFT studies of a disk-like membrane and a pore in a planar membrane were used to extract the line tension of the membrane edge. We have explored the dependence of the elastic properties on the microscopic characteristics of the amphiphiles as specified by the hydrophilic vol- ume fraction fA. Three notable results are: (i) As the hydrophilic volume fraction increases, the line tension de- creases linearly from a positive value to zero and even to a negative value, indicating that big-headed lipids can be used as edge stabilizers. (ii) The Gaussian modulus is a monotonically decreasing function of the hydrophilic 0.0 0.0 5.0 10.0 R/d 15.0 20.0 FIG. 7: The excess free energy for a disk-like bilayer membrane (a), and a pore in a planar bilayer membrane (b). The amphiphilic molecules are symmetric, fA = 0.50. The insets show representative density profiles for the amphiphilic molecules of the corresponding structures. composed of a mixture of short- and long-tail lipids. The line tension for the disk/pore structure can also be estimated from the monolayer properties if we per- ceive the edge as a folded monolayer. In Appendix B, an analytic expression of the line tension is obtained for the two geometries, σ = π¯κM 1 − 4¯c0δ 2δ . (28) The prediction from this expression is plotted in Fig. 8 as dashed line. The analytic result correctly predicts that the line tension is a decreasing function of the hydrophilic volume fraction, but it overestimates the value of the line tension. In this calculation, we used the value δ for a planar bilayer membrane, while we can see from the pro- files (insets of Fig. 7), that the radius near the edge is slightly larger than the planar bilayer. Also, near the edge the monolayer is strongly curved and higher-order contribution to the bending energy become important, volume fraction, whereas the bending modulus is a con- cave function with a maximum around fA = 0.50. The ratio of the Gaussian modulus and the bending modulus changes from 1 to -2 as fA is increased from 0.30 to 0.70. (iii) The quadratic Helfrich model is accurate at small membrane curvatures, but it becomes inaccurate when the radius of curvature is comparable to the membrane thickness. Therefore, higher-order curvature moduli have been introduced and calculated. The elastic properties of a bilayer membrane could also be analyzed within a two-monolayer approximation, which is capable of predicting the properties of bilay- ers from the monolayer properties. Physical insights can be obtained from the relation between the elastic prop- erties of the bilayer and the monolayer. For example, the bending modulus of bilayers is roughly twice that of monolayers, while the Gaussian modulus of a bilayer is a combined effect of the bending modulus, Gaussian mod- ulus and spontaneous curvature of a monolayer. The line tension of a disk/pore structure largely originates from the bending energy (dictated by the monolayer bending modulus) of the monolayer that forms the bilayer edge. This work focuses on the influence of the amphiphilic architecture on the properties of single-component bi- layer membranes. It should be noticed that other fac- tors such as the interaction strength between different molecules can also be important. For example, the Flory- Huggins parameter determines the scale of the curvature moduli to a large extent, because our results are scaled by the interfacial tension between the A/B homopolymers. The present model also has a number of limitations con- cerning the higher-order contributions to the elastic free energy. Ideally, the fourth-order terms in the free energy expression (2) also include terms related to the derivative of the curvatures. For the planar bilayers, cylinders and spheres, the curvatures are uniform thus the derivative terms vanish. For the disk/pore geometries, their con- tribution will change the effective line tension. Further- more, the geometries considered in this work only allow us to obtain two linear combinations of the three fourth- order moduli. To extract these moduli separately, one can consider the cylinders with an ellipse cross-section or an ellipsoid in three-dimension [66]. In many biological membranes, the composition of the inner and outer membrane leaflets are different. The asymmetric composition arises naturally in cell mem- branes when the lipids are generated and initially in- serted into the inner leaflet, and the asymmetry often persists over long periods of time because the sponta- neous flip-flop of lipids between two leaflets is extremely slow [67]. The current SCFT model in general deals with thermodynamical equilibrium thus it is challeng- ing to treat dynamic processes, but one may develop clever constraint schemes to enforce the asymmetric com- position. Another possible cause of the asymmetry is that the fluids on two sides of the membrane have dif- ferent chemical compositions. This can be simulated in the present model by using homopolymers interact- 11 ing differently with the copolymers. Furthermore, multi- component bilayer membranes exhibit interesting fea- tures [68 -- 70], for example, lipids that are capable of sta- bilizing the edge tend to aggregate on the edge, resulting in a smaller line tension. Local curvature effects may also reduce the line tension of phase separated lipid do- mains in multi-component membranes, or stabilize mi- crophase separated "raft" structures [71]. The current model can be extended to multi-component membranes by introducing two or more copolymers, even with differ- ent architectures. The study of the elastic properties of multi-component bilayer membranes with the SCFT will be an appealing and exciting subject of future work. Acknowledgments We are grateful to Marcus Muller and Ashkan Dehghan for valuable discussions. We acknowledge support from the National Basic Research Program of China (Grant No. 2011CB605700), the National Natural Science Foun- dation of China (Grants No. 20874019, 20990231, and 21104010), the Natural Sciences and Engineering Re- search Council (NSERC) of Canada through Discovery and CREATE program, and the Deutsche Forschungs- gemeinschaft (DFG) through SFB 625. This work was made possible by the facilities of the Shared Hierar- chical Academic Research Computing Network (SHAR- CNET:www.sharcnet.ca) and Compute/Calcul Canada, and JGU Mainz (MOGON). Appendix A: Two-Monolayer: Bending and Gaussian Moduli Let us assume that the bilayer is composed of two identical monolayer leaflets, and each monolayer has the same spontaneous curvature ¯c0, bending modulus ¯κM and Gaussian modulus ¯κG. We will use symbols with top bar for the monolayer and symbols without for the bilayer. The bilayer is described by the area A at its midplane, and the mean and Gaussian curvature of the bilayer midplane are denoted by M = (c1 + c2)/2 and G = c1c2, where c1 and c2 are principal curvatures. The areas at the neutral surfaces of the outer and inner mono- layers, Aout and Ain, are related to that of the bilayer midplane by [16, 18, 58, 59] Aout = A(1 + 2M δ + Gδ2), Ain = A(1 − 2M δ + Gδ2), (A1) (A2) where δ is the separation between the monolayer neutral surface to the bilayer midplane. The mean and Gaussian curvatures at the neutral surfaces of the two monolayers 12 y z δ v u r (A3) (A4) (A5) (A6) x are A Aout , Mout = (M + Gδ) Gout = G A Aout , Min = −(M − Gδ) A Ain , Gin = G A Ain . The Helfrich free energy density for the monolayers are fout = 2¯κM (Mout − ¯c0)2 + ¯κGGout, fin = 2¯κM (Min − ¯c0)2 + ¯κGGout. (A7) (A8) The free energy of the bilayer results from a summation of the free energies of the outer and inner monolayer, f A = foutAout + finAin (A9) = h4¯κM M 2 + 2[¯κG − 4¯κM ¯c0δ(1 − 1 2 ¯c0δ)]GiA, where we have neglected the constant term and terms that are of higher than second order in the curvatures. The free energy of the bilayer membrane can be written in terms of the bilayer properties f A =h2κM (M − c0)2 + κGGiA. (A10) From Eqs. (A9) and (A10), we immediately obtain the relations between the bilayer and monolayer properties: c0 = 0, κM = 2¯κM , κG = 2(cid:2)¯κG − 4¯κM ¯c0δ(1 − 1 2 ¯c0δ)(cid:3). (A11) (A12) (A13) Appendix B: Two-Monolayer: Edge Line Tension The surface of a bilayer membrane disk near the edge can be regarded as the outer part of a torus [39, 72]. The torus surface can be parametrized using two angle u and v (see Fig. 9), r(u, v) =  (r + δ cos v) cos u (r + δ cos v) sin u δ sin v   (B1) where r is the radius of the torus center line and the thickness of the disk is 2δ. For the outer side of the torus, the range of the two angles are u ∈ [0, 2π] and v ∈ [−π/2, π/2]. The two principal curvatures of this surface are c1 = c2 = 1 δ , cos v r + δ cos v . (B2) (B3) FIG. 9: A sketch of the surface near the disk edge. The mean and Gaussian curvatures are M = G = 2δ cos v + r 2δ(r + δ cos v) , cos v δ(r + δ cos v) . (B4) (B5) One more quantity we need is the area element dA, dA = √gdudv = δ(r + δ cos v)dudv, (B6) where g is the surface metric. The bending free energy of the disk edge is then F D = Z (cid:2)2¯κM (M 2 − 2M ¯c0) + ¯κGG(cid:3)dA = 2πZ π/2 −π/2(cid:2)2¯κM (M 2 − 2M ¯c0) + ¯κGG(cid:3)√gdv. The contribution from the Gaussian curvature can be easily evaluated FG 2π = 2¯κGZ π/2 −π/2 cos vdv = 2¯κG. (B7) The calculation for the mean curvature is slightly more complicated FM 2π = −4¯κM + 8¯κM ¯c0δ − 2π¯κM ¯c0r + π¯κM r d 1 q1 − ( d r )2(cid:16)1 − 2 π tan−1s 1 − d r(cid:17). r 1 + d The last term in the above equation can be expanded up to order ( d r )2 using 2 d r )2i− 1 h1 − ( tan−1s 1 − d r 1 + d = 1 + 1 2 ( d r )2 + O(( δ R )2) r ≈ tan−1(cid:16)1 − d 2r − π 4 − = 1 4 d r(cid:17) ( d r )2 + O(( δ R )2), (B8) (B9) 13 For the pore structure, the edge surface can be re- garded as the inner part of the torus. A slight modifi- cation to the above derivation is the integration range of parameter v, which changes to [π/2, 3π/2]. The final result is (B10) and the final result is 1 2δ − 2¯c0(cid:3)r + ¯κM(cid:2)3 − 8¯c0δ(cid:3) = π¯κM(cid:2) π + ¯κM(cid:2) 2(cid:3) 4 − + O(( δ R )2). δ r 1 Note that the disk radius is defined as R = r + δ. From Eqs. (B7) and (B10), we get the total energy of the disk edge FM 2π F D 2π 1 2δ − 2¯c0(cid:3)R π 2 = π¯κM(cid:2) + ¯κM(cid:2)(3 − + 2¯κG + ¯κM(cid:2) 2(cid:3) ) + (2π − 8)¯c0δ(cid:3) π 4 − + O(( δ R 1 (B11) δ R )2). We can identify the line tension as F P 2π 1 π = π¯κM(cid:2) + ¯κM(cid:2)( − 2¯κG + ¯κM(cid:2) 2δ − 2¯c0(cid:3)R 2 − 3) + (8 − 2π)¯c0δ(cid:3) 1 + O(( π 4 + δ R 2(cid:3) (B13) δ R )2). σ = π¯κM(cid:2) 1 2δ − 2¯c0(cid:3). (B12) The line tension for the pore is the same as that obtained for the disk geometry. [1] R. Lipowsky, Encyclopedia of Applied Physics 23, 199 [20] M. Hu, J. J. Briguglio, and M. Deserno, Biophys. J. 102, (1998). 1403 (2012). [2] M. B. Schneider, J. T. Jenkins, and W. W. Webb, Bio- [21] H.-T. Jung, S. Y. Lee, E. W. Kaler, B. Coldren, and J. A. phys. J. 45, 891 (1984). Zasadzinski, PNAS 99, 15318 (2002). [3] J. F. Faucon, M. D. Mitov, P. M´el´eard, I. Bivas, and [22] Z.-G. Wang and S. A. Safran, J. Phys. (Paris) 51, 185 P. Bothorel, J. Phys. France 50, 2389 (1989). (1990). [4] E. Evans and W. Rawicz, Phys. Rev. Lett. 64, 2094 [23] Z.-G. Wang and S. A. Safran, J. Chem. Phys. 94, 679 (1990). (1991). [5] W. Pfeiffer, S. Konig, J. F. Legrand, T. Bayerl, D. Richter, and E. Sackmann, Europhys. Lett. 23, 457 (1993). [24] Z.-G. Wang, Macromolecules 25, 3702 (1992). [25] T. Ohta and M. Nonomura, Eur. Phys. J. B 2, 57 (1998). [26] M. Laradji and R. C. Desai, J. Chem. Phys. 108, 4662 [6] J. R. Henriksen and J. H. Ipsen, Eur. Phys. J. E 14, 149 (1998). (2004). [7] G. Illya, R. Lipowsky, and J. C. Shillcock, J. Chem. Phys. [27] M. W. Matsen, J. Chem. Phys. 110, 4658 (1999). [28] M. Muller and G. Gompper, Phys. Rev. E 66, 041805 122, 244901 (2005). (2002). [8] M. C. Rheinstadter, W. Haussler, and T. Salditt, Phys. [29] K. Chang and D. C. Morse, Macromolecules 39, 7397 Rev. Lett. 97, 048103 (2006). (2006). [9] L. Bo and R. E. Waugh, Biophys. J. 55, 509 (1989). [30] M. L. Nunalee, H. Guo, M. O. de la Cruz, and K. R. [10] D. Cuvelier, I. Der´enyi, P. Bassereau, and P. Nassoy, Bio- Shull, Macromolecules 40, 4721 (2007). phys. J. 88, 2714 (2005). [31] K. Katsov, M. Muller, and M. Schick, Biophys. J. 87, [11] R. Goetz, G. Gompper, and R. Lipowsky, Phys. Rev. 3277 (2004). Lett. 82, 221 (1999). [32] T. Baumgart, S. T. Hess, and W. W. Webb, Nature 425, [12] G. Ayton and G. A. Voth, Biophys. J. 83, 3357 (2002). [13] G. Brannigan and F. L. H. Brown, J. Chem. Phys. 122, 074905 (2005). 821 (2003). [33] P. Fromherz, C. Rocker, and D. Ruppel, Faraday Discuss. Chem. Soc. 81, 39 (1986). [14] V. A. Harmandaris and M. Deserno, J. Chem. Phys. 125, [34] D. V. Zhelev and D. Needham, Biochim. Biophys. Acta. 204905 (2006). 1147, 89 (1993). [15] R. S. Millman and G. D. Parker, Elements of Differential Geometry (Prentice-Hall, New Jersey, 1977). [16] M. P. do Carmo, Differential Geometry of Curves and Surfaces (Prentice-Hall, New Jersey, 1976). [35] E. Karatekin, O. Sandre, H. Guitouni, N. Borghi, P.- H. Puech, and F. Brochard-Wyart, Biophys. J. 84, 1734 (2003). [36] M. Muller and M. Schick, J. Chem. Phys. 105, 8282 [17] A. Pressley, Elementary Differential Geometry (Springer, (1996). London, 2001). [37] Z.-J. Wang and D. Frenkel, J. Chem. Phys. 122, 234711 [18] D. P. Siegel and M. M. Kozlov, Biophys. J. 87, 366 (2005). (2004). [38] T. V. Tolpekina, W. K. den Otter, and W. J. Briels, J. [19] S. Semrau, T. Idema, L. Holtzer, T. Schmidt, and Chem. Phys. 121, 8014 (2004). C. Storm, Phys. Rev. Lett. 100, 088101 (2008). [39] J. Wohlert, W. K. den Otter, O. Edholm, and W. J. 14 Briels, J. Chem. Phys. 124, 154905 (2006). sity Press, New York, 2007), 3rd ed. [40] W. K. den Otter, J. Chem. Phys. 131, 205101 (2009). [41] C. Loison, M. Mareschal, and F. Schmid, J. Chem. Phys. 121, 1890 (2004). [58] A. G. Petrov and I. Bivas, Prog. Surf. Sci. 16, 389 (1984). [59] D. Marsh, Chem. Phys. Lipids 144, 146 (2006). [60] E. Kurtisovski, N. Taulier, R. Ober, M. Waks, and [42] F. Y. Jiang, Y. Bouret, and J. T. Kindt, Biophys. J. 87, W. Urbach, Phys. Rev. Lett. 98, 258103 (2007). 182 (2004). [61] I. R. Cooke, K. Kremer, and M. Deserno, Phys. Rev. E [43] J. de Joannis, F. Y. Jiang, and J. T. Kindt, Langmuir 72, 011506 (2005). 22, 998 (2006). [62] I. R. Cooke and M. Deserno, J. Chem. Phys. 123, 224710 [44] W. Shinoda, T. Nakamura, and S. O. Nielsen, Soft Mat- (2005). ter 7, 9012 (2011). [63] H. J. Risselada, S. J. Marrink, and M. Muller, Phys. Rev. [45] J. D. Moroz and P. Nelson, Biophys. J. 72, 2211 (1997). [46] S. May, Eur. Phys. J. E 3, 37 (2000). [47] W. Helfrich, Z. Naturforsch. C 28, 693 (1973). [48] Z.-C. Ou-Yang, J.-X. Liu, and Y.-Z. Xie, Geometric Methods in the Elastic Theory of Membranes in Liquid Crystal Phases (World Scientific, 1999). [49] F. Schmid, J. Phys.: Condens. Matter 10, 8105 (1998). [50] G. H. Fredrickson, V. Ganesan, and F. Drolet, Macro- molecules 35, 16 (2002). Lett. 106, 148102 (2011). [64] J. Li, H. Zhang, F. Qiu, and A.-C. Shi, Phys. Rev. E 88, 012719 (2013). [65] K. Katsov, M. Muller, and M. Schick, Biophys. J. 90, 915 (2006). [66] O. V. Manyuhina, I. O. Shklyarevskiy, P. Jonkheijm, P. C. M. Christianen, A. Fasolino, M. I. Katsnelson, A. P. H. J. Schenning, E. W. Meijer, O. Henze, A. F. M. Kil- binger, et al., Phys. Rev. Lett. 98, 146101 (2007). [51] M. W. Matsen, J. Phys.: Condens. Matter 14, R21 [67] S. Sanyal and A. K. Menon, ACS Chemical Biology 4, (2002). 895 (2009). [52] A.-C. Shi, in Developments in Block Copolymer Science and Technology, edited by I. Hamley (John Wiley & Sons, New York, 2004), chap. 8. [53] M. W. Matsen, in Soft Matter, Volume 1: Polymer Melts and Mixtures, edited by G. Gompper and M. Schick (Wiley-VCH, Weinheim, 2006). [68] E. Gutlederer, T. Gruhn, and R. Lipowsky, Soft Matter 5, 3303 (2009). [69] J. Hu, T. Weikl, and R. Lipowsky, Soft Matter 7, 6092 (2011). [70] J. Li, H. Zhang, and F. Qiu, J. Phys. Chem. B 117, 843 (2013). [54] G. H. Fredrickson, The Equilibrium Theory of Inhomo- [71] S. Meinhardt, R. L. C. Vink, and F. Schmid, PNAS 110, geneous Polymers (Clarendon Press, Oxford, 2006). 4476 (2013). [55] M. W. Matsen, Macromolecules 28, 5765 (1995). [56] J. Zhou and A.-C. Shi, Macromol. Theory Simul. 20, 690 (2011). [57] W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery, Numerical Recipes (Cambridge Univer- [72] S. A. Safran, Statistical Thermodynamics of Surfaces, In- terfaces, and Membranes (Addison-Wesley, New York, 1994).
1603.01571
1
1603
2016-03-04T19:09:06
Growth and Division of Active Droplets: A Model for Protocells
[ "physics.bio-ph", "cond-mat.soft" ]
It has been proposed that during the early steps in the origin of life, small droplets could have formed via the segregation of molecules from complex mixtures by phase separation. These droplets could have provided chemical reaction centers. However, whether these droplets could divide and propagate is unclear. Here we examine the behavior of droplets in systems that are maintained away from thermodynamic equilibrium by an external supply of energy. In these systems, droplets grow by the addition of droplet material generated by chemical reactions. Surprisingly, we find that chemically driven droplet growth can lead to shape instabilities that trigger the division of droplets into two smaller daughters. Therefore, chemically active droplets can exhibit cycles of growth and division that resemble the proliferation of living cells. Dividing active droplets could serve as a model for prebiotic protocells, where chemical reactions in the droplet play the role of a prebiotic metabolism.
physics.bio-ph
physics
Growth and Division of Active Droplets: A Model for Protocells David Zwicker,1, 2, ∗ Rabea Seyboldt,1, ∗ Christoph A. Weber,1 Anthony A. Hyman,3 and Frank Julicher1, † 1Max Planck Institute for the Physics of Complex Systems, 01187 Dresden, Germany 2School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA 3Max Planck Institute of Molecular Cell Biology and Genetics, 01307 Dresden, Germany (Dated: March 7, 2016) Abstract It has been proposed that during the early steps in the origin of life, small droplets could have formed via the segregation of molecules from complex mixtures by phase separation. These droplets could have provided chemical reaction centers. However, whether these droplets could divide and propagate is un- clear. Here we examine the behavior of droplets in systems that are maintained away from thermodynamic equilibrium by an external supply of energy. In these systems, droplets grow by the addition of droplet material generated by chemical reactions. Surprisingly, we find that chemically driven droplet growth can lead to shape instabilities that trigger the division of droplets into two smaller daughters. Therefore, chemi- cally active droplets can exhibit cycles of growth and division that resemble the proliferation of living cells. Dividing active droplets could serve as a model for prebiotic protocells, where chemical reactions in the droplet play the role of a prebiotic metabolism. 6 1 0 2 r a M 4 ] h p - o i b . s c i s y h p [ 1 v 1 7 5 1 0 . 3 0 6 1 : v i X r a ∗ These two authors contributed equally † To whom correspondence should be addressed; Email: [email protected] 1 Introduction Living systems consist of cells that can grow and divide. Cells take up matter from the outside world to grow, they release waste products, and they are able to divide, creating more cells. A fundamental question is to understand how cells arose early in evolution. Early in the origin of life, chemical reaction centers or chemical micro reactors had to form in order to organize chemical reactions in space. These micro reactors had to exchange material with the outside and they had to propagate. Recently, the idea of Oparin and Haldane1,2 that small droplets, which they called coacervates, could organize molecules in micro reactors has resurfaced to prominence3–8. Such droplets are liquid-like aggregates that concentrate molecules that have separated from a complex mixture. Liquid droplets are self-organized structures that coexist with a surrounding fluid7,9. The inter- face separating the two coexisting phases provides them with a well defined surface. The associ- ated surface tension forces them into a spherical shape. Furthermore, many substances can diffuse across the interface. The segregation of components into a droplet concentrates material in a con- fined volume, which may facilitate specific chemical reactions. Thus droplets provide containers in which chemical reactions can be spatially organized. Although the thermodynamics of phase transitions can explain how liquid drops can form, it is unclear how such droplets could propagate by division and subsequent growth, an ability that would be key at the origin of life. Droplets grow by taking up material from a supersaturated environment or by Ostwald ripen- ing9–13. Ostwald ripening describes the exchange of material between droplets by diffusion, usu- ally leading to growth of large droplets while small droplets shrink. Furthermore, droplets can increase in size by fusion of two droplets into a larger one. These processes lead to the formation of droplets of increasing size while the droplet number decreases with time. This behavior is op- posite to that of cells which have a characteristic size and increase their number by division. How could droplets divide and propagate? We have recently shown that droplets that are maintained away from thermodynamic equilib- rium by a chemical fuel can have unusual properties14,15. In particular, in the presence of chemical reactions, Ostwald ripening can be suppressed15 and multiple droplets can stably coexist, with a characteristic size set by the reaction rates15–18. Here, we show that surprisingly, spherical droplets subject to chemical reactions spontaneously split in two smaller daughter droplets of equal size. Therefore, chemically active droplets can grow and subsequently divide and thereby propagate by using up the inflowing material as a fuel. We conclude that droplets can indeed behave similarly to cells in the presence of chemical reactions that are driven by an external fuel reservoir. Such 2 active droplets could represent models for growing and dividing protocells with a rudimentary metabolism which is represented by simple chemical reactions that are maintained by an external fuel. Division of active droplets Droplets can serve as small compartments to spatially organize chemical reactions. The emergence of droplets requires phase separation into two coexisting liq- uid phases of different composition. Phase separation is driven by molecular interactions, where molecules with an affinity for each other lower their energy if they come closely together. A fluid can demix if the energy decrease associated with molecular interactions overcomes the effects of entropy increase by mixing19,20. If those interactions are strong, a sharp interface separates the coexisting phases. Droplets can become chemically active if the material of the droplet is produced and destroyed by chemical reactions. An example that resembles a simple protocell is shown schematically in Fig. 1A. The droplet is formed by a droplet material D that is generated inside the droplet from a high energy precursor N, which plays the role of a nutrient. Droplet material can degrade into a lower energy component W that plays the role of a waste, which leaves the droplet by diffusion. The droplet can survive if N is continuously supplied and W is continuously removed. This can be achieved by recycling N using an external energy source such as a fuel or radiation. Inspired by Oparin21, we discuss the physics of such active droplets using a minimal model with only two components A and B, see Fig. 1B. The droplet material B phase separates from the solvent. It can spontaneously be degraded by a chemical reaction B → A (1) into molecules of type A that are soluble in the background fluid and leave the droplet. The backward reaction A → B is not proceeding spontaneously because B is of higher energy than A. New droplet material B can be produced by the second reaction A + C → B + C(cid:48) , (2) that is coupled to a fuel C. Here C(cid:48) is the low energy reaction product of the fuel molecules. The chemical potential difference ∆µC = µC − µC(cid:48) > 0 provided by the fuel powers the production of high energy B from low energy A. The difference ∆µC can be maintained constant if the concentrations of C and C(cid:48) are set by an external reservoir. In this case, the system is kept away from a thermodynamic equilibrium, see Box 1. 3 FIG. 1. A) Schematic representation of an active droplet as a simple model of a protocell. The droplet (orange) consists of a droplet material D. Nutrients N of high chemical energy can diffuse into the droplet. Inside the droplet, N is transformed to D by chemical reactions. Droplet material D is degraded chemically into low energy waste W that leaves the droplet. B) Minimal model, with droplet material B and soluble component A. C) Sequence of shapes of a dividing droplet at different times as indicated. The dynamic equations of a continuum model corresponding to the situation shown in B) were solved numerically. The droplet shapes are shown as equal concentration contours (black). Parameter values are ν−t0/∆c = 7·10−3, ν+t0/∆c = 1.9 · 10−3, and k±t0 = 10−2, where t0 is a characteristic time of the continuum model (see supplementary information). Indicated times are given in units of 102 t0. The combination of phase separation and non-equilibrium chemical reactions can be studied in a continuum model15–17, see supplemental information. Using this model, we find that spherical droplets that are chemically active can undergo a shape instability and split in two smaller droplets, despite their surface tension, see Fig. 1C. A droplet first grows until it reaches its stationary size15. Then, the droplet starts to elongate and forms a dumbbell shape. This dumbbell splits in two smaller droplets of equal size. The resulting smaller droplets grow again until a new division may 4 BAdropletBABAABBC(cid:38)(cid:183)BAchemical fuelor radiationNWWwastenutrientNdropletDWNDA06121819.22024C occur, reminiscent of living cells. In order to investigate the stability of spherical droplets, we study the droplet shape by an effective droplet model described in Box 215. Fig. 2A shows the behavior of the stationary droplet radius in this model as a function of the supersaturation . This supersaturation is the excess concentration of droplet material far from the droplet, generated by the chemical reaction (2). For  > 0, material diffuses to the droplet and is incorporated. Fig. 2A shows that for a given turnover ν− of droplet material inside the droplet (see Box 2), stationary droplets only exist for sufficiently large supersaturation. Beyond this threshold, droplets smaller than the critical radius (Fig. 2A, black dotted lines) shrink, while larger droplets grow toward the stationary radius (Fig. 2A, black solid line)15. At this stationary radius, the influx of B due to supersaturation outside is balanced by the efflux of material A produced inside the droplet. Thus a larger turnover leads to smaller droplets (Fig. 2A). Droplet division occurs when a spherical droplet becomes unstable and elongates. We per- formed a linear stability analysis of spherical droplets at their stationary radius in the effective droplet model, see supplemental material. We find that for increasing supersaturation , a spher- ical droplet with surface tension undergoes a shape instability when its radius reaches a critical value Rdiv that depends on the reaction rates and droplet parameters, see Fig. 2A. Beyond the ra- dius Rdiv, the spherical shape is unstable and any small shape deformation triggers the elongation of the droplet shape along one axis. The stability analysis of the effective droplet model can be represented in a state diagram, see Fig. 2B. We find three different regions as a function of supersaturation  and turnover of droplet material ν−. A region where droplets do not exist (white), a region in which spherical droplets are stable (blue), and a region in which spherical droplets are unstable (red). In order to study how the shape instability leads to droplet division, we investigated the droplet dynamics beyond the linearized analysis using the continuum model. This model can capture the topological changes of the droplet surface that occur during division. Numerical calculations of the continuum model (see supplemental information) confirm the results of the stability analysis. An example of droplet division is shown in Fig. 1C. The state diagram for the continuum model is shown in Fig. 2C. Comparing the state diagrams Fig. 2B and Fig. 2C reveals that both models exhibit qualitatively the same behaviors. Note that due to simplifications in the effective droplet model, the parameters are different in both models (see supplemental information) and the regions in both diagrams differ slightly. While Fig. 2B only shows where droplets become unstable (red 5 FIG. 2. A) Stationary radii of active droplets. The droplet radius R of spherical droplets is shown as a function of supersaturation  for different values of normalized turnover ν−/ν0 = 0, 1, 3 (from left to right). Radii of stable droplets are shown as solid black lines. Dotted lines indicate states where droplets are unstable with respect to size (black) or shape (red). The results are obtained for the effective droplet model described in Box 2. Parameter values are: k±τ0 = 10−2, c (0) + = 0, β− = β+, D− = D+ and ν0 = 10−2∆c/τ0. Here, w = 6β+γ/∆c, and τ0 = w2/D+ are characteristic length and time scales. B) Stability diagram of active droplets as a function of supersaturation  = ν+/(k+∆c) and turnover ν− of droplet material. Droplets either dissolve and disappear (white region), are spherical and stable (blue region), or undergo a shape instability and typically divide (red region). The lines of instability are obtained for the droplet model described in Box 2 for the same parameters as in A). C) Same stability diagram as in B) but for the continuum model described in the supplemental information. The behavior of droplets is indicated by symbols for different values of ν− and . Parameter values are k±t0 = 10−2 (see supplementary information). The parameter values corresponding to Fig. 1C are indicated (large red circle). 6 00.050.10.150.2Supersaturation0246810StationaryradiusAcritical radiusstable dropletshapeinstability00.10.221.510.50SupersaturationTurnoverB stabledropletsshape instability0.10.221.510.50SupersaturationCnodropletsstabledropletsdivision into 2 dropletsdivision into 3 dropletsdropletelongationTurnover FIG. 3. A) Sequence of droplet divisions at different times as indicated. Droplet configurations obtained from numerical solutions to the continuum model are represented as three dimensional shapes. Parameter ν+t0/∆c = 2 · 10−3. Remaining parameters are the same as in Fig. 1C. B) Schematic representation of the orientation of subsequent division axes. C) Droplet division is oriented along the axis for which diffusion fluxes (orange arrows) are maximal. line), Fig. 2C reveals the behaviors of droplets in the unstable region. We find that droplets typ- ically divide into two daughters (red circles). However, for some parameter values they divide into three droplets (red triangles). Our calculations show that droplets typically divide after they become unstable. However, in some cases division was not seen during the time of calculations (red rectangles). In these cases droplets elongated until they reaches the size of the simulation box. It is unclear whether they would divide in a larger box. 7 364248660748896ABxyzinfluxC Our numerical calculations also reveal that droplets typically undergo multiple divisions, see Fig. 3A and supplemental movie. After a first division, the smaller daughters grow until they di- vide again when they reach the radius Rdiv. Interestingly, the division axes are not independent of each other, see Fig. 3A. In the absence of system boundaries, the division axes of both daughters are perpendicular to the first division axis, see Fig. 3B. Similarly, when the four granddaughters divide, their division axes are perpendicular to both the division axes of the first and the second division. The division axes in subsequent droplet divisions are determined by droplet interac- tions via the concentration fields surrounding the droplets. The two growing daughter droplets effectively compete for droplet material, leading to the depletion of droplet material in the space between them. Therefore, diffusion fluxes and growth rates are larger along axes perpendicular to the previous division axis, see Fig. 3C. This bias due to droplet interactions determines the divi- sion axes. In our numerical calculations, boundary conditions also influence the droplet divisions and slightly modify the division axes, see Fig. 3A. Discussion The question how life first arose on earth has fascinated both scientists and non- scientists since it was understood that modern life emerged by evolution from early precursors. While evolution can be reconstructed to a large extend both from fossil records and from the phy- logenetic analysis of todays genomes, the structure and nature of early life forms remain quite unclear22. How did the first replicating cells emerge from prebiotic precursors? Since replication involves specific chemical reactions, early replicators had to spatially organize chemistry and to concentrate certain molecules to facilitate reactions that would be unlikely in dilute or disorganized situations. Therefore, protocells as containers for chemical reactions had to appear. Alexander Oparin pioneered the idea that macromolecular aggregation could lead to the forma- tion of 'coacervates', liquid droplets that could organize chemistry and provide microreactors in which selected molecules were concentrated for prebiotic chemistry1,23. What types of molecules could have formed such droplets? It is interesting to note that modern day cells possess a number of chemical compartments that are not separated by a membrane from the cell cytoplasm but that form by phase separation from the cytoplasm3,7,24,25. Many of these compartments are liquid and consist of RNA molecules and RNA binding proteins26–29. The RNA world hypothesis suggests that at the origin of life, RNA was both the carrier of genetic information and could have acted as early enzymes30,31. Folded RNA molecules called ribozymes can be catalysts for many reactions including RNA processing32. Combining RNA with other molecules such as simple peptides may have been sufficient to organize RNA in liquid droplets4. 8 The steps from chemically active droplets to the first dividing cells with membranes pose a big challenge to the understanding of early evolution. While it has been suggested that ribozymes that replicate RNA could have formed by molecular evolution31,33, it is unclear how a cell membrane and cell division could have emerged34–37. The possibility that droplets may spontaneously divide has been discussed in the context of either negative surface tension38,39 or in active nematic droplets40. Here we show that simply adding a proto-metabolism to droplets formed by classical phase separation can naturally lead to droplet division despite their surface tension. Membranes or surfactants are therefore not required to achieve division of prebiotic cells. Active droplets are natural systems to organize the chemistry of replicators and to form protocells. Such droplets can in principle form spontaneously by a rare nucleation event. Once they exist, they grow and divide. They provide a container for chemical reactions and they concentrate selected molecules that have an affinity to the droplet phase. The liquid and dynamic nature of active droplets implies that components in the droplet can mix and chemical reactions are facilitated. Protocells formed by active droplets require a constant energy supply, which could have been provided by a chemical fuel, by tides, or by temperature gradients, e.g. in hydrothermal vents on the sea floor2,41–43. The chemical reactions by which new droplet material is formed and subsequently degraded represents an early metabolism. The fact that active droplets tend to become unstable and divide is a very unusual behavior of droplets. Usually, droplets maintain their spherical shape because surface tension tends to reduce the surface area. An instability of the droplet shape requires non-equilibrium conditions. In our models, the chemically driven diffusion fluxes associated with stationary droplets trigger the shape instability. In the absence of chemical reactions and stationary fluxes, the shape instability does not occur. This is reminiscent of other well known shape instabilities of moving interfaces. The droplet instability discussed here that triggers droplet division is related to the Mullins-Sekerka instability often discussed in the context of crystal growth44. In the case of the Mullins-Sekerka instability, an interface advances because of a diffusive influx. Beyond a critical interface velocity, a flat interface becomes unstable with respect to growing spikes called dendrites. The Mullins- Sekerka instability occurs for a moving interface in the absence of chemical reactions, while the active droplet instability discussed here requires reactions but no interface motion. In the limit of large characteristic length scales introduced by the chemical reactions, the conditions for both instabilities become the same (supplemental information). We propose that active droplets that are maintained away from thermodynamic equilibrium by 9 a constant influx of nutrients and a constant efflux of waste are a simple model of membraneless protocells. One can speculate that such droplets could have concentrated RNA molecules together with other molecular species to form early replicators with an early metabolism. It is interesting to envision early ecosystems in which droplets of different type may have had symbiotic relationships if one produces the nutrient of the other. Alternatively one can find predator-prey relationships when a droplet fuses with a different one to harvests its resources. Finally, the possibility that early protocells were active droplets suggest possible scenarios by which cell membranes and cells with a more modern architecture could have emerged. The droplet surface is an interface that will in general attract amphiphilic molecules. Such molecules have neither an affinity for the droplet phase nor for the surrounding fluid. As a result, selected molecules populate the droplet surface and surface chemical reactions could be established. Such surface modifications could improve the resistance of droplets to varying environmental conditions and provide specific surface properties. If lipids were available in the outside fluid, lipid bilayers could be attracted to the specific droplet surface chemistry. Our work shows that active droplets can naturally divide. Therefore protocells could have obtained their membranes long after the first dividing cells had appeared on earth. 1 Oparin, A. I. Origin of Life (Dover Publications, Inc., New York, 1952). 2 Haldane, J. B. S. The origin of life. Rationalist Annual 148, 3–10 (1929). 3 Brangwynne, C. P. et al. Germline P Granules Are Liquid Droplets That Localize by Controlled Disso- lution/Condensation. Science 324, 1729–1732 (2009). 4 Koga, S., Williams, D. S., Perriman, A. W. & Mann, S. Peptide-nucleotide microdroplets as a step towards a membrane-free protocell model. Nat. Chem. 3, 720–724 (2011). 5 Crosby, J. et al. Stabilization and enhanced reactivity of actinorhodin polyketide synthase minimal complex in polymer–nucleotide coacervate droplets. Chem. Commun. 48, 11832 (2012). 6 Sokolova, E. et al. Enhanced transcription rates in membrane-free protocells formed by coacervation of cell lysate. Proc. Natl. Acad. Sci. USA 110, 11692 (2013). 7 Hyman, A. A., Weber, C. A. & Julicher, F. Liquid-liquid phase separation in biology. Annu. Rev. Cell Dev. Biol. 30, 39–58 (2014). 8 Dora Tang, T.-Y., van Swaay, D., DeMello, A., Ross Anderson, J. L. & Mann, S. In vitro gene expression 10 within membrane-free coacervate protocells. Chem. Commun. 51, 11429–11432 (2015). 9 Bray, A. Theory of phase-ordering kinetics. Adv. Phys. 43, 357–459 (1994). 10 Ostwald, W. Studien uber die Bildung und Umwandlung fester Korper. Z. Phys. Chem 22, 289–330 (1897). 11 Lifshitz, I. M. & Slyozov, V. V. The kinetics of precipitation from supersaturated solid solutions. J. Phys. Chem. Solids 19, 35–50 (1961). 12 Binder, K. & Stauffer, D. Statistical theory of nucleation, condensation and coagulation. Adv. Phys. 25, 343–396 (1976). 13 Voorhees, P. W. Ostwald ripening of two-phase mixtures. Annu. Rev. Mater. Sci. 22, 197–215 (1992). 14 Zwicker, D., Decker, M., Jaensch, S., Hyman, A. A. & Julicher, F. Centrosomes are autocatalytic droplets of pericentriolar material organized by centrioles. Proc. Natl. Acad. Sci. USA 111, E2636–45 (2014). 15 Zwicker, D., Hyman, A. A. & Julicher, F. Suppression of Ostwald ripening in Active Emulsions. Phys. Rev. E 92, 012317 (2015). 16 Puri, S. & Frisch, H. Segregation dynamics of binary mixtures with simple chemical reactions. J. Phys. A 27, 6027–6038 (1994). 17 Glotzer, S. C., Stauffer, D. & Jan, N. Monte Carlo simulations of phase separation in chemically reactive binary mixtures. Phys. Rev. Lett. 72, 4109–4112 (1994). 18 Carati, D. & Lefever, R. Chemical freezing of phase separation in immiscible binary mixtures. Phys. Rev. E 56, 3127–3136 (1997). 19 Huggins, M. L. Solutions of long chain compounds. J. Chem. Phys. 9, 440–440 (1941). 20 Flory, P. I. Thermodynamics of high polymer solutions. J. Chem. Phys. 10, 51–61 (1942). 21 Oparin, A. Proiskhozhedenie Zhizni Mosckovskii Rabochii, Moscow. Reprinted and translated in JD Bernal (1967) The Origin of Life London: Weidenfeld and Nicolson (1924). 22 Woese, C. R., Kandler, O. & Wheelis, M. L. Towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya. Proc. Natl. Acad. Sci. USA 87, 4576 (1990). 23 Fox, S. W. The evolutionary significance of phase-separated microsystems. Orig. Life 7, 49–68 (1976). 24 Brangwynne, C. P. Soft active aggregates: mechanics, dynamics and self-assembly of liquid-like intra- cellular protein bodies. Soft Matter 7, 3052–3059 (2011). 25 Toretsky, J. A. & Wright, P. E. Assemblages: Functional units formed by cellular phase separation. J. Cell Biol. 206, 579–88 (2014). 11 26 Weber, S. C. & Brangwynne, C. P. Getting RNA and protein in phase. Cell 149, 1188–1191 (2012). 27 Elbaum-Garfinkle, S. et al. The disordered P granule protein LAF-1 drives phase separation into droplets with tunable viscosity and dynamics. Proc. Natl. Acad. Sci. USA 112, 7189 (2015). 28 Molliex, A. et al. Phase Separation by Low Complexity Domains Promotes Stress Granule Assembly and Drives Pathological Fibrillization Article Phase Separation by Low Complexity Domains Promotes Stress Granule Assembly and Drives Pathological Fibrillization. Cell 163, 123–133 (2015). 29 Lin, Y. et al. Formation and Maturation of Phase-Separated Liquid Droplets by RNA-Binding Proteins Article Formation and Maturation of Phase-Separated Liquid Droplets by RNA-Binding Proteins. Mol. Cell 1–12 (2015). 30 Gilbert, W. Origin of life: The RNA world. Nature 319 (1986). 31 Higgs, P. G. & Lehman, N. The RNA World: molecular cooperation at the origins of life. Nat. Rev. Genet. 16, 7–17 (2015). 32 Fedor, M. J. & Williamson, J. R. The catalytic diversity of RNAs. Nat. Rev. Mol. Cell. Biol. 6, 399–412 (2005). 33 Unrau, P. J. & Bartel, D. P. RNA-catalysed nucleotide synthesis. Nature 395, 260–3 (1998). 34 Hanczyc, M. M., Fujikawa, S. M. & Szostak, J. W. lular Compartments: Encapsulation, Growth, and Division. Experimental Models of Primitive Cel- Science 302, 618–622 (2003). arXiv:http://www.sciencemag.org/content/302/5645/618.full.pdf. 35 Hanczyc, M. M. & Szostak, J. W. Replicating vesicles as models of primitive cell growth and division. Curr. Opin. Chem. Biol. 8, 660–664 (2004). 36 Mac´ıa, J. & Sol´e, R. V. Synthetic Turing protocells: vesicle self-reproduction through symmetry- breaking instabilities. Philos. Trans. R. Soc. Lond. B 362, 1821–9 (2007). 37 Murtas, G. Early self-reproduction, the emergence of division mechanisms in protocells. Mol. Biosyst. 9, 195–204 (2013). 38 Browne, K. P., Walker, D. A., Bishop, K. J. M. & Grzybowski, B. A. Self-division of macroscopic droplets: Partitioning of nanosized cargo into nanoscale micelles. Angew. Chem. Int. Ed. Engl. 49, 6756–6759 (2010). 39 Patashinski, A. Z., Orlik, R., Paclawski, K., Ratner, M. A. & Grzybowski, B. A. The unstable and expanding interface between reacting liquids: Theoretical interpretation of negative surface tension. Soft Matter 8, 1601–1608 (2012). 40 Giomi, L. & DeSimone, A. Spontaneous division and motility in active nematic droplets. Phys. Rev. 12 Lett. 112, 147802 (2014). 41 Baross, J. & Hoffman, S. Submarine hydrothermal vents and associated gradient environments as sites for the origin and evolution of life. Origins Life Evol. B. 15, 327–345 (1985). 42 Martin, W. F. Hydrogen, metals, bifurcating electrons, and proton gradients: the early evolution of biological energy conservation. FEBS Lett. 586, 485–93 (2012). 43 Martin, W. F., Sousa, F. L. & Lane, N. Evolution. Energy at life's origin. Science 344, 1092–3 (2014). 44 Mullins, W. W. & Sekerka, R. F. Morphological stability of a particle growing by diffusion or heat flow. J. Appl. Phys. 34, 323–329 (1963). 45 Atkins, P. & de Paula, J. Atkins' Physical Chemistry (OUP Oxford, 2010). 46 Desai, R. C. & Kapral, R. Dynamics of Self-organized and Self-assembled Structures (Cambridge Uni- versity Press, 2009). 47 Cahn, J. W. & Hilliard, J. E. Free Energy of a Nonuniform System. I. Interfacial Free Energy. J. Chem. Phys. 28, 258–267 (1958). 48 Christensen, J. J., Elder, K. & Fogedby, H. C. Phase segregation dynamics of a chemically reactive binary mixture. Phys. Rev. E 54, R2212–R2215 (1996). 49 Dennis, G. R., Hope, J. J. & Johnsson, M. T. XMDS2: Fast, scalable simulation of coupled stochastic partial differential equations. Comput. Phys. Commun. 184, 201–208 (2013). 50 Zhong-can, O.-Y. & Helfrich, W. Instability and deformation of a spherical vesicle by pressure. Phys. Rev. Lett. 59, 2486–2488 (1987). 13 Box1:ReactionratesandenergysupplyBACC'reservoirABA+CB+C'(1)(2)Figure4:Schematicrepresentationofthereactioncy-cleinvolvingtwopathways(1)and(2).Thedifferencesofthechemicalpotentialsµde-terminethedirectionofthespontaneousreac-tions:CouplingtothechemicalfuelCwithreactionproductC0drivespathway(2)inthedirectionA→Boutsidethedroplet,whileB→AisoccurringintheabsenceofCin-sidethedroplet.ThechemicalreactionA(cid:10)Bconvertssolublepre-cursorsAtodropletmaterialBwithforwardreactionfluxs→andreversefluxs←.Thenetreactionfluxs=s→−s←characterizestheconcentrationperunittimethatisundergoingthereaction.Compatibilitywiththermodynamicsrequires(45)s→s←=exp(cid:18)−∆µkBT(cid:19),(B.1)where∆µisthechemicalfreeenergychangeassociatedwiththeforwardreaction.Thisconditionleadstode-tailedbalanceofforwardandbackwardreactionratesatchemicalequilibrium.Thenetreactionfluxscanthere-forebewrittenass=s←·(cid:20)exp(cid:18)−∆µkBT(cid:19)−1(cid:21).(B.2)Chemicalequilibriumisreachedwhen∆µ=0andthenetreactionfluxvanishes,s=0.Ifasin(1)there-actiondoesnotinvolveotherreactionpartnersorex-ternalenergyinput,thechemicalfreeenergychange∆µ=∆µ(1)isgivenbythedifferenceofthechemi-calpotentials,∆µ(1)=µB−µA.(B.3)SuchareactionleadstospontaneousdegradationofBandformationofAifµB>µAandthus∆µ(1)>0.Thechemicalpotentialsofamolecularspeciesncanbewrittenasµn=kBTln(vncn)+wn,wherevnisamolecularvolumeandcntheconcentrationofspeciesn.Thefirsttermisofentropicoriginwhilethecon-tributionwnismainlyenthalpicandincludesinter-nalmolecularfreeenergiesandinteractionenergiesbe-tweenmolecules.Notethatwngenerallydependsoncomposition.Thenetreactionratecorrespondingtoreactionpathway(1)canthusbewrittenass(1)=s(1)←·(cid:18)cAcBK(1)−1(cid:19)(B.4)whereK(1)=(vA/vB)exp((wA−wB)/kBT)istheequilibriumconstantofreactionpathway(1).Notethatinthecaseofphaseseparation,thisequilibriumconstantcanhavedifferentvaluesinsideandoutsidethedroplet.Ifonlyreactionpathway(1)occurs,dropletsarepas-sivedespitethepresenceofthereactionandthesystemreachesathermodynamicequilibrium.Nodropletdivi-sionsoccur.SuchasystemexhibitsOstwaldripeningandafterlongtimesreachesanequilibriumwhichcon-tainseitherasinglelargedropletornodroplet.Activedropletsrequireanexternalenergysupplythatmaintainsthedropletsawayfromthermodynamicequi-libriumatalltimes.ThereactionA(cid:10)BcanbecoupledtoanexternallysuppliedfuelCwithreactionproductC0withchemicalpotentialdifference∆µC=µC−µC0>0.Thissecondreactionpathway(2)obeysEq.(B.2)with∆µ=∆µ(2)and∆µ(2)=µB−µA−∆µC.(B.5)Thecorrespondingreactionfluxcanbewrittenass(2)=s(2)←(cid:18)cAcBK(2)−1(cid:19)(B.6)withequilibriumconstantK(2)=K(1)exp(∆µC/kBT).Ifbothpathwaysareactiveatthesametime,thenetre-actionfluxiss=s(1)+s(2).InthispaperweconsiderthecasewhereanactivedropletconvertsBtoAinsidethedropletmainlyviathereactionpathway(1)whileoutsidethedropletmaterialAisusedtogenerateBmainlyviathereactionpathway(2)usingtheexternalfuelasanenergysource,seeFig.4.Nochemicalequi-libriumcanbereachedinthiscasebecausetheequi-libriumconstantsK(1)andK(2)implyincompatibleequilibriumconditions.Thedropletisthusactive. Box2:DynamicsofactivedropletsConcentrationReaction fluxA01-10Radial distanceConcentration0B01105Stationary flux0-22468Figure5:A)Chemicalreactionfluxasafunctionofconcentration(solid).Thelinearizedfluxinsideandoutsidethedropletisindicatedasdashedlines.B)Stationaryconcentrationprofileofanactivedroplet(orange)withradius¯Rinbackgroundfluid(blue).Theequilibriumconcentrationsc(0)±andtheconcentrationfarfromthedropletc∞areindicated.Thegreenline(axisontheright)showsthestationaryfluxj=−D±∂rc.ParametervaluesarethesameasinFig.2A,withν−/ν0=1.2,ν+/ν0=0.1.WeconsiderafluidthatcontainsthedropletformingmaterialBatconcentrationc=cB.Thesystemsegre-gatesintotwocoexistingphasesthatareseparatedbyasharpinterface.Weconsiderthelimitofstrongsegrega-tionofphasesbyasharpinterface.Acrosstheinterface,chemicalpotentialsarecontinuous,µ+=µ−,whilethepressureexhibitsajumpP−−P+=2γH,(B.7)knownasLaplacepressure.Here,γdenotessurfaceten-sionandHthelocalmeancurvatureoftheinterface.Thesubscripts−and+refertovaluesattheinterfaceinsideandoutsidethedroplet,respectively.Thesether-modynamicconditionsdeterminetheconcentrationsc−andc+attheinterfacewherebothphasescoexist.Be-causetheLaplacepressuredependsonlocalcurvature,theequilibriumconcentrationsalsodependoncurvatureH.Weexpressthisdependencetolinearorderbyc±'c(0)±+γβ±H,(B.8)wherec(0)±denotetheequilibriumconcentrationsofco-existingphasesataflatinterfaceandwehaveintroducedthecoefficientsβ±todescribetheeffectsofinterfacecurvature.ThedropletmaterialBisproducedbychemicalre-actionswithtotalreactionfluxs,whichisafunctionofconcentration,seeFig.5AandBox1.Thetime-evolutionoftheconcentrationfieldcisthendescribedbythereaction-diffusionequation∂tc=D±∇2c+s,(B.9)whereD+andD−denotethediffusioncoefficientsout-sideandinsidethedroplet,respectively.Theevolutionofthedropletshapeisgovernedbythenormalvelocityofthedropletinterfacevn=j−−j+c−−c+,(B.10)wherej±=−n·D±∇carethenormaldiffusionfluxesattheinterface,insideandoutsidethedroplet,andndenotesthesurfacenormal.Thereactionfluxistypicallypositive(Bisproduced)outsidethedroplet,whileitisnegative(Aisproduced)inside,seeFig.5A.Weexpandthefunctions(c)intermsoftheconcentrationvariationsinsideandoutsidethedroplettolinearorderass(c)'(ν+−k+(cid:0)c−c(0)+(cid:1)outsidethedroplet−ν−−k−(cid:0)c−c(0)−(cid:1)inside.(B.11)Thereactionratesk±characterizetherelaxationtimesinsideandoutsidethedroplet.ThefluxofproductionofBmoleculesattheequilibriumconcentrationsout-sideandinsidethedropletaredenotedν+andν−,re-spectively.Wecallν−turnoverbecauseitisthefluxatwhichBmoleculesdisappearinsidethedroplet.Theconcentrationfieldvariesoverthecharacteristiclengthscales'±=(D±/k±)1/2insideandoutsidethedroplet,respectively.Atlargedistancesr(cid:29)'+fromthedroplet,thenetre-actionfluxs(c)vanishesandtheconcentrationreachestheconstantvaluec∞=c(0)++ν+/k+.Thechemicalreactionsthusgenerateasupersaturation=c∞−c(0)+∆c,(B.12)where∆c=c(0)−−c(0)+.Thissupersaturationdrivesthediffusionfluxj+towardthedropletinterface.Insidethedroplet,dropletmaterialisdegraded,leadingtoaconcentrationprofilewithminimalconcentrationinthedropletcenter.Thiscausesadiffusionfluxj−towardsthecenter,seeFig.5B. Supplement Growth and Division of Active Droplets: A Model for Protocells I. CONTINUUM MODEL FOR ACTIVE DROPLETS A. Free energy function and chemical rates We consider an incompressible fluid containing two components: a component A that forms the background fluid and a droplet material B that forms droplets by phase separation. Chemical reactions convert the two components into each other. The concentration of the droplet material B is denoted by c(r, t) where r is the position and t denotes time. The concentration of the second component can be determined from c using the in- compressibility condition. Therefore, the free energy density f only depends on the concentration c. We use the following double-well free energy function f (c) = , b 2(∆c)2(cid:16)c − c(0)− (cid:17)2(cid:16)c − c(0) + (cid:17)2 + (cid:12)(cid:12). Here, the positive parameter b characterizes molecular (S.1) where we have defined ∆c =(cid:12)(cid:12)c(0)− − c(0) interactions and entropic contributions. This free energy describes the segregation of the fluid in two coexisting phases46: one phase rich in droplet material with c ≈ c(0)− and a diluted phase with c ≈ c(0) + . The state of the system is characterized by the free energy F [c] =(cid:90) (cid:104)f (c) + κ 2(cid:0)∇c(cid:1)2(cid:105)d3r , (S.2) where the integral is over the system volume. Here, the coefficient κ is related to surface tension and the interface width47. The chemical potential ¯µ = δF [c]/δc, which governs demixing, reads ¯µ = b (∆c)2(cid:0)c − c(0) + (cid:1)(cid:0)c − c(0)− (cid:1)(cid:0)2c − c(0)− − c(0) + (cid:1) − κ∇2c . The dynamics of the concentration field is described by the reaction-diffusion equation17,48 ∂tc = m∇2 ¯µ + s(c) . (S.3) (S.4) Here, m is a mobility coefficient of the droplet material. The source term s(c) describes chemical reactions. 16 We choose the function s(c) to be linear in the phases outside and inside the droplet. We connect these linear behaviors by a cubic interpolating polynomial: s(c) = ν+ + k+c(0) + + k+c ν− + k−c(0)− + k−c for c < c+ c for c > c− c , (S.5) p(c) for c+ c < c < c− c c and c− where c+ c are two characteristic concentrations and p(c) = a0 + a1c + a2c2 + a3c3 is a cubic polynomial. The coefficients ai are determined uniquely by the conditions that s(c) and its derivative are continuous functions: p(c+ p(c− p(cid:48)(c+ p(cid:48)(c− c ) = ν+ + k+c(0) + + k+c+ c c ) = ν− + k−c(0)− + k−c− c ) = k+ c c ) = k− . (S.6a) (S.6b) (S.6c) (S.6d) The reaction flux given in Eq. (S.5) describes a situation where an external energy source drives the system away from equilibrium, see Box 1. Eqs. (S.3)–(S.5) define the continuum model of active droplets. B. Relation with the effective droplet model The model described by Eq. (S.4) typically forms distinct phases, which are separated by an interface. Considering a flat interface between two phases with bulk concentrations c = c(0)− and c = c(0) + , the free energy F given in Eq. (S.2) is minimized by the concentration profile c∗(x) = c(0)− + c(0) + 2 + c(0)− − c(0) + 2 tanh x w , (S.7) where x is a coordinate that is normal to the interface and w = 2(κ/b)1/2 is the interface width47. The surface tension, i.e. the free energy per unit area of the interface, is46 γ =(cid:90) ∞ −∞ F [c∗(x)]dx = (∆c)2 6 √κb . (S.8) Two different bulk concentrations c− and c+ coexist across the interface for which the chemical potential is equal on both sides. For a curved interface the pressure difference between the inside 17 and outside of the droplet is the Laplace pressure 2γH, where H is the mean curvature of the interface. These two equilibrium conditions read 0 = ¯µ(c−) − ¯µ(c+) 0 = (c− − c+)¯µ(c−) + f (c+) − f (c−) − 2γH , (S.9a) (S.9b) where c− and c+ denote the concentration at the interface inside and outside the droplet, respec- tively. Using the free energy density as defined in Eq. (S.1), the concentrations that obey Eqs. (S.9) can be expressed to first order in H as c− ≈ c(0)− + βγH c+ ≈ c(0) + + βγH , (S.10a) (S.10b) which is valid for small surface tension, γ (cid:28) R∆c/β. Here, the coefficient β = 2/(b∆c) describes the effect of Laplace pressure on the concentration at the interface. Note that γβ defines a length scale, which is related to the interface width by γβ = w∆c/6. Linearizing Eq. (S.4) at the values c(0) + and c(0)− outside and inside the droplet gives the linear reaction-diffusion equation defined in Box 2, with diffusivity D = mb. We thus can relate the parameters b, κ, and m of the continuous theory to the parameters γ, β±, and D± of the effective droplet model. In particular, β+ = β− = β, and D+ = D− = D. C. Numerical methods We solved Eq. (S.4) with (S.5) and (S.3) numerically using the xmds2 software package (ver- sion 2.2.2)49 using an adaptive Runge-Kutta scheme of order 4/5, with tolerance 10−5. The Laplace operator was evaluated by a spectral method, while the chemical rates were evaluated directly. Nu- merical calculations were performed in a finite volume with no flux boundary conditions. We normalize concentration, length and time by ∆c = c(0)− − c(0) + , w and t0 = w2/D, re- spectively, where the characteristic length scale is w = 2(κ/b)1/2. The relevant dimensionless model parameters are c(0)± /∆c, k±t0, ν±t0/∆c and c± c /∆c. In all numerical calculations, we chose + /∆c = 0, c(0)− /∆c = 1 and k±t0 = 10−2. c(0) 18 1. Stability diagram Using three dimensional calculations in Cartesian coordinates, we observed that droplet con- figurations during the division of isolated single droplets were approximately axisymmetric. To determine the stability diagram shown in Fig. 2C we therefore performed calculations in cylindri- cal coordinates imposing axisymmetry. We used an axisymmetric cylindrical box with length 60w and radius 30w, discretized with 120 and 60 points, respectively. The initial conditions were given by a concentration profile that corresponded to a droplet geometry of a slightly prolate ellipsoid with unequal half axes with length R/w − 0.1 and R/w + 0.1, centered at the box center. The initial droplet size was chosen close to the stationary size in the continuum model. As an estimate for the stationary size we typically chose R/w = 0.9 ¯Rs/ w. Here, ¯Rs is the stationary radius calculated in the effective droplet model and w = 6β+γ/∆c, see Section I B. The concentration field at positions r was initialized by the function c(r) = c∞ + c(0)− 2 + c∞ − c(0)− 2 tanh d(r) w . (S.11) where d(r) is the oriented distance of r to the nearest point on the ellipsoid. The value of d(r) is negative for points inside the droplet and positive for points outside. The concentration far from the droplet is c∞ = ν+/k+ + c(0) + . We calculated the dynamics of the concentration field over a time interval T /t0 = 104, for different values of ν±t0/∆c. The parameters c± related to the chemical reaction in Eq. (S.5) c /∆c = 0.75. Because close to the shape instability the were chosen as c+ dynamics slows down, we may slightly overestimate the region of stability, since we cannot detect the exact instability with the finite time intervals simulated. Contours shown in Fig. 1C correspond to c/∆c = 0.5. c /∆c = 0.25 and c− c 2. Calculations for multiple divisions Several subsequent divisions break cylindrical symmetry. The calculations shown in Fig. 3A were therefore performed in three dimensions using cartesian coordinates. We chose a cubic box with side length L = 50w and an equidistant discretization of 100 points along each dimension. Initial conditions corresponded to a spherical droplet centered at r = (L/4, L/4, L/4). The concentration field was initialized with c = c(0)− inside the droplet and c = c∞ outside. The 19 parameters for the calculations were ν−t0/∆c = 7 · 10−3, ν+t0/∆c = 2 · 10−3 and c+ c− c /∆c = 0.5. Surfaces shown in Fig. 3A correspond to c/∆c = 0.5. c /∆c = II. EFFECTIVE MODEL FOR ACTIVE DROPLETS Using the effective droplet model defined in Box 2, we discuss steady state droplets and perform a linear stability analysis of the spherical droplet shape. We determine conditions for a shape instability towards an elongated shape. A. Droplet dynamics in spherical coordinates Using spherical coordinates r, θ, φ centered on the droplet, the interface defining the droplet surface is positioned at radial distance r = R(θ, φ). Away from the interface, the concentration field c(r, θ, φ) obeys the reaction-diffusion equation ∂tc = D±∇2c + s . (S.12) Here, t denotes time and D+, D− are the diffusion coefficients outside (r > R(θ, φ)) and inside (r < R(θ, φ)) the droplet, respectively. The reaction flux s is given by s = ν+ − k+(c+ − c(0) + ) −ν− − k−(c − c(0)− ) for r > R for r < R . (S.13) Here, reaction rates inside and outside the droplet are denoted by k±, c(0)± denote the equilibrium bulk concentrations of coexisting phases near a planar interface. The reaction fluxes at equilibrium concentrations are denoted by ν±. For the concentration c = c0, with c0 = −ν−/k− + c(0)− , the reaction flux inside the droplet vanishes, while for c = c∞ with c∞ = ν+/k+ + c(0) + the reaction flux outside the droplet vanishes. At the interface at r = R(θ, φ) we impose boundary conditions for the concentration: c(R±) = c(0)± + β±γH(θ, φ) . (S.14) This boundary condition describes a concentration jump at the interface. It corresponds to local thermodynamic equilibrium at a curved interface with surface tension γ. Here, R± denote the 20 limits of approaching the interface at radial distance R(θ, φ) from the outside or the inside, re- spectively. The mean curvature of the interface is denoted H and the coefficients β± describe the change of the equilibrium concentration at the interface due to Laplace pressure 2γH. The normal velocity vn of the interface is proportional to the difference of normal fluxes inside and outside9, vn = n · j− − j+ c(R−) − c(R+) with flux j± = −D±∇c(R±) and unit vector n normal to the interface. The droplet shape R(θ, φ) = R(θ, φ)er, where er denotes the unit vector in radial direction, can be parameterized using the angles θ and φ. The interface velocity can be written as , (S.15) ∂R(θ, φ, t) ∂t = vθe1 + vφe2 + vnn , (S.16) where e1 = ∂R/∂θ and e2 = ∂R/∂φ are the two basis vectors of the tangential plane. Using ∂R/∂t = (∂R/∂t)er, the velocity components vθ and vφ can be obtained from the conditions (∂R/∂t)eθ = 0 and (∂R/∂t)eφ = 0. Here, eθ and eφ are the local normalized basis vectors cor- responding to θ and φ in spherical coordinates. The radial interface velocity ∂R/∂t = (∂R/∂t)·er then reads = vn(cid:34)1 +(cid:18)∂θR R (cid:19)2 R sin θ(cid:19)2(cid:35) 1 +(cid:18) ∂φR 2 , (S.17) ∂R ∂t where vn is given by (S.15). B. Stationary states of spherical droplets Stationary solutions to Eq. (S.12) with spherically symmetric concentration field can be ex- pressed as ¯c(r) = A± + B± er/l± r + C± e−r/l± r , (S.18) where l± = (D±/k±)1/2 are characteristic length scales. Here, the coefficients A± are set by the chemical reactions, A± = ± ν± k± + c(0)± . (S.19) Regular behavior at r = 0 implies C− = −B−. For an infinite system, the concentration far from the droplet reaches a constant value. This implies B+ = 0. Using the boundary conditions (S.14) 21 FIG. S1. Rate of droplet growth dR/dt as a function of droplet radius R in a quasistatic limit in the presence of chemical reactions (red line) and without chemical reactions (blue line). The zeros of R correspond to stationary radii. An unstable critical radius (white circle) and a stable droplet radius (black circle) are indicated. Parameter values are: ν−τ0/∆c = −10−2 (red line) or ν−τ0/∆c = 0 (blue line), ν+τ0/∆c = 2 · 10−3 , k±τ0 = 0.01, c (0) + = 0, β− = β+, D− = D+. Here, w = 6β+γ/∆c, and τ0 = D+/ w2 are characteristic length and time scales. at the interface of a spherical droplet of radius R we obtain the remaining coefficients R − + R C+ =(cid:18) γβ+ B− =(cid:18) γβ− R (cid:18) γβ+ R (cid:18) γβ− D− D+ R R ν+ ν− 2 sinh(R/l−) k+(cid:19) R exp(R/l+) k−(cid:19) k+(cid:19)(cid:18)1 + k−(cid:19)(cid:18)1 − l+(cid:19) R l− ν− R ν+ R − j+(R) = j−(R) = . (S.20a) (S.20b) (S.21a) (S.21b) R l−(cid:19) . coth The normal fluxes at the droplet interface are + 22 05101520−0.10−0.050.000.050.10with chemical reactionswithout chemical reactionsRadius Using these steady state fluxes in Eq. (S.17) and Eq. (S.15) provides a relation between dR/dt = vn and the droplet radius R in a quasi-static limit. Steady state droplets exist for radii R = ¯R for which dR/dt vanishes. These stationary radii thus obey j+( ¯R) = j−( ¯R) . (S.22) Fig. S1 shows an example of dR/dt as a function of R in the presence (red line) and absence (blue line) of chemical reactions. If chemical reactions are present, two steady state radii denoted ¯Rc (white circle) and ¯Rs (black circle) exist, corresponding to a critical nucleation radius and a stationary droplet radius, respectively. Both stationary radii are shown in Fig. 2A in the main text. In the limit of large characteristic lengths l± compared to the droplet radius R, the stationary radii can be approximated as γβ+ , + and ¯Rc ≈ c∞ − c(0) ¯Rs ≈(cid:115)3D+(c∞ − c(0) ν− where we have used R (cid:28) l− and β±γ/[ ¯R(c(0)− − c(0) + )] (cid:28) 1. The latter is obeyed for a sharp interface, see section I B. (S.24) (S.23) + ) , The critical radius estimated by Eq. (S.23) is closely related to the classical expression for the critical nucleation radius of passive droplets. In the case of active droplets, the supersaturation  = (c∞ − c(0) + )/∆c is determined by chemical reactions instead of the amount of material pro- vided. The stationary droplet radius given in Eq. (S.24) describes an inherently non-equilibrium stationary state that is maintained by opposing fluxes15. C. Stability analysis of the spherical droplet shape To analyze the linear stability of the stationary droplets, we linearize the dynamic equations in the vicinity of the stationary state and identify the dynamic eigenmodes. The stationary state is unstable with respect to a dynamic mode if the corresponding growth rate is positive. 1. Linearization at the stationary solution We linearize the dynamic equations (S.12)–(S.15) and (S.17) around a stationary solution ¯c(r), which obeys Eqs. (S.18)–(S.22). Introducing small perturbations δc and δR of the concentration 23 field and the droplet shape, respectively, we write and c(r, θ, ϕ, t) = ¯c(r) + δc(r, θ, ϕ, t) , R(θ, ϕ, t) = ¯R + δR(θ, ϕ, t) . The concentration perturbation then obeys The boundary conditions (S.14) become ∂tδc = D±∇2δc − k±δc . δc( ¯R±) = β±γδH − ¯c(cid:48)( ¯R±)δR , (S.25) (S.26) (S.27) (S.28) where δH = H( ¯R + δR) − H( ¯R). Using Eqs. (S.15) and (S.17), the time dependence of the droplet shape perturbation is described to linear order by (c(0)− − c(0) + )∂tδR = D+∂rδc( ¯R+) − D−∂rδc( ¯R−) +(cid:2)D+¯c(cid:48)(cid:48)( ¯R+) − D−¯c(cid:48)(cid:48)( ¯R−)(cid:3) δR . (S.29) 2. Dynamic modes and relaxation spectrum The linearized dynamics of droplet perturbations near the steady state defines a linear operator L by The operator L has eigenfunctions (ci, Ri)(cid:124) with corresponding eigenvalues µi, where i is the mode index. These modes obey The linear droplet dynamics can thus be decomposed in eigenmodes with amplitude Ai as where the sum is over all eigenmodes. Thus, the eigenfunctions of L correspond to dynamic modes of the system. For µi < 0, the values −µi are relaxation rates. The steady state is stable if all µi < 0. δc δc ∂t δR = L L Ri = µi Ai δR =(cid:88)i δR . Ri . Ri eµit , δc ci ci  ci 24 (S.30) (S.31) (S.32) 3. Determination of eigenmodes We determine the eigenmodes and the spectrum of relaxation rates of a stationary droplet with radius ¯R. Because of the spherically symmetric reference state, we introduce radial and angular indices i = (n, m, l) and use the ansatz where Ylm are spherical harmonics and the corresponding eigenvalues will be denoted µnl. Using Eq. (S.27) with r2∇2Ylm = l(l + 1)Ylm, the radial part of the eigenfunctions obeys cnlm(r, θ, φ) Rnlm(θ, φ) =  (cid:18) 1 r2 cnl(r) nl  Ylm(θφ) , r2 (cid:19) cnl(r) = 0 , ∂ ∂r r2 ∂ ∂r − (λ± nl)2 − l(l + 1) (S.33) (S.34) (S.35) (S.36a) (S.36b) (S.37) (S.38) (S.39) where (λ± nl)2 = k± + µnl D± . The boundary conditions (S.28) at r = ¯R can be written as cnl( ¯R+) = a+ cnl( ¯R−) = a− l nl l nl with a± l = γβ± hl ¯R2 − ¯c(cid:48)( ¯R±) , where50 hl = (l2 + l − 2)/2. From Eqs. (S.36) we obtain a boundary condition at r = ¯R: a+ l a− Using Eq. (S.29), we obtain a second boundary condition cnl( ¯R+) cnl( ¯R−) = . l (cid:16)c(0)− − c(0) + (cid:17) µnl = D+¯c(cid:48)(cid:48)( ¯R+) − D−¯c(cid:48)(cid:48)( ¯R−) + D+a+ l c(cid:48) nl( ¯R+) cnl( ¯R+) − D−a− l c(cid:48) nl( ¯R−) cnl( ¯R−) . The boundary conditions (S.38) and (S.39) provide jump conditions for both the values and the first derivatives of the radial modes cnl(r) at r = ¯R. 4. Radial profiles and relaxation rates of dynamic modes When solving Eq. (S.34) with Eq. (S.35) to determine the dynamic modes of the system, we nl)2 differs. Near have to distinguish the cases µnl < −k± and µnl > −k±, for which the sign of (λ± 25 an instability of the droplet shape, an eigenmode exists for which µnl changes sign. Therefore, to discuss this instability, it is sufficient to consider the case µnl > −k±. In this case, (λ± nl)2 is positive and solutions to Eq. (S.34) are given by modified spherical Bessel functions kl(λ± nlr) and il(λ± nlr). In order to obtain solutions that are finite at r = 0 and which do not diverge for large r, we have where the coefficient Cnl is determined by boundary conditions (S.38) as cnl(r) = nlr) kl(λ+ Cnl il(λ− nlr) for for r > ¯R r < ¯R , a− l kl(λ+ l il(λ− a+ ¯R) nl ¯R) nl (S.40) (S.41) ¯R) ¯R) . (S.42) The boundary condition (S.39) becomes Cnl = (cid:16)c(0)− − c(0) + (cid:17) µnl = D+¯c(cid:48)(cid:48)( ¯R+) − D−¯c(cid:48)(cid:48)( ¯R−) + D+a+ l λ+ nl . k(cid:48) l(λ+ nl kl(λ+ nl ¯R) ¯R) − D−a− l λ− nl l(λ− i(cid:48) il(λ− nl nl Using λ± nl = ((k± + µnl)/D±)1/2, Eq. (S.42) becomes an implicit equation for the unknown eigen- values µnl. This equation typically has either no solution or one solution. We identify the largest eigenvalue for given l with n = 1. nljl(λ± nlr) + D± In order to determine the full spectrum µnl of eigenmodes, we have to consider the case µnl < −k±. We then define (λ± nl)2 = −(k± + µnl)/D± and the solutions to Eq. (S.34) are of the form C± nlyl(λ± nlr), where jl(z) and yl(z) denote spherical Bessel functions, and the coefficients Cnl and Dnl are determined by boundary conditions. The functions jl(z) and yl(z) behave for large r as jl(z) ∼ z−1 sin(z − lπ/2) and yl(z) ∼ z−1 cos(z − lπ/2). Eq. (S.39) now has an infinite set of solutions µnl for n > 1, which we order such that µnl > µn+1,l. In an infinite system, the set µnl approaches a continuous spectrum. 5. Instability of stationary spherical droplets The droplet shape is unstable if at least one mode with µ1l > 0 exists. We can obtain a criterion for this instability by using µnl = 0 in Eq. (S.42). This leads to k(cid:48) l( ¯R/l+) kl( ¯R/l+) − 0 = D+¯c(cid:48)(cid:48)( ¯R+) − D−¯c(cid:48)(cid:48)( ¯R−) + D+a+ l l+ D−a− l− l i(cid:48) l( ¯R/l−) il( ¯R/l−) , (S.43) which is a condition for the radius ¯R at which the shape becomes unstable with respect to a deformation characterized by l. 26 FIG. S2. Eigenvalues µ1l as a function of supersaturation . At the onset of the instability (red dot) the second mode becomes unstable, leading to droplet division. For larger values of , higher modes become unstable as well. The same parameters as in Fig. 2A (main text), with ν−/ν0 = 1. Stationary and stable radii were used (µ10 < 0). Different modes l can become unstable. The case l = 0 corresponds to changes of the radius. A droplet with µ10 < 0 has a stable radius ¯R. For l = 1 there always exists one marginal mode with µ1m = 0, which corresponds to a translation of the steady state and does not lead to an instability. The first mode that becomes unstable and changes the droplet shape is the elongation mode l = 2. Fig. S2 shows numerically determined values of the largest relaxation rate µ1l for l = 0, 1, 2 and 3 as a function of supersaturation  far from the droplet. The figure reveals that µ12 changes sign and becomes positive as  is increased, indicating the shape instability. Eq. (S.42) can be solved numerically. An approximation of the eigenvalues that is valid in the limit of weak chemical reactions R (cid:28) l+ is ∆c ¯R2(cid:34)(cid:0)c∞ − c(0) + (cid:1) − γ µ1l (cid:39) (l − 1) D+ 2 ¯R(cid:18)(4 + 3l + l2)β+ + l(l + 2) β−D− D+ (cid:19)(cid:35) . (S.44) For modes l ≥ 2, the spherical droplet becomes unstable for ¯R > Rl which in this limit is given 27 0.10.2−0.0100.01Supersaturation (4 + 3l + l2)D+β+ + l(l + 2)D−β− . (S.45) 2D+(c∞ − c(0) + ) by Rl ≈ γ This expression shows that the the elongation mode l = 2 is the first mode to become unstable. This provides an approximation for the critical radius of droplet division, Rdiv (cid:39) γ 7β+ + 4β− D− D+ c∞ − c(0) + , (S.46) in the weak reaction limit. Because the limit R (cid:28) l+ corresponds to vanishing chemical re- actions, this approximate expression approaches the instability condition of the Mullins-Sekerka instability44. 28
1703.09515
2
1703
2017-04-11T17:11:31
Spring-damper equivalents of the fractional, poroelastic, and poroviscoelastic models for elastography
[ "physics.bio-ph", "cond-mat.soft" ]
In MR elastography it is common to use an elastic model for the tissue's response in order to properly interpret the results. More complex models such as viscoelastic, fractional viscoelastic, poroelastic, or poroviscoelastic ones are also used. These models appear at first sight to be very different, but here it is shown that they all may be expressed in terms of elementary viscoelastic models. For a medium expressed with fractional models, many elementary spring-damper combinations are added, each of them weighted according to a long-tailed distribution, hinting at a fractional distribution of time constants or relaxation frequencies. This may open up for a more physical interpretation of the fractional models. The shear wave component of the poroelastic model is shown to be modeled exactly by a three-component Zener model. The extended poroviscoelastic model is found to be equivalent to what is called a non-standard four-parameter model. Accordingly, the large number of parameters in the porous models can be reduced to the same number as in their viscoelastic equivalents. As long as the individual displacements from the solid and fluid parts cannot be measured individually the main use of the poro(visco)elastic models is therefore as a physics based method for determining parameters in a viscoelastic model.
physics.bio-ph
physics
Spring-damper equivalents of the fractional, poroelastic, and poroviscoelastic models for elastography Sverre Holm University of Oslo 12th April, 2017 Abstract In MR elastography it is common to use an elastic model for the tissue's response in order to properly interpret the results. More complex models such as vis- coelastic, fractional viscoelastic, poroelastic, or poro- viscoelastic ones are also used. These models appear at first sight to be very different, but here it is shown that they all may be expressed in terms of elementary viscoelastic models. For a medium expressed with fractional models, many elementary spring–damper combinations are added, each of them weighted according to a long- tailed distribution, hinting at a fractional distribution of time constants or relaxation frequencies. This may open up for a more physical interpretation of the frac- tional models. The shear wave component of the poroelastic model is shown to be modeled exactly by a three-component Zener model. The extended poroviscoelastic model is found to be equivalent to what is called a non-standard four-parameter model. Accordingly, the large number of parameters in the porous models can be reduced to the same number as in their viscoelastic equivalents. As long as the individual displacements from the solid and fluid parts cannot be measured individually the main use of the poro(visco)elastic models is therefore as a physics based method for determining parameters in a viscoelastic model. 1 Introduction In modeling of data from MR elastography, it is com- mon to use a simple elastic model for the medium. This is the case even for ultrasound elastography. For more accurate modeling, there are three families of models that are used. These are the linear viscoelastic models such as the Kelvin-Voigt and Zener models, the fractional extensions of these models, and poroelastic models based on the theory of Biot. The linear viscoelastic models are among those that have been used for fitting frequency dependency of shear wave data from MR elastography in the brain [1]. The fractional Kelvin-Voigt model was fitted to breast MR elastography data in [2] and also analyzed in [3] and compared to other models for elastography in [4]. It may be argued that these single-phase models are too simplistic and that in tissue a bi-phasic model which distinguishes between the solid and liquid phases would be more accurate. This potential has al- ready been demonstrated in models of the quasi-static biomechanics of hydrocephalus [5] and of infusion- induced swelling in the brain [6]. The poroelastic model has also been used for elastography. In [7] ultra- sound elastography was modeled with either an elastic model (i.e. without viscosity) or a simplified poroe- lastic model which depended on porosity, composite density and fluid density. Similarly [8] evaluated the full poroelastic model vs. an elastic one for MR elastog- raphy. In [9] they went one step further and compared the poroleastic model with a viscoelastic one with a complex shear modulus. They also tried to use these models for inversion and observed that the viscoelastic model produced better reconstructions at 50 Hz while the poroelastic model was superior at 1 Hz. The challenge in making a reconstruction algorithm based on the poroelastic model is the need for captur- ing the displacement of the solid and the fluid inde- pendently. Chapter 5 in [10] states that due to the voxel size of MR elastography, it cannot detect properties of individual pores, but only sees a homogeneous effec- tive medium. Further, since it is sensitive to signals from hydrogen atoms in the voxel, one cannot separate 1 the signal from the solid and the fluid, even if the solid should contain fluid which is considered not to be in the pores. Likewise ultrasound is scattered from the soft tissue which is mainly part of the solid matrix. Ul- trasound will therefore produce a strain estimate in the solid matrix which is only indirectly influenced by the wave in the fluid [7]. In this paper, the goal is however not to develop a poroelastic model for reconstruction of the MR elastography image, but only to discuss it in the framework of explaining variations in parameters due to physiological and physical parameters. The claim of this paper is that these models are much more similar than they appear to be. The frac- tional models can be developed as sums of ordinary viscoelastic elements weighted in a particular way. The fractional model has its strength in that it gives a par- simonious description of the phenomenon, i.e. one with a minimal number of parameters, especially when power-law variation in frequency is observed. But it is not fundamentally different. Likewise the poroelastic model for the wave mode of interest for elastography, the shear wave, can be described in terms of standard viscoelastic models. In this case it is the viscoelastic model which requires the smallest number of parame- ters. The strength of the poroelastic formulation is that it gives a way of finding how these parameters depend on physical and even physiological parameters. The linear viscoelastic model is therefore first gen- eralized in Sec. 2 from the simple two- and three- component models to chains of spring-damper ele- ments described by time- and frequency-spectral func- tions. These functions are used to show that the frac- tional viscoelastic models in Sec. 3 can be described as sums of ordinary viscoelastic models. A surprising result is that the weighting in the sum follows a long- tailed distribution reminiscent of a fractal. The shear wave solution of the poroelastic theory is then developed in Sec. 4 both from the original for- mulation of Biot [11] and also from that of Stoll [12]. Somewhat unexpectedly, it is found that there is a one to one correspondence between the poroelastic shear wave response and that of a simple spring-dashpot network. 2 Linear viscoelastic models The linear viscoelastic model is expressed in three dif- ferent ways. The different ways are needed in order to generalize from the simple two and three-component models (e.g. Kelvin-Voigt and Zener) to higher order models. In order to illustrate the similarities between models it is sufficient here to express the models in one dimen- sion, although three dimensions are really needed for a complete shear wave description. This section, as well as Sec. 3 builds to a large degree on [13]. 2.1 Three descriptions In linear viscoelasticity there are three different ways of describing the medium's response. The first is the hereditary model of Boltzmann [14, 15] where the con- stitutive equation is a convolution integral [13] (ch. 2): σ(t) = G(t)∗ ∂(t) ∂t . (1) The kernel G(t) is called the relaxation modulus and represents a fading memory, i.e. one where changes in the past have less effect now than more recent changes. In order to ensure causality, the kernel, G(t), has to be zero for non-negative time. The relaxation modulus is the strain response of the system to a step excitation in strain. The second description is a linear differential equa- tion between stress and strain with constant coeffi- cients:(cid:34) 1+ p(cid:88) (cid:35) ak ∂k ∂t k k=1 σ(t) = (cid:35) (cid:34) Ee + q(cid:88) k=1 bk ∂k ∂t k (t). (2) n=0 The third description is in the form of a relaxation spec- trum or Prony expansion: G(t) = Ge +Gτ +G−δ(t), Gτ(t) = N−1(cid:88) En exp(−t/τn). (3) A physically realizable model is obtained if G(t) is mod- eled by a parallel network of pairs of springs and dash- pots in series where the spring constants are En and the viscosity of the dashpots are ηn = τnEn. Further- more, both the spring constants and the viscosities are non-negative. This is the Maxwell-Wiechert model of Fig. 1. The left-hand spring leads to the constant Ge = Ee, the equilibrium modulus, and if there is a dashpot directly across the terminals (e.g. if E1 = 0), the response will have an impulse at time zero as well given by G−. This is the case in the Kelvin-Voigt model which will soon be discussed. Not all the coefficient sets of (2) will lead to a fading memory model, so the linear differential equation is 2 dynamic modulus is: E(ω) = Ee + i ωη = Ee(1+ i ωτ) (7) where τ = η/Ee. The even simpler elastic model, which is often the reference in elastography as noted in the Introduction, is obtained by setting the viscosity, η, to 0. The Zener model adds one more term to the linear differential equation of the Kelvin-Voigt model: σ(t)+ a1 ∂σ(t) ∂t = Ee (t)+ η ∂(t) ∂t , and the relaxation modulus is: τ τσ G(t) = Ee + Ee( − 1)e −t/τσ. (8) (9) Compared to the Kelvin-Voigt model, the effect of the second spring is to 'soften' the impulse in (6) into a falling exponential. In this model p = q = 1 and b1 = η in (2) and (4). The time constants are τ as in the Kelvin-Voigt model and τσ = a1. The parameters of the model relate to the physical parameters of Fig. 1 as follows: τσ = η/E ≤ τ = η/E (cid:48) , 1 E(cid:48) = 1 Ee For this model the dynamic modulus is: (4) E(ω) = Ee 1+ i ωτ 1+ i ωτσ + 1 E . (10) (11) The addition of the extra spring leads to frequency de- pendent denominator in the dynamic modulus which gives more degrees of freedom in fitting this model to data. The dynamic modulus can also be included in a dis- persion relation when propagating waves are involved. In [16], Eqs. (13-15), it is shown that the complex wave number in that case is: (cid:181) k (cid:182)2 = ρκ(ω) = ρ E(ω) (12) where ρ is the density and κ(ω) is the dynamic com- pressibility, the inverse of the dynamic modulus. This result is important in the subsequent analysis of the poroelastic model. Note that the roles of τ and τσ are reversed here compared to [16]. Here the convention of [13] is followed instead. Figure 1: Maxwell-Wiechert model consisting of paral- lel networks of spring-damper models Figure 2: Kelvin-Voigt (left) and Zener models. As shown in this paper, the three-component Zener model also models the shear wave in the Biot poroe- lastic model the most general model. Also, a model described as a relaxation spectrum, (3), can always be described with a linear differential equation, but not vice versa. In addition, a relaxation spectrum model, (3), always results in a fading memory model as it consists of sums of positively weighted falling exponentials. The frequency domain equivalent of (2) is also use- ful: E(ω) = σ(ω) (ω) = Ee +(cid:80)q 1+(cid:80)p k=1 bk(i ω)k k=1 ak(i ω)k The dynamic modulus, E(ω), is often called G(ω) in elastography, but since G here means the relaxation modulus in this paper, E is used instead. 2.2 Elementary linear viscoelastic models The two simplest viscoelastic models are the Kelvin- Voigt and the Zener models as shown in Fig. 2. The Kelvin-Voigt model is found by just keeping Ee and η1 in Fig. 1. The linear differential equation is σ(t) = Ee (t)+ η ∂(t) ∂t , (5) ω and the relaxation modulus is: G(t) = Ee + ηδ(t). (6) Thus p = 0, q = 1, a1 = 0, and b1 = η in (2) and (4). Further Ge = Ee, G− = η, and Gτ(t) = 0 in (3). The 3 EeE1η1η1EN−1ηN−1ηN−1EeηEeEη A slightly rewritten version of the compressibility based on factoring of (11) will be needed when analyz- ing the poroelastic model: κ(ω) = 1 E(ω) = 1 Ee 1+ ω2ττσ − i ω(τ − τσ) 1+ ω2τ2  (13) Judging from Fig. 1, a Zener model is characterized by two springs and a damper. These three parameters may be expressed in different ways, and one common way is via the low frequency asymptote of the propaga- tion speed, c0, and the two cross-over frequencies, ω and ωσ. The two frequencies express approximately (cid:112) where the phase velocity starts rising and where it ap- proaches its asymptotic value c∞ = c0 ωσ/ω [16]. = Ee/ρ and thus de- Taking into consideration that c2 0 pends on both the shear modulus and the density, then a Zener medium will depend on four independent pa- rameters: Ee, ρ, ω and ωσ. 2.3 Time- and frequency-spectral func- tions Gτ(t) = b −t/τd τ, (cid:90) ∞ The continuous generalization of (3) is given in [17, 18, 13]: Rσ(τ)e (14) where b = Gτ(0) is a non-negative constant which for the Zener model is b = E(τ/τσ − 1) [18]. Rσ(τ) is a non-negative relaxation spectrum. 0 There is a corresponding frequency-spectral func- tion which is Sσ(Ω) = b Rσ(1/Ω) Ω2 By substituting Ω = 1/τ so d Ω = −d τ/τ2 this gives (15) Gτ(t) =(cid:90) ∞ 0 Sσ(Ω)e −Ωt d Ω. (16) The relaxation modulus is, as noted, the stress re- sponse of the system to a step excitation in strain. The relaxation spectrum could also have been expressed with the creep response, which is the strain response to a step excitation in strain. It is denoted by J(t) and plays the same role as G(t) in (1) if σ(t) and (t) are interchanged. The time-spectral function is: −t/τ)d τ (17) where a = Jτ(∞) is a non-negative constant which for the Zener model is a = (1/E)(1− τσ/τ) [18]. This cor- responds to a decomposition in a sum of elementary R(τ)(1− e Jτ(t) = a (cid:90) ∞ 0 Figure 3: Kelvin model consisting of series networks of spring-damper models. It is the conjugate of the Maxwell-Wiechert model of Fig. 1 models in series as shown in Fig. 3. This is the con- jugate of the Maxwell-Wiechert model of Fig. 1, i.e. a different configuration of springs and dash-pots which has the same characteristics [19]. It follows that the conjugate of the Zener model in the right-hand part of Fig. 2 is the top three elements of Fig. 3 also. The frequency-spectral function for the creep is found by a transformation which is analogous to that for the relaxation: S(Ω) = a R(1/Ω) Ω2 and the frequency-spectral function is: J(t) = −(cid:90) ∞ 0 Rσ()(1− e −t/τ)d τ (18) (19) This frequency spectral function will be used to find the spring-damper equivalent of the fractional models. 3 Fractional models Both the Kelvin-Voigt model and the Zener model can be generalized by introducing non-integer, fractional derivatives of order α in the constitutive equation (2). For the Zener model this is: = E σ(t)+ τα (t)+ τα ∂ασ(t) (20) (cid:183) (cid:184) σ ∂α(t) ∂t α  ∂t β The dynamic modulus is now: E(ω) = Ee 1+ (i ωτ)α 1+ (i ωτσ)α (21) 4 EgE1η1η1EN−1ηN−1ηN−1 Figure 4: Fractional Kelvin-Voigt model with spring characterized by shear modulus, E, and spring-pot given by shear viscosity, η, and fractional order, α while in the simpler Kelvin-Voigt model shown in Fig. 4, the constant τσ = 0: σ(t) = E (t) (22) (cid:183) (cid:184) . (t)+ τα ∂α ∂t α (cid:161)1+ (i ωτ)α(cid:162) The dynamic modulus is: E(ω) = Ee In the limit as the frequency and/or viscosity is very large the dynamic modulus approaches E(ω) → (i ωη)α. This is equivalent to the case where the spring of Fig. 4 can be neglected. That seems to be the case often for elastography data [20, 21]. The relaxation modulus of the fractional Kelvin- Voigt model is expressed by a power law function of time, while for the fractional Zener model it follows a Mittag-Leffler function. That function is a general- ization of the exponential function with power-like behavior. The creep compliance of both models also follows a Mittag-Leffler function. That means that the creep time- and -frequency-spectral functions of the two models will be similar also. 3.1 Fractal time- and frequency-spectral function In [17, 18, 13] it has been shown that for the fractional Zener model, the creep time-spectral function of (17) is: R(τ) = 1 πτ sin απ (τ/τ)α + (τ/τ)−α + 2cos απ (24) It was plotted in [18, 13] with linear axes and is a de- creasing function of τ for small α and gets a more and more pronounced peak as α approaches 1. For the non-fractional case, α = 1 it is a delta function at τ/τ = 1 showing that it is equivalent to a single relax- ation process in that case. Figure 5: Time-spectral function for the fractional Zener and Kelvin-Voigt models. Solid line α = 0.8, dashed line α = 0.01 (23) Here the function is plotted on a log log scale in Fig. 5. That brings out the properties of this particular function in a different way than if it is plotted in a linear plot. The log log plot fixes the attention on the asymptotes, rather than the peak, and they are: (cid:40)τα−1 τ−α−1 R(τ) ∝ for τ/τ (cid:191) 1, for τ/τ (cid:192) 1 (25) Both of them are power laws. As long as the exponent of the high-frequency tail falls off slower than τ−2, i.e. for α < 1, the variance of the distribution will not ex- ist. Such long-tailed distributions are scale-invariant and therefore may indicate fractal properties. This indicates that the distribution of elementary spring- damper models in the medium, as given in Figs. 1 and 3, may be a multi-fractal with two different fractal or- ders. The references above didn't derive the frequency- spectral function, but it can be found by substituting (18) and using the normalizing constant a = Jτ(∞) = (1/E)(1− (τσ/τ)α) for the fractional Zener model. The creep frequency-spectral function of (19) is then: S(Ω) = 1 πE  − τα (τα σ)Ωα−1 sin απ (Ωτ)2α + 1+ 2(Ωτ)α cos απ . (26) This result was first given in [22] in the form above. There it was derived with a starting point in the mul- tiple relaxation theory of [23]. It should be noted that the frequency-spectral function is exactly the same as the time-spectral function except for a scaling factor. The plot for α = 0.8 and α = 0.01 in Fig. 6 shares many properties with the plot of the time-spectral 5 Eη,α10-310-210-1100101102103104105/10-910-810-710-610-510-410-310-210-1100 = 0.8 = 0.01 Thus an alternative way to ensure causality is to em- ploy a fractional Kelvin-Voigt model with parameter α =  = 0 + . Hysteresis is particularly interesting in the context of fractional operators because the result for the limiting case α → 0 of the previous section is similar to that re- cently found for hysteresis in [26]. There it was shown that the hysteresis model is the same as a sum of re- laxation processes weighted with a long-tailed power law Ω−1+. This parallels our discussion of the asymp- totic result of (27) for α =  and Parker's result can be interpreted as a special case of (27). Figure 6: Frequency-spectral function for the fractional Zener model for τσ = 1000τ. Solid line α = 0.8 as in [22], dashed line α = 0.01. It is also valid for the Kelvin- Voigt model except for an amplitude scaling 4 Poroelastic model function. The asymptotes are also very similar: (cid:40)Ωα−1 Ω−α−1 S(Ω) ∝ for Ωτ (cid:191) 1, for Ωτ (cid:192) 1 (27) Interestingly, the relaxation spectrum approaches a single fractal for the limiting case α → 0, where both the low- and high-frequency parts will approach Ω−1. See the dash-dotted line for α =  = 0.01 in Fig. (6) with asymptotes Ω−1+ for low frequencies and Ω−1− for high frequencies. 3.2 Relationship with hysteresis The results found here have some resemblance with those derived from hysteresis. Hysteresis is a hypothet- ical loss element with a constant phase lag between stress and strain at all frequencies [24] resulting in at- tenuation that increases linearly with frequency [25]. The dynamic modulus for hysteresis is: E(ω) = K + i H (28) where K and H are constants. The model leads to a noncausal model and therefore [24] is concerned with making a bandlimited approximation where the constants are allowed to vary with frequency, K (ω) and H(ω), in order to ensure causality. By comparing the dynamic moduli, (28) and (23), it is evident that hysteresis can be viewed as the limit- ing case as α → 0 for the fractional Kelvin-Voigt model. The Biot poroelastic model deals with a saturated porous medium with a solid phase and a fluid phase. Wave propagation in such a medium is described by a set of coupled vector wave equations as given in Eq. (4.2) in [11]. The variables are the displacement vectors u and U for the displacement in the solid and the fluid respectively. That paper assumes that as the fluid moves in the pores, the flow is laminar and that losses are given by Darcy's law and proportional to the relative velocity, ∂(u− U)/∂t. The theory predicts three solutions, two compressional waves and one shear wave. In a biological porous medium like cancellous bone (bone with a low volume fraction of solid, less than 70%), all three waves have been detected [27]. This model is much more complex than the vis- coelastic models of the previous sections. It also seems more appropriate for a complex medium like brain, liver, or bone. Here a surprising exact relationship be- tween the poroelastic and viscoelastic models will be shown for the shear wave solution. 4.1 Biot's original formulation As elastography is only concerned with shear waves, we restrict the analysis here to that mode. Then the following dispersion relation was derived in [11], Eqs. (7.5)–(7.6): (cid:181) k ω (cid:182)2 = ρ(κr − i κi ), (29) 6 10-510-410-310-210-1100101102103 10-610-510-410-310-210-1100101 = 0.8 = 0.01 where κr = 1 µ (cid:179) f (cid:180)2 (cid:180)2 (cid:180)2(cid:179) f fc 12 fc 1+ γ22 1+(cid:179) γ11γ22−γ2 (γ12+γ22)2 γ22 γ12+γ22 , κi = 1 µ f fc 1+(cid:179) (cid:180)2(cid:179) f γ12 + γ22 γ12+γ22 γ22 fc (30) This expression depends on three normalized densi- ties: γ11 = ρ11 , γ22 = ρ22 , γ12 = ρ12 , (31) ρ ρ ρ and the aggregate or composite density which is ρ = φρ f + (1− φ)ρs. (32) These formulas depend on the porosity, φ, and the fluid and solid densities. The parameter ρ12 represents a negative mass coupling density between fluid and solid, ρ11 represents the total effective mass of the solid moving in the fluid, and ρ22 is the total effective mass of the fluid. There is also a characteristic frequency fc = b 2π(ρ12 + ρ22) , b = ηφ2 B (33) where η is the fluid viscosity, and B is the permeability. The effective characteristic frequency of (30) is how- ever a lower frequency: γ12 + γ22 (cid:48) c = fc f (34) since ρ22 = αφρ f [28] where α is a structure constant related to tortuosity. γ22 = 1 2π ηφ B αρ f Comparison with (12) and (13) shows that (29) and (30) are exactly the same as those of the Zener model. This is a remarkable result which shows that for shear waves, the poroelastic model is also a linear viscoelas- tic model. What distinguishes it from the ordinary linear viscoelastic models is that it provides a sophisti- cated way of determining the parameters from physi- cal considerations. 4.2 Biot-Stoll formulation The parameters of the original Biot theory are often considered to be hard to find in practice and the the- ory has been rewritten in terms of the relative dis- placement between fluid and solid, u− U, or the vol- ume of fluid that has flowed in or out of an element, ζ = φ∇(u− U), in combination with the solid displace- ment u. The material parameters are then transformed to a new set of parameters. 7 (cid:180)2 In that case Eq. (16.77) in [29] gives the shear disper- sion of the low frequency Biot model. It can also be found from Eq. (12) of [12]: ρ )− i 1+ i ω (cid:182) B η (cid:181) k η ωB (cid:182)2 = ρ (cid:181) ρc − ρ2 1+ i ω ρc B f ρ (ρc − ρ2 ρc − i = ρ µ ω µ f η ωB η (35) where the sign of ω has been reversed compared to the original articles due to a different definition of the Fourier transform than here (see also Appendix A of [30]). The mass coupling density is ρc = αρ f /φ (36) where α is the tortuosity. Comparing (35) to (12) shows that the dynamic modulus is: (cid:179) ω (cid:180)2 = µ k E(ω) = ρ (cid:181) 1+ i ω ρc B ρc − ρ2 f ρ η (cid:182) 1+ i ω (37) B η As expected this is also equivalent to that of a Zener model as can be seen by comparison with (11). The constants when ω = 1/τ and ωσ = 1/τσ are: Ee = µr , ω = η ≥ ω, ρc B = µ ρ , ωσ = ω 1− ρ2 f ρρc (38) Insertion of the expression for the mass coupling den- sity, (36), in the equation for the effective characteristic frequency, (34), shows that this frequency also is the same as ω/(2π). c2 0 . 4.3 Redundant parameters in the Biot shear wave model The low frequency Biot model is characterized by the ten parameters shown in Table 1. Seven of them affect shear wave propagation as indicated in the right-hand column, but several of them are connected as (38) in- dicates. The parameter ω depends on the mass coupling density ρc . In addition it depends on the ratio of the viscosity, η, and the permeability, B. They do not in- fluence any of the other parameters ωσ and c0 so it is clear that it is their ratio which matters. Likewise, the sound velocity, c0, depends on the ratio of the shear modulus, µr and the aggregate density, ρ. Furthermore for the third parameter, ωσ, the de- nominator in the expression is usually larger than [0 ... 1] [kg/m3] [kg/m3] [Pa] [Pa] Porosity Solid density Fluid density Bulk modulus, solid Bulk modulus, fluid Biot model parameters Bulk properties: φ ρs ρ f Ks K f Fluid parameters: η B α Rigid frame response parameters: Kr µr Viscosity Permeability Tortuosity Bulk modulus Shear modulus [Pa s] [m2] [1 ... 3] [Pa] [Pa] Shear? Y Y Y N N Y Y Y N Y Figure 7: The equivalent to the shear wave poro- viscoelastic model called the non-standard four- parameter model in [19] (left) with its conjugate to the right . Table 1: Biot-Stoll parameters and their effect on the shear wave. The pore radius parameter is not included as the flow in the pores is assumed to be laminar. The right-hand column shows whether the parameter af- fects the shear wave or not 0.8 so the ratio of ωσ and ω is not very sensitive to changes in the densities, ρ, ρc , or ρ f . That means that neither is it very sensitive to changes in tortuosity, and it will mainly be ω which depends on α in the ratio η/(αB). In this way three of the seven parameters can be said to be redundant and the seven parameters from Table 1 (η, B, α, ρ f , ρs, φ, and µr ) may be reduced to four by combining the three first ones into one, η/(αB). Alternatively, the four parameters may be stated as η/α, ρc , ρ, and µr . In case one wants to compute ωσ exactly, a fifth parameter, ρ f , also needs to be included. In this way the number of parameters in the Biot shear model with some approximation is four as in the Zener medium model, and five in the exact case. 4.4 Extension to poroviscoelasticity In the sediment acoustics field, the Biot model has been amended in order to extend the model from a rigid porous frame, like a porous rock or bone, to one where the grains are allowed to move. This is a model for a saturated sediment and it is not unlikely that it may be more appropriate for tissue than the rigid frame implied in the Biot model also. In this case viscosity is introduced in the solid frame in addition to in the flowing liquid. The model is called the Biot squirt flow and viscous drag (BICSQS) model [31]. This viscosity is added by allowing a relaxation model for the shear modulus, thus allowing for a fre- quency dependent complex shear modulus or a dy- namic shear modulus, described by a characteristic frequency ωµ: (39) (cid:181) 1+ i (cid:182) ω ωµ µ(cid:48) = µ When the relaxation is included in (37) and (11), the result is E(ω) = µ (1+ i ω/ωµ)(1+ i ω/ω) 1+ i ω/ωσ (40) In the limit as frequency goes to zero the dynamic modulus approaches E(0) = µ and as the frequency ap- proaches infinity it approaches E(∞) = ∞. This model is called the non-standard Four-parameter model in [19] (Chap. 3.4.2). Its spring–damper realization is shown in Fig. 7 with the equivalent conjugate model to the right. Therefore, even in this case, an equivalent viscoelastic model may be found. 5 Discussion and Conclusion The concept of the time- and frequency-spectral de- compositions of viscoelastic systems has been devel- oped and applied to the fractal Zener and Kelvin-Voigt models. This summation of multiple elementary mod- els is an idea which is independent of the fractional models. In fact it is the idea behind the relaxation spec- trum models of Kelvin and Wiechert dating from 1888 8 E1η1E2η2E0Eηη0 and 1893 respectively [15]. In addition to the cover- age in [13], the book [19] devotes several sections to it (Chap. 3.5-3.6). Similar ideas of model fitting have been used in biomechanics and elastography also. In the white matter model of [32] the shear relaxation modulus was for instance modeled with three very slow relaxation terms (0.01 Hz and slower) and a simi- lar model with two terms was used in [33]. The idea behind the particular weightings implied in the time-spectral and frequency-spectral functions found here is however to make the sum approximate the behavior of the fractional models, i.e. in terms of power laws in the frequency domain. Therefore it par- allels the modeling of arbitrary power law attenuation in medical ultrasound over a limited frequency range with a few terms in [34]. The surprising result is that the weighting has a long- tailed distribution. This is a property which is associ- ated with a fractal geometry. Here it characterizes the distribution of individual relaxation processes both in the time and the frequency domains. An interesting special case is the hysteresis model recently proposed by [26] and this could possibly shed some light on the interpretation of the distribution. The fractal distri- bution is an explanation that seemingly is based on a completely different physical mechanism than the link shown between the fractional damper and a linearly time-varying viscosity in [35]. But perhaps the link is even deeper and suggests that such a time-varying vis- cosity in some way is associated with a medium with a fractal distribution of relaxation processes? The connection between the fractal distribution and the fractional constitutive equation of an absorbing medium is also a result that complements recent re- sults for the alternative attenuation mechanism due to scattering rather than absorption. An example is [36, 37] where it is shown that a fractal distribution of random scatterers leads to attenuation that varies with frequency according to a power law. The fractal – fractional derivative connection can also be compared to the geometrical and physical interpretation of frac- tional integration and differentiation in [38], but the result here is more specific to one particular applica- tion, namely viscoelasticity. The poroelastic model depends on ten independent parameters, of which seven influence the shear wave solution, and seems to build on an entirely different foundation than the viscoelastic models. Despite that it has been shown to be equivalent to a Zener model. This was done by comparing dispersion relations, and not unsurprisingly the result is the same whether the dispersion relations from the original Biot theory or those from the Biot-Stoll theory are analyzed. This re- sult is an extension of [39] where it was shown that in the low loss/low frequency approximation, an equiva- lent can be found between the poroelastic model and a Kelvin-Voigt model. That was done by comparing ap- proximate expressions for the velocity and attenuation. As the Zener model in the low loss/low frequency case is equivalent to the Kelvin-Voigt model [16], the result found here also agrees with that result. It is also shown that the seven parameters of the poroelastic model can be reduced to five, and even four as in the Zener model if a small approximation is allowed. When the Biot model is extended to include viscosity in the frame as in [31], an extra damper has to be added to the Zener model making it into what is called the non-standard four-parameter model in [19]. References [1] Klatt D, Hamhaber U, Asbach P, Braun J, Sack I. Noninvasive assessment of the rheological behav- ior of human organs using multifrequency MR elastography: a study of brain and liver viscoelas- ticity. Phys. Med. Biol.. 2007;52:7281–7294. [2] Sinkus R, Siegmann K, Xydeas T, Tanter M, Claussen C, Fink M. MR elastography of breast lesions: Understanding the solid/liquid du- ality can improve the specificity of contrast- enhanced MR mammography. Magn. Res. in Med.. 2007;58:1135–1144. [3] Holm S, Sinkus R. A unifying fractional wave equa- tion for compressional and shear waves. J. Acoust. Soc. Am.. 2010;127:542–548. [4] Zhang W, Holm S. Estimation of shear modulus in media with power law characteristics. Ultrasonics. 2016;64:170–176. [5] Nagashima T, Tamaki N, Matsumoto S, Hor- witz B, Seguchi Y. Biomechanics of hydro- cephalus: A new theoretical model. Neurosurgery. 1987;21:898–904. [6] Basser PJ. Interstitial pressure, volume, and flow during infusion into brain tissue. Microvasc. Res.. 1992;44:143–165. 9 [7] Konofagou EE, Harrigan TP, Ophir J, Krouskop TA. Poroelastography: imaging the poroelastic prop- erties of tissues. Ultras Med Biol. 2001;27:1387– 1397. [20] Zhang M, Castaneda B, Wu Z, et al. Congruence of imaging estimators and mechanical measure- ments of viscoelastic properties of soft tissues. Ultrasound Med. Biol.. 2007;33:1617–1631. [8] Perriñez PR, Kennedy FE, Van Houten EE, Weaver JB, Paulsen KD. Modeling of soft poroelastic tissue in time-harmonic MR elastography. IEEE Trans Biomed Eng. 2009;56:598–608. [21] Sinkus R, Lambert S, Abd-Elmoniem KZ, et al. Rheological determinants for simultaneous stag- ing of hepatic fibrosis and inflammation in pa- tients with chronic liver disease. submitted. 2017. [9] McGarry M, Johnson C, Sutton B, et al. Suitability of poroelastic and viscoelastic mechanical mod- els for high and low frequency MR elastography. Med. Phys.. 2015;42:947–957. [10] Hirsch S, Braun J, Sack I. Magnetic Resonance Elastography: Physical Background and Medical Applications. Wiley-VCH 2016. [11] Biot MA. Theory of propagation of elastic waves in a fluid-saturated porous solid. I. Low- frequency range. J. Acoust. Soc. Am.. 1956;28:168– 178. [12] Stoll RD. Acoustic waves in ocean sediments. Geo- physics. 1977;42:715–725. [13] Mainardi F. Fractional Calculus and Waves in Lin- ear Viscoelesticity: An Introduction to Mathemati- cal Models:1–347. London, UK: Imperial College Press 2010. [14] Boltzmann L. Zur theorie der elastischen nach- wirkung (On the theory of hereditary elastic ef- fects). Ann. Phys. Chem. 1876;Bd. 7:624–654. [15] Markovitz H. Boltzmann and the beginnings of linear viscoelasticity. Trans. Soc. Rheol. (1957- 1977). 1977;21:381–398. [16] Holm S, Näsholm SP. A causal and fractional all-frequency wave equation for lossy media. J. Acoust. Soc. Am.. 2011;130:2195–2202. [17] Gross B. On creep and relaxation. J Appl Phys. 1947;18:212–221. [18] Caputo M, Mainardi F. Linear models of dissipa- tion in anelastic solids. La Rivista del Nuovo Ci- mento (1971-1977). 1971;1:161–198. [19] Tschoegl NW. The phenomenological theory of linear viscoelastic behavior: An introduction. Springer-Verlag Berlin 1989. Reprinted in 2012. [22] Näsholm SP, Holm S. Linking multiple relaxation, power-law attenuation, and fractional wave equa- tions. J. Acoust. Soc. Am.. 2011;130:3038–3045. [23] Nachman AI, Smith III JF, Waag RC. An equa- tion for acoustic propagation in inhomogeneous media with relaxation losses. J. Acoust. Soc. Am.. 1990;88:1584–1595. [24] Parker KJ. Real and causal hysteresis elements. J. Acoust. Soc. Am.. 2014;135:3381–3389. [25] Stoll RD, Bryan GM. Wave attenuation in J. Acoust. Soc. Am.. saturated sediments. 1970;47:1440–1447. [26] Parker KJ. Could linear hysteresis contribute to shear wave losses in tissues? Ultrasound Med. Biol.. 2015;41:1100–1104. [27] Hosokawa A, Otani T. Ultrasonic wave propaga- tion in bovine cancellous bone. J. Acoust. Soc. Am.. 1997;101:558–562. [28] Berryman JG. Confirmation of Biot's theory. Appl. Phys. Lett.. 1980;37:382–384. [29] Hovem JM. Marine Acoustics: The Physics of Sound in Underwater Environments. Peninsula publishing 2012. [30] Holm S, Näsholm SP. Comparison of fractional wave equations for power law attenuation in ultra- sound and elastography. Ultrasound. Med. Biol.. 2014;40:695–703. [31] Chotiros NP, Isakson MJ. A broadband model of sandy ocean sediments: Biot–Stoll with contact squirt flow and shear drag. J. Acoust. Soc. Am.. 2004;116:2011–2022. [32] Cheng S, Bilston LE. Unconfined compression of white matter. J Biomech. 2007;40:117–124. 10 [33] Caenen A, Shcherbakova D, Verhegghe B, et al. A versatile and experimentally validated finite ele- ment model to assess the accuracy of shear wave elastography in a bounded viscoelastic medium. IEEE Trans. Ultrason. Ferroelectr., Freq. Control. 2015;62:439–450. [34] Yang X, Cleveland RO. Time domain simulation of nonlinear acoustic beams generated by rectangu- lar pistons with application to harmonic imaging. J. Acoust. Soc. Am.. 2005;117:113–123. [35] Pandey V, Holm S. Linking the fractional derivative and the Lomnitz creep law to non- Newtonian time-varying viscosity. Phys. Rev. E. 2016;94:032606. [36] Fouque JP, Garnier J, Papanicolaou G, Sølna K. Wave propagation and time reversal in randomly layered media. Springer Science & Business Me- dia 2007. [37] Lambert SA, Näsholm SP, Nordsletten D, et al. Bridging three orders of magnitude: Multiple scat- tered waves sense fractal microscopic structures via dispersion. Phys. Rev. Lett.. 2015;115:094301. [38] Podlubny I. Geometric and physical interpreta- tion of fractional integration and fractional differ- entiation. Fract. Calc. Appl. Anal.. 2002:367–386. [39] Bardet JP. A viscoelastic model for the dynamic be- havior of saturated poroelastic soils. J Appl Mech- T ASME. 1992;59:128–135. 11
1111.2756
2
1111
2011-11-14T16:57:54
Comparing ion conductance recordings of synthetic lipid bilayers with cell membranes containing TRP channels
[ "physics.bio-ph" ]
In this article we compare electrical conductance events from single channel recordings of three TRP channel proteins (TRPA1, TRPM2 and TRPM8) expressed in human embryonic kidney cells with channel events recorded on synthetic lipid membranes close to melting transitions. Ion channels from the TRP family are involved in a variety of sensory processes including thermo- and mechano-reception. Synthetic lipid membranes close to phase transitions display channel-like events that respond to stimuli related to changes in intensive thermodynamic variables such as pressure and temperature. TRP channel activity is characterized by typical patterns of current events dependent on the type of protein expressed. Synthetic lipid bilayers show a wide spectrum of electrical phenomena that are considered typical for the activity of protein ion channels. We find unitary currents, burst behavior, flickering, multistep-conductances, and spikes behavior in both preparations. Moreover, we report conductances and lifetimes for lipid channels as described for protein channels. Non-linear and asymmetric current-voltage relationships are seen in both systems. Without further knowledge of the recording conditions, no easy decision can be made whether short current traces originate from a channel protein or from a pure lipid membrane
physics.bio-ph
physics
Comparing ion conductance recordings of synthetic lipid bilayers with cell membranes containing TRP channels Katrine R. Laub1, Katja Witschas2, Andreas Blicher1, Søren B. Madsen1, Andreas Luckhoff2 and Thomas Heimburg1,∗ 1Niels Bohr Institute, University of Copenhagen, Blegdamsvej 17, 2100 Copenhagen Ø, Denmark 2Institute of Physiology, Medical Faculty, RWTH Aachen University, D-52057 Aachen, Germany ABSTRACT In this article we compare electrical conductance events from single channel recordings of three TRP channel proteins (TRPA1, TRPM2 and TRPM8) expressed in human embryonic kidney cells with channel events recorded on synthetic lipid membranes close to melting transitions. Ion channels from the TRP family are involved in a variety of sensory processes including thermo- and mechano-reception. Synthetic lipid membranes close to phase transitions display channel-like events that respond to stimuli related to changes in intensive thermodynamic variables such as pressure and temperature. TRP channel activity is characterized by typical patterns of current events dependent on the type of protein expressed. Synthetic lipid bilayers show a wide spectrum of electrical phenomena that are considered typical for the activity of protein ion channels. We find unitary currents, burst behavior, flickering, multistep-conductances, and spikes behavior in both preparations. Moreover, we report conductances and lifetimes for lipid channels as described for protein channels. Non-linear and asymmetric current-voltage relationships are seen in both systems. Without further knowledge of the recording conditions, no easy decision can be made whether short current traces originate from a channel protein or from a pure lipid membrane. ∗corresponding author, [email protected] -- K. R. Laub and K. Witschas con- tributed equally to this work. Keywords: TRP channels; TRPA1; TRPM2; TRPM8; lipid membrane pores; phase transition Abbreviations: DMPC - 1,2 dimyristoyl-3-sn-phosphatidylcholine; DMPG - 1,2 dimyristoyl-3-sn-phosphatidylglycerol; AITC - allylisothiocyanate ; pdf - probability distribution function; HEK cell - human embryonic kidney cell; TRP channel - transient receptor potential channel; ADPR - adenosine diphos- phate ribose Introduction The cell membrane forms a diffusion barrier for hydrophilic substances, and its integrity is crucial for the viability of living organisms. The established view is that ion channels selectively regu- late the flow of ions across the membrane, permitting the pas- sage of certain ions while excluding others. In addition to their selectivity, protein ion channels have been reported to be functionally characterized by many properties such as con- ductance, voltage dependence, ligand activation, temperature- and mechanosensitivity. The synthetic lipid bilayer has been a model system for the biological membrane since the early days of membrane research. Biological membranes show a large compositional heterogeneity and complexity. They consist of lipid mixtures with varying amounts of unsaturation, different acyl and head groups, and they contain a multitude of imbed- ded or associated proteins. Lipid bilayers display chain-melting transitions. In the melting regime, the physical state of the membranes is influenced by temperature, pressure, pH, drugs including anesthetics and some neurotransmitters, and the pres- ence of proteins (1 -- 3). Electrophysiological membrane models generally assume that the lipid membrane is an insulator and capacitor. This assumption is important because a significant non-specific ion conductance of the lipid membrane itself would be inconsistent with the specific conductances required by current models of biomembranes. However, membranes are not generally insu- lators. Close to their chain melting temperatures, membranes become quite permeable to ions and small molecules (4). Tem- perature changes of only a few degrees have been reported to alter permeation rates for fluorescence markers by several or- ders of magnitude (5 -- 7). More strikingly, electrophysiologi- cal transmembrane current recordings using black lipid mem- branes (BLMs) or patch pipettes reveal quantized conduction events that resemble those from biological membrane prepara- tions (7 -- 10, 10 -- 16) (reviewed in (4)). Antonov and collabo- rators (9, 10) first described conduction events through mem- branes made of synthetic lipids eliminating all sources of cur- rent fluctuations other than those of the lipid membrane itself. Other investigations show that recordings of the conductance of synthetic membranes convincingly demonstrate its quantized nature (eg.,(4, 7, 15 -- 17)). The existence of lipid pores with diameters on the order of 1 nm has long been discussed in the field of electropora- tion (18 -- 20). Several molecular dynamics studies have demon- strated the generation of such lipid pores by voltage (21, 22). The transient generation of lipid pores by large voltage pulses is used in clinical praxis, e.g. to transport cytostatic drugs into cancer cells or to transfect cells with DNA or RNA segments (20, 23). It seems likely that the quantized conductance events observed in synthetic membranes are a phenomenon closely re- lated to electroporation. Nevertheless, the quantized nature of the ion currents across lipid membranes is surprising and is not really understood. It suggests that pores or defects in the lipid membrane have a well-defined size. While we believe that the explanation for fixed pore size remains an open issue, the de- 1 pendence of the pore formation rate on changes in temperature, pressure and other intensive thermodynamic variables is well understood. Its theory is based on the fluctuation-dissipation theorem that treats the couplings of fluctuations in enthalpy, volume, area and other extensive variables with susceptibilities such as heat capacity and compressibility (4). Thus, the oc- currence of the lipid membrane channels responds to changes in temperature, pressure, voltage (4) and (for charged mem- branes) on pH and Ca2+ (24). As a consequence, they are ther- mosensitive, mechanoreceptive, voltage dependent and pH and Ca2+ sensitive. Furthermore, the fluctuation-dissipation theo- rem implies a connection between the magnitude of the fluc- tuations and the fluctuation time scale (3, 25). Fluctuations in the membrane state are especially large close to melting tran- sitions. Therefore, the mean conductance of the membrane is larger and the channel open lifetimes are longer in the tran- sition regime (16). This is particularly important since many biological membranes are in fact found in a state close to a melting transition, e.g., E.coli and bacillus subtilis membranes, lung surfactant (2, 26, 27), and rat central nerve membranes (unpublished data from 2011 by S. B. Madsen and N. V. Olsen, Copenhagen), suggesting that these phenomena might play a role under physiological conditions. In this publication we compare the channel events in syn- thetic lipid membranes with ion channel protein conductances in biomembranes. The transient receptor potential (TRP) chan- nel family has recently attracted considerable interest due to their involvement in sensing processes. Members of this familiy of ion channels have been reported to respond to environmen- tal stimuli such as temperature, membrane tension, pH, osmo- larity, pheromones, and intracellular stimuli such as Ca2+ and phosphatidylinositol signal transduction pathways (28). They may also be involved in detecting the taste sensations of sweet, sour and umami (29, 30). Here, we focus on the activity of three TRP channels, namely TRPA1, TRPM2 and TRPM8. They are considered to form ho- motetrameric proteins with six transmembrane segments (S1- S6) in each subunit with cytosolic N- and C-termini. TRPA1 is a nociceptive channel and polymodal receptor activated by pun- gent or irritant chemicals such as AITC (allyl isothiocyanate) from wasabi, acrolein (in smoke), allicin and diallyl disulfide from garlic (28, 31 -- 33). There is evidence for a cold-sensitivity of TRPA1 (30, 31). Furthermore, it is sensitive to depolarisa- tion and to Ca2+ (32). Pore properties of TRPA1 vary dynam- ically depending on the presence of Ca2+ and agonist stimu- lation (33). The TRPM2 channel is a member of the transient receptor potential melastatin family that was first identified in cancer cells. TRPM2 currents show a linear I-V relationship indicating that this channel is not voltage-dependent. It is ac- tivated by intracellular ADP-ribose and Ca2+ in a synergistic manner (34), by heat, and by hydrogen peroxide, and this points to a role of the channel in oxidative stress signaling cascades (28, 35). A single channel conductance of ∼62 pS has been reported for TRPM2 (28). The closely related TRPM8 channel is a cold receptor (36, 37) and a voltage-gated channel show- ing strong outward rectification. It is activated by substances that cause the sensation of cold such as menthol or menthol derivatives (e.g. WS-12). The "super-cooling" agent icilin is not structurally related to menthol activates TRPM8 in the pres- ence of Ca2+ (38). TRPM8 channel activity is modulated by pH (39) and the presence of polyunsaturated fatty acids and lysophospholipids (40). A single channel conductance of ∼ 81 pS was reported for TRPM8 (28). All three of these channels are selective for cations but show little selectivity for particular cations. Moreover, both TRP channels and lipid membrane channels are influenced by chan- ges in intensive thermodynamical variables such as temperature and membrane tension (lateral pressure). Therefore, it is tempt- ing to compare the characteristic properties of lipid and protein channel activity. In the present paper we discuss the similarities and differences of TRP channel conductance and conduction events due to synthetic lipid membrane pores. Since menthol is an agonist for human TRPM8 and TRPA1, we also study the influence of menthol on lipid phase behavior. The underlying question is whether membrane pores and protein channels are related or synergetic, and whether it is possible that they are governed by the same physical laws. Materials and Methods Synthetic membranes Calorimetry: Heat capacity profiles were recorded using a VP- DSC (MicroCal, Northhampton/MA, USA) with a scan rate of 5◦/h. Electrophysiological recordings on synthetic lipid membra- nes: We used the droplet method, whereby the planar lipid bi- layer membranes is formed on the tip of a patch-clamp glass pipette that has been filled with electrolyte solution (41). The tip is in contact with the surface of a beaker filled with the same electrolyte solution. Lipids are dissolved in a hexane/ethanol mixture and are then brought into contact with the outer sur- face of the glass pipette. When the solution flows down the pipette, a membrane is formed spontaneously at the tip of the pipette. The solvent was allowed to evaporate for at least 30 seconds before the pipette was lowered 2 -- 5 mm below the bath surface. The main advantage of this method is that the resulting membrane is practically solvent free. Pipettes were pulled from 1.5 mm / 0.84 mm (outer diame- ter / inner diameter) borosilicate glass capillaries (#1B150F-3, World Precision Instruments, USA) with a vertical PC-10 puller from Narishige Group, Japan. A two-step pulling procedure was used, where the first pull was 8mm and the heater was set to 80% of the instrument's maximum output. For the second pull, the heating coil was lowered 4mm and the heater setting was reduced to 45%. This produced fairly short and thick pipettes with an hourglass-like taper. For some experiments the pipette was subsequently fire polished using a Narishige MF-900 Mi- 2 croforge. This created pipettes that had a pipette opening in the range from 5 -- 15 µm in diameter. As a general rule, larger openings made it more difficult to create a stable membrane, with 20 µm being the practical upper limit. In most experi- ments it was chosen to use the pipette as-is without fire polish- ing. These pipettes had openings of less than a 1 µm. Unpol- ished pipettes were use in order to minimize the variation be- tween experiments. The electrodes were made of high-purity, chlorinated silver wires, which were frequently re-chlorinated to avoid baseline drift and additional noise. Pipettes were al- ways freshly prepared immediately before use. The buffers in the pipette and in the medium were identical. Current recordings were made using an Axopatch 200B patch clamp amplifier (Axon Instruments Inc., Union City/CA, USA). The pipette and electrodes were mounted on a cooled capacitor feedback integrating headstage amplifier (Headstage CV 203BU, Axon Instruments Inc.). The headstage itself was mounted on a micro-manipulator (model SM1, Luigs and Neu- mann, Germany), allowing for careful and precise control of the pipette position relative to the bath surface. Lastly, the head- stage and micro-manipulators were wrapped in a finely meshed metal cloth that acted as a Faraday cage. Data traces were recorded with Clampex 9.2 software (Axon Instruments) using the Whole Cell (headstage gain, β = 1), voltage clamp mode. The sampling frequency was either 10kHz or 20kHz, and the signal was filtered by the patch clamp amplifier's analog 4-pole lowpass Bessel filter with the cut-off (-3dB) frequency set to 2kHz. Cell membranes Molecular biology and cell culture: The cDNAs of human TRPM2, TRPM8 and TRPA1 were subcloned into pIRES- hrGFP-2a vectors (Stratagene, USA). Wild-type channels were stably expressed in HEK-293 cells (ATCC,USA) as described previously (42). Solutions: Standard bath solution (BP1) contained 140mM NaCl, 1.2mM MgCl2 , 1.2mM CaCl2, 5mM KCl, 10mM HE- PES, pH 7.4 (NaOH). Pipette solution contained 145mM ce- sium glutamate, 8mM NaCl, 2mM MgCl2, 10mM HEPES, pH 7.2 (CsOH) and the Ca2+ concentration was adjusted to ei- ther <10 nM (P10, 10 mM Cs-EGTA), or to 1 µM (P10*) as calculated with the MAXC-program (0.886mM CaCl2 and 1 mM Cs-EGTA). For the stimulation of TRPM2 currents in the inside-out configuration, ADPR (Sigma, 100 mM stock solu- tion in distilled water) was added to the intracellular (bath) so- lution containing 1 µM Ca2+ , yielding a final concentration of 0.1 mM ADPR. Alternatively, TRPM2 currents were evoked in cell- attached configuration by application of hydrogen per- oxide (Merck, 30% stock solution) to the bath solution, and after the appearance of single channel openings, a small mem- brane patch was excised by lifting up the pipette to reach inside- out configuration. TRPM8 currents were induced with icilin (Cayman Chemical, 10 mM stock solution in DMSO) or WS- 12 (Tocris, 30 mM stock solution in DMSO) by application to the bath (final concentrations as indicated in the experiments). For stimulation of TRPA1, allylisothiocyanate (Sigma, 30 mM stock solution in DMSO) was added directly to the bath solu- tion. All chemicals were tested on native and vector-transfected HEK-293 cells and did not evoke single channel currents in the given range of concentrations without the presence of a TRP channel protein. Electrophysiology: Single-channel recordings were routinely performed with pipettes made of borosilicate glass (outer diam- eter 1.8 mm / inner diameter 1.08 mm / length 75 mm, Hilgen- berg, Germany). Alternatively, microelectrodes were made from borosilicate glass with slightly different specifications (outer diameter 1.5 mm / inner diameter 0.86 mm / length 100 mm, Harvard Apparatus, USA) to ensure that biophysical proper- ties of ion channels were not dependent on the type of pipette used. Pipettes were fabricated in a two-step procedure and fire- polished with a programmable DMZ-Universal puller (Zeitz- Instrumente GmbH, Germany). Pipettes had a tip diameter in the range of 1 µm and resistances between 5 and 7 MΩ. To reduce thermal noise, pipette tips were coated with dental wax (Moyco Industries Inc., USA). Microelectrodes were used on the day of preparation and stored in a container for the preven- tion of dust deposits. Electrophysiological signals were recorded with an Axopatch 200B amplifier in combination with Digidata 1440A AD/DA converter controlled by the pClamp10 software suite (Axon CNS, USA). The CV 203 BU headstage (Axon CNS, USA) was mounted on a micro-manipulator (model SM1, Luigs and Neu- mann, Germany) inside a Faraday cage. Transfected cells were visually identified with an Axiovert 200 inverted microscope (Carl Zeiss MicroImaging GmbH, Germany) and a blue LED lamp (model CREE XP-E, wavelength λ = 460 nm) serving as a light source for excitation of green fluorescence from GFP (green fluorescent protein). The Axopatch 200B amplifier was run in resistive whole-cell voltage-clamp mode (β=1) to avoid contamination of traces with reset glitches due to capacitor dis- charge. The output gain (α) was set to ×100 when recording single-channel currents. A gap-free acquisition mode was used with a sample rate of 10 or 20 kHz and analogous filtering at 2 or 5 kHz performed with a build-in 4-pole Bessel filter (-3 db). All experiments were done at room temperature (21◦C- 23◦C). Voltage signals were not corrected for liquid junction potentials. In order to facilitate comparison of lipid and protein traces, current and voltage signals are depicted as recorded with Clampex software from pClamp10 program suite. It is com- mon in physiology to invert single-channel currents recorded in the cell-attached and inside-out configuration so that the inward movement of ions is represented as downward deflection. The reason for this lies in the convention that the net movement of positive ions in the direction of the outer to the inner membrane is, by definition, an inward current, but as positive ions are leav- ing the headstage and patch pipette this would be recorded as positive (or upward) current (43, 44). The patch-clamp com- 3 Figure 1: Left: Heat capacity profiles of multilayered vesicles of a DMPC:DLPC=10:1 mixture in 150mM NaCl (1mM EDTA, 2mM HEPES, pH 7.4) and DMPG:DMPC=6:4 vesicles in 100 mM NaCl (1mM EDTA, 10mM HEPES, pH 7.4). The DMPC:DLPC mixture displays a pronounced maximum at 22.1 ◦C with wings towards higher and lower temperatures. The DMPG:DMPC mixture displays a peak at 23.2 ◦C and a broad pretransition peak at 16.7 ◦C. Center: DMPC membranes in the absence and presence of menthol (150mM KCl, 3mM EDTA, 3mM Hepes, pH 7.4). Different molar ratios are shown: 10mM DMPC in the absence of menthol (top), 10mM DMPC in the presence of 0.5mM menthol and 2.5mM DMPC in the presence of 0.5mM menthol in the buffer (before mixing). One recognizes that the presence of menthol shifts melting profiles towards lower temperatures and broadens the profiles. The bottom two traces have been amplified by a factor of 2 and baselines been shifted for better visibility. Right: 10mM DMPC membranes in the absence (top) and presence of 2mM and 10 mM AITC in the buffer (before mixing). As in the menthol experiments, the presence of AITC broadens and shifts the heat capacity profiles towards lower temperatures. The bottom two traces have been amplified by a factor of 2 and 4, respectively. mand voltage is positive if it increases the potential inside the micropipette. In physiology, it is common usage to report the transmembrane potential (Vm), i.e. the potential at the inside of the cell minus the potential at the outside instead of pipette po- tential. In cell-attached and inside-out configuration where the pipette is connected to the outside of the membrane, a positive command voltage causes the transmembrane potential to be- come more negative, therefore it is hyperpolarizing (43). In the inside-out and cell-attached configuration, the transmembrane potential is inversely proportional to the command potential, in cell-attached configuration Vm is additionally shifted by the resting membrane potential of the cell (43, 44). Results The results section contains data regarding the thermodynamic properties, channel-like ion conductance events, and current- voltage relationships of synthetic lipid membranes. We then compare conductance events from the synthetic membranes with biological membranes containing various transient receptor po- tential (TRP) ion channels. Fig. 1 shows the heat capacity profiles of two lipid mix- tures, DMPC:DLPC=10:1 and DMPG:DMPC = 6:4 in a 150 mM NaCl or KCl buffer. We have used these two mixtures for the conductance recordings that are documented below. Both mixtures display a maximum close to room temperature where our conductance studies were performed. In the transition, the compressibility of the membrane is at maximum (26, 45). This effect results in a maximum of the membrane permeability close to the transition temperature (5, 46) and the maximum proba- bility of finding lipid pores or lipid ion channel events (7, 16). This effect is central to our description in the following fig- ures. Menthol is an agonist of TRP channel activity. For this reason we investigated the influence of menthol on lipid mem- brane phase behavior. The center panel of Fig. 1 shows the melting profile of DMPC membranes with and without men- thol. The right hand panel shows the same membrane in the presence of AITC. AITC is an activator of the TRPA1 chan- nel. It can be seen that small amounts of menthol (500 µM) and AITC (2 mM) have a significant effect on the melting pro- file both with respect to peak position and width. In particular, menthol lowers the melting peak and broadens the profile sig- nificantly. The effect of both menthol and AITC is quite similar to that of a general anesthetic molecule on lipid membranes (2). The presence of both menthol and AITC should result in mea- surable effects on lipid membrane permeability for ions if it is measured close to the transition temperature. Fig. 2A shows a channel recording of a DMPC:DLPC=10:1 mixture in 150 mM KCl at 50 mV at 30◦C. Here and in all following experiments on synthetic membranes, the membrane was deposited on a patch pipette tip using a method described 4 Figure 2: Patch clamp recording of a synthetic lipid membrane (DMPC:DLPC=10:1 mol:mol, T=30◦C, 150mM KCl, 1mM EDTA, 2mM HEPES, pH 7.4) at U=+50 mV. Panel A: 8 representative 2.5 second segments out of a 30 minute recording. Panel B: Current histogram for the quantized steps of the left hand panel (closed channel state set to zero). Panel C: current-voltage relation for the currents of the membrane shown in the left hand panel (only positive voltages were recorded). The channel conductance is about 330pS. Panel D: Probability distribution function (pdf) for the open times of the channels in panel A. It is well fitted by an biexponential profile in the range up to 100ms open time. Panel E: The double-logarithmic representation of the data in panel D shows that for long open times the pdf is better described by a power law with a critical exponent close to -2. by (41). This trace was stable over the complete recording time (about 30 minutes). The current histogram shows two well- defined peaks for baseline and single channel opening with an amplitude of about 18 pA at 50 mV (Fig. 2B). Also shown is the linear current-voltage relation of the single channel currents with the baseline subtracted. The single channel conductance resulted in a conductance of about 330 pS (Fig. 2C). No single- channel events were detected for negative voltages. Thus, the single-channel current-voltage relationship is only shown for positive voltages. Panels D and E show the probability distri- bution function (pdf) for the channel open times in linear and double-logarithmic representation. In the linear representation (Fig. 2D), it seems as if the pdf is well fitted by a bi-exponential curve with characteristic open lifetimes of 2.5 ms and 14.4 ms (the fit in the range up to 100 ms is shown as a dashed line that is hardly visible because it overlaps the pdf-profile). In the double-logarithmic representation, however, it is obvious that the bi-exponential fit underestimates the long open time proba- bilities, and the long open times are better described by a power law with a fractal exponent between 1.5 and 2 (Fig. 2E). Simi- lar power-law behavior was found for the closed time pdf (data not shown). As single channel events we consider stepwise conductance changes on top of a baseline. While the baseline current is usu- ally believed to be related to leaks introduced by bad seals, this is not what is typically found for the synthetic membranes. We have consistently found that the background conductance re- flects the thermodynamic properties of the membrane. For in- stance, the conductance of the overall membrane reflects the heat capacity profile (16). The conductance of single channels is strictly linear in the experiments presented here but chan- nel open likelihoods increase with voltage (15) leading to a non-linear mean conductance. Within a certain voltage range, the current-voltage relation of the baseline is also linear. It is possible that the baseline current consists in part of unre- solved conduction events that are of shorter duration than the recording resolution and the time constant of the low pass fil- In Montal-Muller type black lipid membranes, we have ter. often found non-linear but symmetric current voltage relation- ships. This is to be expected for a fully symmetric bilayer. An example is given in (7, 15) for a DOPC:DPPC= 2:1 mixture (150 mM KCl, pH = 6.5 at T =19◦C). In experiments on patch pipettes we often find non-symme- tric current-voltage profiles. In Fig. 3 the average conductance of a membrane with identical composition as in Fig. 2 (in- cluding the baseline conductance) is shown as a function of voltage. We find a nonlinear current-voltage relation with out- ward rectification shown here over an interval from −150 mV 5 Figure 4: Single channel recordings from synthetic lipid mem- branes and from two TRP channels with the corresponding his- tograms. Top: DMPC:DLPC=10:1, 150mM KCl, T=20.6◦C and U = +200 mV with a current of about 3.5pA. Center: hTRPM2 channel activated by H2O2 at U= -60 mV (pipette voltage, inside-out configuration, pipette solution P10∗, bath solution BP1 + 5mM H2O2). Bottom: hTRPM8 channel acti- vated by 5 µM icilin at a pipette voltage of U=-60 mV (inside- out configuration, pipette solution BP1, bath solution P10∗ + 5 µM icilin, low pass filter 5 kHz). age) indicating that the membrane is not symmetric. One way to describe rectified current-voltage relationships is the transi- tion state (Eyring) model (48), which assumes a free energy barrier for the ions, ∆G0, inside of the membrane (49). The relative position of this barrier within the membrane is given by a parameter δ, with 0 ≤ δ ≤ 1. For a symmetric membrane Figure 5: Conduction bursts in pure synthetic and in a cell membrane with TRP channels, and the corresponding current histograms. Top: DMPC:DLPC=10:1 (150mM KCl, T=20.6◦C and U = + 200 mV). Bottom: Burst from hTRPM8 expressed in HEK 293 cells at a pipette potential of U=-60 mV after addition of the menthol derivative WS-12. Cell-attached configuration, pipette solution P10, bath solution BP1 + 7.5 µM WS-12. Figure 3: The current-voltage relationship of the total mem- brane (DMPC:DLPC=10:1 mol:mol, T=30◦C, 150mM NaCl, 1mM EDTA, 2mM HEPES, pH 7.4) shows an outward rectifica- tion that is probably due to asymmetries of the patch setup (e.g. induced by slight suction or by pipette shape). The dotted line is a guide to the eye. The solid line is a fit for the Eyring transition state model using a barrier position of δ = 0.86 and equal ion concentration on both sides of the membrane. The fit was from -130 mV to 70 mV. The insert shows the outward rectified pro- files of the TRPM8 channel at two temperatures (data adapted from (47)). to +150 mV. For comparison, we include an insert showing the current-voltage profiles of the TRPM8 channel at two temper- atures which were adapted from (47). The I/V relationships from the synthetic membrane and the biological channels show striking similarity regarding their strong outward rectification properties. Since the TRPM8 channel is temperature sensitive, one finds different profiles at different temperatures. Because the mean conductance of lipid membranes also displays tem- perature dependence (16), we expect temperature sensitivity also for the current-voltage relationships. This issue has not been studied systematically here. It should be added that we always find outward rectified curves with a significant varia- tion in magnitudes found in individual experiments. We have never seen an inward-rectified profile for any synthetic mem- brane. This is probably a consequence of the asymmetry of our experimental setup. We cannot exclude that conditions exist for which inward rectification indeed exists. In the present manuscript we have exclusively used patch pipettes to monitor currents through synthetic membranes. Pi- pette apertures are much smaller than the holes in the teflon film of BLMs. Fig. 3 shows a conductivity profile with out- ward rectification (conductivity increases with increasing volt- 6 δ is equal to 0.5. The electrical potential is assumed to change linearly across the membrane, which means that a uniform di- electric constant within the membrane is assumed. In the pres- ence of an applied voltage, the height of the kinetic barrier has different values, and one finds a current-voltage relationship of the following form: I = zF βk0 [C]in exp (cid:104) (cid:19) (cid:18) δF U z − (1 − δ)F U z RT RT (cid:18) −[C]out exp (cid:19)(cid:105) (1) where z is the charge of the respective ion, F is Faraday's con- stant, βk0 is a rate constant reflecting the height of the barrier and the solubility of ions in the membrane. [C]in and [C]out are the ion concentrations inside and outside of the membrane. For equal ion concentration outside and inside ([C]in = [C]out), the current I is zero at U = 0, as measured in Fig. 3. For a monovalent salt (z = 1) eq. 1 simplifies to (cid:20) (cid:18) δF U (cid:19) RT I ∝ exp − exp (cid:18) − (1 − δ)F U (cid:19)(cid:21) RT (2) This model describes the movement of gating particles in the Hodgkin-Huxley gate model for channel proteins (49). It has also been used to describe the voltage-sensitive conductivity of membrane patches containing TRP channels (50). If the barrier is placed at a position δ = 0.5 (the center of the membrane), the current-voltage relationship is symmetric under sign changes of the voltage. The transition state model is expected to work only if the height of the barrier is greater than the free energy of the ion in the field (∆G0 ≥ δF U z and ∆G0 ≥ (1 − δ)F U z, see the derivation of eq. 1 in (49)), indicating that even under idealized conditions the relationship can only be used within a certain voltage regime. In Fig. 3 we show a fit of the total membrane conductance over the voltage range of −130 mV to +70 mV using a barrier position of δ = 0.86 that describes the conductance of the membrane reasonably well. The position of the barrier at δ = 0.86 indicates that the membrane is asym- metric. Although we cannot offer a conclusive explanation of this result, a number of possibilities present themselves. One is the application of suction in our patch experiments, which can lead to a membrane curvature that could contribute to the current-voltage asymmetry. The patch pipette itself introduces an asymmetry since it contacts only one side of the membrane. It is known that the contact of membranes with glass surfaces influences their phase behavior (51). This effect is of little rel- evance in the whole-cell patch-clamp configuration which was employed to record the current-voltage relationship of TRPM8 depicted in the insert of Fig. 3. In whole-cell configuration, the membrane is ruptured under the patch providing access to the intracellular space of the cell. The pipette contacts only a small area of the membrane surface while recording the current flowing over the membrane of the entire cell. Figure 6: Flickering of a synthetic and a cell membrane con- taining TRP channels. Top: DMPC:DLPC=10:1 (150mM KCl, T=20.8◦C and U = +350 mV). Bottom: hTRPA1 in HEK293 cell in cell attached configuration at a pipette potential of U= +60 mV. Pipette solution BP1, bath solution P10. Comparison of current recordings from synthetic lipid mem- branes and biological preparations containing TRP chan- nels: The current recordings from artificial membranes showed various behaviors including single channel openings, conduc- tance bursts, and flickering. For some samples and tempera- tures we found several of these phenomena in the same record- ings. The occurrence of these lipid channel events differed somewhat for different pipettes and different conditions. In contrast, TRP channel activity was independent on the type of pipette used and reproducible under experimental conditions. In the following we compare some typical findings selected from artificial membranes with recordings from HEK293 cells containing TRP channels. All voltages are given as pipette po- tentials. Fig. 4 (top) shows single channel events in a DMPC: DLPC=10:1 membrane with single steps of about 3.3 pA at a voltage of +200 mV (corresponding to ∼17 pS) at a temper- ature of 20.6◦ C. This experiment shows that one finds sin- gle current events not only at 330 pS (Fig. 2 recorded at 30 ◦C) but also much smaller values depending on conditions. We compare this with the current from a HEK293 cell with over-expressed hTRPM2 channels activated by 5 mM H2O2 at a voltage of −60 mV in inside-out configuration (Fig. 4, cen- ter). We found single steps of about 3.1 pA corresponding to a conductance of 52 pS. Fig. 4 (bottom) shows the hTRPM8 channel in a HEK293 cell activated by 5 µM icilin at a voltage of −60 mV in inside-out configuration. Single-channel open- ings have a current amplitude of 5.1 pA corresponding to a conductance of 85 pS. Time axis scaling was adapted for the three traces to depict channel openings with a similar appear- ance. While the openings of TRPM2 are very long compared to TRPM8, both TRP channels show relatively short openings compared to the above trace from the synthetic membrane. Pe- riods of repetitive channel activity separated from each other by long closures are called "bursts". The conductance events 7 Figure 7: Spikes from synthetic membranes and from recombi- nantly expressed TRP channels. Top: Spikes in a recording from a DMPC:DLPC=10:1 mol:mol membrane (T=23◦C, 150mM KCl, 1mM EDTA, 2mM HEPES, pH 7.4) at U=+150 mV. Bot- tom: Spikes from hTRPM8 in HEK293 cells in cell-attached configuration at a pipette potential of U = +60 mV. Pipette so- lution P10, bath solution BP1. Figure 8: Multistep conductance in synthetic membranes and in a cell transfected with TRP channels. Top: DMPC:DLPC =10:1 mol:mol membrane (T=20.6◦C, 150mM KCl, 1mM EDTA, 2mM HEPES, pH 7.4) at U=+200 mV. Bottom: hTRPA1 in HEK293 cells after addition of AITC in inside-out configura- tion at U= -60 mV (pipette voltage). Bath solution P10*, pipette solution BP1, post-recording low pass filtering with 500 Hz. In both traces the lowest conductance level has been set to zero. of both synthetic membranes and biomembranes often occur in such bursts. Fig. 5 (top) shows a burst in a synthetic DMPC: DLPC=10:1 membrane recorded at 200 mV and T=20.6◦ C. The histogram indicates approximately evenly spaced conduc- tance levels with an individual step size of 6.8 pA (34 pS). The total burst lasts for about 7 seconds. Fig. 5 (bottom) shows a burst of an HEK293 cell membrane containing the hTRPM8 channel activated by 7.5 µM WS-12 at a voltage of −60 mV in cell attached configuration. The bursting behavior of TRPM8 occured for several minutes after the activator was given, only a short section is shown in Fig. 5. The histogram consists of approximately equally spaced steps of -3.5 pA (58 pS). Thus, the bursts of the synthetic membrane and the cell membrane display very similar characteristics both in the conductance of individual steps and in the appearance of the current histogram. Again, the traces by themselves appear nearly indistinguishable in both preparations. Flickering resembles a burst with a single conductance step. Examples of this phenomenon are depicted in (Fig. 6). The top trace (Fig. 6, top) shows a flickering event from a syn- thetic DMPC:DLPC=10:1 membrane, the lower trace (6, bot- tom) shows the hTRPA1 channel in a HEK293 cell membrane recorded in cell-attached configuration. The flickering event in the artificial membrane occured at 350 mV and 20.8◦C and lasts for about 6 seconds. The step size is 10.2 pA corresponding to a single channel conductance of 23 pS. The current trace of a cell preparation expressing TRPA1 was recorded at 60 mV and displays a unitary current of 7.4 pA corresponding to a single- channel conductance of 123 pS. Except for small differences in detail, the overall appearance of the traces does not allow us to distinguish easily between the artificial membrane events from the cell preparation on the basis of the traces alone. In Fig. 7 we show channel openings which are so brief that they are mostly not fully resolved, leading to a pattern of spike-like appearance with variable single-channel ampli- tudes truncated by low-pass filtering. The top trace is for a DMPC:DLPC=10:1 membrane recorded at 150 mV and 23◦ C and shows spikes with an amplitude of 4 -- 5 pA corresponding to about 30 pS. The bottom trace is for a cell-attached recording of an HEK293 cell containing the hTRPM8 channel at U=60 mV. Here, the spike has an amplitude of about 2 pA corresponding to about 33 pS. The limited time resolution of the recording sys- tem may have lead to an underestimation of the mean current related to the spikes. Fig. 8 shows a comparison of multi-step conductances in synthetic membranes and in a biological preparation containing the hTRPA1 channel activated by 30 µM allyl isothiocyanate (AITC). The synthetic membrane display at least four clearly distinguishable and equally spaced conduction levels with a conductance of 34 pS. The histogram of the biological prepara- tion shows four visible steps. Since the two small peaks of the lower histogram are diffuse, two interpretations are feasible: In accordance with the upper histogram, four identical steps may be identified with a mean conductance per step of approxi- mately 28 pS. Alternatively, this pattern of channel activity may be interpreted as long-lived openings of endogenous channels from HEK293 cells (32) or a subconductance state of TRPA1 with a small amplitude (2 pA) and overlying larger amplitudes of approximately 5 pA stemming from the main conductance state of TRPA1 with flickery openings (52). The overall ap- pearance of the traces, the life times, and the individual con- ductances are once again very similar in both systems. While the appearance of the traces is different from the conduction bursts in Fig. 5, the conduction histograms display some simi- larities both for synthetic and cell membranes. 8 brane tension) can always lead to conductance traces in syn- thetic membranes that are indistinguishable from recordings of biological preparations containing particular proteins. The experimental conditions are not necessarily the same but are comparable to those used in biomembrane experiments. This similarity is sufficiently pronounced that an inspection of short traces may fail to identify whether recordings are from syn- thetic membranes or from cells. Protein ion channel activity in patch clamp experiments was characterized by typical and repetitive current events which we- re visually distinctive among different members of the TRP channel family. Those fingerprint openings from TRP channels with typical conductance and lifetime were absent under con- trol conditions in patches of native HEK293 cells or of vector- transfected cells. In many experiments on pure lipid bilay- ers, strikingly similar traces were observed, although synthetic membranes did not exhibit the fingerprint-like behavior of chan- nels with their highly reproducible and predictable reactions. Lipid membranes displayed various responses and the condi- tions for any specific response could not be controlled exactly. We selected several types of events found in sections of cur- rent recordings from artificial membranes for comparison with representative traces of TRP channels pointing phenomenolog- ically at their striking similarities. We found similar single- channel events, multistep conductance, conduction bursts, flick- ering, conduction spikes and staircase behavior in synthetic membranes and in HEK293 cells containing three different TRP channels. Conductances and current histograms as well as life- time distributions are all similar. Further, we found asymmet- ric non-linear current-voltage relationships in synthetic mem- branes that could be described by an Eyring-type kinetic barrier model that has also been used to describe the current-voltage re- lationship of TRP channels (50). The question arises whether, in spite of their similarity, channel events in the lipid membrane and protein channels can be considered as completely indepen- dent. Some findings in the literature suggest that these similar- ities are not merely accidental and point at a common origin of the conduction events in the two systems. Seeger et al. (53) found that both the mean conductance and the open lifetimes of the KcsA potassium channel exactly follow the heat capac- ity profile of the membrane into which the protein is reconsti- tuted. The protein shows precisely the behavior expected for its host lipid membrane, indicating that lipid membrane properties dominate the experimental characteristics of the system. It was suggested that physical properties of the lipid bilayer influence ion channel activity via a fine-tuning of protein conformation, but it was also pointed out that the lipid membrane itself dis- plays a similar behavior (53). It was also found that the mean conductance is proportional to the protein concentration. Fur- ther, conduction events could still be blocked by tetraethyl am- monium (TEA), which is a KcsA blocker. Thus, the conduction events are clearly correlated with the presence of the proteins. A similar correlation of channel activity with lipid membrane phase behavior was found for the sarcoplasmic reticulum cal- Figure 9: "Stairs" (subsequent opening of several channels) in synthetic membranes and in a biological membrane with TRP channels. Top: DMPG:DMPC=60:40 mol:mol membrane (T=23.8◦C, 50mM KCl, 1mM EDTA, 10mM HEPES, pH 7.4) at U=+80 mV. Bottom: hTRPM2 channel in HEK293 cells ac- tivated with ADP-ribose at a pipette potential of U= -60 mV in inside-out configuration. Bath solution P10∗ + 0.1 mM ADP- ribose, pipette solution BP1. In practically all synthetic lipid membrane preparations we find channel-like events under appropriate conditions. In par- ticular, when voltage is increased the membranes eventually break at a threshold voltage, probably due to the growth of a very large pore. Below this threshold one typically finds quan- tized channel events. The threshold voltage is different in dif- ferent preparations and depends on temperature, pipette suc- tion, and probably on the properties of the pipette itself. In- creasing the voltage to threshold values leads to rupture visi- ble by a stepwise increase in membrane conductivity as can be seen in Fig. 9 (top) for a synthetic DMPG:DMPC=6:4 mix- ture. This onset of the rupture process can be reversed if the voltage is lowered. Biological preparations can also display such staircase-like behavior as shown in Fig. 9 (bottom) for the TRPM2 channel activated by 0.1mM ADP-ribose in the pres- ence of 1 µM Ca2+. In contrast to membrane rupture, which is mostly an irreversible process, activation of TRPM2 by ADPR is reversible as is obvious from the current going downstairs in a stepwise fashion until finally the baseline is reached (not shown). Discussion In this paper we have compared conductance traces from bi- ological (HEK293) cells over-expressing TRP channels with recordings from synthetic lipid bilayers. We have demonstrated that suitable adjustments of temperature and voltage (or mem- 9 cium channel (54), where the highest conduction activity and the longest open times were found close to the phase bound- aries of the host lipid matrix. These findings suggest a strong correlation and possibly mutual modulation of the lipid proper- ties with protein channel activity. While the importance of protein channels and receptors is widely accepted, the finding of quantized conductance events in synthetic membranes is striking but little known (e.g., (7 -- 10, 14 -- 16)). Such events occur more frequently close to the melting transition of the lipid membranes. Mean conductances and lifetimes are in agreement with the fluctuation-dissipation theorem that defines the coupling between the amplitude of area fluctuations and the lifetime of the fluctuations (4). The mean open and closed lifetimes of a lipid membrane channel is closely related to the fluctuation lifetime (16). The conse- quences for lipid membranes can be summarized in the follow- ing manner: In the vicinity of the lipid melting transition, area fluctuations and the likelihood of finding pores (or channels) are maximal. Simultaneously, the fluctuation relaxation time is at maximum as is the mean lifetime of open channels. Differ- ent phenomenological behavior as shown in Figs. 4-9 probably reflects the properties of different points in the phase diagram of the lipid mixtures. Pores in lipid membranes have been discussed for more than 25 years (5, 19, 21, 22, 46). Elastic theory suggests the existence of stable pores of a diameter of approximately 1 nm. This is close to the suggested size of aqueous pores in ion chan- nel proteins. The experimentally measured pores, i.e., channel events in synthetic membranes, have a diameter similar to that suggested by theory. (7) aqueous pores of 0.7 nm diameter were described, close to the theoretical value and close to the aqueous pore sizes suggested for protein chan- nels. In Blicher et al. In our synthetic membrane experiments we typically ob- serve coherent trends. For instance, it is possible to induce channel events in lipid membranes if one increases the voltage above some threshold value. This threshold probably varies with temperature. The same is true for increasing pipette suc- tion. Further, channel events are more likely in the melting regime. However, the details of these phenomena varied from experiment to experiment. We believe that this is partially due to strong influences of the patch pipette on the phase behavior of lipid membranes. It is known that the contact of membranes with glass or mica can shift the lipid melting transition by as much as to 5 degrees towards higher temperatures (51). For this reason, the melting behavior of our synthetic membrane patches may be influenced by the patch pipette and vary as a function of the exact surface features of the pipette perimeter. Further, typical domain sizes of lipid mixtures in the transi- tion regime can be several micrometers large (55). With pipette diameters on the order of 1µm (smaller than some domains) one may therefore observe considerable variation in the rela- tive amounts of fluid and gel domains in different experiments. We have found more reproducible patterns in the larger mem- brane patches in BLM experiments or in macroscopic perme- ation events that do not involve interfaces (7, 15). One of the most obvious differences of lipid membrane channels and bi- ological protein preparations is that in the latter case the con- ductances seem to be more reproducible and stable. Thus, cells transfected with particular channel proteins display characteris- tic electrical properties whereas lipid membranes do not exhibit such fingerprint-like features. In Fig. 2E we show that the open time pdf displays power law behavior. The same was found to be true for the closed time pdf (data not shown). Recently we have demonstrated that this applies for other synthetic membranes as well (17). Power law behavior implies that very long closed times occur with a much higher frequency than in stochastic Markovian kinetics. It leads to the possibility of long silent periods between burst of activity. For protein channels this fractal kinetics has been long discussed (56 -- 59) as an alternative to multi-state Marko- vian kinetics (60). In this context it is interesting to note that power law behavior is also a natural consequence of critical be- havior of membranes close to transitions (61). Fig. 5 shows that the bursts from TRPM8 in HEK293 cells display very similar phenomenological behavior compared to synthetic membranes close to phase transitions both in respect to channel amplitudes, lifetimes and conduction histograms. This is also true for flick- ering activity as shown in Fig. 6. Obviously, the temporal pat- terns of both classes of events obey similar kinetics. Channels in synthetic lipid membranes can be influenced by drugs if these drugs affect the melting behavior of the mem- brane. For instance, anesthetics lower and broaden the melt- ing transition of membranes (2) and thereby have been shown to be able to "block" lipid membrane channel events without binding to macromolecular receptors (7). Many other drugs af- fect the melting behavior and permeability of lipid membranes, e.g., the pesticide lindane (6) or the neurotransmitter seroto- nine (3). We have also shown that the anesthetic octanol and serotonine can influence the fluctuation lifetimes of lipid mem- branes and thereby probably the lipid membrane channel life- times. This implies that the response of channels to drugs is not necessarily the consequence of specific binding of the drugs to a receptor. This result could rather be a more general con- sequence of thermodynamic changes in the surrounding lipid matrix. Here we have shown that menthol, an agonist of hu- man TRPM8 and TRPA1 channels, and AITC, an agonist of the TRPA1 channel are also able to influence the state of the lipid membrane (Fig. 1). In particular, both menthol and AITC lower and broaden the cooperative melting events of membranes in a manner very similar to general anesthetics. This means that menthol and AITC partition in the lipid membrane and are much more soluble in the fluid phase than in the gel phase. In fact, menthol has been compared to general anesthetics (e.g., propofol) (62). Interestingly, no specific binding site of men- thol is known for TRP channels, in contrast to AITC in the case of TRPA1. For human TRPM8, the amino acids implied in menthol sensitivity are conserved in menthol-insensitive 10 TRPM2 (63, 64). This would be in agreement with the idea that the effects of menthol on TRP channels are related to its in- fluence on the host lipid matrix. Above the melting transition, menthol should act as an antagonist of lipid channel formation (because the transition is shifted away from experimental tem- perature), and below the transition as an agonist (because the transition is shifted towards experimental temperature). Pre- cisely this effect has been observed for general anesthetics such as octanol (7). Interestingly, menthol was reported to have a bi- modal action on mouse TRPA1 (65). At high concentrations of menthol an inhibitory effect was found while low concen- trations lead to channel activation. In contrast, human TRPA1 is exclusively activated by menthol (66). The authors of this study note that there is no direct evidence that menthol specifi- cally binds to TRPA1, instead the possibility of indirect effects of menthol was considered, e.g. involving modifications of the lipid bilayer. Since effects of menthol on the physical proper- ties of model membranes are better characterized than its bind- ing to TRP channels, it may be speculated that these effects are governed by interaction at the membrane interface. The same may apply for the sensitivity to temperature, a hallmark of this family of thermosensitive channels. We have shown that many biological membranes display transitions slightly below body temperature, including E. coli membranes, bacillus subtilis membranes (2, 27), bovine lung surfactant (26, 27) and the membranes from the central nerve of rat brains (S. B. Madsen and N. V. Olsen, recent unpublished results). Fur- ther, we have shown that channel conductance in synthetic lipid membranes displays a strong temperature dependence close to the melting transition of the membranes (7, 16). Transition in biomembranes are found about 10 degrees below physiologi- cal temperature (27) and display a half width of order 10 de- grees, corresponding to a van't Hoff enthalpy ∆H of about 300 kJ/mol and an entropy ∆S of about 1000 J/mol K. Sim- ilar values have been found for the temperature activation of TRP channels. Talavera et al. reported activation enthalpies on the order of 200 kJ/mol for TRPM4, TRPM5, TRPM8 and TRPV1 channels (29). Further, they report ∆S ≈ 500 J/mol K for several TRP channels. While the origin of the very large ∆H and ∆S for the TRP channels remains mysterious, the or- der of magnitude is just in the range of the melting transition of the biological membrane. For this reason, it has been specu- lated in the past if the temperature dependence of TRP channels may originate from transitions in the surrounding membrane (67). Conclusion In the present study, we have explored the possibility that lipid bilayers, in addition to their role as electrical insulators and solvents for membrane proteins, provide relevant ion perme- abilities as well as changes of permeabilities, by themselves and without the presence of channel-forming proteins. Indeed, lipid membranes exhibited non-linear and asymmetric current- voltage profiles in pipette experiments. We document a wide spectrum of electrical phenomena in synthetic membranes such as bursts, spikes, flickering and multi-step openings that are normally considered typical for the activity of protein ion chan- nels. Furthermore, we show the similarity of current events from lipid bilayers with single-channel recordings of TRP chan- nels. Thus, electrical properties of lipid bilayers may contribute to membrane excitability generated by protein ion channels. On the other hand, electrical properties of pure lipid bilayers also display clear differences as compared to those of cell mem- branes containing channel proteins, such as the different regu- lation by specific ligands. Importantly, the observed electrical phenomena in synthetic membranes lack stability that would warrant reproducible responses in the presence of slightly chang- ing conditions. This is illustrated by the problems to obtain strictly reproducible electrical events with tip-dipping using glass pipettes on pure lipid bilayers, in contrast to the single-channel recordings with similar pipettes on cell membranes. The response of lipid membranes to electrical stimuli demon- strated in this study may broaden our view how ion channels in biological membranes are regulated and how channel proteins and their lipid matrix may cooperate in signal transduction in excitable cells. Acknowledgments: We thank Prof. Andrew D. Jackson for a careful reading of the manuscript and valuable comments. References 1. Heimburg, T., 2007. Thermal biophysics of membranes. Wiley VCH, Berlin, Germany. 2. Heimburg, T., and A. D. Jackson. 2007. The thermodynamics of general anesthesia. Biophys. J. 92:3159 -- 3165. 3. Seeger, H. M., M. L. Gudmundsson, and T. Heimburg. 2007. How anesthetics, neurotransmitters, and antibiotics influence the relaxation pro- cesses in lipid membranes. J. Phys. Chem. B 111:13858 -- 13866. 4. Heimburg, T. 2010. Lipid ion channels. Biophys. Chem. 150:2 -- 22. 5. Papahadjopoulos, D., K. Jacobson, S. Nir, and T. Isac. 1973. Phase tran- sitions in phospholipid vesicles. fluorescence polarization and permeabil- ity measurements concerning the effect of temperature and cholesterol. Biochim. Biophys. Acta 311:330 -- 340. 6. Sabra, M. C., K. Jørgensen, and O. G. Mouritsen. 1996. Lindane suppresses the lipid-bilayer permeability in the main transition region. Biochim. Biophys. Acta 1282:85 -- 92. 7. Blicher, A., K. Wodzinska, M. Fidorra, M. Winterhalter, and T. Heimburg. 2009. The temperature dependence of lipid membrane permeability, its quantized nature, and the influence of anesthetics. Biophys. J. 96:4581 -- 4591. 8. Yafuso, M., S. J. Kennedy, and A. R. Freeman. 1974. Spontaneous con- ductance changes, multilevel conductance states and negative differential resistance in oxidized cholesterol black lipid membranes. J. Membr. Biol. 17:201 -- 212. 9. Antonov, V. F., V. V. Petrov, A. A. Molnar, D. A. Predvoditelev, and A. S. Ivanov. 1980. The appearance of single-ion channels in unmodified lipid 11 bilayer membranes at the phase transition temperature. Nature 283:585 -- 586. 10. Antonov, V. F., E. V. Shevchenko, E. T. Kozhomkulov, A. A. Molnar, and E. Y. Smirnova. 1985. Capacitive and ionic currents in BLM from phos- phatidic acid in Ca2+-induced phase transition. Biochem. Biophys. Re- search Comm. 133:1098 -- 1103. 28. Wu, L.-J., T.-B. Sweet, and D. E. Clapham. 2010. International union of basic and clinical pharmacology. LXXVI. Current progress in the mam- malian TRP ion channel family. Pharmacol. Rev. 62:381 -- 404. 29. Talavera, K., K. Yasumatsu, T. Voets, G. Droogmans, N. Shigemura, Y. Ni- nomiya, R. F. Margolskee, and B. Nilius. 2005. Heat activation of TRPM5 underlies thermal sensitivity of sweet taste. Nature 438:1022 -- 1025. 11. Kaufmann, K., and I. Silman. 1983. Proton-induced ion channels through lipid bilayer membranes. Naturwissenschaften 70:147 -- 149. 30. Venkatachalam, K., and C. Montell. 2007. TRP channels. Annu. Rev. Biochem. 76:387 -- 417. 12. Kaufmann, K., and I. Silman. 1983. The induction by protons of ion channels through lipid bilayer membranes. Biophys. Chem. 18:89 -- 99. 13. Gogelein, H., and H. Koepsell. 1984. Channels in planar bilayers made from commercially available lipids. Pflugers Arch. 401:433 -- 434. 31. del Camino, D., S. Murphy, M. Heiry, L. B. Barrett, T. J. Earley, C. A. Cook, M. J. Petrus, M. Zhao, M. D'Amours, N. Deering, G. J. Brenner, M. Costigan, N. J. Hayward, J. A. Chong, C. M. Fanger, C. J. Woolf, A. Patapoutian, and M. M. Moran. 2010. TRPA1 contributes to cold hypersensitivity. J. Neurosci. 30:15165 -- 15174. 14. Antonov, V. F., A. A. Anosov, V. P. Norik, and E. Y. Smirnova. 2005. Soft perforation of planar bilayer lipid membranes of dipalmitoylphos- phatidylcholine at the temperature of the phase transition from the liquid crystalline to gel state. Eur. Biophys. J. 34:155 -- 162. 15. Wodzinska, K., A. Blicher, and T. Heimburg. 2009. The thermodynamics of lipid ion channel formation in the absence and presence of anesthetics. blm experiments and simulations. Soft Matter 5:3319 -- 3330. 32. Doerner, J. F., G. Gisselmann, H. Hatt, and C. H. Wetzel. 2007. Transient receptor potential channel a1 is directly gated by calcium ions. J. Biol. Chem. 282:13180 -- 13189. 33. Banke, T. G., S. R. Chaplan, and A. D. Wickenden. 2010. Dynamic changes in the TRPA1 selectivity filter lead to progressive but reversible pore dilation. Am. J. Physiol. Cell Physiol. 298:C1457-C1468, 2010. 298:C1457 -- C1468. 16. Wunderlich, B., C. Leirer, A. Idzko, U. F. Keyser, V. Myles, T. Heimburg, and M. Schneider. 2009. Phase state dependent current fluctuations in pure lipid membranes. Biophys. J. 96:4592 -- 4597. 34. Csanady, L., and B.Torocsik. 2009. Four Ca2+ ions activate TRPM2 channels by binding in deep crevices near the pore but intracellularly of the gate. J. Gen. Physiol. 133:189 -- 203. 17. Gallaher, J., K. Wodzinska, T. Heimburg, and M. Bier. 2010. Ion-channel- like behavior in lipid bilayer membranes at the melting transition. Phys. Rev. E 81:061925. 18. Winterhalter, M., and W. Helfrich. 1987. Effect of voltage on pores in membranes. Phys. Rev. A 36:5874 -- 5876. 19. Glaser, R. W., S. L. Leikin, L. V. Chernomordik, V. F. Pastushenko, and A. I. Sokirko. 1988. Reversible breakdown of lipid bilayers: Formation and evolution of pores. Biochim. Biophys. Acta 940:275 -- 287. 20. Neumann, E., S. Kakorin, and K. Taensing. 1999. Fundamentals of electro- porative delivery of drugs and genes. Bioelectrochem. Bioenerg. 48:3 -- 16. 35. Bari, M. R., S. Akbar, M. Eweida, F. J. P. Kuhn, A. J. Gustafsson, A. Luckhoff, and M. S. Islam. 2009. H2O2-induced Ca2+ influx and its inhibition by N-(p-amylcinnamoyl) anthranilic acid in the beta-cells: involvement of TRPM2 channels. J. Cell Mol. Med. 13:3260 -- 3267. 36. McKemy, D. D., W. M. Neuhausser, and D. Julius. 2002. Identification of a cold receptor reveals a general role for TRP channels in thermosensation. Nature 416:52 -- 58. 37. Peier, A. M., A. Moqrich, A. C. Hergarden, A. J. Reeve, D. A. Andersson, G. M. Story, T. J. Earley, I. Dragoni, P. McIntyre, S. Bevan, and A. Pat- apoutian. 2002. A trp channel that senses cold stimuli and menthol. Cell 108:705 -- 715. 21. Tieleman, D. P., H. Leontiadou, A. E. Mark, and S. J. Marrink. 2003. Simulation of pore formation in lipid bilayers by mechanical stress and electric fields. J. Am. Chem. Soc. 125:6382 -- 6383. 38. Chuang, H.-H., W. M. Neuhausser, and D. Julius. 2004. The super- cooling agent icilin reveals a mechanism of coincidence detection by a temperature-sensitive TRP channel. Neuron 43:859 -- 869. 22. Bockmann, R., R. de Groot, S. Kakorin, E. Neumann, and H. Grubmuller. 2008. Kinetics, statistics, and energetics of lipid membrane electropora- tion studied by molecular dynamics simulations. Biophys. J. 95:1837 -- 1850. 23. Gehl, J. 2003. Electroporation: theory and methods, perspectives for drug delivery, gene therapy and research. Acta Physiol. Scand. 177:437 -- 447. 39. Andersson, D. A., H. W. N. Chase, and S. Bevan. 2004. Trpm8 activation by menthol, icilin, and cold is differentially modulated by intracellular ph. J. Neurosci. 24:5364 -- 5369. 40. Andersson, D. A., M. Nash, and S. Bevan. 2007. Modulation of the cold- activated channel trpm8 by lysophospholipids and polyunsaturated fatty acids. J. Neurosci. 27:3347 -- 3355. 24. Trauble, H., M. Teubner, P. Woolley, and H. Eibl. 1976. Electrostatic inter- actions at charged lipid membranes. i. effects of ph and univalent cations on membrane structure. Biophys. Chem. 4:319 -- 342. 41. Hanke, W., C. Methfessel, U. Wilmsen, and G. Boheim. 1984. Ion chan- nel reconstruction into lipid bilayer membranes on glass patch pipettes. Bioelectrochem. Bioenerg. 12:329 -- 339. 25. Grabitz, P., V. P. Ivanova, and T. Heimburg. 2002. Relaxation kinetics of lipid membranes and its relation to the heat capacity. Biophys. J. 82:299 -- 309. 42. Kuhn, F. J. P., K. Witschas, C. Kuhn, and A. Luckhoff. 2010. Contribu- tion of the S5-pore-S6 domain to the gating characteristics of the cation channels TRPM2 and TRPM8. J. Biol. Chem. 285:26806 -- 26814. 26. Ebel, H., P. Grabitz, and T. Heimburg. 2001. Enthalpy and volume changes in lipid membranes. i. the proportionality of heat and volume changes in the lipid melting transition and its implication for the elastic constants. J. Phys. Chem. B 105:7353 -- 7360. 27. Heimburg, T., and A. D. Jackson. 2005. On soliton propagation in biomembranes and nerves. Proc. Natl. Acad. Sci. USA 102:9790 -- 9795. 43. Maertz, W. E., editor, 2008. The Axon Guide, A Guide to Electrophys- iology & Biophysics Laboratory Techniques 2500-0102 Rev. C. MDS Analytical Technologies, 3rd edition. 44. Wyllie, D. J. A., 2007. Single-channel recording. In Patch-Clamp analysis: Advanced techniques (Walz, W., editor). Humana Press, New Jersey, 69 -- 119. 12 45. Heimburg, T. 1998. Mechanical aspects of membrane thermodynam- ics. Estimation of the mechanical properties of lipid membranes close to the chain melting transition from calorimetry. Biochim. Biophys. Acta 1415:147 -- 162. 63. Bandell, M., A. E. Dubin, M. J. Petrus, A. Orth, J. Mathur, S. W. Hwang, and A. Patapoutian. 2006. High-throughput random mutagenesis screen reveals trpm8 residues specifically required for activation by menthol. Na- ture Neurosci. 9:493 -- 500. 46. Nagle, J. F., and H. L. Scott. 1978. Lateral compressibility of lipid mono- and bilayers. Theory of membrane permeability. Biochim. Biophys. Acta 513:236 -- 243. 64. Voets, T., G. Owsianik, A. Janssens, K. Talavera, and B. Nilius. 2007. Trpm8 voltage sensor mutants reveal a mechanism for integrating thermal and chemical stimuli. Nature Chem. Biol. 3:174 -- 182. 47. Voets, T., G. Droogmans, U. Wissenbach, A. Janssens, V. Flockerzi, and B. Nilius. 2004. The principle of temperature-dependent gating in cold- and heat-sensitive TRP channels. Nature 430:748 -- 754. 65. Karashima, Y., N. Damann, J. Prenen, K. Talavera, A. Segal, T. Voets, and B. Nilius. 2007. Bimodal action of menthol on the transient receptor potential channel trpa1. J. Neurosci. 27:9874 -- 9884. 66. Xiao, B., A. E. Dubin, B. Bursulaya, V. Viswanath, T. J. Jegla, and A. Pat- apoutian. 2008. Identification of transmembrane domain 5 as a critical molecular determinant of menthol sensitivity in mammalian trpa1 chan- nels. J. Neurosci. 28:9640 -- 9651. 67. Voets, T., K. Talavera, G. Owsianik, and B. Nilius. 2005. Sensing with TRP channels. Nat. Chem. Biol. 1:85 -- 92. 48. Tsien, R. W., and D. Noble. 1969. A transition state theory approach to the kinetics of conductance changes in excitable membranes. J. Membr. Biol. 1:248 -- 273. 49. Johnston, D., and S. M. S. Wu, 1995. Cellular Neurophysiology. MIT Press, Boston. 50. Nilius, B., K. Talavera, G. Owsianik, J. Prenen, G. Droogmans, and T. Voets. 2005. Gating of TRP channels: a voltage connection? J. Physiol. 567:35 -- 44. 51. Yang, J., and J. Appleyard. 2000. The main phase transition of mica- supported phosphatidylcholine membranes. J. Phys. Chem. B 104:8097 -- 8100. 52. Nagata, K., A. Duggan, G. Kumar, and J. Garc´ıa-Anoveros. 2005. No- ciceptor and hair cell transducer properties of TRPA1, a channel for pain and hearing. J. Neurosci. 25:4052 -- 4061. 53. Seeger, H. M., A. Alessandrini, and P. Facci. 2010. KcsA redistribution upon lipid domain formation in supported lipid bilayers and its functional implications. Biophys. J. 98:371a. 54. Cannon, B., M. Hermansson, S. Gyorke, P. Somerharju, and J. A. Virtanen. 2003. Regulation of calcium channel activity by lipid domain formation in planar lipid bilayers. Biophys. J. 85:933 -- 942. 55. Korlach, J., J., P. Schwille, W. W. Webb, and G. W. Feigenson. 1999. Characterization of lipid bilayer phases by confocal microscopy and fluo- rescence correlation spectroscopy. Proc. Natl. Acad. Sci. USA 96:8461 -- 8466. 56. Liebovitch, L. S., J. Fischbarg, J. P. Koniarek, I. Todorova, and M. Wang. 1987. Fractal model of ion-channel kinetics. Biochim. Biophys. Acta 896:173 -- 180. 57. Liebovitch, L. S. L., J. Fischbarg, and J. P. Koniarek. 1987. Ion channel kinetics: a model based on fractal scaling rather than multistate Markov processes. Math. Biosci. 84:37 -- 68. 58. Liebovitch, L. S. 1989. Analysis of fractal ion channel gating kinetics: kinetic rates, energy levels, and activation energies. Math. Biosci. 93:97 -- 115. 59. Liebovitch, L., D. Scheurle, M. Rusek, and M. Zochowski. 2001. Fractal methods to analyze ion channel kinetics. Methods 24:359 -- 375. 60. Millhauser, G. L., E. E. Salpeter, and R. E. Oswald. 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. 61. Nielsen, L. K., T. Bjørnholm, and O. G. Mouritsen. 2000. Critical phe- nomena - fluctuations caught in the act. Nature 404:352. 62. Watt, E. E., B. A. Betts, F. O. Kotey, D. J. Humbert, T. N. Griffith, E. W. Kelly, K. C. Veneskey, N. Gill, K. C. Rowan, A. Jenkins, and A. C. Hall. 2008. Menthol shares general anesthetic activity and sites of action on the gaba(a) receptor with the intravenous agent, propofol. Eur. J. Pharmacol. 590:120 -- 126. 13
1911.12694
1
1911
2019-11-28T13:18:09
Biophysical characterization of membrane phase transition profiles for the discrimination of Outer Membrane Vesicles (OMVs) from Escherichia coli grown at different temperatures
[ "physics.bio-ph" ]
Dynamic Light Scattering (DLS), Small Angle X-ray Scattering (SAXS) and Transmission Electron Microscopy (TEM) are physical techniques widely employed to characterize the morphology and the structure of vesicles such as liposomes or human extracellular vesicles (exosomes). Bacterial extracellular vesicles are similar in size to human exosomes, although their function and membrane properties have not been elucidated in such detail as in the case of exosomes. Here, we applied the above cited techniques, in synergy with the thermodynamic characterization of the vesicles lipid membrane using a turbidimetric technique to the study of vesicles produced by Gram-negative bacteria (Outer Membrane Vesicles, OMVs) grown at different temperatures. This study demonstrated that our combined approach is useful to discriminate vesicles of different origin or coming from bacteria cultured under different experimental conditions. We envisage that in a near future the techniques employed in our work will be further implemented to discriminate complex mixtures of bacterial vesicles, thus showing great promises for biomedical or diagnostic applications.
physics.bio-ph
physics
Biophysical characterization of membrane phase transition profiles for the discrimination of Outer Membrane Vesicles (OMVs) from Escherichia coli grown at different temperatures Angelo Sarra,1,§ Antonella Celluzzi,2,§ Stefania Paola Bruno,2 Caterina Ricci,3 Simona Sennato,4 Maria Grazia Ortore,3 Stefano Casciardi,5 Paolo Postorino,4 Federico Bordi4,#,* and Andrea Masotti2,#,* 1) Department of Science, University of Roma Tre, Via della Vasca Navale, 00146, Rome, Italy 2) Children's Hospital Bambino Gesù-IRCCS, Research Laboratories, V.le di San Paolo 15, 00146, Rome -- Italy 3) Department of Life and Environmental Sciences, Marche Polytechnic University, Via brecce bianche, Ancona, I-60131, Italy 4) CNR-ISC UOS Sapienza and Department of Physics, Sapienza University of Rome, Pz.le Aldo Moro, 00185, Rome, Italy; 5) National Institute for Insurance against Accidents at Work (INAIL), Department of Occupational and Environmental Medicine, Epidemiology and Hygiene, 00078 Monte Porzio Catone, Italy. § These two authors equally contributed to the work # These two authors equally contributed to the work Corresponding authors: Andrea Masotti, Bambino Gesù Children's Hospital-IRCCS, Research Laboratories, 00146 Rome, Italy. E-mail: [email protected]. Tel: +39 06 6859 2650. Federico Bordi, Department of Physics, Sapienza University of Rome, Pz.le Aldo Moro, 00185, Rome, Italy. E-mail: [email protected]. Tel: +39 06 4991 3503. ABSTRACT Dynamic Light Scattering (DLS), Small Angle X-ray Scattering (SAXS) and Transmission Electron Microscopy (TEM) are physical techniques widely employed to characterize the morphology and the structure of vesicles such as liposomes or human extracellular vesicles (exosomes). Bacterial extracellular vesicles are similar in size to human exosomes, although their function and membrane properties have not been elucidated in such detail as in the case of exosomes. Here, we applied the above cited techniques, in synergy with the thermodynamic characterization of the vesicles lipid membrane using a turbidimetric technique to the study of vesicles produced by Gram-negative bacteria (Outer Membrane Vesicles, OMVs) grown at different temperatures. This study demonstrated that our combined approach is useful to discriminate vesicles of different origin or coming from bacteria cultured under different experimental conditions. We envisage that in a near future the techniques employed in our work will be further implemented to discriminate complex mixtures of bacterial vesicles, thus showing great promises for biomedical or diagnostic applications. Keywords: Outer membrane vesicle (OMV); Dynamic light scattering (DLS); Transmission electron microscopy (TEM); Small-angle X-ray scattering (SAXS); Gram negative bacteria. INTRODUCTION In all domains of life Eukarya, Archaea and Bacteria produce and release membrane vesicles for reasons that are still not completely understood (Deatherage and Cookson, 2012). In humans, many cells such as dendritic cells, lymphocytes, and tumor cells actively release (i.e., by exocytosis) small (30-100 nm of diameter) membrane vesicles, referred as exosomes, into biofluids (i.e., plasma/serum, urine, cerebrospinal fluid and saliva). These vesicles are powerful cell-to-cell messengers as they transfer lipids, proteins, DNA and ribonucleic acids (i.e., mRNA, microRNA, lncRNA and other RNA species) between cells (Simpson et al., 2009; Chen et al., 2012; Valadi et al., 2007). In the last few years, exosomes and their inner content have been also exploited as innovative and effective biomarkers for the diagnosis of many different diseases (i.e., tumors, obesity, gastrointestinal disorders, fibromyalgia, etc.) (Logozzi et al., 2009; Taylor and Gercel-Taylor, 2008; Masotti et al., 2017; Felli, Baldassarre and Masotti, 2017). Similarly to human cells, either Gram-negative and Gram-positive bacteria produce extracellular vesicles, referred as Outer-membrane vesicles (OMVs) and membrane vesicles (MVs), respectively. The production and release of vesicles by bacteria is a natural process and it is necessary for inter-species (bacteria-bacteria) and inter-kingdom (bacteria-host) interactions (Leitão and Enguita, 2016). One of the main structural differences between Gram-positive and Gram-negative bacteria relies on the composition of their cell envelope. In both cases, the bacterial envelope comprises the plasma membrane and a layer of peptidoglycans. However, in the case of Gram-negative bacteria an additional layer (i.e., the outer membrane) is also observed. The space between the cytoplasmic membrane and the outer membrane is the periplasmic space or periplasm. The outer membrane is made by negatively charged phospholipids and lipopolysaccharides that confer to the Gram-negative wall an overall negative charge. Owing to the presence of these lipopolysaccharides on the outer membranes, many gram- negative bacteria are often pathogenic. Gram-negative bacteria (e.g. Escherichia coli, Pseudomonas aeruginosa, etc.) spontaneously secrete OMVs, small (20-100 nm of diameter) spherical bi-layered vesicles, delivered in a variety of environments, including planktonic cultures, fresh and salt water, biofilms, inside eukaryotic cells and within mammalian hosts (Beveridge, 1999; Beveridge et al., 1997; Biller et al., 2014; Hellman et al., 2000). Gram positive bacteria as well, such as Staphylococcus aureus, Bacillus subtilis, Bacillus anthracis, Streptomyces coelicolor, Listeria monocytogenes, Clostridium perfringens, Streptococcus mutants, and Streptococcus pneumoniae spontaneously produce MVs (Lee et al., 2009; Rivera et al., 2010; Schrempf et al., 2011; Lee et al., 2013; Jiang et al., 2014; Liao et al., 2014). Recently, it has been observed that also non-pathogenic bacteria (i.e., probiotics), such as Lactobacilli, release extracellular vesicles. Given the importance and the positive impact that probiotic effects have on human health, the study of Lactobacilli MVs may be an interesting opportunity for various applications, from vaccines to therapeutic delivery (Dean et al., 2017; van der Pol, Stork and van der Ley, 2015). In this context, a detailed physicochemical characterization of MVs from Lactobacillus acidophilus ATCC 53544, Lactobacillus casei ATCC 393, and Lactobacillus reuteri ATCC 23272, was recently reported (Dean et al., 2019; Grande et al., 2017). Among the Lactobacilli, the Lactobacillus rhamnosus GG (LGG), is emerging as an important probiotic strain due to its validated effects in both treating and preventing some gastro-intestinal diseases (Wolvers et al., 2010). Behzadi and colleagues revealed the cytotoxic and inhibitor role of LGG-derived extracellular vesicles on hepatic cancer cells, but the characterization of these vesicles has not still been reported (Behzadi, Mahmoodzadeh Hosseini and Imani Fooladi, 2017). Moreover, owing to the recent employ of OMVs and MVs as adjuvants in vaccines (van der Pol, Stork and van der Ley, 2015; Bottero et al., 2016) or as novel vaccines platform (Wang et al., 2018) to regulate host activity processes (Grandi et al., 2017; Chen et al., 2017) or other pathogenic processes (Ellis and Kuehn, 2010), an in depth biophysical characterization (i.e., identification, discrimination and quantification) of these types of vesicles is highly desired for biomedical applications. Among the most employed biophysical techniques available to characterize such small particles in aqueous suspension, Dynamic Light Scattering (DLS) represents one of the most versatile ones (see for example (Berne and Pecora, 1976)). DLS is generally used to determine vesicles size but at the same time it provides information about the reorganization of lipid molecules of the membrane bilayer with a turbidimetric method (Michel et al., 2006). Actually, the intensity of light scattered at a fixed angle by particles in a suspension depends on their size and on their optical properties (hence on the structure and on the refractive index of the components). By measuring the time-averaged intensity of scattered light as a function of the temperature, this experiment allows the determination of thermotropic lipid phase transitions (with characteristic transition temperatures Tc), related to the lipid organization and their remodeling within the bilayer, which affects the membrane optical properties (refractive index) (Aleandri et al., 2012). Consequently, if different thermotropic behaviours feature different kind of vesicles, DLS experimental set-up can be used as a tool to discriminate them, too. In fact, it was previously known that E. coli outer membranes show a phase transition due to conformational changes of their lipid components (Trauble and Overath, 1973). Lipid composition modifications have been also observed in their outer membranes when the environmental conditions are changed (i.e. at growing temperatures) (Morein et al., 1996; Mika et al., 2016; Marr and Ingraham, 1962). A similar behavior can be expected also for OMVs produced by E. coli as well. Therefore, aimed at providing a useful tool to characterize better OMVs and MVs, we propose a solid set of experimental physical techniques, including this turbidimetric approach that is able to discriminate extracellular vesicles originating from bacteria grown at different environmental conditions. MATERIALS AND METHODS Bacterial strains and culturing conditions Escherichia coli (ATCC 8739) was cultured at three different temperatures: 20°C, 27°C and 37°C in essential M9 microbial growth medium consisting in Na2HPO4 (6.8 g/L), KH2PO4 (3 g/L), NH4Cl (1 g/L) and NaCl (0.5 g/L). To M9 medium, a solution containing D-glucose (4 g/L), MgSO4 (241 mg/L) and CaCl2·2H2O (15 mg/L) was added and the resulting solution was adjusted to pH=7. Bacteria were cultured overnight until the culture reached an OD600 of approximately 1. Lactobacillus rhamnosus LGG (ATCC 53103) (Dicoflor 60 ®, DICOFARM) was cultured overnight in a De Man, Rogosa and Sharpe medium (SigmaAldrich, MRS broth, 51 g/L) at 37°C in anaerobic conditions up to an OD600 of approximately 2. Aliquots (5 ml) of the bacterial cultures (both E. coli and L. rhamnosus) were retained for further analysis and to be used as controls. To isolate either OMVs or MVs, bacteria were removed from culture media by centrifugation (Beckman Avanti J-25 centrifuge, JA-10 rotor, 6000 × g, 15 min), and the supernatant was filtered through a 0.45 μm filter unit (Sartorius). The supernatant, which contains vesicles, was concentrated by ultrafiltration (Vivaflow 200, Sartorius) up to small volume (50 mL), then filtered through a 0.45 μm filter. Isolation of bacterial extracellular vesicles Bacterial vesicles (OMVs or MVs) were separated by ultracentrifugation (107,000g for 3h at 4°C) (Optima™ XPN-100 Ultracentrifuge, Beckman Coulter, SW 40 Ti-rotor) of the ultrafiltered solution obtained previously. The pellet was washed once in sterile phosphate buffered saline (Dulbecco's Phosphate-Buffered Saline -- PBS) and suspended in the same buffer. At the end of the isolation protocol, a drop of the purified solution was cultured on agar plates at 37°C for one day to exclude the presence of residual bacteria. The suspension obtained with this method contains a concentrated amount of vesicles (OMVs or MVs) that have been characterized by dynamic light scattering and turbidimetric analyses. Dynamic Light Scattering and turbidimetric analyses DLS experimental set-up has been employed to provide vesicles size and to characterize the bilayer organization because of its sensitivity to the sample refractive index (Berne and Pecora, 1976). In this case, we use the term 'turbidimetric' measurements for this unconventional way of using a light scattering apparatus, although this term is not strictly appropriate, because we observe changes in the refractive index of the lipid membranes, concerning the scattered light and not the transmitted light, as in conventional turbidimetric measurements. DLS and turbidimetric measurements were performed employing a MALVERN Nano Zetasizer apparatus equipped with a 5 mW HeNe laser (Malvern Instruments LTD, UK). This system uses backscatter detection, i.e. the scattered light is collected at 173°. To obtain the size distribution, the measured autocorrelation functions were analyzed using the CONTIN algorithm (Provencher, 1982). Decay times are used to determine the distribution of the diffusion coefficients D0 of the particles, which in turn can be converted in a distribution of apparent hydrodynamic diameter, Dh, using the Stokes-Einstein relationship Dh = kBT/3D0, where kB is the Boltzmann constant, T the absolute temperature and  the solvent viscosity (Berne and Pecora, 1976). The values of the radii shown in this work correspond to the average values on several measurements and are obtained from intensity weighted distributions (Provencher, 1982; De Vos, Deriemaeker and Finsy, 1996). The thermal protocol used for both OMVs, MVs and bacteria in DLS measurements consists of an ascending ramp from 10°C to 45°C with temperature step of 1°C. At each step, samples were left to thermalize at the target temperature for few minutes. In particular, for samples grown at 37°C, 27°C and 20°C the thermalization times were 360 s, 720 s and 1080 s, respectively. For the turbidimetric determination of vesicles' membranes thermotropic behavior, we measured the mean count rate of scattered photons (i.e. the time-averaged intensity of the scattered light), I, versus temperature, T, and fitted the data to a Boltzmann sigmoidal curve: 𝐼 = 𝐼0 + 𝐼1−𝐼0 1+𝑒𝑥𝑝 𝑇−𝑇0 ∆𝑇 (Equation 1) where the fitting parameters are I0 and I1, the minimum and maximum intensity respectively, the transition temperature T0 and the so called 'slope' ΔT that describes the steepness of the curve. ΔT roughly measures the transition width. Small Angle X-ray Scattering SAXS experiments were performed at the Austrian SAXS beamline of the Elettra Synchrotron in Trieste, Italy (Amenitsch, Bernstorff and Laggner, 1995). SAXS images were recorded with a 2D pixel detector Pilatus3 1M spanning the Q-range between 0.1 and 6 nm−1, with Q=4πsin(2Θ)/λ, where 2Θ is the scattering angle and λ = 0.0995 nm the X-ray wavelength. The image conversion to one-dimensional SAXS pattern was performed with FIT2D (Hammersley, 2004). Detector calibration was performed with silver behenate powder (d-spacing =0.05838 nm). Samples were held in a 1 mm glass capillary (Hilgenberg, Malsfeld, Germany). Experiments were carried out between 10 and 45°C, by using the same heating rate already adopted in DLS experiments. Each measurement was performed for 1 s and followed by a dead time of at least 10 s in order to avoid radiation damage. We obtained a set of SAXS curves resulting from the average of several overlapping SAXS spectra obtained at about the same temperature (±1°C), in order to analyze data with an improved signal-to-noise ratio. SAXS data were fitted according to a core-shell model (Berndt, Pedersen and Richtering, 2005) by GENFIT software (Spinozzi et al., 2014), and taking into account the vesicles polydispersion, in agreement with the DLS results. The fitting parameters are OMVs radius R and thickness d, as well as the electron densities of the core and of the shell, ρi and ρe, respectively. Transmission Electron Microscopy Transmission electron microscopy (TEM) was used for the morphological characterization of OMVs and MVs. TEM images were performed at Department of Occupational and Environmental Medicine, Epidemiology and Hygiene of INAIL-Research in Rome. Samples for TEM measurements were prepared at room temperature by depositing 20 µl of vesicles suspensions on a 300-mesh copper grid for electron microscopy covered by a thin amorphous carbon film. A negative staining was realized by addition of 10 μl of 2 % aqueous phosphotungstic acid (PTA) solution (pH-adjusted to 7.3 using 1 N NaOH). Measurements were carried out by using a FEI TECNAI 12 G2 Twin (FEI Company, Hillsboro, OR, USA), operating at 120 kV and equipped with an electron energy loss filter (Biofilter, Gatan Inc, Pleasanton, CA, USA) and a slow-scan charge-coupled device camera (794 IF, Gatan Inc, Pleasanton, CA, USA). Scattering cross section simulations To better understand the behaviour of phase transitions in OMVs and bacteria, Mie scattering simulations were performed using the MiePlot 4.6 simulator by P.Laven (Laven, 2003). This simulator allows to calculate the scattered intensity as a function of scattering angle for inhomogeneous spheres where the refractive index is a function of the radius. For simplicity sake, both OMVs and bacteria were described as multiple-shelled spheres. In all calculations, we assumed a wavelength of 633 nm for the incident light and a scattering angle of 173°. For the refractive index of the aqueous core (of both vesicles and bacteria) we used the value ncytoplasm = 1.367 according to the literature (Choi et al., 2007; Marquis, 1973), whereas for the bacterial cell wall we used ncellwall = 1.455 (Marquis, 1973). Finally, we assumed that the dependence of the refractive index as a function of the temperature of the lipid membrane (for vesicles and bacteria) was described by a Boltzmann sigmoidal curve (Equation 1). RESULTS AND DISCUSSION Isolation of OMVs and MVs The overall workflow followed for the isolation of bacterial extracellular vesicles comprehensive of the quality control steps has been depicted in Figure 1. Briefly, to have an enriched solution of bacterial vesicles, we started from a large amount of culture medium that has been first centrifuged and filtered to remove bacteria. Then the tangential flow ultrafiltration procedure allowed us to concentrate the sample up to 100 times and reduce the operating time compared to traditional ultrafiltration techniques. The ultrafiltered solution undergone ultracentrifugation to obtain a transparent pellet of bacterial vesicles that after a washing step afforded a very clean sample without residual proteins contaminants. Our extraction and purification protocol allowed us to obtain pure and concentrated samples of OMVs from cultures of E. coli grown at three different temperatures (i.e., 20°C, 27°C and 37°C). Similar procedures were adopted for the isolation of MVs from LGG. Morphological characterization DLS intensity weighed distribution (Figure 2) and TEM analysis of OMVs from E. coli grown at 37°C (inset of Figure 2) revealed that vesicles have a spherical shape with a mean hydrodynamic diameter of 48±3 nm (FWHM=24±2 nm). Notably, the DLS analysis revealed that the vesicle size decreases as the growth temperature is lowered (Table 1). OMVs grown at 27°C and 20° C have a mean diameter of 37±4 nm (FWHM=32±2 nm) and 24±2 nm (FWHM=20±3 nm), respectively. Typical TEM images and DLS intensity weighed distributions for vesicles growth at 27°C and 20°C are reported in Supporting Materials. It is known that E. coli can undergo remarkable modifications (i.e., conformational and/or compositional changes of the lipid components) of the outer membrane as a function of the different environmental conditions such as the growth temperature. Therefore, to investigate if the size reduction of OMVs as a function of the growing temperature could be related to a different conformational state of the lipids that form the vesicle itself or their remodeling within the bilayer, we performed turbidimetric measurements on OMV samples extracted from bacteria grown at the same temperatures. Turbidimetric experiments The intensity of light scattered at a fixed angle by the particles in a suspension depends on their size, geometry and optical properties (structure and refractive index of the components). In particular, the refractive index of a lipid bilayer depends on the conformational state of the lipids and on their ordering in the bilayer. Hence, by measuring the time-averaged intensity of scattered light as a function of the temperature, the instrument allowed us to determine the thermotropic lipid phase transitions (with characteristic transition temperatures Tc), related to the lipid organization and their remodeling within the bilayer (Michel et al., 2006; Aleandri et al., 2012). We used the term 'turbidimetric measurements' for this unconventional way of using a light scattering apparatus although this term is not strictly appropriate, since in the present case to reveal possible changes in the refractive index of the lipid membranes as a function of temperature, the scattered light and not the transmitted light, as in conventional turbidimetric measurements, is monitored. Turbidimetric measurements were performed in the temperature range 10°C-45°C on OMVs samples and on bacteria grown at 37°C, 27°C and 20°C. Characterization of OMVs extracted from E. coli cultured at 37°C In Figure 3 the results obtained from turbidimetric measurements for vesicles (Figure 3a) and E. coli (Figure 3b) (as comparison) grown at 37°C have been reported. As far as the vesicles are concerned, the mean scattering intensity (Figure 3a, upper panel) decreased significantly (about 25%) when the temperature was increased. We determined a transition temperature by fitting a Boltzmann sigmoidal curve [Equation 1] to the experimental data (red line in Figure 3a, upper panel). The fitting procedure provided a transition temperature of 32 ± 2°C, where the uncertainty is the fitted slope ΔT. The observed change of the scattered intensity could be ascribed to different factors (Berne and Pecora, 1976) (i.e., to a decrease of the concentration and/or the size of the suspended particles) or to a variation of vesicles refractive index (Michel et al., 2006). However, we did not observe any phase separation (flocculation or precipitation of the particles) and the hydrodynamic diameter of the particles determined by DLS (lower panel of Figure 3a) did not show any appreciable dependence on temperature. Hence, we ascribed the observed decrease of the scattered intensity to a change of the refractive index of the particles, likely due to a phase transition of the membrane lipid bilayer or, more in general, to a structural reorganization of the membrane (Berne and Pecora, 1976). Figure 3b shows the results obtained from a suspension of E. coli. Again, also with bacteria a rather steep variation of the scattered intensity was observed within the same temperature range observed in the case of OMVs suggesting the occurrence of a similar membrane transition. However, in the case of E. coli the scattered intensity increased, whereas in the case of OMVs decreased. Although the scattered intensity measurements were unreliable above 35°C (owing to the proliferation of bacteria that altered the optical properties of the suspension), a transition temperature of 30 ± 3°C was obtained (Figure 3b). We attributed the opposite behavior of the scattered intensity to the different structures of the bacteria cell wall and, as a consequence, of the OMVs membrane. Characterization of OMVs extracted from E. coli cultured at 27°C and 20°C Similar results were obtained also for samples grown at 27°C and 20°C. At these growth temperatures, OMVs were smaller (diameters were 37±4 nm and 24±2 nm, respectively, compared to 48±3 nm at 37°C) and their concentration was found to be systematically lower than that obtained for the samples grown at 37°C. As a consequence, also the mean scattering intensity was significantly lower. Nevertheless, also in these conditions the trend of the registered data points suggested the presence of a phase transition. In Figure 4 we compared the scattering intensity of vesicles (Figure 4a) and bacteria (Figure 4b) grown at different temperatures. For each sample we reported the normalized scattering intensity, the fitting curve (solid line) and the value of the fitted transition temperature. All the samples analyzed shown a similar behavior in terms of phase transitions. Interestingly, for both OMVs and bacteria a strong correlation between the growth temperature at which E. coli was cultured and the transition temperature of the membrane was found (Figure 5). From previous E. coli studies, we know that by reducing the temperature of the growth medium the content of unsaturated lipids, which confer a higher flexibility to the membrane, increases consequently. Moreover, bacteria membranes show thermotropic phase transitions whose characteristic transition temperature depends on the lipid composition (Trauble and Overath, 1973) which in turn is related to the growth temperature (Morein et al., 1996). It is reasonable to think that OMVs, that originate from the bacterial outer membrane, may have a similar composition and similar thermotropic behaviors. Simulations of bacteria and OMVs scattering cross section To better understand the different optical response observed for vesicles and bacteria, we calculated the scattering cross section of different structures simulating vesicles and bacteria by using the Mie theory (Laven, 2003). Vesicles and bacteria have been represented and modeled as shown in Figure 6. In particular, OMVs have been modeled as a single-shelled sphere, where the core is the cytoplasm (assumed homogeneous) and the shell represents the lipid membrane, whereas in the case of E. coli also a second (internal) shell is added, which represents the cell wall, mainly composed of peptidoglycans. To take into account the thermotropic behavior and evaluate the refractive index n(T) of the lipid bilayer, we assumed the Boltzmann function dependence on the temperature (see Equation 1) with the T0 and ΔT values determined from experimental turbidimetric measurements. We also imposed the ratio of the asymptotic values of the refractive index (n0/n1) equal to the ratio of the measured intensities (I0/I1). Finally, we constrained the Boltzmann curve to assume the value of 1.44 at T=23 °C according to the literature (White et al., 2000). The results of the calculations obtained using MiePlot 4.6 simulator were reported in Figure 7. In particular, Figure 7a showed the transitions of OMV models, whereas Figure 7b the corresponding curves calculated for the E. coli models. Notably, the simulations clearly confirmed not only the temperature-dependent transition trend but also the opposite scattering intensity behavior of OMVs and bacteria, as we previously observed experimentally. Simulations gave us the possibility to identify the reason behind the transition differences observed in the case of OMVs and bacteria and ascribe this behavior to the presence of a second 'layer' beneath the outer membrane (the cell wall comprising the peptidoglycan layer) with a fixed refractive index. In fact, this is the only difference between these two models (at least from the optical point of view). This strongly support the concept the trend of the transition temperature may represent a characteristic 'fingerprint' of OMVs and this behaviour could be employed as a characteristic 'marker' to distinguish vesicles from the bacteria that originated them. Small Angle X-ray Scattering experiments To investigate in more details the effects of the temperature on OMVs structure we performed SAXS experiments. Due to the low concentration of vesicles obtained at 27°C and 20°C, we performed these measurements only on vesicles grown at 37°C. We reported SAXS patterns in Figure 8a. The overall SAXS pattern, which is typical of shelled spheres, did not change significantly when the temperature increased. These data confirmed that OMVs size did not vary in the temperature range investigated and no membrane damages were reported. However, the smooth peak centered at Q ≈ 1 nm−1 (inset of Figure 8a) is related to the thickness of the membrane and has been already detected by SAXS on Alix-positive exosome-like small extracellular vesicles (Romancino et al., 2018). This peak significantly changed the shape and moved towards higher Q values at increasing temperatures, although this variation was modest. In order to better understand the significance of this peak position shift, SAXS curves were analysed considering the simplified model of OMVs reported in Figure 6. Although OMVs membrane contain a mixture of different lipids (i.e., sterols, polyisoprenoids, etc.) (Bramkamp and Lopez, 2015) and proteins, this simple model appeared to be able to reproduce reasonably our SAXS data. The thickness as a function of temperature obtained from the fitting procedure was reported in Figure 8b. The radius and the electron densities as a function of temperature were reported in the Supporting Material. The thickness of the OMVs membrane changed significantly as the temperature increased, whereas vesicles size was not affected by temperature, remained intact in this temperature range as indicated by the core electron density as well as the shell's one that remained constant. These findings suggested that temperature was able to induce a general rearrangement of the membrane structural motifs and that these motifs are linked to the reorganization of the lipid membrane and/or to the conformational changes of its components. OMVs have a distinct phase transition behaviour compared to other MVs In order to assess the feasibility of using the phase transition profile to discriminate extracellular vesicles coming from different bacteria (i.e., Gram-negative and Gram-positive bacteria) and potentially use this information for diagnostic purposes in the microbiology field, we decided to measure the phase transition properties of MVs from the Gram-positive LGG comparing the results with those obtained with OMVs from the Gram-negative E. coli. In this specific case, we studied only the phase transition of MVs from LGG by turbidimetric measurements as this technique is easy to reproduce, does not require expensive equipment, is cheap and also relatively fast. All of these characteristics are highly desirable when thinking to techniques for diagnostics applications. In Figure 9 we reported the results of the scattering intensity measurements in the temperature range between 10°C and 45 °C. By examining the profile of MVs from LGG, we observed the presence of multiple transitions, two of them at a T<35°C and another one likely around 37°C. DLS analysis confirmed that the diameter remained constant (as show in Supporting Material) throughout all the temperature range. Similarly to what observed for OMVs, flocculation or precipitation of MVs did not occur, confirming that the transitions are related only to membrane modifications. The intensity data were fitted with a double Boltzmann function in the region at T<35°C, providing two transition temperatures of 16±2°C and 28±3°C. Interestingly, the scattering profile of MVs was completely different than that of OMVs, suggesting that MVs and OMVs have a different lipid composition of the membrane. CONCLUSIONS In this work the thermotropic behavior of OMVs and bacteria grown at different temperatures have been studied. The analysis of mean scattering intensity as a function of the temperature has shown the presence of peculiar membrane phase transitions characteristic of both OMVs and bacteria. In particular, the transition temperature that we measured allowed to discriminate not only vesicles of different composition (i.e., OMVs and MVs) but also to correlate them to bacteria that originated them and to their culture conditions (i.e., temperature). The results we obtained have shown that it is possible to distinguish vesicles coming from different bacteria (i.e., gram-positive and gram-negative bacterial) by simply studying the phase transitions of their membranes. In a close future, by implementing this technique, it could be even possible to perform microbiological analysis by discriminating different bacteria in complex mixtures or matrices (i.e., biological fluids, stool, blood, tissues, etc.) More generally, this technique could be also employed in fields such as oncology, for example to discriminate human exosomes originating from tumors from that produced by normal cells and help clinicians to diagnose different form of tumors. ACKNOWLEDGMENTS Dr. Federica Del Chierico is gratefully acknowledged for her fundamental role in suggesting the best culturing conditions to grow E. coli at different temperatures using the M9 minimal medium and for having critically revised and edited the manuscript draft. Prof. Heinz Amenitsch is acknowledged for his precious assistance in SAXS experiments. REFERENCES Aleandri, S., Bonicelli, M.G., Bordi, F., Casciardi, S., Diociaiuti, M., Giansanti, L., Leonelli, F., Mancini, G., Perrone, G., and Sennato, S. (2012). How stereochemistry affects the physicochemical features of gemini surfactant based cationic liposomes. Soft Matter 8, 5904-5915 Amenitsch, H., Bernstorff, S., and Laggner, P. (1995). High‐flux beamline for small‐angle x‐ ray scattering at ELETTRA. Rev. Sci. Instrum. 66, 1624-1626 Behzadi, E., Mahmoodzadeh Hosseini, H., and Imani Fooladi, A.A. (2017). The inhibitory impacts of Lactobacillus rhamnosus GG-derived extracellular vesicles on the growth of hepatic cancer cells. Microb. Pathog. 110, 1-6 Berndt, I., Pedersen, J.S., and Richtering, W. (2005). Structure of multiresponsive "intelligent" core-shell microgels. J. Am. Chem. Soc. 127, 9372-9373 Berne, B. and Pecora, R. (1976). Dynamic Light Scattering. New York: John Wiley Beveridge, T.J. (1999). Structures of gram-negative cell walls and their derived membrane vesicles. J. Bacteriol. 181, 4725-4733 Beveridge, T.J., Makin, S.A., Kadurugamuwa, J.L., and Li, Z. (1997). Interactions between biofilms and the environment. FEMS Microbiol. Rev. 20, 291-303 Biller, S.J., Schubotz, F., Roggensack, S.E., Thompson, A.W., Summons, R.E., and Chisholm, S.W. (2014). Bacterial vesicles in marine ecosystems. Science 343, 183-186 Bottero, D., Gaillard, M.E., Zurita, E., Moreno, G., Martinez, D.S., Bartel, E., Bravo, S., et al. (2016). Characterization of the immune response induced by pertussis OMVs-based vaccine. Vaccine 34, 3303-3309 Bramkamp, M. and Lopez, D. (2015). Exploring the existence of lipid rafts in bacteria. Microbiol. Mol. Biol. Rev. 79, 81-100 Chen, F., Cui, G., Wang, S., Nair, M.K.M., He, L., Qi, X., Han, X., Zhang, H., Zhang, J.R., and Su, J. (2017). Outer membrane vesicle-associated lipase FtlA enhances cellular invasion and virulence in Francisella tularensis LVS. Emerg. Microbes Infect. 6, e66 Chen, X., Liang, H., Zhang, J., Zen, K., and Zhang, C.Y. (2012). Horizontal transfer of microRNAs: molecular mechanisms and clinical applications. Protein Cell. 3, 28-37 Choi, W., Fang-Yen, C., Badizadegan, K., Oh, S., Lue, N., Dasari, R.R., and Feld, M.S. (2007). Tomographic phase microscopy. Nat. Methods 4, 717-719 De Vos, C., Deriemaeker, L., and Finsy, R. (1996). Quantitative Assessment of the Conditioning of the Inversion of Quasi-Elastic and Static Light Scattering Data for Particle Size Distributions. Langmuir 12, 2630-2636 Dean, S.N., Leary, D.H., Sullivan, C.J., Oh, E., and Walper, S.A. (2019). Isolation and characterization of Lactobacillus-derived membrane vesicles. Sci. Rep. 9, 877-018- 37120-6 Dean, S.N., Turner, K.B., Medintz, I.L., and Walper, S.A. (2017). Targeting and delivery of therapeutic enzymes. Therapeutic Delivery 8, 577-595 Deatherage, B.L. and Cookson, B.T. (2012). Membrane vesicle release in bacteria, eukaryotes, and archaea: a conserved yet underappreciated aspect of microbial life. Infect. Immun. 80, 1948-1957 Ellis, T.N. and Kuehn, M.J. (2010). Virulence and immunomodulatory roles of bacterial outer membrane vesicles. Microbiol. Mol. Biol. Rev. 74, 81-94 Felli, C., Baldassarre, A., and Masotti, A. (2017). Intestinal and Circulating MicroRNAs in Coeliac Disease. Int. J. Mol. Sci. 18(9), 1907 Grande, R., Celia, C., Mincione, G., Stringaro, A., Di Marzio, L., Colone, M., Di Marcantonio, M.C., et al. (2017). Detection and Physicochemical Characterization of Membrane Vesicles (MVs) of Lactobacillus reuteri DSM 17938. Front. Microbiol. 8, 1040 Grandi, A., Tomasi, M., Zanella, I., Ganfini, L., Caproni, E., Fantappie, L., Irene, C., et al. (2017). Synergistic Protective Activity of Tumor-Specific Epitopes Engineered in Bacterial Outer Membrane Vesicles. Front. Oncol. 7, 253 Hammersley, A.P. (2004). Fit2d. http://www.esrf.eu/computing/scientific/FIT2D/ Hellman, J., Loiselle, P.M., Zanzot, E.M., Allaire, J.E., Tehan, M.M., Boyle, L.A., Kurnick, J.T., and Warren, H.S. (2000). Release of gram-negative outer-membrane proteins into human serum and septic rat blood and their interactions with immunoglobulin in antiserum to Escherichia coli J5. J. Infect. Dis. 181, 1034-1043 Jiang, Y., Kong, Q., Roland, K.L., and Curtiss, R.,3rd (2014). Membrane vesicles of Clostridium perfringens type A strains induce innate and adaptive immunity. Int. J. Med. Microbiol. 304, 431-443 Laven, P. (2003). Simulation of rainbows, coronas, and glories by use of Mie theory. Appl. Opt. 42, 436-444 Lee, E.Y., Choi, D.Y., Kim, D.K., Kim, J.W., Park, J.O., Kim, S., Kim, S.H., et al. (2009). Gram-positive bacteria produce membrane vesicles: proteomics-based characterization of Staphylococcus aureus-derived membrane vesicles. Proteomics 9, 5425-5436 Lee, J.H., Choi, C.W., Lee, T., Kim, S.I., Lee, J.C., and Shin, J.H. (2013). Transcription factor sigmaB plays an important role in the production of extracellular membrane-derived vesicles in Listeria monocytogenes. PLoS One 8, e73196 Leitão, A.L. and Enguita, F.J. (2016). Non-coding RNAs and Inter-kingdom Communication. Switzerland: Springer International Publishing Liao, S., Klein, M.I., Heim, K.P., Fan, Y., Bitoun, J.P., Ahn, S.J., Burne, R.A., Koo, H., Brady, L.J., and Wen, Z.T. (2014). Streptococcus mutans extracellular DNA is upregulated during growth in biofilms, actively released via membrane vesicles, and influenced by components of the protein secretion machinery. J. Bacteriol. 196, 2355-2366 Logozzi, M., De Milito, A., Lugini, L., Borghi, M., Calabro, L., Spada, M., Perdicchio, M., et al. (2009). High levels of exosomes expressing CD63 and caveolin-1 in plasma of melanoma patients. PLoS One 4, e5219 Marquis, R.E. (1973). Immersion refractometry of isolated bacterial cell walls. J. Bacteriol. 116, 1273-1279 Marr, A.G. and Ingraham, J.L. (1962). Effect of Temperature on the Composition of Fatty Acids in Escherichia Coli. J. Bacteriol. 84, 1260-1267 Masotti, A., Baldassarre, A., Fabrizi, M., Olivero, G., Loreti, M.C., Giammaria, P., Veronelli, P., Graziani, M.P., and Manco, M. (2017). Oral glucose tolerance test unravels circulating miRNAs associated with insulin resistance in obese preschoolers. Pediatr. Obes. 12, 229-238 Michel, N., Fabiano, A.S., Polidori, A., Jack, R., and Pucci, B. (2006). Determination of phase transition temperatures of lipids by light scattering. Chem. Phys. Lipids 139, 11-19 Mika, J.T., Thompson, A.J., Dent, M.R., Brooks, N.J., Michiels, J., Hofkens, J., and Kuimova, M.K. (2016). Measuring the Viscosity of the Escherichia coli Plasma Membrane Using Molecular Rotors. Biophys. J. 111, 1528-1540 Morein, S., Andersson, A., Rilfors, L., and Lindblom, G. (1996). Wild-type Escherichia coli cells regulate the membrane lipid composition in a "window" between gel and non- lamellar structures. J. Biol. Chem. 271, 6801-6809 Provencher, S.W. (1982). A constrained regularization method for inverting data represented by linear algebraic or integral equations. Comput. Phys. Commun. 27, 213- 227 Rivera, J., Cordero, R.J., Nakouzi, A.S., Frases, S., Nicola, A., and Casadevall, A. (2010). Bacillus anthracis produces membrane-derived vesicles containing biologically active toxins. Proc. Natl. Acad. Sci. U. S. A. 107, 19002-19007 Romancino, D.P., Buffa, V., Caruso, S., Ferrara, I., Raccosta, S., Notaro, A., Campos, Y., et al. (2018). Palmitoylation is a post-translational modification of Alix regulating the membrane organization of exosome-like small extracellular vesicles. Biochim. Biophys. Acta Gen. Subj. 1862, 2879-2887 Schrempf, H., Koebsch, I., Walter, S., Engelhardt, H., and Meschke, H. (2011). Extracellular Streptomyces vesicles: amphorae for survival and defence. Microb. Biotechnol. 4, 286- 299 Simpson, R.J., Lim, J.W., Moritz, R.L., and Mathivanan, S. (2009). Exosomes: proteomic insights and diagnostic potential. Expert Rev. Proteomics 6, 267-283 Spinozzi, F., Ferrero, C., Ortore, M.G., De Maria Antolinos, A., and Mariani, P. (2014). GENFIT: software for the analysis of small-angle X-ray and neutron scattering data of macro-molecules in solution. J. Appl. Crystallogr. 47, 1132-1139 Taylor, D.D. and Gercel-Taylor, C. (2008). MicroRNA signatures of tumor-derived exosomes as diagnostic biomarkers of ovarian cancer. Gynecol. Oncol. 110, 13-21 Trauble, H. and Overath, P. (1973). The structure of Escherichia coli membranes studied by fluorescence measurements of lipid phase transitions. Biochim. Biophys. Acta 307, 491- 512 Valadi, H., Ekstrom, K., Bossios, A., Sjostrand, M., Lee, J.J., and Lotvall, J.O. (2007). Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell Biol. 9, 654-659 van der Pol, L., Stork, M., and van der Ley, P. (2015). Outer membrane vesicles as platform vaccine technology. Biotechnol. J. 10, 1689-1706 Wang, X., Thompson, C.D., Weidenmaier, C., and Lee, J.C. (2018). Release of Staphylococcus aureus extracellular vesicles and their application as a vaccine platform. Nat. Commun. 9, 1379-018-03847-z White, G.F., Racher, K.I., Lipski, A., Hallett, F.R., and Wood, J.M. (2000). Physical properties of liposomes and proteoliposomes prepared from Escherichia coli polar lipids. Biochim. Biophys. Acta 1468, 175-186 Wolvers, D., Antoine, J.M., Myllyluoma, E., Schrezenmeir, J., Szajewska, H., and Rijkers, G.T. (2010). Guidance for substantiating the evidence for beneficial effects of probiotics: prevention and management of infections by probiotics. J. Nutr. 140, 698S- 712S Table 1. Mean hydrodynamic diameter and Full Width Half Maximum (FWHM) of DLS intensity weighed size distribution obtained by CONTIN analysis of correlation function measured at 23°C of OMVs by Escherichia Coli growth at different temperature; reported errors are the standard deviation of three measurements. growth temperature (°C) mean diameter (nm) FWHM (nm) 37 27 20 48 ± 3 37 ± 4 24 ± 2 24 ± 2 32 ± 2 20± 3 Figure 1. Workflow of the procedures followed to isolate and purify bacterial vesicles for downstream measurements. E.Colicultureatdifferenttemperatures: 20, 27 and 37 °CCentrifugationto removebacteriaand collectthe supernatantUltrafiltrationand UltracentrifugationBACTERIAL CULTUREOD600monitoringFiltration0.45 mmVESICLE ISOLATIONWashingof vesiclepelletwith PBSCONCENTRATION OF THE SUPERNATANTPURIFICATIONQualityControl steps Figure 2. (a) Dynamic Light Scattering intensity-weighed distribution of vesicles by E. coli grown at 37°C. (b) Transmission Electron Microscopy image of vesicles by E. coli grown at 37°C obtained without negative staining. In the inset is shown a detail of a TEM image obtained with negative staining. Figure 3. (a) Mean scattering intensity (upper panel) and hydrodynamic diameter (lower panel) resulting from DLS measurements at increasing temperatures between 10°C and 45°C of vesicles grown at 37°C; red lines results from Boltzmann fit and linear fit respectively while the blue dashed line highlights the transition temperature. (b) Mean scattering intensity of a DLS measurement at increasing temperatures between 10°C and 35°C of E. coli bacteria grown at 37°C. Inset shows the mean scattering intensity up to 45°C (above 35°C the bacteria start to grow). Figure 4. Normalized scattering intensity as a function of temperature: comparison between membrane phase transitions of OMVs (a) and E.Coli bacteria (b) grown at 37°C, 27°C and 20°C; vertical dashed lines highlights the transition temperatures. Figure 5. Comparison between phase transitions temperature of OMVs and E. coli bacteria grown at 37°C, 27°C and 20°C; the error bars represent the transition widths whereas the blue ovals are guide to eyes that highlight the clear separation between transitions from different membranes. Figure 6. Outer membrane vesicle and E. coli models representing core and shell structures of different densities (as for SAXS analysis) or different refractive index (scattering intensity simulations). Figure 7. Normalized scattered intensity as a function of the temperature: comparison between simulated membrane phase transitions of OMVs (a) and E. coli bacteria (b) grown at 37°C, 27°C and 20°C. Figure 8. (a) SAXS spectra of vesicles maintained at temperatures between 10 and 45°C as in the legend. (b) Vesicles bilayer thickness as resulting from SAXS data fitting procedure, as a function of temperature. Figure 9. Mean scattered intensity of a DLS measurement of MVs by Lactobacillus rhamnosus LGG grown at 37°C monitored at temperatures between 10 and 45°C.
1309.4853
1
1309
2013-09-19T04:37:59
Bilayer registry in a multicomponent asymmetric membrane : dependence on lipid composition and chain length
[ "physics.bio-ph", "cond-mat.soft" ]
A question of considerable interest to cell membrane biology is whether phase segregated domains across an asymmetric bilayer are strongly correlated with each other and whether phase segregation in one leaflet can induce segregation in the other. We answer both these questions in the affirmative, using an atomistic molecular dynamics simulation to study the equilibrium statistical properties of a 3-component {\em asymmetric} lipid bilayer comprising an unsaturated POPC (palmitoyl-oleoyl-phosphatidyl-choline), a saturated SM (sphingomyelin) and cholesterol with different composition ratios. Our simulations are done by fixing the composition of the upper leaflet to be at the coexistence of the liquid ordered ($l_o$) - liquid disordered ($l_d$) phases, while the composition of the lower leaflet is varied from the phase coexistence regime to the mixed $l_d$ phase, across a first-order phase boundary. In the regime of phase coexistence in each leaflet, we find strong transbilayer correlations of the $l_o$ domains across the two leaflets, resulting in {\it bilayer registry}. This transbilayer correlation depends sensitively upon the chain length of the participating lipids and possibly other features of lipid chemistry, such as degree of saturation. We find that the $l_o$ domains in the upper leaflet can {\em induce} phase segregation in the lower leaflet, when the latter is nominally in the mixed ($l_d$) phase.
physics.bio-ph
physics
Bilayer registry in a multicomponent asymmetric membrane : dependence on lipid composition and chain length Anirban Polley1, Satyajit Mayor2, and Madan Rao1,2∗ 1Raman Research Institute, C.V. Raman Avenue, Bangalore 560080, India 2National Centre for Biological Sciences (TIFR), Bellary Road, Bangalore 560065, India E-mail: [email protected],[email protected] Abstract A question of considerable interest to cell membrane biology is whether phase segregated domains across an asymmetric bilayer are strongly correlated with each other and whether phase segregation in one leaflet can induce segregation in the other. We answer both these questions in the affirmative, using an atomistic molecular dynamics simulation to study the equilibrium statistical properties of a 3-component asymmetric lipid bilayer comprising an un- saturated POPC (palmitoyl-oleoyl-phosphatidyl-choline), a saturated SM (sphingomyelin) and cholesterol with different composition ratios. Our simulations are done by fixing the composi- tion of the upper leaflet to be at the coexistence of the liquid ordered (lo) - liquid disordered (ld) phases, while the composition of the lower leaflet is varied from the phase coexistence regime to the mixed ld phase, across a first-order phase boundary. In the regime of phase coexistence in each leaflet, we find strong transbilayer correlations of the lo domains across the two leaflets, resulting in bilayer registry. This transbilayer correlation depends sensitively upon the chain ∗To whom correspondence should be addressed 1 length of the participating lipids and possibly other features of lipid chemistry, such as degree of saturation. We find that the lo domains in the upper leaflet can induce phase segregation in the lower leaflet, when the latter is nominally in the mixed (ld) phase. Introduction Cell membranes are composed of many different lipid species and exhibit both lateral heterogene- ity1 -- 3 and bilayer asymmetry in their lipid composition.4 While there have been many in-vitro studies of lateral phase segregation in multicomponent giant unilamellar vesicles (GUVs)5 -- 7 and suspended membranes,5,6 it is only recently that attention has turned to membranes with asymmet- ric bilayers.8,9 One of the issues highlighted in these experiments and relevant to cell membrane biology, is the extent of correlation or registry of phase segregated domains in the two leaflets of the bilayer. This has inspired theoretical10,11 and coarse-grained computer simulation12 studies of bilayer registry of domains in asymmetric bilayers. A well studied multicomponent model system is the 3-component lipid mixture comprising an unsaturated lipid (POPC), a saturated lipid (PSM) and cholesterol (Chol) which exhibits liquid-ordered (lo) - liquid-disordered (ld) phase coexistence. Since the extent of bilayer registry is likely to be sensitive to lipid chemistry, in this paper we study the transbilayer coupling and extent of bilayer registry of the phase domains across the membrane, using an atomistic molecular dynamics (MD) simulation of an asymmetric lipid bilayer membrane comprising POPC/PSM/Chol. The registry of lipids across the bilayer suggest a mechanism by which outer leaflet lipids may couple with inner leaflet lipids and vice versa. This is important in the construction of membrane domains, and more generally in transducing information across the bilayer by lipidic receptors such as GPI-anchored proteins (GPI-APs)13 or glycolipids14 and other lipid species. Our moti- vation for this work also comes from a series of experiments that study the spatial organization and dynamics of GPI-APs, on the surface of living cells. A variety of experimental strategies such as Fluorescence Resonance Energy Transfer (FRET),4,15 -- 18 near-field microscopy (NSOM) and 2 electron microscopy, have revealed that both the organization and dynamics of the outer-leaflet GPI-APs are regulated by cholesterol, sphingolipids and cortical actin and myosin at the inner leaflet. The question is how do the outer-leaflet GPI-APs couple to cortical actin that abuts the inner leaflet of the cell membrane.17,18 Since the interaction across the bilayer must be indirect, are there other lipids, such as cholesterol and sphingolipids, that are involved in this linkage ? Do specific inner leaflet lipids that interact with actin, participate in this transbilayer coupling18 ? This naturally brings up the issue of bilayer registry in the cell membrane and its dependence on the specificity of lipids and its chemistry. We have been addressing these important issues using both experiments on live cells and atomistic molecular dynamics simulations on multicomponent model membranes composed of so-called 'raft-components'.1 The article is organized as follows : we first describe the details of the atomistic molecular dynamics (MD) simulation of the 3-component bilayer. Next we present our results on lateral compositional heterogeneity, extent of bilayer registry and mismatch area across the bilayer, as a function of the concentration of the saturated lipid (PSM). We also study how changes in lipid chain length of SM affect bilayer registry. We end with a short summary of the results and conclusions. Methods Model membrane : We study the phase segregation and bilayer registry of a symmetric and asym- metric 3-component bilayer membrane embedded in an aqueous medium by atomistic molec- ular dynamics simulations (MD) using GROMACS. We prepare the bilayer membrane at 23◦C at different relative concentrations of palmitoyl-oleoyl-phosphatidyl-choline (POPC), long chain palmitoyl-sphingomyelin (SM-16:0) (PSM) and cholesterol (Chol). All multicomponent bilayer membranes have 512 lipids in each leaflet (with a total 1024 lipids) and 32768 water molecules (such that the ratio of water to lipid is 32 : 1) so as to completely hydrate the simulated lipid bilayer. For the symmetric bilayer, the relative concentration (in percentage, x) of PSM and Chol in the upper and lower leaflet is varied from x (in%) = 33.3,19.9,12.5,10.0,9.1,7.1,5.8,2 and 1%, with 3 POPC contributing to the rest of the lipid content. For the asymmetric bilayer, the upper leaflet has POPC /PSM /Chol in the ratio 1 : 1 : 1 (i.e., the relative concentration of each component is 33%). We vary the relative composition in the lower leaflet; denoting x as the relative concentration (in percentage) of PSM and Chol in the lower leaflet, we run through the values x (in%) = 33.3,25.0,19.9,14.3,12.5, 10.0,9.1,7.1,5.9 and 4.5%, with POPC contributing to the rest of the lipid content. With this choice of compositions and temperature, the upper leaflet is in the putative lo-ld phase coexistence regime (see 1A, for the ternary phase diagram at 23◦C, taken from Ref. [17]), while in the lower leaflet the compositions straddle the phase boundary allowing us to go from the lo-ld phase coexistence regime to the ld phase, 2A. To study the role of lipid chemistry, we repeat the above set of simulations with PSM in the lower leaflet replaced by the short chain sphingomyelin, SM-14:0 (MSM). We vary the concentra- tion x of MSM (Chol) across the range x (in%) = 4.5,5.9,7.1,9.1,10.0,12.5 and 14.35%. Figure 1: (A) Ternary phase diagram of POPC, PSM and CHOL at T = 23◦C taken from Ref. [25]. Triangle (orange) represents the composition 1 : 1 : 1 which is deep in the phase coexistence region; Dot (red) is a point on the phase boundary xc = 10%. (B) We have verified this phase diagram by doing simulations at different compositions along the blue line in (A). Panel shows the probability distribution of the phase segregation order parameter φ (main text) for the symmetric bilayer at different values of x : (i) - (iii) shows that for x < 10%, the bilayer is in the ld phase, while (v) - (vi) shows that for x > 10%, the membrane is at lo-ld coexistence. The phase transition is clearly at (iv) xc = 10%. 4 Force fields : The force field parameters for POPC, PSM and Chol are taken from the previous validated united-atom description.19 -- 21 We construct the force field parameters for MSM (SM- 14:0) from the parameters of PSM and POPC. We use the improved extended simple point charge (SPC/E) model to simulate water molecules, having an extra average polarization correction to the potential energy function. Initial configurations : We generate the initial configurations of the asymmetric multicomponent bilayer membrane using PACKMOL.22 For all simulation runs, we choose two sets of initial con- ditions : (i) where the components in each leaflet are homogeneously mixed and (ii) where the ternary components are completely phase segregated in lo-ld domains.19 Choice of ensembles and equilibration : The asymmetric bilayers are equilibrated for 50 ps in the NVT ensemble using a Langevin thermostat to avoid bad contacts arising from steric constraints and then for 160 ns in the NPT ensemble (T = 296 K (23◦C), P = 1 atm). The simulations are carried out in the NPT ensemble for the first 20 ns using Berendsen thermostat and barostat, then for 20 ns using Nose-Hoover thermostat and the Parrinello-Rahman barostat to produce the correct ensemble. Rest of the simulations are performed in the NPT ensemble using Berendsen thermo- stat. We use a semi-isotropic pressure coupling with compressibility 4.5 × 10−5 bar−1 for the simulations in the NPT ensemble. The long-range electrostatic interactions are incorporated by the reaction-field method with cut-off rc = 2 nm, while for the Lennard-Jones interactions we use a cut-off of 1 nm.19,21,23 For each initial configuration, we run the simulations for 200 ns before computing the desired physical quantities. To ensure that the bilayer membrane is well equilibrated, we monitor the area per lipid throughout the simulations (Supplementary Figure S1-S2). We calculate the lateral pres- sure profiles in the bilayer using Irving-Kirkwood contour and grid size 0.1 nm. We calculate the pairwise forces by rerunning the trajectory with cut-off 2 nm for electrostatic interactions using LINCS algorithm to constrain the bond lengths24 and the SETTLE algorithm to keep the water 5 molecules rigid25 so that integrator time step of 2 fs can be used. We generate pressure profiles from trajectories over 20 ns using SHAKE algorithm26 to constrain bond lengths. Computation of deuterium order parameter : We calculate the spatial distribution of the deuterium order parameter S from the selected carbon atoms (C5−C7) of each acyl chain (including SN1 and SN2 chains) of the PSM and POPC lipids.19,21 Here, S is defined for every selected CH2 group in the chains as, S ≡ 1 (cid:104)3cos2 θ − 1(cid:105) where θ is the angle between a CH-bond and the normal to 2 the plane of the membrane (z-axis). This is then coarse-grained (binned) over a spatial scale of 0.5 nm for last 20 ns of the trajectory of the simulations. We use our previous estimation of the deuterium order parameter S of the lo-ld domains of the bilayer membrane,19 to declare a region to be liquid-ordered (lo) when the value of S ≥ 0.35. Figure 2: (A) For the simulations of the asymmetric bilayer, we hold the composition of the up- per leaflet at 1 : 1 : 1, while in the lower leaflet the composition of PSM and Chol is varied from x = 33.3% (orange dot) to 0% (green square), with POPC contributing to the rest. The compo- sition where the simulations are carried out, denoted by black dots (see Methods), are indicated against the reference ternary phase diagram of the symmetric bilayer (blue line in Fig. 1A). Tri- angle (orange) represents the composition 1 : 1 : 1 which is deep in the phase coexistence region; Dot (red) is a point on the phase boundary xc = 10%. (B) Snapshot of equilibrium configuration of the asymmetric bilayer when x = 25%, with POPC (gray), PSM (orange), Chol (yellow) and water (cyan). (C) Lateral pressure profile π(z) for the same bilayer at equilibrium. Computation of mismatch area : We calculate the coarse-grained spatial profile of the deuterium order parameter, S in each leaflet using grid size 0.5nm. We then use the above cutoff in S to declare a region as liquid ordered. We compute the area and perimeter of the lo domains in each leaflet using the cluster algorithm available in Image Processing Toolbox, MATLAB 2009. This 6 is used to calculate the overlap and mismatch area of the domains across the bilayer (see, Section Mismatch area and interfacial tension). Results and Discussion We compute the local stress profile of the bilayer membrane from the virial, and use this to cal- culate the net surface tension, force and torque. We ensure that the prepared bilayer membrane is mechanically stable, with both the net force and torque balanced. In addition we ensure that the surface tension is zero to within numerical error. The details of the mechanically stable sym- metric bilayer have appeared in an earlier publication.19 1A shows the phase diagram of the sym- metric bilayer comprising POPC, PSM and Chol at 23◦C taken from Ref. [25]. We have simu- lated the symmetric bilayer membrane composed of POPC, PSM and Chol with concentration, x = 1%,2%,5.8%,7.1%,9.1%,10%,12.5%,19.9% and 33.3% of the PSM (Chol). We have plot- ρPSM − ρPOPC ρPSM + ρPOPC ted P(φ ) with different x for the symmetric bilayer where, φ is defined as, φ = (1B). For details of the mechanical stability of the asymmetric bilayer, see Supplementary Tables S1- S6, where we record the net force, torque and surface tension at each composition of the asymmet- ric bilayer in tabular form. Here, we show a snapshot of the ternary asymmetric bilayer membrane composed of POPC, PSM and Chol and its lateral pressure profile π(z), 2 B and C, respectively (profiles at other concentrations are displayed in Supplementary Figure S3-S4). We perform simulations on our model asymmetric bilayer at varying concentrations x of PSM and Chol in the lower leaflet, whilst maintaining the upper leaflet at a composition 1 : 1 : 1, which is deep in the lo-ld phase coexistence region. The simulations done at various values of x along the line shown in 2A, traverses across the phase boundary at xc = 10% into the ld phase. We perform a similar study when the lower leaflet PSM is replaced by the short chain sphin- gomyelin, MSM. 7 Figure 3: Spatial profile of the coarse-grained number density of PSM (color bar) in the upper leaflet (top panel) at different compositions of PSM in the lower leaflet (as indicated at the top of the panels) The middle panel shows the corresponding profile of the number density of PSM (color bar) in the lower leaflet. The bottom panel shows the extent of correlation between the PSM-rich domains across the bilayer, as measured by the joint probability distribution, JPD (color bar, see text). For x > xc = 10%, the JPD show strong transbilayer correlations, while for x (cid:28) xc, the correlations are poor. Lateral compositional heterogeneity The coarse-grained spatial profile of the lipid number density is calculated with a grid size 1.3 nm. As stated in Methods, the composition in the upper leaflet is fixed at 1 : 1 : 1, while the composition of PSM/Chol in the lower leaflet is varied from 33% to 4.5%. The top and middle panels in 3 show the spatial profile of the number density of PSM in the upper and lower leaflets, respectively, at 4 representative compositions on either side of the phase boundary, xc = 10%. The lower panel, described in the next section, shows the joint correlation of the PSM rich domains across the bilayer. 4 shows a similar study done when PSM in the lower leaflet is replaced by short chain MSM. 8 Figure 4: Spatial profile of the coarse-grained number density of PSM (color bar) in the upper leaflet (top panel) at different compositions of MSM in the lower leaflet (as indicated at the top of the panels) The middle panel shows the corresponding profile of the number density of MSM (color bar) in the lower leaflet. The bottom panel shows the extent of correlation between the PSM-rich domain in the upper leaflet and the MSM-rich domain in the lower leaflet, as measured by the joint probability distribution, JPD (color bar, see text). As in 3, for x > xc = 10%, the JPD show strong transbilayer correlations, while for x (cid:28) xc, the correlations are poor. Domain registry across bilayer We have studied the extent of registry of lo-ld domains across the bilayer of an asymmetric multi- component membrane as a function of varying composition and lipid chemistry. We measure the extent of transbilayer registry by computing the joint probability distribution (JPD) of the coarse- grained number density of PSM in the upper and lower leaflets at the same coarse-grained spatial location (x,y). The lower panel of 3 shows the JPD at different values of the concentration x of PSM in the lower leaflet. In the lower panel in 3, the JPD shows a distinct peak along the diagonal when x > xc = 10%, which is clear evidence of bilayer registry of lo-domains. The off-diagonal peak in the JPD is merely an indication of the relative abundance of PSM in the upper leaflet. On the other hand, for 9 x (cid:28) xc, this diagonal peak in the JPD is absent, indicating lack of bilayer registry. A similar conclusion regarding the bilayer registry can be drawn when the lower leaflet PSM is replaced by the short chain MSM (4). These observations suggest that the configurations of the two leaflets mutually influence each other. As stated in the Abstract, we can ask whether the segregation of lipids in the upper leaflet can induce a phase segregation in lower leaflet, i.e., can the composition in the upper leaflet act as a local "field" for the composition in the lower leaflet. To study this, we define a 'transbilayer order-parameter' from the normalized transbilayer correlation (r denotes the 2d coordinate (x,y)), C(ρu PSM(r),ρd PSM(r)) = (cid:113)(cid:104)ρu (cid:104)ρu PSM(r)ρd PSM(r)2(cid:105)−(cid:104)ρu PSM(r)(cid:105)−(cid:104)ρu PSM(r)(cid:105)2(cid:113)(cid:104)ρd PSM(r)(cid:105)(cid:104)ρd PSM(r)(cid:105) PSM(r)2(cid:105)−(cid:104)ρd PSM(r)(cid:105)2 (1) averaged over space (denoted by Cud) and compute this as a function of the relative concentration x of PSM/Chol. Supplementary Figure S5 shows the transbilayer order-parameter Cud as a function of x for a symmetric bilayer. Cud jumps from a high value in the lo − ld phase coexistence region to a low value in the ld phase. The jump in Cud coincides with the phase boundary xc = 10% (1). For the asymmetric bilayer, we compute the transbilayer order parameter as a function of x, the concentration of PSM (or MSM) in the lower leaflet (5). The transbilayer order parameter Cud is very nearly zero for x (cid:28) xc and rises sharply to ∼ 1 at x = xPSM c < xc, showing the influence of the upper leaflet on the phase segregation of the lower. This transbilayer influence is stronger for the long chain PSM than for the short chain MSM, as seen by the fact that xMSM = 9.09% > xPSM c = c 5.88%. There is thus a shift in the phase boundary from its value of xc = 10% for the ternary symmetric − xc for both the long chain bilayer of POPC-PSM-Chol. This shift is plotted as ∆ = xPSM/MSM PSM (∆PSM = 4.12%) and short chain MSM (∆MSM = 0.91%) in the lower leaflet (inset 5). c The above phenomenology can be understood within a mean-field theory of phase transitions,28 with a Helmholtz free-energy functional written in terms of φu and φd, where φu (φd) is the relative concentration of the lo and ld species in the upper (lower) leaflet. The form of the free-energy 10 Transbilayer order parameter Cud defined from the transbilayer correlation Figure 5: C(ρu(r),ρd(r)) (see text) between the density of upper leaflet PSM and lower leaflet PSM (red circle) or MSM (black triangle) versus x, the concentration of PSM or MSM in the lower leaflet. Color panel below drawn for reference, denotes the values of x at which Cud has been evaluated. The value of Cud is zero for small x and jumps sharply at xPSM/MSM < xc = 10% (red dot in color panel), indicating a first-order phase transition. The phase transition point for the long chained PSM, xPSM = 9.09%. Inset shows c the shift in the transition ∆ (see text) as a function of lipid chain length. c = 5.88% is smaller than that of the short chained MSM xMSM c functional for the asymmetric bilayer can be written as, (2) F[φu,φd] = d2x [ fu(φu) + fd(φd) + fud(φu,φd)] u + vφ 4 u (cid:90) 1 2 u + uφ 3 fu = Cu(∇φu)2 − ruφ 2 fd = Cd(∇φd)2 + rdφ 2 d fud = −Aφuφd + Bφuφ 2 d + Dφ 2 11 u φd c),rd ∼ (x− xd where ru ∼ (x− xu c ) > 0 reflects the fact that the upper leaflet is in the lo-ld phase coexistence regime, (cid:104)φu(cid:105) (cid:54)= 0, and the isolated lower leaflet is in the ld phase, (cid:104)φd(cid:105) = 0. The coefficient A > 0 to account for the fact that the local transbilayer coupling is attractive. We first minimize F with respect to φu : setting δ F/δφu = 0, and keeping terms to linear order, we get Cu∇2φu + ruφu = Aφd , whose Fourier transform, lends itself to a useful interpretation, φu(q) = Aφd(q) −Cuq2 + ru , (3) (4) namely a spatially varying φu can induce a spatially varying φd. Nonlinearities in the free-energy that we have neglected, reinforce this and will lead to bilayer registry. Plugging this expression back into Eq. (3), we obtain an effective free-energy functional in terms of φd alone, which shows that the coefficient of the quadratic term gets reduced by rd → rd −A2/ru, which for large enough A can become negative. This shows that the segregation in the upper leaflet can induce a segregation in the lower, by shifting the phase transition point. This mean field analysis is entirely consistent with our MD simulations. Mismatch area and interfacial tension When there is perfect bilayer registry, the area of the lo domain in the upper leaflet will completely overlap with the area in the lower leaflet (6A). Any mismatch in the overlap area will cost energy proportional to the mismatch area A, defined as A = Au lo are the areas of the lo-domains in the upper and lower leaflets and Ao is the overlap area between the lo-domains in the upper and lower leaflets (6B). The proportionality constant is a tension γ or a mismatch free lo −2Ao, where Au lo +Ad lo and Ad energy per unit area, and is a measure of the domain overlap, a larger value of γ implies a more complete overlap. This tension γ acts as a driving force for inter-leaflet registration of the phase domains across the bilayer. In principle, the value of the tension γ is affected by short wavelength 12 curvature and protrusion fluctuations, which we have ignored in our computation of the area - this will typically go to reduce the value of γ. The linear dependence of the energy on the mismatch area A holds as long as the mean size of the mismatch region (cid:104)R(cid:105) is larger than its root mean square fluctuation w = δ R = R(θ ) − (cid:104)R(cid:105), and R(θ ) is the distance from the domain centre to the domain boundary at the angular position θ. There are strong corrections to this leading behaviour, of order (w/(cid:104)R(cid:105))2, when the domains are small or ramified. Given that the lateral dimension of the model membrane (cid:113)(cid:104)δ R2(cid:105), where is 15.6 nm, this is likely the case in our atomistic MD simulations. To check this, we have plotted the perimeter per area of the mismatch region (Supplementary Figure S6) versus area, at different values of x, the relative concentration of PSM in the lower leaflet - this shows that the mean domain shapes deviate from circularity, especially for small values of x. With this caveat, we have estimated the domain interfacial tension γ by computing the proba- bility distribution of the mismatch area A of the lo-domains between the two leaflets of the bilayer, and equating it to the Boltzmann form, P(A) ∝ exp(−γA), where γ is measured in units of kBT . In Supplementary Figure S7, the plot of the probability distribution of A at various values of x, shows a distinct peak at the most probable value of A; in a semi-log plot 6C, we fit the distribution to the Boltzmann form to extract the value of the tension γ. These values, at x well within the coexistence region, for instance γ = 0.146±0.02 kBT/nm2 at x = 33%, are consistent with those obtained from other coarse-grained simulations.29,30 Given the systematic errors in such a computation and the caveats mentioned above, we should regard this computed value of γ with some caution. Notwith- standing, the qualitative trend showing γ decrease with x, with a sharp drop to zero at x (cid:39) xPSM (6D), is reassuring. c Conclusion We have analyzed the equilibrium properties of a ternary component, asymmetric bilayer mem- brane using atomistic molecular dynamics study. Our central goal was to study the conditions 13 Figure 6: Schematic showing (A) domains in complete registry or overlap across the bilayer and (B) domains in partial overlap with a defined mismatch area (see text). (C) Semi-log plot of the probability distribution of the mismatch area, −lnP(A) vs. A (red dots), at different values of x, the concentration of PSM in the lower leaflet (indicated in the panel). The straight lines in the high A regime are fits to the Boltzmann form (see text), from which we extract the value of the tension γ. Error bars are indicated. (D) Tension γ as a function of x shows a sharp drop to zero at x (cid:39) xPSM . c under which bilayer registry takes place in an asymmetric, multicomponent membrane. To sum- marize, our main results are: (i) lo phase domains formed in the two leaflets are registered across the bilayer membrane, (ii) phase segregation in upper leaflet can induce segregation in the lower, thus the composition on the upper leaflet acts as a "field" which couples linearly to the composi- tion in the lower leaflet and (iii) the strength of the transbilayer coupling and the extent of bilayer registry depends sensitively on the lipid chain length and is greater for longer chain lipids. The registry of the phase domains across the two leaflets of the bilayer membrane has an impor- tant implication to the sorting and signaling in live cell membrane. The cell membrane is inherently asymmetric with both lateral and transverse lipid heterogeneity. Recent experiments on live cells, using Fluorescence Resonance Energy Transfer (FRET)16 -- 18 show that outer leaflet GPI-APs or- 14 ganized as monomers and cholesterol-sensitive nanoclusters are regulated by the active dynamics of cortical actin (CA) and myosin. The present work forms the basis for further investigation of the transbilayer interaction between lateral heterogeneities of the outer leaflet GPI-anchored pro- teins, PSM and cholesterol with saturated, long chain lipids at the inner leaflet whose organization depends on the actin and actin remodeling proteins. Acknowledgements SM is a JC Bose Fellow (DST, Govt of India) and acknowledges support from an HFSP grant. This work was partially supported by a grant from Simons Foundation. References (1) Simons, K.; Ikonen, E. Nature 1997, 387, 569 -- 72. (2) Lingwood, D.; Simons, K. Science 2010, 327, 46 -- 50. (3) Simons, K.; Toomre, D. Nat Rev Mol Cell Biol 2000, 1, 31 -- 39. (4) Mayor, S.; Rao, M. Traffic 2004, 5, 231 -- 40. (5) Veatch, S. L.; Keller, S. L. Phys. Rev. Lett. 2002, 89, 268101. (6) Veatch, S. L.; Keller, S. L. Biophys J. 2003, 85, 3074 -- 83. (7) Baumgart, T.; Hess, S.; Webb, W. Nature 2003, 425, 821 -- 4. (8) Wan, C.; Kiessling, V.; Tamm, L. Biochemistry 2008, 47, 2190 -- 8. (9) Collins, M.; Keller, S. Proc Natl Acad Sci USA 2008, 105, 124 -- 128. (10) Allender, D. W.; Schick, M. Biophys J. 2006, 91, 2928 -- 35. 15 (11) Putzel, G.; Uline, M.; Szleifer, I.; Schick, M. Biophys J. 2011, 100, 996 -- 1004. (12) Risselada, H.; Marrink, S. Proc Natl Acad Sci USA. 2008, 105, 17367 -- 17372. (13) Kusumi, A.; Koyama-Honda, I.; Suzuki, K. Traffic. 2004, 5, 213-230. (14) Hakomori, S. I. Biochim. Biophys. Acta. 2008, 1780, 325-346. (15) Hancock, J. F. Nat. Rev. Mol. Cell Biol. 2006, 7, 456 -- 462. (16) Sharma, P.; Varma, R.; Sarasij, R.; Ira; Gousset, K.; Krishnamoorthy, G.; Rao, M.; Mayor, S. Cell 2004, 116, 577 -- 89. (17) Goswami, D.; Gowrishankar, K.; Bilgrami, S.; Ghosh, S.; Raghupathy, R.; Chadda, R.; Vishwakarma, R.; Rao, M.; Mayor, S. Cell 2008, 135, 1085 -- 97. (18) Gowrishankar, K.; Ghosh, S.; Saha, S.; Rumamol, C.; Mayor, S.; Rao, M. Cell 2012, 149, 1353 -- 67. (19) Polley, A.; Vemparala, S.; Rao, M. J Phys Chem B. 2012, 116, 13403 -- 10. (20) Tieleman, D. P.; Berendsen, H. J. Biophys J. 1998, 74, 2786 -- 2801. (21) Niemelä, P. S.; Ollila, S.; Hyvönen, M. T.; Karttunen, M.; Vattulainen, I. PLoS Comput Biol. 2007, 3, e34. (22) Martínez, L.; Andrade, R.; Birgin, E.; Martínez, J. J Comput Chem. 2009, 30, 2157 -- 64. (23) Patra, M.; Karttunen, M. J. Phys. Chem. B 2004, 108, 4485 -- 4494. (24) Hess, B.; Bekker, H.; Berendsen, H. J. C.; Fraaije, J. G. E. M. J Comput Chem. 1997, 18, 1463 -- 1472. (25) Miyamoto, S.; Kollman, P. A. J. Comput. Chem. 1992, 13, 952 -- 962. (26) Ryckaert, J.-P.; Ciccotti, G.; Berendsen, H. J. J. Comput. Phys. 1977, 23, 327 -- 341. 16 (27) de Almeida, R. F. M.; Fedorov, A.; Prieto, M. Biophys J. 2003, 85, 2406 -- 2416. (28) Chaikin, P. M.; Lubensky, T. C. Principles of Condensed Matter Physics; Cambridge Uni- versity Press: 2000. (29) Watkins, E. B.; Miller, C. E.; Majewski, J.; Kuhl, T. L. Proc Natl Acad Sci USA. 2011, 108, 6975 -- 6980. (30) Garbés Putzel, G.; Uline, M. J.; Szleifer, I.; Schick, M. Biophys J. 2011, 100, 996 -- 1004. 17
1301.2996
2
1301
2013-04-25T19:37:18
Energetic Pulses in Exciton-Phonon Molecular Chains, and Conservative Numerical Methods for Quasi-linear Hamiltonian Systems
[ "physics.bio-ph", "math.NA" ]
The phenomenon of coherent energetic pulse propagation in exciton-phonon molecular chains such as $\alpha$-helix protein is studied using an ODE system model of Davydov-Scott type, both with numerical studies using a new unconditionally stable fourth order accurate energy-momentum conserving time discretization, and with analytical explanation of the main numerical observations. Physically natural impulsive initial data associated with the energy released by ATP hydrolysis are used, and the best current estimates of physical parameter values. In contrast to previous studies based on a proposed long wave approximation by the nonlinear Schr\"odinger (NLS) equation and focusing on initial data resembling the soliton solutions of that equation, the results here instead lead to approximation by the third derivative nonlinear Schr\"odinger equation, giving a far better fit to observed behavior. A good part of the behavior is indeed explained well by the linear part of that equation, the Airy PDE, while other significant features do not fit any PDE approximation, but are instead explained well by a linearized analysis of the ODE system. A convenient method is described for construction the highly stable, accurate conservative time discretizations used, with proof of its desirable properties for a large class of Hamiltonian systems, including a variety of molecular models.
physics.bio-ph
physics
Energetic Pulses in Exciton-Phonon Molecular Chains, and Conservative Numerical Methods for Quasi-linear Hamiltonian Systems Department of Mathematics, College of Charleston, South Carolina∗ Brenton LeMesurier (Dated: March 10, 2021) The phenomenon of coherent energetic pulse propagation in exciton-phonon molecular chains such as α-helix protein is studied using an ODE system model of Davydov-Scott type, both with numerical studies using a new unconditionally stable fourth order accurate energy-momentum conserving time discretization, and with analytical explanation of the main numerical observations. Physically natural impulsive initial data associated with the energy released by ATP hydrolysis are used, and the best current estimates of physical parameter values. In contrast to previous studies based on a proposed long wave approximation by the nonlinear Schrodinger (NLS) equation and focusing on initial data resembling the soliton solutions of that equation, the results here instead lead to approximation by the third derivative nonlinear Schrodinger equation, giving a far better fit to observed behavior. A good part of the behavior is indeed explained well by the linear part of that equation, the Airy PDE, while other significant features do not fit any PDE approximation, but are instead explained well by a linearized analysis of the ODE system. A convenient method is described for construction the highly stable, accurate conservative time discretizations used, with proof of its desirable properties for a large class of Hamiltonian systems, including a variety of molecular models. PACS numbers: 87.10.Ed, 87.10.Hk, 87.14.Ex, 87.14.et, 87.15.Ax, 87.15.ap Keywords: Davydov-Scott system, anharmonic, conservative time-discretization 3 1 0 2 r p A 5 2 ] h p - o i b . s c i s y h p [ 2 v 6 9 9 2 . 1 0 3 1 : v i X r a ∗ Currently visiting CSCAMM, University of Maryland, and the Department of Mathematical Sciences, George Mason University; [email protected]; http://blogs.cofc.edu/lemesurierb/ I. INTRODUCTION 2 Exciton-phonon systems of ODEs are used to model a variety of molecules in which mobile quantum excitations are present along with mechanical degrees of freedom. A. Davydov [Dav71, DK73] introduced such a model to study energy propagation in α-helix protein, present for example in the myocins, kenesins and actin involved in muscular contraction, in chains up to 2000 residues long. A modified version of Davydov's original equations, is used here, incorporating changes suggested by A. Scott [Sco84] and by Davydov and A. Zolotariuk in [DZ84]: i dψn dt − E0ψn + J(ψn−3 + ψn+3) − L(ψn−1 + ψn+1) = χ(qn+3 − qn)ψn. M d2qn dt2 = V (cid:48)(qn+3 − qn) − V (cid:48)(qn − qn−3) + χ(ψn2 − ψn−32), This will be called the Anharmonic Davydov-Scott system. Related exciton-phonon systems arise in other molecular models, such as the system i dψn dt + J(ψn−1 + ψn+1) = χ(qn+1 − qn−1)ψn, M d2qn dt2 = V (cid:48)(qn+1 − qn) − V (cid:48)(qn − qn−1) + χ(ψn+12 − ψn−12), (1) (2) (3) (4) used to model the conducting polymer polydiacetylene in [BECHO00]. This differs in having two-sided (symmetrical) form of the coupling, and only nearest neighbor interactions, but as should become clear below, the results herein adapt easily to differences such as these. We will consider in particular pulses in the exciton variables ψn that are generated by initial excitation at one end of the chain. It will be seen that the phenomena are well modeled by a subsonic limit leading to a Helically Coupled Discrete Nonlinear Schrodinger equation [HDNLS] i dψn dt + J(ψn−3 + ψn+3) − L(ψn−1 + ψn+1) + 2κψn2ψn = 0. (5) Further, an important part (but not all) of the pulse propagation can be described with a new long wave PDE approx- imation; not the nonlinear Schrodinger [NLS] model previously proposed by Davydov and considered in numerous subsequent papers, but a third derivative NLS equation ∂ψ ∂t + ∂3ψ ∂x3 + 2iκψ2ψ = 0 (6) also seen in related work of D. Pelinowsky and V. Rothos [PR05]. Section II introduces the various mathematical models and their Hamiltonian structures, symmetries and conserved quantities, explaining the successive approximations involved. Section III introduces the accurate, energy and momen- tum conserving numerical methods used; these are hopefully useful for a wide variety of similar Hamiltonian systems, due to advantages over the symplectic methods often used for such systems. Section IV presents numerical results, including demonstration of the high degree of accuracy of the successive model simplifications, and the inapplicability (for the present choices of initial data) of the NLS approximations used in various previous studies. Section V gives an analytical explanation for many of the phenomena observed, and ends by proposing some ideas for further study. II. MODELING EXCITON PROPAGATION IN α-HELIX PROTEIN AND OTHER POLYMERS A. The Anharmonic Davydov-Scott ODE System The primary mathematical model used here is the above Anharmonic Davydov-Scott system of ODEs (1,2), which modifies Davydov's original ODE model of α-helix protein by adopting a one-sided form for the exciton-phonon coupling (proposed by A. Scott [Sco84] based on the observations of V. Kuprievich and V. Kudritskaya [KK82]) and using a nonlinear force for the hydrogen bonds (as introduced by A. Davydov and A. Zolotariuk in [DZ84], and resembling the familiar FPU model). The helical structure of this protein has roughly three residues per twist, with 3 hydrogen bonds connecting third-nearest neighbors into nearly straight spines: spatial proximity leads to attractive exciton coupling along spines in addition to repulsive coupling between neighbors along the molecular backbone as the two dominant exciton interactions. Aside: many previous publications group residues into unit cells of three residues labelled m, with the residues within each unit cell labelled by a spine index σ = 1, 2, 3, but here a single index is more convenient, with explicit third-nearest neighbor interactions. The variables and parameters in this system are as follows. • Index n labels amino acid residues. • The exciton variables ψn give the probability of excitation of the amide-I mode in residue n, governed by a second quantization Hamiltonian Hex = E0ψ∗ nψn − J(ψ∗ nψn+3 + ψ∗ n+3ψn) + L(ψ∗ nψn+1 + ψ∗ n+1ψn), where J measures the (attractive) interaction between excitons in residues that are adjacent along a spine and L measures the (repulsive) interaction between excitons in residues that are adjacent along the molecular backbone. • The phonon variables qn are the displacements of the residues from rest position in the direction of the axis of the helix (that is, along spines), with momenta pn = M qn: these are associated to the phonon Hamiltonian (cid:88) n Hph = p2 n 2M + V (qn+3 − qn) with M the effective mass of each amino acid residue and V (r) a potential modeling the hydrogen bond force. • Parameter χ measures the effect of bond-stretching on the excitons through interaction Hamiltonian Hint = χ(qn+3 − qn)ψ∗ nψn. In fact the E0 term can be eliminated with the transformation ψn → eiE0tψn, so this is done from here on. Also, the only anharmonic potential considered is the cubic V (r) = K 2 r2 − γ 3 r3, (7) and in fact it will be demonstrated that for the situation studied herein, it is quite adequate to approximate with the r2, K = V (cid:48)(cid:48)(0), as indeed was done by Davydov originally. This leads to the original harmonic potential V (r) = "harmonic" version of the Davydov-Scott system as proposed by A. Scott in [Sco84], with phonon equation K 2 M d2qn dt2 = K(qn−3 − 2qn + qn+3) + χ(ψn2 − ψn−32). Either form of the system is Hamiltonian, with H = Hex + Hph + Hint and i dψn dt = dqn dt = ∂H ∂ψ∗ n ∂H ∂pn , , i dψ∗ n dt = − ∂H ∂ψn , dpn dt = − ∂H ∂qn . (8) (9) (10) Parameter Values. As the results herein are quite robust under variations in the parameter values within the likely range for α-helix protein, it is for the most part sufficient to use the values reported in [Sco82, Sco84], which facilitates comparisons to numerous other publications that use those values. The exciton couplings are best expressed through the frequencies J = J/ ≈ 1.47 THz, L = J/ ≈ 2.33 THz. 4 The linear stiffness of the hydrogen bond is K ≈ 13 N/m. The effective residue mass M is less precisely known, due in part to potential dependence on the particular sequence of amino acids, but it is sufficient to use the typical value M ≈ 0.127 zg, which leads to a typical phonon frequency ω0 =(cid:112)K/M ≈ 10.1 THz, because it will be seen that the only importance here is that this frequency is substantially larger than the above exciton frequencies. This puts us in the subsonic regime: exciton pulses travel at distinctly lower speeds than the phonons. As a further consequence, it will be seen in Section IV that the subsonic limit M → 0 (so also ω0 → ∞) gives the above HDNLS equation (5), and this approximation is seen in numerical studies to be highly accurate for any physically relevant value of M . Variation of the interaction coefficient χ has more significant effects, and despite the precise computed value of 34 pN cited by [Sco82] and various subsequent papers, there is still substantial uncertainty as to its value: the best current estimate appears to be the broad range of experimental values χ ≈ 35 -- 62 pN, with computed values subject to far greater uncertainty, even as to its sign [FMC10]. Thus the effect of varying this parameter will be studied: fortunately, it will be seen that the results herein depend only mildly on this value, with even the linearization χ = 0 giving useful information. Boundary Conditions. The boundary conditions at the ends of the chain depend on if and how the helix is connected to other parts of the molecule, but here the simplest, unbound form is assumed: "out of bounds" values of ψn and of the bond-stretchings rn := qn+3 − qn are effectively neglected in the Hamiltonian so for such index values ψn = 0, rn = 0. (11) For constructing simplified PDE models via a long wave approximation, it is also convenient to consider an infinite chain with n ∈ N and ψn → 0, rn → 0 as n → ∞. Initial Data. The initial data considered will be the physically plausible cases for an initial excitation caused by the energy release in ATP hydrolysis: primarily initial excitation at one residue. The most interesting phenomena will be seen to arise from excitation at one end of the chain, so ATP hydrolysis can also excite a pair of neighboring residues, so there will be brief comments on the variant ψ1(0) = ψ2(0) = 1. An initially still chain is used: qn(0) = pn(0) = 0. ψ1(0) = 1, ψn(0) = 0 for n > 1. (12) The equations above have a conserved exciton number E =(cid:80) n ψ∗ B. Momenta (conserved quantities other than the Hamiltonian) nψn. This is related to the probability density of quantum mechanics, but as noted above, it need not be unity, due to the possibility of multiple initial excitations. This invariant is associated via Noether's Theorem with a linear symmetry group action, the gauge symmetry The Davydov-Scott system also has a conserved momentum Pσ on each spine: Pσ =(cid:80) ψn → eisψn, ψ∗ n → e−isψ∗ n. m p3m+σ associated with spine translation symmetries q3m+σ → q3m+σ + sσ. However, conservation of linear momentum is respected by almost any reasonable time discretization (for example, any Runge-Kutta method) so no more will be said about this. (13) C. Approximation by a Helically Coupled Nonlinear Schrodinger Equation frequency ω0 =(cid:112)K/M is considerably higher than the exciton coupling frequencies J and L, and in practice exciton The Davydov-Scott system has several disparate scales in both space and time, and these can be used to derive simpler approximations. The first is that for physically relevant initial data, it will be seen in the numerical results of Section IV that the bond-stretchings rn are of small amplitude so that the linearized force −Kr is an adequate approximation, corresponding to harmonic potential V (r) = K 2 r2. Next is the subsonic limit approximation: the phenomena are on an even slower scale, so that variation in the amplitude ψn is far slower that that of the mechanical variables qn. For small M , solving Eq. (8) by variation of parameters gives rn = qn+3 − qn = − χ K ψn2 + oscillations of characteristic frequency ω0 and it is plausible that the excitons respond primarily to the slowly varying moving average part, which is given by setting M = 0 in (8). Using this moving average approximation ψn2 (14) rn ≈ − χ K in the exciton equation (1) eliminates the mechanical variables, reducing the model to the Helically Coupled Discrete Nonlinear Schrodinger [HDNLS] equation (5), with 5 This has Hamiltonian (cid:88) n H = −J(ψ∗ κ := χ2 2K ≈ 0.45 -- 1.4 THz. nψn+3 + ψ∗ n+3ψn) + L(ψ∗ nψn+1 + ψ∗ n+1ψn) − κ(ψ∗ nψn)2. (15) The validity of this approximation is demonstrated numerically in Section IV below. III. ENERGY-MOMENTUM CONSERVING TIME DISCRETIZATIONS To study these systems and assess the adequacy of the above HDNLS approximation, some numerical solutions should be considered. For that, the necessary numerical methods will now be described, and this is done for a general Hamiltonian system = J ∇yH(y) = J ∂H ∂y dy dt (y) (16) with J an anti-symmetric matrix. Notation. We will focus on the time advance map for single time step, from a time t to t + δt. For any scalar variable y or vector y, we use the variable's name alone to denote its value at time t, t+ = t+δt, y+ = y(t+) = y(t+δt), δy = y+ − y, and y = y + y+ . 2 A. Discrete Gradient Methods for Exact Energy Conservation Exact conservation of invariants has been seen to be a desirable feature of numerical methods for Hamiltonian systems; see for example [HLW06]. Following ideas originating in the work of O. Gonzalez and J. Simo [Gon96, GS96], the first step is to ensure conservation of the Hamiltonian (energy) by approximating such a system by a discrete Hamiltonian system using a suitable discrete gradient approximation = J ∇yH(y, y+) δy δt that satisfies the Discrete Chain Rule ∇yf (y, y+) ≈ ∇yf (y). δf = ( ∇yf )(y, y+) · δy. This condition is assumed from now on, along with linearity and the consistency condition (17) (18) (19) (20) ( ∇yf )(y, y+) = ∇yf (y). lim y+→y (cid:69) (cid:68) Dy1 f (y, y+), . . . Component notation like ∇yf (y, y+) = (continuous) Hamiltonian systems: using in succession (19), (17), and the anti-symmetry of J , will occasionally be used. Conservation of energy is easily shown for a discrete gradient method by mimicking the familiar argument for δH δt = ( ∇yH)(y, y+) · δy δt = ( ∇yH)(y, y+) · J ( ∇yH)(y, y+) = 0. B. Choosing a Discrete Gradient that Also Respects Quadratic Momenta 6 Many such "energy conserving" discrete gradients can be found, but conserving other invariants (here all called momenta) requires an appropriate choice of the gradient approximation. It will be seen that there is a natural limitation to quadratic (including linear) momenta, but this is sufficient for most systems of physical relevance. Here the approach introduced in [LeM12b, LeM12a] is followed, based on three facts: 1. There is a unique discrete gradient for functions of a single variable y  ∇yf (y, y+) := y+ (cid:54)= y , (y), y+ = y δf δy df dy (21) following from the chain rule requirement (19). For polynomials, this simplifies in a way that avoids the division by zero issue, via 2. There is a unique time-reversal symmetric discrete gradient for a product of two variables ∇yyp+1 = yn + yn−1y+ ··· + (y+)n. δ(yjyk) = yjδyk + ykδyj which corresponds to evaluating the true gradient at the midpoint: ∇(yjyk)(y, y+) = ∇(yjyk)(y). In fact this extends to a discrete product rule based on δ(f g) = f δg + gδf. (22) (23) (24) (25) Thus linear terms in the equations, corresponding to quadratic terms in the Hamiltonian, are discretized exactly as for the implicit midpoint rule, which is a popular momentum conserving symplectic method for Hamiltonian systems. The only differences are for nonlinearities, which for the systems of interest herein are those coming from the Hamiltonian terms χ(qn+3 − qn)ψ∗ nψn, γ 3 (qn+3 − qn)3 for Eq's (1,2) , κ(ψ∗ nψn)2 for Eq. (5). (26) 3. Many physically relevant Hamiltonian systems with conserved momenta have a natural form in which all the momenta are quadratic (including linear) functions of the state variables, and are related through Noether's theorem to a group of affine symmetries of the Hamiltonian H, with invariance of H manifested by the fact that it can be expressed as a composition where each component Qm of the new state vector Q is a quadratic myj, (each Am := {Ajk bj m} symmetric) H(y) = H(Q), (cid:88) Qm = 1 2 (cid:88) Ajk m yjyk + j,k j (27) (28) that is invariant under the symmetry group. For example, with the systems seen herein, the invariant quadratics with which the Hamiltonian can be expressed are the exciton products en,m = ψ∗ nψm and the bond-stretchings rn. In particular, the nonlinear terms seen here are χrnen,n, γ 3 r3 n, and κe2 n,n. (29) The discrete Jacobian of this change of variables is given by the true Jacobian evaluated at the midpoint: DyQ(y, y+) = DyQ(y). These facts and the above chain rule requirement naturally lead to: ∇yH = Dm H(Q, Q+) ∇yQm(y, y+), = Dm H(Q, Q+) ∇yQm(y). (cid:88) (cid:88) 7 (30) m m For the nonlinearities herein, the discrete gradients are now determined by the factorizations in (29) through simple forms: Dr(re) = e, De(re) = r, De(e2) = 2e, Dr(r3) = r2 + rr+ + (r+)2, (31) using (22) for the last. Using such a discrete gradient, energy and momenta will be conserved with any choice for the factors Dm H(Q, Q+). In practice, the above rules for single variable functions, products, compositions, and linearity are generally enough to construct a suitable discrete gradient for H. Theorem 1 For a Hamiltonian system as described above, and thus with a discrete gradient = J ∇yH(y), H(y) = H(Q) Dm H(Q, Q+) ∇yQm(y), dy dt ∇yH = (cid:88) = J(cid:88) m m solving numerically by the corresponding discrete gradient method y+ − y δt Dm H(Q, Q+) ∇yQm (cid:18) y + y+ (cid:19) 2 (32) conserves the Hamiltonian and all the quadratic momenta. Proof of Theorem 1 Energy conservation is already established above, so consider conservation of an invariant Q. Such quadratics are in fact invariant for any Hamiltonian H = H(Q) constructed from the quadratic forms Qm as in (27), including the alternative choices Hm := Qm, and invariance of Q on each of those Hamiltonian flows means that 0 = dQ dt = ∇Q · J ∇Hm = ∇Q · J ∇Qm, so that we have vanishing of the Poisson brackets {Q, Qm}(y) := ∇Q(y) · J ∇Qm(y) = 0. (33) (34) Mimicking (33) for the discrete flow and using the fact from (24) that discrete gradients of quadratics are given by the true gradients at the midpoint gives = ∇Q · J ∇H = ∇Q(y) · J(cid:88) δQ δt (cid:88) ( Dm H)∇Qm(y) = ( Dm H(Q, Q+)){Q, Qm}(y). m m Evaluating (34) at y = y gives {Q, Qm}(y) = 0, so δQ = 0. C. Practical Implementation: an Iterative Solution Method The system of equations will be nonlinear (unless the Hamiltonian system itself is linear), so we need an iterative solution method. To exploit the quasi-linearity of the system to preserve linear stability properties and exact mo- mentum conservation without the cost of a full quasi-Newton method, we proceed as follows: construct successive approximations y(k) of y+ by solving Dm H(Q, Q(k−1)) ∇yQm(y(k)), (35) = J(cid:88) m y(k) − y δt 8 where y(k) = (y + y(k))/2 and Q(k−1) = Q(y(k−1)), and initialization can be with y(0) = y or some other suitable approximation of y+. That is, the nonlinear part ∇Q H is approximated using the current best available approximation y(k−1) of y+, while the linear terms are left in terms of the unknown y(k) to be solved for. This equation is linear in the unknown y(k), making its solution straightforward, particularly with the narrow coupling bandwidth of the coupling in the systems studied here. Much as above, we have: Theorem 2 Each iterate y(k) given by the above scheme (35) conserves all quadratic first integrals that are conserved by the original discrete gradient scheme (32). The proof is as for Theorem 1 except that the Poisson brackets are evaluated at(cid:0)y + y(k)(cid:1) /2. This approach to iterative solution also gives unconditional linear stability, since as noted above, for a linear system it is the same as the A-stable implicit midpoint method, and indeed only a single iteration is needed in that case. Energy is of course only conserved in the limit k → ∞, but iterating until energy is accurate within machine rounding error is typically practical: if this take too many iterations, it is better for overall accuracy to reduce the time step size δt to speed the convergence. D. Time Discretization for the Davydov-Scott System Applying the above results to the anharmonic Davydov-Scott system is mostly a matter of separating linear terms from nonlinear, applying the implicit midpoint rule to the former and using Eq. (31) in Eq. (35) for the latter: (cid:16) δψ(k) n δt i + J ψ (k) n−3 + ψ (k) n+3 (cid:17) − L(cid:0)ψn−1 + ψn+1 (cid:1) = (cid:16) χ  (cid:17) q(k−1) n+3 − q(k−1) n ψ (k) n , δp(k) n δt M δq(k) n δt = p(k) n , (cid:17) 2 − ψ(k−) n−3 2 q(k−1) n+3 − q(k−1) n =K(q(k) n ) + χ (cid:20)(cid:16) (cid:16)ψ(k−1) (cid:17)2 (cid:16) n+3 − q(k) (cid:17)2(cid:21) q(k−1) n+3 − q(k−1) (cid:16) n+3 − (q+)(k−1) − γ 3 (q+)(k−1) + + n n n . (cid:17)(cid:16) (q+)(k−1) n+3 − (q+)(k−1) n (cid:17) (36) (37) (38) E. Higher Order Accuracy by Symmetric Step Composition The methods seen so far are only second order accurate in time. Fortunately, the method of symmetric step composition, (developed by M. Creutz, A. Gocksch, E. Forest, M. Suzuki and H. Yoshida [CG89, For89, Suz90, Yos90] for use with symplectic methods, and reviewed by E. Hairer, C. Lubich, and G. Wanner in the book [HLW06]) gives a systematic way to construct methods of any higher even order while preserving all the interesting properties: conservation of the Hamiltonian and quadratic invariants, time-reversal symmetry, and unconditional stability. Numerical results are computed below by combining the above discrete gradient method with the fourth-order accurate Suzuki form of step composition [HLW06, Example II.4.3, p. 45]: compose five discrete gradient steps of lengths ρjδ, ρ1 = ρ2 = ρ4 = ρ5 = √ 1 4 − 3 4 ≈ 0.41, ρ3 = 1 − 4ρ1 ≈ −0.66. F. Comparisons to Other Methods The most commonly used conservative methods for Hamiltonian systems are symplectic methods, which can conserve momenta but cannot in general conserve energy, as described by a theorem of Z. Ge and J. Marsden [GM88]. In the present situation with stiff systems of ODEs and Hamiltonian not of purely mechanical form H(q, p) = K(p) + U (q), the preferred choices of symplectic method are the implicit midpoint method, higher order Diagonally Implicit Runge-Kutta [DIRK] methods, and fully implicit Gaussian Runge-Kutta methods. All DIRK symplectic methods are cognates of the energy-momentum methods described here, given by applying the same step composition procedures to the implicit midpoint method instead of to the discrete gradient method. It has been illustrated in [LeM12b, LeM12a] that the basic discrete gradient method can handle qualitative features of solutions better than the midpoint method, though this has not been tested directly when step composition is applied to each method. Gaussian symplectic methods can be desirable when the time step size is small enough to allow their solution by simple fixed point iteration, but are not cost effective for stiff systems, where an unconditionally stable iterative method such as that above is highly desirable. 9 IV. NUMERICAL RESULTS As the initial excitation due to ATP hydrolysis will occur at at most two residues, the initial state is very far from the slowly varying form assumed in long wave approximations by PDE's. Thus one question addressed here, as in earlier work like [Sco82, Sco84], is whether solutions with such initial data evolve into a form that can be well-approximated at later times by a smooth function of position, leading to a hopefully more tractable PDE model. Time Step Choice. The choice of time steps here is always cautiously constrained by (cid:32) (cid:33) , δt ≤ min 1 2( J + L) , 1 ω0 which satisfies the natural accuracy and stability requirements for explicit methods, and for convergence of basic fixed point iterative solution of the nonlinear schemes. However it is confirmed that accurate solutions, in the sense that all graphs of exciton data are completely indistinguishable from results with smaller time steps, are given for any time step size δt ≤ 1 2( J + L) depending only on the time scale manifested in the exciton evolution equation. Thus the time discretization is effectively handling any faster time scales in the mechanical variables in the innocuous way that one hopes for stiff modes to be handled by an unconditionally stable method, with no adverse effect on the accuracy of the more slowly evolving (exciton) variables. A. Numerical Observations for the Davydov-Scott and HDNLS Systems We first solve the Anharmonic Davydov-Scott system (1,2) with 1000 residues, hydrogen bond nonlinearity of cubic form (7) with γ = 4, and initial excitation at one end as in (12). Figure 1 is for χ = 35, the minimum of the likely range cited above, showing the exciton amplitude ψn at times t = 20 and 40. It reveals a dominant leading pulse of speed about 13.3 residues per unit time that is slowly varying in n, and a secondary pulse of speed about 6.4 with no slow spatial variation. The time evolution is very similar in all cases, so it is sufficient to compare at a single time t = 40 from now on. Figure 2 repeats the above data at that time, and Figure 3 is the same except for χ = 62, the other extreme of the likely range of values. Although a significant quantitative difference is seen, the qualitative description above still holds for the stronger nonlinearity, and it will be seen soon that other key features are also unchanged. (The latter is also similar to what is seen of [Sco82], which however used the two-point initial impulse form ψ1(0) = ψ2(0) = 1, ψn(0) = 0 for n > 2, and χ = 34.) The slow variation of exciton amplitude suggests the possibility of a long wave PDE approximation for this part of the solution, as proposed by Davydov and others. However, slow variation is not seen in ψn as a whole, due to rapid phase variation, and this is true even if one restricts to individual spines. Instead, the phase advances by a factor of approximately −i at each step along the chain, and thus by factor of i at each step along a spine. This is best revealed by studying wn := inψn: the real and imaginary parts of this are shown in Figures 3 and 4 for the two cases above. Next, it can observed that the nonlinearity of the hydrogen bonds is of little significance, due to the magnitude of rn staying quite small: less than about 0.3. This is indicated by Figure 6 for the harmonic case γ = 0, with χ = 35. 10 FIG. 1. Anharmonic Davydov-Scott system, γ = 4, χ = 35: ψn2 at times t = 20, 40. FIG. 2. Anharmonic Davydov-Scott system, as in Figure 1 except at t = 40 only: ψn2. 11 FIG. 3. Anharmonic Davydov-Scott system, as in Figure 2 except with χ = 62: ψn2. FIG. 4. Real and imaginary parts of wn = inψn for χ = 35. 12 FIG. 5. Real and imaginary parts of wn = inψn for χ = 62. FIG. 6. Harmonic Davydov-Scott system, χ = 35: ψn2. However, this point is made more emphatically by considering the next level of approximation, by the subsonic limit of HDNLS (5). Even for the harder case of χ = 62, the exciton form is little changed, as seen in Figure 7, and it is much the same over the full range of likely χ values. 13 FIG. 7. HDNLS system, χ = 62: ψn2. B. The Linear Approximation χ → 0 A final approximation worth considering is χ → 0, which for either the Davydov-Scott or HDNLS systems gives a linear equation for the excitons alone: i dψn dt + J(ψn+3 + ψn−3) − L(ψn+1 + ψn−1) = 0. (39) This will be the starting point for the analysis below, but first it can be observed that at least some main qualitative features of the above solutions are retained in this linear model, as seen in Figures 8 and 9. The form of wn might now be recognized as resembling the Airy function Ai, and this will be explained in the analysis of the next section. C. Brief Remarks on Other Cases Some brief observations for other choices of initial data and parameter values. 1. For an initial impulse at other locations, one has exciton self-trapping, with most of the signal staying at the initial location. There are weaker pulses propagating in each direction, which are well explained by the analysis of linearized equations given in the next section. 2. For larger values of χ, about 100 and up, there is again strong exciton self-trapping, with little signal propagation. 3. For the double excitation initial data ψ1(0) = ψ2(0) = 1, ψn(0) = 0 for n > 2 as considered in [Sco82], the behavior is similar to that discussed here, though with somewhat stronger nonlinear effects. 14 FIG. 8. Linearization χ = 0: ψn2. FIG. 9. Linearization χ = 0: wn, which is now real-valued. 15 4. For a simplified exciton-phonon model such as in equations (3,4) and/or with the symmetric coupling form seen in (4), the main phenomena are similar, though with the pulse velocity of course changed to 2 J. V. ANALYSIS, AND THE THIRD DERIVATIVE NLS APPROXIMATION Previous studies have proposed a PDE approximation based on the assumption that ψn(t) varies slowly in n, leading to PDEs related to the nonlinear Schrodinger equation, and thus to the study of solutions related to its traveling wave solutions of hyperbolic secant form. However, it is seen above that for the impulsive initial data considered herein, ψn does not become slowly varying in phase. Instead, slow variation along the chain is seen in the transformed quantity wn = inψn, for which the Davydov-Scott exciton evolution equation (2) becomes and HDNLS becomes dwn dt + J(wn+3 − wn−3) + L(wn+1 − wn−1) = −i χ  (qn+3 − qn)wn, dwn dt + J(wn+3 − wn−3) + L(wn+1 − wn−1) = 2iκwn2wn. (40) (41) Recalling that J ≈ 1.47 THz and L = 2.33 THz whereas κ ≈ 0.45 -- 1.4 THz, and that our initial data ensures wn ≤ 1 with far smaller values typical, it appears likely that the linearization dwn dt + J(wn+3 − wn−3) + L(wn+1 − wn−1) = 0 (42) is a useful first approximation. One initial observation is that for the initial data considered herein, this has real valued solutions, fitting with the observed phase behavior. Following the approach of D. Pelinowsky and V. Rothos [PR05], we seek solutions of the form ψn = ei(βn+ωt)w(τ, z), z = n − vt, τ = t,  (cid:28) 1 (43) where the fast spatial and time scales are isolated in an exponential factor, leaving a slowly varying envelope w(τ, z). In the limit χ → 0, these should relate to "discrete traveling wave" solutions of the linearization (39), with This has dispersion relation and thus group velocity with maximum occurring for β = −π/2, ω = 0, so that w(τ, z) = eikz, k (cid:28) 1. ω(k) = kv + 2 J cos(3(β + k)) − 2 L cos(β + k), v = 6 J sin(3β) − 2 L sin(β), v = vmax = 6 J + 2 L ≈ 13.48 ψn = (−i)nw(z), (44) (45) (46) (47) (48) the same transformation suggested above based on numerical observations. course, excluded by the initial data used here.) (There is a left going counterpart of One way to see this is that from initially impulsive initial data with a wide range of wave numbers present, there is a clustering of signals of various wave numbers at critical numbers of group velocity dv/dβ = 0, in particular at β = −π/2, which gives the maximum velocity. There are in fact six critical numbers, with the other two that correspond to right-going pulses forming a supplementary pair β(cid:48) ≈ 0.15π, β(cid:48)(cid:48) = π − β(cid:48) with the same velocity v(cid:48) ≈ 6.60, fitting well the velocity of about 6.4 observed for the second slower pulse above. This double root allows pulses with spatial dependence given by the real and imaginary parts sin(β(cid:48)n), cos(β(cid:48)n) which explains the break-down of slow amplitude variation seen for that second pulse. Returning to the Davydov-Scott system, we now seek solutions similar to this. Nonlinearity requires an amplitude 16 scaling, so we use wn(t) ≈ √  u(z, τ ), where v = 6 J + 2 L, τ = t,  = (1/3)ak3, a = (27 J + L). This gives and in the subsonic limit of HDNLS, wτ + wzzz = i χqzw, wτ + wzzz + i χw2w = 0, which is sometimes called the third derivative nonlinear Schrodinger equation. For either of these equations, the linearization is the Airy PDE wτ + wzzz = 0, and the impulsive initial conditions considered here can be associated with its fundamental solution (cid:18) (cid:19) . w(z, τ ) = 1 (3τ )(1/3) Ai Converting back gives approximate solution wn(t) ≈ √ k a1/6 √ 3 t1/3 Ai z (3τ )(1/3) (cid:18) n − vt (cid:19) (at)1/3 . (49) (50) (51) (52) (53) (54) (55) Proposals for further analysis. For the related case of discrete NLS equations of the form i dψn dt + ψn−1 + ψn+1 + χf (ψn−1, ψn, ψn+1) with cubic nonlinearities f having the gauge symmetry (13), Pelinowsky and Rothos [PR05] showed that solutions of the nonlinear equation for small χ bifurcate from solutions of the linearization at certain points, in particular the one k = 0, ω = 0, β = −π/2 seen above. It seems likely that a similar analysis would apply here. Beyond that, what the numerical results suggest, and which should be analyzed further, is that the nonlinearity provides some "dispersion management", preventing the leading pulse from spreading as fast as in the linearization, and making it more dominant compared to the following oscillation train. VI. CONCLUSIONS 1. The sustained traveling exciton pulses seen in Davydov-style exciton-oscillator models of energy propagation in α-helix protein are well approximated by the subsonic, small mass approximation, giving a variant of the discrete NLS equation. 2. As noted by other authors, the main part of the pulse has magnitude ψn that varies slowly, suggesting a long wave PDE approximation. However, the phase of the ψn varies rapidly in index n, by about a quarter turn at each step, and thus the slow spatial variation is instead in wn = inψn. This leads to a new PDE approximation by the third derivative nonlinear Schrodinger equation wτ + wzzz = iw2w, which indeed gives solutions fitting well to the fastest moving part of the solutions. 3. Linearization of this to the Airy PDE wτ + wzzz = 0 also gives a good qualitative fit to many features such as pulse speed, with the main nonlinear deviation being in the most intense front-most part of the pulse. 4. Analysis of the linearized discrete system also explains a good part of the observed behavior: it is not nearly as accurate as the above nonlinear PDE in describing the leading part of the pulse, but explains the second, slower pulse for which the PDE is not applicable. 5. Evidence of nonlinear "self-trapping" effects are seen, in that the leading hump of the pulse remains stronger and narrower as time increases than those of the linearization, supporting more sustained propagation than a linear model would predict. 6. The higher order exactly energy-momentum conserving time-discretization method used is seen to handle well the stiffness that can arise in such systems, making it a good candidate for similar problems, including spatial discretization of various stiff nonlinear dispersive PDE's. 17 [BECHO00] Larissa Brizhik, Alexander Eremko, Leonor Cruzeiro-Hansson, and Yulia Olkhovska. Physical Review B, 61:1129, 2000. [CG89] M. Creutz and A. Gocksch. Higher order hybrid Monte-Carlo algorithms. Phys. Rev. Lett., 63:9 -- 12, 1989. [Dav71] Alexandr S. Davydov. Theory of Molecular Excitations. Plenum press, New York, 1971. [DK73] Alexandr S. Davydov and N. I. Kislukha. Solitary excitations in one-dimensional molecular chains. Phys. Status Solidi B, 59:465 -- 470, 1973. [DZ84] Alexandr S. Davydov and A. V. Zolotariuk. Physica Scripta, 30:426, 1984. [FMC10] Holly Freedman, Paulo Martel, and Leonor Cruzeiro. Mixed quantum-classical dynamics of an amide-I vibrational excitation in a protein alpha-helix. Phys. Rev. B, 82(17):174308, 2010. [For89] E. Forest. Canonical integrators as tracking codes. AIP Conference Proceedings, 184:1106 -- 1136, 1989. [GM88] Z. Ge and J. E. Marsden. Lie-Poisson Hamilton-Jacobi theory and Lie-Poisson integrators. Phys. Lett. A, 133:134 -- 139, 1988. [Gon96] O. Gonzales. Time integration and discrete Hamiltonian systems. Journal of Nonlinear Science, 6:449 -- 467, 1996. [GS96] O. Gonzales and Juan C. Simo. On the stability of symplectic and energy-momentum algorithms for nonlinear Hamil- tonian systems with symmetry. Comput. Methods Appl. Mech. Eng., 134:197 -- 222, 1996. [HLW06] Ernst Hairer, Christian Lubich, and Gerhard Wanner. Geometric Numerical Integration: Structure Preserving Algo- rithms for Ordinary Differential Equations. Springer, 2nd edition, 2006. [KK82] V. A. Kuprievich and V. Kudritskaya. Preprint ITP-82-64E, Institute for Theoretical Physics, Kiev, 1982. [LeM12a] B. LeMesurier. Conservative unconditionally stable discretization methods for Hamiltonian equations, applied to wave motion in lattice equations modeling protein molecules. Physica D, 241(1):1 -- 10, January 2012. Published online 1 Oct 2011. [LeM12b] Brenton LeMesurier. Studying Davydov's ODE model of wave motion in alpha-helix protein using exactly energy- momentum conserving discretizations for Hamiltonian systems. Mathematics and Computers in Simulation, 82(7):1239 -- 1248, 2012. Published online 30 December 2010. [PR05] D. Pelinovsky and V. Rothos. Bifurcations of travelling wave solutions in the discrete NLS equation. Physica D, 202:16 -- 36, 2005. [Sco82] Alwyn C. Scott. The vibrational structure of Davydov solitons. Physica Scripta, 25:651 -- 658, 1982. [Sco84] Alwyn C. Scott. Launching a Davydov soliton: I. Soliton analysis. Physica Scripta, 29:279 -- 283, 1984. [Suz90] M. Suzuki. Fractal decomposition of exponential operators with applications to many-body theories and Monte-Carlo simulations. Phys. Lett. A, 135:319 -- 323, 1990. [Yos90] H. Yoshida. Construction of higher order symplectic integrators. Phys. Lett. A, 150:262 -- 268, 1990.
1805.08860
1
1805
2018-05-22T20:49:18
Statistical parameter inference of bacterial swimming strategies
[ "physics.bio-ph" ]
We provide a detailed stochastic description of the swimming motion of an E.coli bacterium in two dimension, where we resolve tumble events in time. For this purpose, we set up two Langevin equations for the orientation angle and speed dynamics. Calculating moments, distribution and autocorrelation functions from both Langevin equations and matching them to the same quantities determined from data recorded in experiments, we infer the swimming parameters of E.coli . They are the tumble rate ${\lambda}$, the tumble time $r^{-1}$ , the swimming speed $v_0$ , the strength of speed fluctuations ${\sigma}$, the relative height of speed jumps ${\eta}$, the thermal value for the rotational diffusion coefficient $D_0$ , and the enhanced rotational diffusivity during tumbling $D_T$ . Conditioning the observables on the swimming direction relative to the gradient of a chemoattractant, we infer the chemotaxis strategies of E.coli . We confirm the classical strategy of a lower tumble rate for swimming up the gradient but also a smaller mean tumble angle (angle bias). The latter is realized by shorter tumbles as well as a slower diffusive reorientation. We also find that speed fluctuations are increased by about 30% when swimming up the gradient compared to the reversed direction.
physics.bio-ph
physics
Statistical parameter inference of bacterial swimming strategies Maximilian Seyrich,1, ∗ Zahra Alirezaeizanjani,2 Carsten Beta,2 and Holger Stark1, † 1Institut fur Theoretische Physik, Technische Universitat Berlin, Hardenbergstrasse 36, 10623 Berlin, Germany 2Institut fur Physik und Astronomie, Universitat Potsdam, Karl-Liebknecht-Strasse 24/25, D-14476, Germany (Dated: May 24, 2018) We provide a detailed stochastic description of the swimming motion of an E.coli bacterium in two dimension, where we resolve tumble events in time. For this purpose, we set up two Langevin equations for the orientation angle and speed dynamics. Calculating moments, distribution and autocorrelation functions from both Langevin equations and matching them to the same quantities determined from data recorded in experiments, we infer the swimming parameters of E.coli . They are the tumble rate λ, the tumble time r−1, the swimming speed v0, the strength of speed fluctuations σ, the relative height of speed jumps η, the thermal value for the rotational diffusion coefficient D0, and the enhanced rotational diffusivity during tumbling DT . Conditioning the observables on the swimming direction relative to the gradient of a chemoattractant, we infer the chemotaxis strategies of E.coli . We confirm the classical strategy of a lower tumble rate for swimming up the gradient but also a smaller mean tumble angle (angle bias). The latter is realized by shorter tumbles as well as a slower diffusive reorientation. We also find that speed fluctuations are increased by about 30% when swimming up the gradient compared to the reversed direction. I. INTRODUCTION One of the most prominent model swimmers in the field of biological microswimmers is the gut bacterium E.coli equipped with peritrichous flagella [1]. Its well known run-and-tumble swimming motion and chemotaxis strat- egy has been thoroughly studied [2–7]. Nowadays, mod- ern imaging techniques allow for high-throughput record- ing of bacterial trajectories [8–15]. The method of la- beling flagella by fluorescent markers allows to unravel the diverse swimming mechanisms of microorganisms [16, 17]. These refined techniques require an appropriate theoretical modeling of the bacterium's stochastic swim- ming path, including the dynamics of tumbling. They also require a rational and effcient method how to an- alyze the recorded data in experiments. In this article we provide such a theoretical framework and illustrate it for the model bacterium E.coli . Thereby, we also re- veal some new and detailed insights into its chemotaxis strategy. The E.coli bacterium resides in the run phase, when all of its flagella form a bundle and rotate counterclockwise. The bacterium swims along a straight line, only thermal rotational diffusion affects its persistence. When at least one of the flagella reverses its sense of rotation, it leaves the bundle and the bacterium is in the tumble phase, where it strongly reorients [18, 19]. Typically, the tumble phase is much shorter than the run phase [1]. Therefore, in theoretical models tumbling is considered as instanta- neous and a single event is described by a tumble angle drawn from a distribution [8, 20–22]. However, a recent and instructive work by Saragosti et al. showed that re- ∗Email: [email protected] †Email: [email protected] orientation during tumbling can be modeled by enhanced rotational diffusion [23]. In order to analyze large amounts of data from recorded trajectories, specialized computer algorithms, called tumble recognizers, have widely been used to iden- tify tumble events [3, 8, 9]. In order to distinguish runs from tumbles, these automated tumble recognizers com- pare turning rate and speed to threshold parameters. They are necessary to distinguish variations of speed and turning rate due to the ubiquitous noise from a real tum- ble event. The threshold parameters have to be chosen a-priori and adjusted until results from the automatized tumble recognition agree with a visual inspection of the trajectories. There is no general rule how to set these pa- rameters and indeed they vary quite substantially [3, 8]. In an earlier work [22], we presented a parameter in- ference technique that allows to quantify the swimming behavior of bacteria without the need of setting param- eters a priori. Kramers-Moyal coefficients were calcu- lated from a suitable stochastic model for the dynam- ics of the orientation angle and matched to the coeffi- cients determined from experimental data. In particular, the stochastic model treated tumble events as instanta- neous. This procedure provided the main characteristics of E.coli and the bacterium Pseudomonas putida: tum- ble rate, distribution of tumble angles, and the thermal rotational diffusivity. For E.coli it also confirmed an an- gle bias during chemotaxis reported earlier [8]: the mean tumble angle is larger when swimming against a chemical gradient compared to moving along it. Other parameter inference techniques use the framework of Bayesian in- ference [9, 24]. However, they pose a complex numerical challenge as one has to maximize a likelihood function that contains the data of all the recorded trajectories. In this article we considerably extend our earlier work by resolving tumble events in time and by incorporating a stochastic process for the speed dynamics (see Fig. I). 8 1 0 2 y a M 2 2 ] h p - o i b . s c i s y h p [ 1 v 0 6 8 8 0 . 5 0 8 1 : v i X r a 2 II. MODEL AND METHOD A. Stochastic model for the random walk of E.coli A typical trajectory of bacteria such as E.coli is de- scribed by a run-and-tumble random walk. During the run phase the bacterium moves forward along a nearly straight line, only rotational thermal noise affects its per- sistence. During the tumble phase the bacterium's speed is reduced and it reorients strongly into a new direction. The angle between the orientations before and after the tumble event is the tumble angle β. We express the ve- locity of the bacterium in two dimensions as the product of speed v(t) and unit vector e(t) = (cos Θ, sin Θ), r(t) = v(t)e(t) , (1) where the orientation angle Θ is measured with respect to the x axis. We set up two overdamped Langevin equations for speed and orientation angle, which fully describe the bac- terial motion, v(t) = r [v0 − v(t)] + ξsp(t) + q(t), Θ(t) =(cid:112)2Drot(t) ξan(t) . (2) (3) We introduce both Langevin equations in more detail. (1) The equation for speed v(t) contains three terms, which are associated with drift, diffusion, and jumps. We start with the last term, q(t) = − N λ(cid:88) ηv(t)δ(t − ti) . (4) i=1 It initiates each tumble event at time ti by a shot-noise process, while the occurrence of times ti follows a Pois- son process with tumble rate λ. At the beginning of each tumble, the bacterial speed is reduced by the rel- ative jump height η to (1 − η)vt and N λ is the ac- tual number of tumble events. The first and second term represent a conventional Ornstein-Uhlenbeck pro- cess. After a tumble event the speed relaxes with relax- ation rate r towards the swimming speed v0 of the run phase. Thus r−1 is the mean duration of a tumble event, which we call tumble time in the following. The Gaus- sian white noise term is fully determined by (cid:104)ξsp(cid:105) = 0 and (cid:104)ξsp(s)ξsp(t)(cid:105) = σ2δ(t − s), where we introduce the white noise strength σ. It describes the ubiquitous noise due to internal noise of the swimming mechanism, varia- tions between individual bacteria, and measurement er- rors. Note that the actual tumble time of a bacterium is exponentially distributed. In our model the white noise term also induces stochastic fluctuations in the duration of the tumble events as visible in Fig. 2(b). Altogether, the stochastic speed process is determined by five param- eters: {λ, r, v0, σ, η}. (2) The stochastic equation for the orientation an- gle Θ is fully described by rotational diffusion, where Figure 1: E.coli with swimming velocity v(t) = v(t)e(t) = v(t)[cos Θ(t), sin Θ(t)]. A tumble event occurs between the times t + ∆t and t + 3∆t. A possible chemical gradient is indicated. The dynamics of the orientation angle is diffusive, where the rotational diffusivity switches via a telegraph pro- cess [25] between its thermal (run phase) and enhanced value (tumble phase). The dynamics of the speed con- tains a shot-noise process [26, 27]. It initiates a tumble event by decreasing the speed value, which then relaxes back to the swimming speed according to an Ornstein- Uhlenbeck process [28]. Calculating moments, distribu- tion and autocorrelation functions for orientation angle as well as speed and matching them to the same quan- tities calculated from experimental data, we are able to infer the swimming parameters of E.coli . Their values are in good agreement with the parameters determined using a tumble recognizer. Compared to the Bayesian framework, our method of parameter inference consider- ably lowers the efforts of the numerical optimization. To explore the chemotaxis strategy of E.coli , we con- dition [29, 30] moments and autocorrelation functions on the swimming direction relative to the chemical gradi- ent and infer the swimming parameters as a function of the orientation angle. Besides the well-known chemo- taxis strategy (modulation of the tumble rate), we con- firm the recently discovered angle bias [8]. We show that the increased angular persistence when swimming up the gradient is caused by both shorter tumbles as well as smaller rotational diffusivity. Moreover, for the same swimming direction we identify larger fluctuations in the speed value. The article is organized as follows. In Sect. II we intro- duce the two Langevin equations of our stochastic model and calculate moments, distribution functions, and au- tocorrelation functions for speed and orientation angle. Section III reviews details of the experiments. Section IV first explains the inference method and then presents our results in a uniform buffer solution (control experiment) and in the gradient of a chemoattractant. We close with a summary and an outlook in Sect. V. tt+1Δtt+2Δtt+3Δtt+4Δt the white noise process is defined by (cid:104)ξan(cid:105) = 0 and (cid:104)ξan(s)ξan(t)(cid:105) = δ(t − s). Following Ref. [31], we model tumbles as a random walk on a unit sphere with enhanced rotational diffusion. Thus, the rotational diffusion co- efficient Drot(t) is no longer a constant but alternates between two values: the thermal rotational diffusion co- efficient D0 during run phases and an enhanced value DT during tumble phases. We describe each transition between the two states by a Poisson process and thus obtain a telegraph process. The transition rate from the run to the tumble phase is the tumble rate λ, whereas the transition rate in the opposite direction is the speed relaxation rate r or the inverse tumble time. A full def- inition and basic properties of the telegraph process are given in the appendix B 2 or can be found in [25]. To link the telegraph process to the shot-noise process for the speed value in Eq. (4), the diffusion coefficient switches at the same times ti from the thermal (D0) to the enhanced (DT ) value. Note, while the speed process allows a second tumble although the first one is not fin- ished yet, this is not possible in the telegraph process for rotational diffusion. However, for bacteria like E.coli the time between tumble events is typically one order of magnitude larger than the tumble time r−1. This makes these double events very rare and tumble events in both speed and angular processes coincide. All in all, we have four parameters governing the stochastic process for the orientation angle: {λ, r, D0, DT}. Figure 2 shows a typical simulated trajectory (a) and the corresponding time series for speed and angular dis- placement ∆Θ during time step ∆t = 0.1 s (b). It has to be compared to the experimental time series of both quantities in (c). Note that ∆Θ ∆t represents the turning rate of the bacterium. In the following we will always work with the angular displacement ∆Θ. B. Basics of the inference method In this section we state moments, stationary distribu- tions, and time autocorrelation functions for the stochas- tic processes of speed and orientation angle in Eqs. (2) and (3). They depend on the swimming parameters in- troduced above. Matching the theoretical expressions of these quantities to the values determined by averaging over all individual tracks of the experiments, we are able to infer the mean swimming parameters of an E.coli pop- ulation. We refer to appendix B for details of the deriva- tions and only state the final expressions in the following. 1. Speed The moments mV n = (cid:104)v(t)n(cid:105) of Eq. (2), where the aver- age is taken over all times t and all tracks in the long-time limit, can be calculated as a function of the reduced pa- rameter set(cid:0)λ/r, η, v0, σ2/r(cid:1). For the first moment, the 3 Figure 2: (a) Simulated run-and-tumble trajectory of a bac- terium using the stochastic equations (2) and (3). It starts at the green and ends at the red triangle. (b) Initial part (from green triangle to the black diamond) of the corresponding time series for speed v(t) and angular displacement ∆Θ(t) during time step ∆t = 0.1 s. Tumble initiations are marked in orange. (c) Experimental time series for v(t) and ∆Θ(t). mean speed, we obtain (cid:19) = v0 1 + ηλ/r . (5) mV 1 , η, v0, σ2 r r (cid:18) λ (cid:19) (cid:18) λ The mean speed is smaller than the swimming speed v0 since during the tumble phase speed is reduced by a fac- tor η. More generally, a recursive formula for the nth moment is given by mV n , η, v0, σ2 r r v0 mV = n−1 + 1 1 + λ nr − λ 2 (n − 1) σ2 nr (1 − η)n r mV n−2 , (6) where the zeroth moment is m0 = 1 due to normaliza- tion. We now have access to all the speed moments. As an example, Fig. 3(a) shows a histogram for the distri- bution of speed values recorded in an experiment, from which the speed moments can be calculated. The orange line represents the distribution obtained from numeri- cally solving the speed equation (2) using the actual pa- rameters inferred from this experiment. The two distri- butions nicely agree, which is an a-posteriori verification of our Langevin equation. From the moments we can only infer the ratios λ/r and σ2/r. In order to determine the full set of parameters of (a)01020v[µm/s]v0010t[s]0π2∆Θ(b)Simulation01020v[µm/s]010t[s]0π2∆Θ(c)Experiment 4 Figure 4: (a) Histogram showing the distribution of angu- lar displacements ∆Θ in time step ∆t for the same data set as in Fig. 3. The orange line shows the distribution p(∆Θ) from Eq. (8) using the inferred parameters. In- (b) Semi- set: Semi-logarithmic plot of the distribution. logarithmic plot of the corresponding directional autocorrela- tion function gΘ(τ ). Green line: linear fit with negative slope αΘ = 0.33 s−1; orange line: exponential fit with relaxation rate αΘ = 0.32 s−1. region β > π/2 only represents roughly 3% of all an- gular displacements. There are two possible reasons for this deviation: First, we record angular displacements Θ = π +  as a displacement −(π − ) since we cannot distinguish between tumbles to the right and left during one time step. Second, it is also possible that the dif- fusion model for tumbling does not apply for such large angles. For completeness we also give the nth moment of the n = (cid:104)∆Θn(cid:105). It fol- absolute angular displacement, m∆Θ lows directly from the probability distribution of Eq. (8): (cid:18) (2D0∆t) n (cid:40)(cid:113) 2 1 + λ/r 2 · π 1 (cid:19) + (2DT ∆t) n 1 + r/λ 2 (n − 1)!! if n is odd if n is even , (9) Figure 3: (a) Histogram showing the distribution of speed values for a dataset recorded for E.coli in a control experi- ment moving in a buffer medium without any chemical gra- dient. The orange line shows the distribution from the simu- lated process using the inferred parameters. (b) Correspond- ing speed autocorrelation function gV (τ ) of the same dataset. The orange line shows an exponential fit with relaxation rate αV = 5.1 ± 0.2 s−1. Inset: Semi-logarithmic plot of gV (τ ). 1 ][v(t) − mV Eq. (2), we also use the speed autocorrelation function for our model. It has an exponential form with relaxation rate r + ηλ, gV (τ ) = (cid:104)[v(t + τ ) − mV 1 ](cid:105) = ∆2ve−(r+ηλ)τ , (7) where we have introduced the variance ∆2v = (cid:104)(v − 1 )2(cid:105). Figure 3(b) shows the autocorrelation function mV for the experimental data of E.coli . Indeed, the curve is well-fitted by an exponential over two decades up to τ (cid:39) 1s, which is around half the mean track length. This agreement supports the validity of our stochastic descrip- tion of the speed process in Eq. (2). 2. Angle m∆Θ n (λ, r, D0, DT ) = (cid:112) Here, we work directly with the steady-state proba- bility distribution p(∆Θ) for the absolute angular dis- placement ∆Θ during a finite time step ∆t. We deter- mine p(∆Θ) from Eq. (3) for the orientation angle as a function of the reduced parameter set (λ/r, D0, DT ). In the long-time limit the probability distribution p(∆Θ) becomes stationary and is given by λ N (0, p(∆Θ) = r N (0, λ + r 2D0∆t)+ 2DT ∆t) (8) where N (0, σ) denotes the normal distribution with zero mean and standard deviation σ. For our parameter in- ference we use the same time step ∆t = 0.1s−1 as in Ref. [31]. λ + r Figure 4(a) presents a histogram for all angular dis- placements in time step ∆t recorded in the experiment. It shows a deviation from the theoretical distribution of Eq. (8) in the tail at angles larger than π/2, which is visible only in the semi-logarithmic plot. Note that the (cid:112) where n!! denotes the double factorial. Similar to the speed process, we can only infer the ratio λ/r from fits to the probability distribution p(∆Θ) of Eq. (8). In order to determine the full set of parameters of Eq. (3), we use again the autocorrelation function of our model, now for the swimming direction e(t). Numerical investigations of our model (see appendix B 3) suggest that it has a simple exponential form with relaxation rate αΘ for parameters relevant to the experiments: gΘ(τ ) = (cid:104)e(t + τ ) · e(t)(cid:105) ∝ e−αΘτ (10) Analytically, we are not able to calculate this exponential form. However, in the time interval (λ + r)−1 < τ < (cid:104)Drot(cid:105)−1 relevant to the experiments, we can derive the linear approximation gΘ(τ ) ≈ 1 − αΘτ ≈ 1 − (cid:104)Drot(cid:105) − ∆2Drot λ + r τ (11) (cid:19) (cid:18) (a)(b)(a)(b) and thereby obtain an expression for the relaxation rate αΘ. Here we have introduced the respective mean (cid:104)Drot(cid:105) and variance ∆2Drot of the telegraph process Drot(t), D0 DT + (cid:104)Drot(cid:105) = 1 + λ/r 1 + r/λ ∆2Drot = (cid:104)(Drot − (cid:104)Drot(cid:105))2(cid:105) = (D0 − DT )2λ/r (1 + λ/r)2 .(12) Figure 4(b) shows the directional autocorrelation func- tion for the experimental data of E.coli moving in a uni- form buffer medium. Indeed, the curve is well-fitted by an exponential up to τ (cid:39) 5 s excluding the first point. This agreement supports the validity of our stochastic description of the angle process in Eq. (3). The devia- tion in the experimental data for the first point is caused by the offset for angular displacements larger than π/2, where the experimental distribution function in Fig. 4(a) deviates from theory. For two and more time steps the influence of this offset becomes smaller and smaller. III. EXPERIMENTAL MATERIALS AND METHODS A. Cell culture E.coli AW405 strain was cultured overnight in liquid Tryptone Broth (TB) (10 g/l Difco BactoTM-Tryptone and 5 g/l NaCl) at 37 ◦C on a rotary shaker at 300 rpm. The cell suspension was diluted 1:100 into fresh TB, and grown to mid-exponential phase (OD600 = 0.5). Then the bacterial suspension was washed and resuspended in motility buffer (11.2 g/l K2HPO4, 4.8 g/l KH2PO4, 3.93 g/l NaCl, 0.029 g/l EDTA and 0.5 g/l glucose; pH 7.0). Afterward, the cell suspension was divided into two fractions. One was centrifuged and resuspended in the same motility buffer, and the other was centrifuged and resuspended in motility buffer supplemented with the chemoattractant α-methyl-aspartate (Sigma-Aldrich, USA) in a final concentration of 0.5 mM. In both cases, the final OD600 of the cell suspensions was 0.07 before filling them into chemotaxis chambers. B. Chemotaxis assay In this study, a µ-Slide Chemotaxis 3D (ibidi, Mar- tinsried, Germany) was used in order to maintain a stable linear gradient of the chemoattractant α-methyl- aspartate. This chemotaxis chamber consists of two large reservoirs connected to a central observation area. For the chemotaxis assay, the cell suspension with chemoat- tractant was filled into the reservoir on the right hand side and the chemoattractant-free cell suspension into the reservoir on the left hand side. The central obser- vation area was filled with motility buffer (see appendix A). A stable linear chemoattractant gradient is generated 5 by diffusion in the observation area and maintained for several hours [32]. For the control assay, both reservoirs were filled with chemoattractant-free cell suspension. In this case, a homogeneous environment without any gra- dient was established in the observation area. C. Cell imaging and tracking An IX71 inverted microscope with a 20× UPLFLN-PH objective (both Olympus, Germany) in phase contrast mode was used for imaging cell trajectories. Five im- age sequences were taken with 10 min intervals between them using a Orca Flash 4.0 CMOS camera (Hamamatsu Photonics, Japan). For each sequence, the images were acquired at 20 frames per second for 30 s. The field of the view was placed in the center of the gradient region at 30 µm above the bottom of the chamber (total height in the observation area was 70 µm). A custom Matlab program based on the Image Pro- cessing Toolbox (version R2015a, The MathWorks, USA) was used to process the image sequences automatically. For each image sequence, a background image was cal- culated by pixel-wise time average projection. It was subtracted from each frame to eliminate non-motile ob- jects and shading effects. The built-in Matlab function imerode was then applied for morphological erosion (with a disk of radius 0.6 µm) to reduce the background noise. The putative bacterial cells are distinguished from back- ground using the maximum entropy thresholding algo- rithm by Kapur et al. [33]. The threshold was calculated for each image in the sequence separately. The median of all threshold values was used to segment the whole se- quence. The binary images were further processed with the morphological operations, imopen and imclose (with a disk of radius 0.3 µm) to eliminate any noise caused by segmentation. The built-in function bwconncomp was used to find all connected objects in the binary images. Size and centroid of the objects were determined using the regionprops function. Afterwards, particles with an area between 1 µm2 to 15.6 µm2 were considered as single bacterial cell. Finally, trajectories were obtained employ- ing the tracking algorithms by Crocker and Grier [34]. To avoid tracking artifacts caused by tumble events when cells enter and leave the focal plane, the first and last 0.5 s of each track were removed. Highly curved tracks as well as tracks with a total displacement < 10 µm were eliminated, as they most likely result from damaged flagella. The minimal track length is 0.5 s and the maxi- mal length is 19.35 s. The control data set consists of 769 tracks with a total length of 1629 s. The gradient data set consists of 3498 tracks with a total length of 7206 s. D. Heuristic run-tumble analysis The trajectories were smoothed using a second-order Savitztky–Golay filter with a window size of 5 data points corresponding to 250 ms [35]. Instantaneous speed v = ∆t , direction of propagation θ, and turning rate ω = ∆θ ∆s ∆t were evaluated on the smoothed tracks. The tumble events were detected as described previously [9, 22, 36]. Briefly, in the time series of speed and turning rate, local minima and local maxima were detected, respec- tively, to identify tumble events. Four parameters, two for the speed and two for the turning rate, were adjusted such that the recognition of tumble events was correct as checked by visual examination (threshold parameters α = 3 and β = 6.5 and tumble duration parameters 0.55×∆v and 0.65×∆ω, see the Supporting Information S5 in Ref [22]). IV. RESULTS We are now equipped to infer the swimming param- eters from experimental data for different experimental settings. We first illustrate the inference method by ap- plying it to a control experiment, where E.coli swims in a homogeneous buffer solution. We validate the infer- ence method by comparing the inferred parameters to their values determined by a heuristic tumble recognizer. Then we demonstrate that our method also reveals the chemotaxis strategy of E.coli when moving in a chemical gradient. In particular, we apply it to data, which was recorded in a linear gradient of α-methyl-aspartate. A. Inferring the swimming parameters for E.coli in a uniform environment Figures (3) and (4) show distributions and autocor- relation functions for speed and angular displacements recorded for E.coli when swimming in a homogeneous buffer without any chemical gradient. Note that speed and angle inference are performed separately from each other but they are linked by the tumble rate λ and the inverse tumble time r. (cid:80)Ti i=0 T −1 i n N−1(cid:80)N Speed inference: From the histogram of the recorded speed values in Fig. 3(a) we determine the moments of the experimental speed data: mv,exp := t=0 [vi(t)]n. The sums are taken over all tracks i = 0, . . . , N and all times t, where Ti is the length of track i. Figure 3(b) shows the exponential fit to the speed auto-correlation function, which yields the experimental relaxation rate αV = −5.1 ± 0.2 s−1. Note that the error estimate and all the following ones are ob- tained by the method of bootstrapping (see appendix C for more details). We match the first eight speed mo- ments and the relaxation rate to the their theoretical expressions of Eqs. (5)-(7) and obtain 9 non-linear equa- tions for the speed swimming parameters. We solve these equations numerically using a simplex-downhill optimiza- tion algorithm from the python package scipy. 6 Angle inference: Independently, we match the the- oretical distribution function for the angular displace- ment [given in Eq. (8)] to the experimentally recorded histogram in Fig. 4(a) and thereby extract the param- eters D0, DT , and λ/r. We perform the fit up to ∆Θ = π/2 to avoid the offset for angular displacements larger than π/2. Last, by matching the experimental re- laxation rate αΘ of the directional autocorrelation func- tion to the theoretical expression of Eq. (11), we obtain the full set of parameters [see also Eqs. (B33) and (B34) in appendix B 2]. Figure 4(b) shows the linear fit with relaxation rate αΘ = 0.33 s−1 ± 0.02 (green line) and the exponential fit with rate αΘ = 0.32 s−1 ± 0.01 (orange line) in a semi-logarithmic plot. Inferred Parameters: Table I gives an overview of the inferred swimming parameters for the two stochastic processes for speed and angle. The two inferred tumble rates λ are very close together and the inverse tumble times r agree within the error bars. Our results are in good agreement with tumble rate λ = 0.84 s−1 and swim- ming velocity v0 = 20.7 µms−1 determined with a heuris- tic tumble recognizer (see Sect. III D and Ref. [22]). This validates our inference method. Moreover, our findings are in good agreement with previously measured tumble rates [3, 9, 22] and swimming speeds [37]. The inferred value for the thermal rotational diffusivity D0 agrees with previously reported values in the literature, which range from 0.06 s−1 [20, 22] to 0.18 s−1 [38]. We use the enhanced rotational diffusion coefficient DT = 2.31 s−1 and the inverse tumble time r = 3.81 s−1 of the angle stochastic process to determine the distri- bution function of absolute tumble angles, P (β), by recording the angular displacement for exponentially dis- tributed tumble times with mean r−1. The correspond- ing three-dimensional distribution function is obtained by multiplying the two-dimensional quantity with sin β from the solid angle element. The resulting distribution is shown in orange in Fig. 5 for β < π. It has a max- imum at βmax = 0.78 = 45° and the mean tumble an- gle is (cid:104)β(cid:105) = 1.06 = 61°, which are remarkably close to the values βmax = 45° and (cid:104)β(cid:105) = 62° from Ref. [3]. The shape of the distribution function is similar to the one obtained with the heuristic tumble recognizer (blue bars). Also, the maximum values are very close. While the main characteristics of the two curves agree well, the heuristic tumble recognizer determines more tumbles for angles close to π. As a result, it finds a larger mean tum- ble angle (cid:104)β(cid:105) = 1.43 = 82°. This might be explained as follows. Some tumbles occur only in one time interval, where one cannot distinguish between a leftward tum- ble angle β and a rightward tumble β − 2π. Thus, the heuristic tumble recognizer chooses always the smaller angle and, therefore, the distribution of tumble angles close to π is enhanced. In contrast, our inference for the angle process only uses angular displacements up to π/2 in Fig. 4. Thus, it gives a more correct account of the distribution. Speed λ r v0 (cid:113) η σ2 r 0.83 ± 0.04 s−1 4.41 ± 0.30 s−1 20.8 ± 0.2 µms−1 5.11 ± 0.07 µms−1 0.85 ± 0.01 Angle λ r D0 DT 0.84 ± 0.02 s−1 3.81 ± 0.30 s−1 0.090 ± 0.002 s−1 2.31 ± 0.12 s−1 7 Table I: Inferred parameters for the stochastic processes of speed and angle for E.coli moving in a buffer medium without a chemical gradient (control experiment). Figure 5: Comparison of the two tumble angle distributions P (β) measured by the heuristic tumble recognizer (blue bars) and determined from the stochastic process for the ori- entation angle using the inferred parameters DT and r from table I (orange line). The distribution determined from the- ory has a maximum at βmax = 0.78 = 45° and the mean tumble angle is (cid:104)β(cid:105) = 1.06 = 61°. Compared to literature we define the tumble time dif- ferently by setting τt = r−1. Usually, one employs a tum- ble recognizer and identifies the tumble state when the angular displacement (per time step) exceeds a thresh- old value [3, 7, 8, 16]. The duration of this period is then the tumble time [see also Fig. 6(a)], for which val- ues of τt = 0.12 s and 0.14 s were measured using different thresholds [3, 8]. However, this procedure underestimates the duration of a tumble event, which starts when a flag- ellum leaves the bundle and ends when it returns to the bundle. At the beginning and end of this period the an- gular displacement (per time step) can of course be below the given threshold value. Indeed, Ref. [16] showed that the duration of a tumble event obtained from visualiz- ing the flagellar dynamics during tumbling is significantly larger than the time determined by tumble recognizers. In contrast to tumble recognizers, our method defines the tumble time as the inverse relaxation rate τt = r−1. This is a more rational quantification of the tumble time without the need of an a-priori threshold value. Tumbles are initiated when the speed jumps below the swimming speed and they end when the speed has relaxed back to Figure 6: (a) Usually, the tumble time τt is defined as the period where the angular displacement per time step exceeds an a-priori threshold value. (b) In our method the tumble time is the inverse speed relaxation rate r−1. the swimming speed. We argue that the higher value τt = 0.23 s obtained by our method describes the tumble process more precisely. B. Chemotaxis Next, we apply our method to experimental data of E.coli recorded in a constant gradient of a chemoattrac- tant concentration. Conditioning the analysis on the swimming direction, we are able to determine how the swimming parameters depend on the orientation or swim- ming angle θ. Thus, we divide the experimental data into eight subsets or sectors each spanning a range of orienta- tion angles centered at θn = 2πn/8 for n = {0, 1, ..., 7}. Here, θ = 0, 2π means swimming up the gradient and θ = π against the gradient. In practice, instead of di- viding the data for the orientation angle into 8 disjunct sectors, we use smooth weighting based on Gaussian ker- nels as in Ref. [22] (for further details see appendix D). Figure 7 shows the results from applying our inference method to the moments of speed and to the distribution of angular displacements. Graph (a) plots the tumble bias λ/r, the ratio of tumble time to run time, versus orientation angle. It is lowered when swimming up the gradient (θ = 0, 2π) and increased when swimming down the gradient (θ = π). This confirms the classical chemo- taxis strategy. The curves from angle inference (orange) and speed inference (blue) show good agreement. Again, we recognize that both inference strategies give coherent results, even though they are performed independently from each other. In Fig. 7(b) the rotational diffusion co- efficient DT during tumbling also depends on the swim- ming direction. It is lowered when swimming up the gradient and increased when swimming down the gradi- ent. This suggests angular persistence or a reduced mean tumble angle, when swimming in a favorable direction, as a chemotaxis strategy. It was already reported in Refs. [8, 22]. We will comment more on this strategy in the following. Adding the speed autocorrelation function to the pa- 0βmaxπ/2πβ0.00.20.40.6FrequencyInferenceHeuristic(a)(b) 8 Figure 7: (a) Tumble bias λ/r conditioned on the swimming angle θ and determined either by angle inference (orange) or speed inference (blue) for E.coli in a linear gradient of chemoattractant (α-methyl-aspartate). (b) Rotational diffu- sion coefficient DT during tumbling conditioned on the ori- entation angle θ. The bacterium swims up the gradient for θ = 0, 2π and down the gradient for θ = π. rameter inference, we investigate whether tumble rate λ and tumble time r−1 are separately modulated during chemotaxis. Figure 8 shows the results for the speed parameters λ, r, v0, σ. Indeed, we recover the classical chemotaxis strategy in plot (a) with a strong reduction of the tumble rate when swimming up the chemical gra- dient. The tumble rate for θ = 0 is less than half of the tumble rate for θ = π. The same trend occurs for the tumble time r−1, which increases when swimming down the gradient. This bias in tumble time together with the same trend for the diffusion coefficient DT found above confirms a bias in the mean tumble angle (cid:104)β(cid:105). It is enhanced when swimming in an unfavorable direction, which confirms the alternative chemotaxis strategy iden- tified in Refs. [8, 22]. No significant modulations are vis- ible for the swimming speed v0 plotted in (c). So there is no chemokinesis. The same applies to the jump height η, which is shown in Fig. 11 of appendix E. In the last plot (d) we identify a novel bias in speed fluctuations. The swimming speed is significantly more volatile when swimming up a chemical gradient compared to swimming against it. To the best of our knowledge, this has not been reported yet. V. CONCLUSIONS AND OUTLOOK In this article, we provide a detailed stochastic descrip- tion of the swimming motion of an E.coli bacterium in two dimensions, where we also resolve tumble events in time. We set up an overdamped Langevin equation for the speed dynamics, which contains three terms associ- ated with drift, diffusion, and jumps that initiate a tum- ble event. A second Langevin equation for the angular dynamics describes rotational diffusion of the orientation angle, where the diffusion coefficient alternates between its thermal value during run phases and an enhanced value during tumbling. The transition between both phases is described by a telegraph process. An analysis Figure 8: Inferred parameters of the speed process condi- tioned on the swimming angle θ and inferred from the same experiment as in Fig. 7. We recover the bias of tumble rate in (a), find a bias in tumble time r−1 in (b), no chemokinesis in (c), and a novel fluctuation bias in (d). The bacterium swims up the gradient for θ = 0, 2π and down the gradient for θ = π. of experimental data verifies our description a-posteriori: distribution and autocorrelation functions for both speed and orientation angle agree with theoretical predictions from our model and with numerically determined func- tions using the inferred swimming parameters. We considerably extent earlier work [22] by resolving tumble events in time and by incorporating a stochas- tic process for the speed dynamics. Based on moments as well as distribution and autocorrelation functions, we provide a robust methodology for inferring the full set of swimming parameters that characterize the run-and- tumble motion. The inferred swimming parameters are the tumble rate λ, the tumble time r−1, the swimming speed v0, the strength of speed fluctuations σ, the jump height η, the thermal value for the rotational diffusion coefficient D0, and the enhanced coefficient during tum- bling DT . Although the inference of angle and speed parameters are carried out completely independent from each other, they show good and very good agreement for the two common swimming parameters, r and λ, respec- tively. We validated our results by comparing the swimming parameters to the results of a heuristic tumble recognizer and obtained good agreement. However, our approach of inferring parameters has three advantages. First, it does not need to set a-priori threshold parameters for speed and angular displacement. Second, it is able to infer the strength σ of speed fluctuations and the thermal rotational diffusion coefficient D0. Third, it provides a (a)(b)(a)(b)(d)(c) more rational and precise choice for the tumble time that encompasses the whole tumble event instead of just the part which is determined by threshold parameters. The inference method allows to condition the swim- ming parameters on a specific situation and monitor how they change with the situation by dividing the full data set into subsets. In particular, while conditioning on the swimming direction, we are able to confirm the classical chemotaxis strategy, which modulates the tumble rate λ when changing the swimming direction relative to the chemical gradient. We also confirm the recently discov- ered modulation of the mean tumble angle (angle bias) [8]. Resolving the tumble event in time, we realize that this angle bias is due to modulations of both the tum- ble time and the enhanced rotational diffusivity during tumbling. This has not been reported so far. As the tum- ble rate we expect the tumble time to be determined by the internal chemotaxis machinery of E.coli , which mon- itors the changing chemoattractant concentration dur- ing swimming. The higher rotational diffusivity during a tumble phase, which follows swimming against the gradi- ent, may be caused by more flagella leaving the flagellar bundle, as argued in [23]. Finally and also not reported so far, we show that speed fluctuations are larger by 30% when E.coli swims up the chemical gradient. for example, Our method of conditioning can be applied to other quantities, the concentration c of the chemoattractant. In particular, the tumble rate of a bac- terium, which is adapted to a chemoattractant, should not depend on the concentration c [39]. In an earlier analysis of experiments we already verified this for E.coli and Pseudomonas putida [40]. Other possible conditions explore the biological variability in properties such as the swimming speed v0 of a bacterium or its size. In the following we mention some further directions, where our method of inference can be applied or needs to be extented. Recent experimental techniques allow to record tracks of length of the order of 100 s [12, 15]. Such long tracks provide enough data to apply our method to a single track and thereby measure swimming parameters for individual bacteria. This can then reveal and quantify heterogeneities in a bacterial population. To apply the method of inference to other bacterial swimming mechansim, the Langevin equations (2) and (3) need to be modfied. For example, run-reverse bac- teria such as the soil bacterium Pseudomonas putida, possess a tumble angle distribution with a sharp peak centered around π [36]. The marine bacteria Vibrio al- ginolyticus has a bimodal distribution of tumble angles with two maxima as measured in Ref. [41]. In both cases, rotational diffusion with an enhanced diffusivity cannot reproduce such distributions. A possibility to address these cases is to extend the approach of Ref. [22]. There, instantaneous tumbling was modeled by a shot noise pro- cess with a delta-peaked angular turning rate and tumble angles drawn from an appropriate distribution. Broad- ening the delta function to a Gaussian function with the tumble time τt as standard deviation, one can again re- 9 solve the tumble event in time. Furthermore, an elab- orate model of the speed dynamics for Pseudomonas putida should include the alternating swimming speeds reported in Ref. [36], which belong to different swimming modes [17]. Once such models are established, the inference method provides a rational way of analyzing experimen- tal data in order to determine the relevant swimming parameters and to understand important processes such as chemotaxis by conditioning the available data on sub- sets. Thus, in this article we have introduce a powerful methodology for analyzing properties of bacterial popu- lations, which can handle large amounts of experimental data. ACKNOWLEDGMENTS Fruitful discussions with J.-T. Kuhr, J. Blaschke, M. Hintsche and A. Kulik are acknowledged. This work was supported by the German Research Foundation (DFG) within the research training group GRK 1558. Appendix A: Chemotaxis chamber Figure 9 presents a layout of the chemotaxis device used to quantify the chemotactic response of E.coli . Appendix B: Derivation of moments, autocorrelation and distribution functions We present detailed derivations of stochastic proper- ties of the two Langevin equations (2) and (3), which we mention in the main text. First, we derive expres- sions for the moments and the autocorrelation function of the speed process. Second, we present the probability distribution function (pdf), the moments, and the ap- proximation for the directional autocorrelation function of the angle process. 1. Speed In order to perform the derivations, we rewrite Langevin equation (2) as a stochastic differential equa- tion (SDE) using mathematical notation: dvt = r(v0 − vt)dt + σdWt − ηvtdN λ t . (B1) Here, we define the Poisson Process where dN λ t = 1 oc- curs with probability λdt for each time step indicating the start of a tumble and dN λ t = 0 otherwise. Moreover, we introduce the Wiener process dWt. Integrating Eq. (B1) and splitting the Poisson process into a determin- istic part and a fluctuating part dN λ [42] yields t = λdt + d N λ t 10 Taking the long-time limit t → ∞, we recover the equa- tion (5) from the main text. Next, we calculate the n-th moment mn = (cid:104)vn(cid:105). Using Ito's lemma [43], we first formulate a SDE for an arbi- trary function f (vt) of the speed variable: (cid:19) f(cid:48)(cid:48)(vt)σ2 f(cid:48)(vt)r(v0 − vt) + + f(cid:48)(vt)σdWt + [f (vt− − ηvt−) − f (vt−)] dN λ t . (B6) 1 2 dt (cid:18) df (vt) = (cid:90) t (cid:18) 0 (cid:104)vn t (cid:105) = Here, vt− denotes the value right before a jump. Set- ting f (vt) = vn t , integrating Eq. (B6), and taking the expectation value on both sides yields: (cid:105) + (λ [(1 − η)n − 1)] − nr)(cid:104)vn s (cid:105) nrv0(cid:104)vn−1 n(n − 1) s + 2 (cid:19) σ2(cid:104)vn−2 s (cid:105) ds , (B7) where we again extracted the deterministic part of the Poisson process and all martingales dropped out. Taking the time derivative on both sides, we obtain an ODE, which also contains the lower-order moments mn−1 and mn−2: dmn dt = nrv0 mn−1+ λ [(1 − η)n − 1] − nr mn (B8) n(n − 1) 2 + σ2 mn−2. (cid:19) (cid:18) The solution of this ODE in the long-time limit t → ∞, where dmn/dt = 0, yields Eq. (6) in the main text, mn = v0 mn−1 + 1 2 (n − 1) σ2 nr (1 − η)n nr − λ r mn−2 1 + λ . (B9) Finally, we calculate the speed autocorrelation func- tion g(s, t) = (cid:104)(vs− m1)(vt− m1)(cid:105) of Eq. (B1). We define the probability distributions for the speed process P (v(cid:48)) and the conditional probability P (v, tv(cid:48), s) of having v at time t given that we have v(cid:48) at time s and obtain [(v − m1)(v(cid:48) − m1)P (v, tv(cid:48), s)P (v(cid:48))] dvdv(cid:48) (cid:90) (cid:90) (cid:90) (cid:90) (cid:104)(cid:16) (v(cid:48) − m1)e−(r+ηλ)t−s(cid:17) = = ∆v2 e−(r+ηλ)t−s , [(cid:104)v(t) − m1[v(cid:48), s](cid:105)(v(cid:48) − m1)P (v(cid:48))] dv(cid:48) (cid:105) (v(cid:48) − m1)P (v(cid:48)) dv(cid:48) (B10) where we have have used Eq. (B5) with C = v(cid:48) in the second last step. We recover Eq. (7) after setting s = t + τ . Identifying the relaxation rate αV , we can write Figure 9: Layout of the chemotaxis device. The chemotaxis chamber consists of two large reservoirs connected to a cen- tral observation area. In this study, both right and left reser- voirs were filled with bacterial cell suspension. The chemoat- tractant α-methyl-aspartate was added to the right hand side reservoir. A linear, stable chemoattractant concentration pro- file was established across the central gradient region marked in blue. The bacteria were observed by video microscopy in the field of view marked in red. Figure was adapted from Ref. [22]. (cid:90) t (cid:90) t (cid:90) t (cid:90) t σdWs− ηvsλds− r(v0−vs)ds+ 0 0 0 vt = ηvsd N λ t (B2) Note that the second and fourth term on the RHS are martingales [42]. Thus, their expectation values vanish. We will use this property when calculating the moments and autocorrelation function of the speed variable. Tak- ing the expectation value (cid:104). . .(cid:105) on both sides, we obtain the first moment: 0 (cid:90) t m1 = (cid:104)vt(cid:105) = rv0t − (r + ηλ)(cid:104)vs(cid:105)ds . (B3) 0 To ease the notation, we dropped the superscript V from the main text. Taking the time derivative on both sides, we obtain a non-homogeneous ordinary differential equa- tion (ODE): g(s, t) = = dm1 dt = rv0 − (r + ηλ)m1 (B4) Its full solution with initial value C at time t0 reads m1(t) = v0 1 + η λ r + e−(r+ηλ)(t−t0) C − v0 1 + η λ r . (B5) (cid:32) (cid:33) chemoattractant-freecell suspensionchemoattractant-containing cell suspensionfield of viewgradient region660μm1000μm the following formulas for λ and r: given by λ = r = αV , 1 + ηλ/r η + (λ/r)−1 . αV 2. Angle (cid:112) We rewrite the Langevin equation (3) from the main text as a SDE using mathematical notation: dΘt = 2DtdWt (B13) 11 (B20) (B21) (B11) (B12) (cid:104)D(cid:105) = ∆2D = D0r + DT λ , (D0 − DT )2λr λ + r (λ + r)2 . The mean value of Dt for any time t with initial condition C at time t0 is given by (cid:104)Dt(cid:105) = (cid:104)D(cid:105) + e−(λ+r)(t−t0) (C − (cid:104)D(cid:105)) . (B22) We can use the pdf p(D) to calculate the first factor on the RHS of Eq. (B14) in the long time limit, (cid:10)[2Dt] 2(cid:11) = n 2 (2D0) n 1 + λ/r + 2 (2DT ) n 1 + r/λ . (B23) the tele- The SDE contains two stochastic processes: graph process Dt, where we drop here the subscript rot used in the main text, and the white noise process dWt. These two processes are stochastically independent of each other. Thus, the moments for the angular displace- ment during time step ∆t factorize into contributions from each process, (cid:104)∆Θn(cid:105) =(cid:10)[2Dt] 2(cid:11)(cid:104)∆Wtn(cid:105) . n (B14) Inserting this expression and Eq. (B16) in Eq. (B14) leads to Eq. (9) stated in the main text. The pdf of the absolute angular displacement p(∆Θ) can be calculated straightforwardly. Using the indepen- dence of the two stochastic processes and combining Eqs. (B15), (B17), and (B18), we obtain p(∆Θ) = (cid:112) (cid:112) N (0, N (0, 2D0∆t)+ λ r λ + r λ + r 2DT ∆t) . (B24) The probability distribution function (pdf) p(∆Wt) and the absolute moments of the white noise increments ∆Wt during time step ∆t are given by p(∆Wt) = N (0, (cid:104)∆W (t)n(cid:105) = (∆t) √ ∆t) , n 2 (n − 1)!! (cid:40)(cid:113) 2 π 1 (B15) if n is odd if n is even , (B16) where N (0, σ) denotes the normal distribution with zero mean and standard deviation σ and n!! denotes the dou- ble factorial. For the telegraph process Dt with states D0 and DT , the two probabilities for being in one of the states at time t obey the following master equations: ∂tP (D0, tC, t0) = −λP (D0, tC, t0) + rP (DT , tC, t0) , ∂tP (DT , tC, t0) = λP (D0, tC, t0) − rP (DT , tC, t0) . Here, λ is the transition rate from D0 to DT and r the transition rate for the reverse process. The variable C indicates the initial condition at time t0. We first state the pdf p(D) in the long-time limit t → ∞ as well as the auto-correlation function (cid:104)DtDs(cid:105) from literature [25]: p(D0) = r λ + r , λ p(DT ) = (cid:104)DtDs(cid:105) = (cid:104)D(cid:105)2 + ∆2D e−(λ+r)t−s . λ + r , (B19) In the last equation we have introduced the mean (cid:104)D(cid:105) and the variance ∆2D in the long time limit. They are This agrees with Eq. (8) from the main text. Finally, we calculate the directional autocorrelation function g(τ ) = (cid:104)e(τ ) · e(0)(cid:105) = (cid:104)cos (Θ(τ ) − Θ(0))(cid:105). In- tegrating Eq. (3) and using the real part (cid:60) of the Euler identity eix = cos(x) + i sin(x) yields: g(τ ) = (cid:60) (cid:68) ei(cid:82) τ 0 √ (cid:69) 2DsdWs (B25) (cid:82) τ √ The term in the real part operator can be interpreted as the characteristic function of the random variable X(τ ) = 2DsdWs for wavenumber k = 1. Using the moment 0 representation of the characteristic function, we obtain g(τ ) = (cid:60) (B26) where we have defined the moments mn = (cid:104)X n(cid:105). For symmetry reasons, the odd moments vanish, mn(τ ) , n=0 ∞(cid:88) in n! m2n+1 = 0 , (B27) and the real part operator can be skipped. First, we calculate m2, where we use again the independence of the two stochastic processes dWt and Dt in the second line, (cid:90) τ (cid:112)2Ds2 dWs2(cid:105) (cid:112)2Ds2(cid:105)(cid:104)dWs1 dWs2(cid:105) (cid:112)2Ds2(cid:105)δ(s1 − s2)ds1ds2 0 0 (cid:90) τ (cid:112)2Ds1dWs1 (cid:90) τ (cid:90) τ (cid:104)(cid:112)2Ds1 (cid:90) τ (cid:90) τ (cid:104)(cid:112)2Ds1 (cid:90) τ 0 0 0 0 (cid:104)D(cid:105)ds1 = 2 = 2(cid:104)D(cid:105)τ 0 (B28) m2(τ ) = (cid:104) (B17) (B18) = = Next, we calculate the fourth moment m4, where we use the correlation function of Eq. (B19) in the fourth line: 12 (cid:90) τ (cid:90) τ 0 0 (cid:104)(cid:112)2Ds1 (cid:104)(cid:112)2Ds1 (cid:112)2Ds2 (cid:112)2Ds2 (cid:112)2Ds3 (cid:112)2Ds3 (cid:112)2Ds4(cid:105)(cid:104)dWs1dWs2dWs3 dWs4(cid:105) (cid:112)2Ds4(cid:105)3δ(s1 − s2)δ(s3 − s4)ds1ds2ds3ds4 m4(τ ) = = 0 0 0 0 (cid:90) τ (cid:90) τ (cid:90) τ (cid:90) τ (cid:90) τ (cid:90) τ (cid:90) τ (cid:90) τ (cid:90) t (cid:90) t (cid:18) 0 0 0 0 = 12 (cid:104)D(cid:105)2 + = 12(cid:0)(cid:104)D(cid:105)2τ 2 + I(cid:1) 0 0 = 12 (cid:104)D(cid:105)2τ 2 + 2 = 12 (cid:104)Ds1Ds2(cid:105)ds1ds2 (D0 − DT )2rλ (λ + r)2 e−(λ+r)s1−s2ds1ds2 ∆2D λ + r τ + e−(λ+r)(τ−0) λ + r − 1 λ + r (cid:21)(cid:19) Replacing 0 . . . ds1 0 . . . ds1, the integral I is calculated as follows: integral double the in by for r and λ, when r/λ is known from the analysis of the pdf p(∆Θ). Solving the equation for αΘ for either λ or r, we obtain the formulas (cid:20) (cid:19) 2(cid:82) s2 ∆2De−(λ+r)(s2−s1)ds1ds2 (cid:82) τ (cid:12)(cid:12)(cid:12)(cid:12)s1=s2 (cid:12)(cid:12)(cid:12)(cid:12)s2=τ s1=0 s2=0 (cid:21) − 1 λ + r e−(λ+r)(s2−s1) 1 − e−(λ+r)s2 ds2 e−(λ+r)s2 λ + r e−(λ+r)τ λ + r 0 (cid:90) s2 (cid:20) ∆2D (cid:90) τ (cid:20) (cid:20) λ + r 0 τ + s2 + (cid:90) τ (cid:90) τ 0 0 ∆2D λ + r ∆2D λ + r ∆2D λ + r I = 2 = 2 = 2 = 2 = 2 ds2 (B30) Truncating the sum of Eq. (B26) for n > 4 , we finally obtain: gΘ(τ ) =1 − (cid:104)D(cid:105)τ + (cid:104)D(cid:105)2τ 2/2 e−(λ+r)τ λ + r ∆2D λ + r τ + + (cid:18) − 1 λ + r (B31) . This form suggests a slope −(cid:104)D(cid:105) of the correlation func- tion for times τ < (λ + r)−1, which in our case means τ < 0.2 s and is just valid for the very initial time range of the correlation function. From Eq. (B31) we can ex- tract another linear approximation by concentrating on the time range (λ + r)−1 < τ < (cid:104)D(cid:105)−1. It gives Eq. (11) from the main text, (cid:18) (cid:19) gΘ(τ ) = 1 − (cid:104)D(cid:105) − ∆2D λ + r τ , (B32) from which we obtain an expression for the relaxation rate αΘ measured in experiments. It is determined by (cid:104)D(cid:105) and the second term in the brackets is a correc- tion. But it is sufficient to determine separate values (B29) λ = r = ∆2D (1 + (λ/r)−1)((cid:104)D(cid:105) − αΘ) ∆2D (1 + λ/r)((cid:104)D(cid:105) − αΘ) . , (B33) (B34) 3. Numerical investigations of the directional autocorrelation function The directional autocorrelation function gΘ(τ ) = (cid:104)e(τ ) · e(0)(cid:105) has an exponential form in experiments up to ca. 2s (see Fig. 4). Here, we validate this depen- dence by numerically solving Eq. (3) with the inferred parameters of table I. The semi-logarithmic plot in Fig. 10(a) shows the resulting autocorrelation function (blue data points). It is in good agreement with the exponen- tial decay of Eq. (10) using the relacation rate αΘ from Eq. (11), which we derived in the previous section in Eq. (B32). This validates our proposition for the relaxation rate. Moreover, we can further validate the exponential fit to the experimental directional autocorrelation function using the theoretical value for the relaxation rate. After having inferred the reduced parameter set (λ/r, D0, DT ) as described in the main text using the pdf p(Θ), we de- termine the directional autocorrelation function by sim- ulating the angle process with the reduced parameter set for different values of the parameter λ. Figure 10(b) shows the mean squared error Σ of the simulated autocor- relation function compared to the experimental function plotted versus the tumble rate λ. The best match is for a λ very close to the value shown in table I, which was determined using the theoretical prediction of Eq. (11) 13 vision, we use the whole data set for each sector but weight the data by a Gaussian kernel similar to Ref. [22]. The speed moments for N experimental trajectories when conditioning on a specific swimming angle θ are then cal- culated according to (cid:80)N (cid:80) (cid:80)N i=1 i=1 (cid:80) (cid:17) (cid:16)− [Θi(t−2∆t)−θ]2 (cid:16)− [Θi(t−2∆t)−θ]2 (cid:17) 2∆θ2 2∆θ2 , (D1) (cid:104)mV n (cid:105) = t vi(t)n exp t exp where we have introduced the width of a section, ∆θ = 0.125π, and their centers θ. Note that we use the actual orientation angle Θi(t− 2∆t) of the second previous time Figure 11: Jump height η conditioned on swimming angle θ. step to calculate the moments. Tumble events have a finite duration of around 2∆t and this ensures that the whole tumble is connected to the condition of the pre- vious run. The same Gaussian kernels are applied when we calculate the histogram of angular displacements and the autocorrelation functions for speed and direction. Appendix E: Jump height conditioned on swimming angle θ. Figure 10: Left: Semi-logarithmic plot of the directional au- tocorrelation function from a numerical solution of Eq. (3) using the inferred parameters from table I (blue data points). The orange line shows an exponential decay with the relax- ation rate from Eq. (11). Right: Mean squared deviation between the simulated directional autocorrelation function gΘ(τ ) = (cid:104)e(t + τ ) · e(t)(cid:105) and the experimental curve for dif- ferent tumble rates λ. The global minimum at λ = 0.81 s−1 verifies the use of the theoretical expression (11) for the re- laxation rate. for the relaxation rate αΘ. Appendix C: The method of bootstrapping Bootstrapping allows to derive an estimate of the stan- dard deviation of the inferred parameters without the need of repeated experiments [44]. Similar to Ref. [22], we create synthetic ensembles by randomly mixing sub- sets of the original data set. Let T0 = {t1, ...., tN} be the set of original trajectories. Pulling N random tra- jectories of this set and laying them back after each pull, one obtains a bootstrap sample T1 = {t1, ..., tN}, where single trajectories can appear several times. We create K = 100 of these bootstrap samples, apply our inference technique to each sample, and obtain a distribution of values for each swimming parameter. The error bars in the main text are the standard deviation from the mean of each swimming parameter. Appendix D: Smooth weighting of data Figure 11 shows the relevant plot. There is no system- atic dependence of η on the swimming angle. The conditioning of section IV B needs the division of Instead of a discrete di- the data in different sectors. [1] H. C. Berg, E. coli in Motion (Springer Science & Busi- phys. J. 111, 630 (2016). ness Media, 2008). [2] J. Adler, Science 166, 1588 (1969). [3] H. C. Berg and D. A. Brown, Nature 239, 500 (1972). [4] S. M. Block, J. E. Segall, and H. C. Berg, Cell 31, 215 (1982). [5] M. J. Schnitzer, Phys. Rev. E 48, 2553 (1993). [6] V. Sourjik and N. S. Wingreen, Curr. Opin. Cell Biol. 24, 262 (2012). [8] J. Saragosti, V. Calvez, N. Bournaveas, B. Perthame, A. Buguin, and P. Silberzan, PNAS 108, 16235 (2011). [9] J.-B. Masson, G. Voisinne, J. Wong-Ng, A. Celani, and M. Vergassola, PNAS 109, 1802 (2012). [10] F. C. Cheong, C. C. Wong, Y. Gao, M. H. Nai, Y. Cui, S. Park, L. J. Kenney, and C. T. Lim, Biophys. J. 108, 1248 (2015). [11] M. Wu, J. W. Roberts, S. Kim, D. L. Koch, and M. P. [7] L. Turner, L. Ping, M. Neubauer, and H. C. Berg, Bio- DeLisa, Appl. Environ. Microbiol. 72, 4987 (2006). 024 [s]101100g()SimTheo0123tumble rate 0.020.040.060.08Deviation 0π2πθ−0.8−0.7η 14 [12] Y. S. Dufour, S. Gillet, N. W. Frankel, D. B. Weibel, and [28] G. E. Uhlenbeck and L. S. Ornstein, Phys. Rev. 36, 823 T. Emonet, PLoS Comput. Biol. 12, e1005041 (2016). (1930). [13] S. M. Vater, S. Weisse, S. Maleschlijski, C. Lotz, F. Kos- chitzki, T. Schwartz, U. Obst, and A. Rosenhahn, PloS one 9, e87765 (2014). [14] N. Figueroa-Morales, T. Darnige, C. Douarche, V. Mar- tinez, R. Soto, A. Lindner, and E. Cl´ement, arXiv preprint arXiv:1803.01295 (2018). [29] W. Graham and D. McLaughlin, Water Resour. Res. 25, 2331 (1989). [30] F. M. Bandi and T. H. Nguyen, J. Econom. 116, 293 (2003). [31] J. Saragosti, P. Silberzan, and A. Buguin, PloS one 7, e35412 (2012). [15] K. Taute, S. Gude, S. Tans, and T. Shimizu, Nat. Comm. [32] P. Zengel, A. Nguyen-Hoang, C. Schildhammer, R. Zantl, 6, 8776 (2015). V. Kahl, and E. Horn, BMC Cell Biol. 12, 21 (2011). [16] L. Turner, W. S. Ryu, and H. C. Berg, J. Bacteriol. 182, [33] J. Kapur, P. Sahoo, and A. Wong, Comput. Vis. Graph. 2793 (2000). 29, 273 (1985). [17] M. Hintsche, V. Waljor, R. Grossmann, M. J. Kuhn, K. M. Thormann, F. Peruani, and C. Beta, Sci. Rep. 7, 16771 (2017). [34] J. C. Crocker and D. G. Grier, J. Coll. Interfat. Sci. 179, 298 (1996). [35] A. Savitzky and M. J. Golay, Anal. Chem. 36, 1627 [18] S. H. Larsen, R. W. Reader, E. N. Kort, W.-W. Tso, and (1964). J. Adler, Nature 249, 74 (1974). [36] M. Theves, J. Taktikos, V. Zaburdaev, H. Stark, and [19] R. Macnab and D. Koshland Jr, J. Mol. Biol. 84, 399 C. Beta, Biophys. J. 105, 1915 (2013). (1974). [37] K. Son, F. Menolascina, and R. Stocker, PNAS 113, 8624 [20] H. C. Berg, Random walks in biology (Princeton Univer- (2016). sity Press, 1993). [21] J. Tailleur and M. Cates, Phys. Rev. Lett. 100, 218103 (2008). [22] O. Pohl, M. Hintsche, Z. Alirezaeizanjani, M. Seyrich, C. Beta, and H. Stark, PLoS Comput. Biol. 13, e1005329 (2017). [38] A. Celani and M. Vergassola, PNAS 107, 1391 (2010). [39] R. M. Macnab and D. Koshland, PNAS 69, 2509 (1972). [40] O. Pohl, Chemotaxis of self-phoretic active particles and bacteria (TU Berlin, 2016), PhD Thesis, URL http:// dx.doi.org/10.14279/depositonce-5409. [41] L. Xie, T. Altindal, S. Chattopadhyay, and X.-L. Wu, [23] J. Saragosti, P. Silberzan, and A. Buguin, PloS one 7, PNAS 108, 2246 (2011). e35412 (2012). [42] E. Cinlar, Introduction to stochastic processes (Courier [24] G. Rosser, A. G. Fletcher, D. A. Wilkinson, J. A. de Beyer, C. A. Yates, J. P. Armitage, P. K. Maini, and R. E. Baker, PLoS Comput. Biol. 9, e1003276 (2013). [25] B. Lindner, Stoch. Meth. Neurosci. p. 1 (2009). [26] C. Van Den Broeck, J. Stat. Phys. 31, 467 (1983). [27] J. Strefler, W. Ebeling, E. Gudowska-Nowak, and L. Schimansky-Geier, Eur. Phys. J. B 72, 597 (2009). Corporation, 2013). [43] L. C. G. Rogers and D. Williams, Diffusions, Markov processes and martingales: Volume 2, Ito calculus, vol. 2 (Cambridge university press, 1994). [44] B. Efron and R. J. Tibshirani, An introduction to the bootstrap (CRC press, 1994).
1709.05727
1
1709
2017-09-17T23:32:08
Molecular Mechanism of Transition from Catch-Bond to Slip-Bond in Fibrin
[ "physics.bio-ph" ]
The lifetimes of non-covalent A:a knob-hole bonds in fibrin probed with the optical trap-based force-clamp first increases ("catch bonds") and then decreases ("slip bonds") with increasing tensile force. Molecular modeling of "catch-to-slip" transition using the atomic structure of the A:a complex reveals that the movable flap serves as tension-dependent molecular switch. Flap dissociation from the regulatory B-domain in $\gamma$-nodule and translocation from the periphery to knob `A' triggers the hole `a' closure and interface remodeling, which results in the increased binding affinity and prolonged bond lifetimes. Fluctuating bottleneck theory is developed to understand the "catch-to-slip" transition in terms of the interface stiffness $\kappa =$ 15.7 pN nm $^{-1}$, interface size fluctuations 0.7-2.7 nm, knob `A' escape rate constant $k_0 =$ 0.11 nm$^2$ s$^{-1}$, and transition distance for dissociation $\sigma_y =$ 0.25 nm. Strengthening of the A:a knob-hole bonds under small tension might favor formation and reinforcement of nascent fibrin clots under hydrodynamic shear.
physics.bio-ph
physics
Molecular Mechanism of Transition from Catch-Bond to Slip-Bond in Fibrin Rustem I. Litvinov1,2, Olga Kononova3,4, ¶, Farkhad Maksudov3, Artem Zhmurov4, Kenneth A. Marx3, John W. Weisel1,*, and Valeri Barsegov3,4,* 1Department of Cell and Developmental Biology, University of Pennsylvania School of Medicine, Philadelphia, PA 19104, USA 2Institute of Fundamental Medicine and Biology, Kazan Federal University, Kazan 420008, Russian Federation 3Department of Chemistry, University of Massachusetts, Lowell, MA 01854, USA 4Moscow Institute of Physics and Technology, Moscow Region 141700, Russian Federation *[email protected], valeri [email protected] ¶Current address: Department of Material Science and Engineering, University of California, Berkeley, Berkeley, CA, 94720, USA ABSTRACT The lifetimes of non-covalent A:a knob-hole bonds in fibrin probed with the optical trap-based force-clamp first increases ("catch bonds") and then decreases ("slip bonds") with increasing tensile force. Molecular modeling of "catch-to-slip" transition using the atomic structure of the A:a complex reveals that the movable flap serves as tension-dependent molecular switch. Flap dissociation from the regulatory B-domain in γ-nodule and translocation from the periphery to knob 'A' triggers the hole 'a' closure and interface remodeling, which results in the increased binding affinity and prolonged bond lifetimes. Fluctuating bottleneck theory is developed to understand the "catch-to-slip" transition in terms of the interface stiffness κ = 15.7 pN nm −1, interface size fluctuations 0.7-2.7 nm, knob 'A' escape rate constant k0 = 0.11 nm2 s−1, and transition distance for dissociation σy = 0.25 nm. Strengthening of the A:a knob-hole bonds under small tension might favor formation and reinforcement of nascent fibrin clots under hydrodynamic shear. Introduction Fibrin is the end product of blood clotting that constitutes a proteinaceous 3D network providing a filamentous mechanical scaffold of clots and thrombi. Formation of fibrin is essential for hemostasis, thrombosis, and antimicrobal host defense; fibrin is also widely used as a biomaterial.1 Fibrin is formed from a blood plasma protein, fibrinogen, which converts to monomeric fibrin that polymerizes to form soluble fibrin protofibrils. These further elongate and aggregate laterally into insoluble thick fibers that branch to form the space-filling network (Fig. 1a-c). During and after polymerization fibrin is covalently cross-linked by a plasma transglutaminase, factor XIIIa, that makes fibrin stiff and resistant to enzymatic lysis (Fig. 1a-c).2 Mechanical stability of nascent blood clots in response to forces imposed by the blood flow, contracting platelets, and other dynamic factors is determined by the strength of the knob-hole interactions prior to cross-linking by factor XIIIa. Consequently, the dynamics of association-dissociation transitions in knob-hole complexes govern formation of fibrin, influence the final structure and mechanical stability of clots and thrombi, including clot rupture (embolization) and shrinkage (contraction, retraction). Impaired knob-hole interactions result in loose, weak, unstable clots and are associated with the tendency to bleed. Dense fibrin networks show increased stiffness, higher fibrinolytic resistance and mechanical resilience, which may predispose individuals to thrombotic cardiovascular diseases, such as heart attack and ischemic stroke.3 Fibrinogen, the soluble fibrin precursor, is a 340-kDa protein comprising three pairs of polypeptide chains, Aα, Bβ , and γ, arranged into one central globular E region and two lateral globular D regions connected with α-helical coiled coils (Fig. 1a). Thrombin splits off two pairs of fibrinopeptides A and B from the N-termini of the fibrinogen's Aα and Bβ chains, respectively, in the central E region. This results in the exposure of binding sites (knobs) 'A' and 'B' that interact with constitutively accessible complementary sites (holes) 'a' and 'b' located in the γ- and β -nodules, respectively, of the lateral D regions of another fibrin molecule (Fig. 1a-c).2, 4, 5 X-ray crystallographic studies of fibrinogen fragments revealed that holes 'a' and 'b' interact with the peptides GPRP and GHRP that mimic knobs 'A' and 'B', respectively, which comprise the newly exposed N-terminal motifs of the α and β chains of fibrin.4, 6 Binding of knobs 'A' to holes 'a' is necessary for fibrin polymerization, while the B:b bonds play a secondary role.2 We employed single-molecule forced unbinding assays to probe the strength of A:a knob-hole bond7, 8 which revealed unusual strengthening with the increasing pulling force applied to disrupt the bond, but the nature of this finding remained unclear. Such counterintuitive behavior has been described in the literature for a number of receptor-ligand pairs as the "catch" bond9, 10 in contrast to the commonly known "slip" bond that dissociates faster with the increasing force. Interestingly, a non-covalent bond can behave as a catch bond at low forces (typically <30-40 pN) and as a slip bond at higher forces, thus displaying the dual "catch-slip" character. Several receptor-ligand complexes showing the catch-slip transition have been characterized including coupled cell adhesion molecules and glycoprotein ligand-1,10 E-cadherin dimer,11 integrin α5β 1 and fibronectin,12 bacterial adhesin FimH,13, 14 von Willebrand factor and receptor GP1b,15, 16 actomyosin,17 microtubule- kinetochore,18 and microtubule-dynein19, 20 complexes. The catch-slip phenomenon was studied experimentally14, 15, 21–24 and computationally.14, 25–30 Theoretical models have been purposed including the two-state model,31, 32 two-pathway model,33 sliding-rebinding model,34 hydrogen bond network model,35 and other models.36, 37 Yet, the atomic-level structural basis underlying the catch-slip transition in receptor-ligand complexes has eluded detailed characterization9. Here, we combined the single-molecule forced unbinding experiments in vitro and in silico using nanomechanical measurements and Molecular Dynamics (MD) simulations to resolve the structural mechanism underlying the dual catch-slip response of fibrin polymers to tension. We show that the strength of A:a knob-hole bonds first increases with tensile force up to f ≈ 30-35 pN (catch bond) and then decreases with force at f > 35 pN (slip bonds). Forced dissociation assays in silico revealed dynamic remodeling of the A:a association interface, which results in a manifold of bound states with tension-dependent binding affinity. We developed new fluctuating bottleneck theory to model the experimental distributions of bond lifetimes and average bond lifetime as a function of force. The results provide a comprehensive structure-based interpretation of the experimentally observed catch-slip dynamics of dissociation of the A:a knob-hole bonds, the strongest non-covalent interactions in early stages of fibrin polymerization (Fig. 1). Our theory can be extended and generalized to characterize biomolecular interactions in receptor-ligand pairs that form deep binding pockets. Results Dissociation kinetics of the A:a knob-hole bonds under constant tensile force Bond lifetime of the fibrin-fibrinogen complex as a function of tensile force: The A:a knob-hole interactions were repro- duced at an interface during repeated touching of two microscopic beads coated covalently with fibrinogen or monomeric fibrin. The fibrin molecule is a source of knobs 'a' and fibrinogen molecule is a source of holes 'a'. By touching the fibrin-coated surface with the fibrinogen-coated surface, we allowed these molecules to associate forming the A:a knob-hole complex (binding phase). Next, the fibrin-coated and fibrinogen-coated surfaces were retracted by a constant tensile (pulling) force to dissociate the A:a knob-hole bond (unbinding phase). By repeating the binding-unbinding cycles, we were able to measure the times-to-dissociation (lifetimes) for the fibrin-fibrinogen complexes and to probe the dependence of A:a knob-hole bond lifetimes on a constant pulling force f . We varied the pulling force in the 5-60-pN-range and collected for each force value the distributions of bond lifetimes P(τ; f ), which were then used to calculate the average bond lifetimes (cid:104)τ(cid:105). We found that the fibrin-fibrinogen interactions display a non-monotonic dependence of (cid:104)τ(cid:105) on f : (cid:104)τ(cid:105) first increased with f up to f = 30-35 pN and then decreased at f > 40 pN (Fig. 2a). To distinguish the A:a knob-hole binding from other interactions, we carried out a number of control experiments. Weak non-specific surface-to-surface adhesion events produced interaction signals with the bond lifetimes <0.03s. Specificity was also tested by coating the interacting surfaces with an inert protein lacking knobs and holes (bovine serum albumin, BSA) or with a relevant protein lacking knobs 'a' and 'b' (fibrinogen). The control interactions (fibrinogen/BSA, fibrin/BSA, fibrinogen/fibrinogen interfaces) revealed mostly (93-97%) short-lived attachment signals with bond lifetimes <0.03s and a small fraction (0.6-2.5%) of more stable interactions lasting 0.04-0.5 s (Table S1). The bond lifetimes of control interactions did not show any dependence on force (Fig. 2a). To distinguish the A:a knob-hole binding from other associations, we measured the fibrinogen-fibrin interaction in the absence and presence of the GPRPam peptide – specific competitive knob 'a' inhibitor. GPRPam suppressed interactions lasting >0.5 s (Fig. S1, Table 1), which were, therefore, considered to reflect the specific A:a knob-hole interactions. The binding probability for these interactions lasting >0.5s was several-fold higher in the fibrinogen-fibrin system than in control non-specific interactions described above, including fibrin-fibrinogen interactions inhibited by GPRP peptide (Fig. S2, Table S1). To test whether the intermediately strong interactions with bond lifetimes 0.03s < τ < 0.5s contribute to the observed dependence of (cid:104)τ(cid:105) on f , we also compared the bond lifetimes obtained either by including or excluding these intermediate- strength interactions from the data sets. The profiles of (cid:104)τ(cid:105) vs. f (Fig. S3) as well as the cumulative binding probability vs. f (Fig. S4) were quite similar with a corresponding small shift to shorter bond lifetimes (Fig. S4) and to a higher binding probability (Fig. S3). Based on these findings, we concluded that the bond lifetimes τ > 0.5s represent specific fibrin-fibrinogen or A:a knob-hole interactions. We included these data into subsequent statistical analyses and modeling, while the bond lifetimes τ < 0.5s were excluded from the data analysis. 2/29 Bond lifetimes of the D:E complex as a function of tensile force: To minimize the role of non-specific interactions due to large size of fibrin and fibrinogen molecules, we replaced the full-length fibrinogen with its smaller proteolytic fragment D containing mainly a lateral globular portion. Monomeric fibrin was replaced with fragment E comprising mainly the central globule. Fragment D bears one constitutively open hole 'a'. Fragment E can exist in three variant, namely: i) it can bear both knobs 'a' and 'b' after cleaving of fibrinopeptides A and B with thrombin (called fragment desAB-E); ii) it can bear only knobs 'a' after cleaving of fibrinopeptide A with batroxobin (called fragment desA-E) and iii) it can have no knobs if remains untreated (fragment E) (Fig. 1d). We found that specific interactions between fragments D and desAB-E also displayed dual catch-slip profile. The average bond lifetimes increased with force up to f = 30-40 pN and then decreased at f > 40 pN (Fig. 2b). The bell-like curve of (cid:104)τ(cid:105) characteristic of the catch-to-slip transition was quite similar to the one obtained for the fibrinogen:fibrin interactions (Fig. 2b) albeit with shorter average bond-lifetimes. Importantly, the dual catch-slip behavior was largely suppressed with addition of GPRP (Fig. S5) and it disappeared altogether following a substantial reduction in surface density of fragment D (Fig. S6). These results strongly indicate that the D/desAB-E interactions are specific and that they reflect formation-dissociation of the knob-hole bonds. Exposure of knobs 'a' in fragment desA-E (without knobs 'b') preserves their ability to interact with fragment D in the catch-slip fashion. However, the average bond lifetimes were substantially shorter and the force-dependence of bond lifetimes was less pronounced (Fig. 2b). Intact fragment E with uncleaved fibrinopeptides A and B was non-reactive with fragment D (Fig. 2b). This provides yet another evidence that the measured knob-hole interactions of fragment D with fragments desAB-E and desA-E was specific and reflected the A:a knob-hole interactions. Molecular Modeling of A:a knob-hole interactions Dynamic force measurements in silico: We focused on the isolated A:a knob-hole complex using the structural model of A:a knob-hole complex as a part of reconstructed fibrin-fibrin D:E:D interactions based on the crystallographic data (Fig. 1e,g). Potentially, A:b interactions (knobs 'a' binding to holes 'b') and B:a interactions (knobs 'b' binding to holes 'a') might also affect the fibrinogen-fibrin complex lifetime, but our previous studies showed that the B:a interactions are unlikely to exist38 and that formation of the A:b knob-hole bonds is structurally and thermodynamically unfavorable.39 Fibrinogen molecule has two pairs of strong Ca2+-binding sites2 and the equilibrium dissociation constant for the high-affinity Ca2+-binding sites in fibrinogen (Kd ∼ 1 µM) implies that these sites are fully occupied with Ca2+ in plasma environment. In our experiments, the catch-slip transition was observed both with and without Ca2+ (data not shown), and so we did not include calcium ions in the computational modeling. We probed the strength of A:a knob-hole interactions using ramped force f (t) (see Methods) with the pulling velocity ν f = 103 and 104 µm/s. The distinct Pathway 1 (Pathway 2) of A:a knob-hole bond dissociation (Fig. S7a,b) is characterized by faster (slower) dissociation at lower (higher) molecular force F. For the slower ν f = 103 µm/s, the bonds yielded at F ≈ 60 pN in 50% of 10 simulation runs (Pathway 1) and at F ≈ 90 pN in 25% of runs (Pathway 2), while in the remaining 25% of trajectories the A:a knob-hole bonds dissociated at F ≈ 70 pN (Fig. S7a). For the faster ν f = 104 µm/s, the A:a knob-hole bonds yielded at F ≈ 90 pN (Pathway 1) and at F ≈ 130 pN (Pathway 2) 70% and 30% of 10 trajectories, respectively (Fig. S7b). Binding affinity and maps of binding contacts: Next, we analyzed the simulations output. We monitored the dissociation dynamics by projecting the total number of persistent binding contacts Q between the residues in hole 'a' (in the γ-nodule ) and residues in knob 'a' (in the α chain) as a function of time t. A pair of amino acids i and j is said to form a binary contact if the distance between the center-of-mass of their side chains ri j < 6.5 A persists for more than 10 ns. Q is related to the binding affinity (i.e. more contacts means stronger binding) and it reflects instantaneous changes in the bond strength. Results are displayed in Fig. S8a (see also Fig. S7c, d). The profiles of Q vs. t show that at the beginning Q ≈ 20 contacts (native bound state of the A:a complex), but with force ramping up Q changes. For Pathways 1 resulting in the lower dissociation force F, Q decays to zero with time. The moment of time t = τ, at which Q(τ) = 0, marks complete dissociation of the A:a knob-hole bond. However, for Pathway 2 corresponding to higher values of F, the dynamics of Q is non-monotonic (Fig. S8a and Fig. S7c, d): Q initially increased to 30-35 contacts and then decreased to zero at longer times. Hence, the higher values of F corresponding to Pathway 2 of the A:a knob-hole complex dissociation are directly correlated with the higher number of contacts Q, which is proportional to the knob-hole binding affinity. To gather the residue-level information about the high- and low-affinity bound states of the A:a knob-hole complex, we analyzed entire maps of knob-hole contacts at the binding interface (residues γSer240–γLys380). The maps of binding contacts for the native bound state and for the intermediate states right before dissociation are compared in Fig. S8c,d for Pathway 1 and 2 trajectories. The bottom corner in the map corresponds to the native state, and the top corner corresponds to the intermediate state (t = 0.1 µs). Dissociation along Pathway 1 is accompanied by a slight increase in contacts density between the movable flap (residues γPhe295–Thr305) and β -sheet stack of the B-domain (residues γIle242–γGly283; compare the areas circled by solid and dashed ovals in Fig. S8c) and no substantial change in contacts density between the receptor and knob 'a'. This correlates with decay in Q for this trajectory (green curve in Fig. S8a). For Pathway 2, additional binding contacts form between the movable flap and knob 'a', while the contacts between movable flap and β -sheet stack are disrupted (compare the areas 3/29 circled by solid and dashed ovals in Fig. S8d). This is reflected by an increase in Q for this trajectory (Fig. S8a). Hence, under the influence of pulling force, hole 'a' transitions between two conformation types: one is characterized by weak interactions of movable flap with the β -sheet in B-domain leading to faster dissociation of the knob 'a' at a lower unbinding force F (Pathway 1); and the other type of conformation favors formation of additional binding contacts between the movable flap and knob 'a', hence, prolonging the bond lifetime (Pathway 2). Low-affinity vs. high-affinity bound states: Analysis of binding affinity showed that in Pathway 2 the forced dissociation occurs from the high-affinity bound state compared to the low-affinity bound state observed in Pathway 1. We analyzed the simulation output to characterize conformational transitions that occur in the A:a knob-hole binding interface under tension. Fig. 3 shows the time-dependent profiles of F and Q and corresponding atomic structures of the A:a knob-hole complex observed along dissociation Pathways 1-2. We found that the transition of the receptor-ligand complex from the low- to high-affinity bound states is controlled by the movable flap, which serves as a tension-sensitive "molecular switch", which triggers opening and closing of the binding interface. In Pathway 1 (Fig. 3), the moveable flap is far away from the ligand (snapshots 2a,3a in Fig. 3). Negatively charged residues γAsp291, γAsp294 and γAsp297 interact with positively charged residues γArg256 and γArg275 of the β -sheet stack in B-domain. In some trajectories, we also observed a structural transition in the movable flap from the random coil to the β -strand, which then associates with the β -sheet stack of B-domain. As a result, the binding interface opens, facilitating dissociation of knob 'a' from the low-affinity bound state. Corresponding to these observations an increase in the contact density between the movable flap and B-domain is manifest on the contact map (circled area in Fig. 3, left). In Pathway 2 (Fig. 3), the movable flap extends and translocates toward the ligand, forming additional binding contacts between residues γAsp297, γAsp298, γPro299, γSer300, γAsp301, γLys302, γPhe303, γPhe304 in the flap and residues αGly17, αPro18, αArg19, αVal20, αGlu22, αTrp33 in knob 'a', which stabilize the high-affinity bound state (snapshots 2b,3b in Fig. 3). This flap's displacement also triggers loop I and interior region straightening, which results in formation of additional binding contacts between residues γLys321, γPhe322, γGlu323 in loop I and residues αGlu22, αArg23, αHis24 in knob 'a', and between residues γAsn337, γCys339, γHis340, γAla341, γAsn361, γGly362, γTyr363 in interior region and residues αGly17, αArg19, αGlu22, αArg23, αHis24, αGln25 in knob 'a' (Fig. S8d). The binding pocket shrinks and binding interface narrows, which results in the prolongation of bonds lifetimes. Role of movable flap: Next, we correlated the dynamics of binding contacts for the knob 'a' and entire binding interface Q and for the knob 'a' and movable flap QMF. The results for f = 30 pN (Fig. S8b) reveal the non-monotonic behavior of Q and QMF due to reversible tilting back and forth of the movable flap concomitant with its elongation and contraction. This results in the flap repeatedly switching off and on interactions with the β -sheet stack of B-domain. Additional (cryptic) binding contacts formed between the residues in movable flap and knob 'a' under tension (γPro299–αVal20, γSer300–αGly17, γSer300–αPro18, γAsp301–αArg19, γLys302–αGlu22, γPhe303–αPro18, γPhe303–αTrp33, γPhe304–αVal20) are essential for the increased strength of the A:a knob-hole bond at higher forces (Fig. 3). We quantified the movable flap elongation DMF by calculating the distance between movable flap and the β -sheet stack of B-domain (Fig. S8b). We selected three pairs of residues in movable flap and β -sheet stack: γGly296–γAla282, γAsp297–γGly283, and γAsp298–γGly284. For each pair, we calculated DMF as a function of time (force) and correlated changes in DMF with changes in Q. The results in Fig. S8b show that the increase in Q from 17 to 25 contacts is accompanied by ∼1.5-fold increase in DMF from 1.5 nm to 2.3 nm, which corresponds to the moveable flap translocation from the periphery to the knob 'a'. The decrease in Q from 25 to 17 contacts is correlated with the decrease in DMF from 2.3 to 1.3 nm, which corresponds to the movable flap tilting back (Fig. S8d). Similar results were observed for f = 35 and 40 pN (data not shown). Therefore, the binding affinity increase and the movable flap translocation are positively correlated. Hence, the A:a knob-hole complex transformation to the high-affinity conformation is controlled by tension-dependent translocation of the movable flap (Fig. 3). Dynamics remodeling of A:a association interface: We estimated the interface width X using the radius of cross-sectional area formed by connecting residues γAsp297, γGlu323, and γAsn361 in the movable flap, loop I and interior region, respectively. The results are presented in Fig. 4a,b, which shows that in Pathway 1 (low-affinity bound state) X fluctuates between 1.3 nm and 2.5 nm (interface wide open), whereas in Pathway 2 (high-affinity bound state) X decreases from 1.3 nm to 0.8 nm (interface closes). Hence, the movable flap translocation also results in the hole 'a' closing as well as decreasing of the interface width. Next, we analyzed the entire probability distributions of X, P(X) presented in Fig. 4c and d. The profiles of P(X) in Fig. 4c, corresponding to the trajectory of X from Fig. 4b move towards smaller X, thus, displaying continuous dynamics of the A:a knob-hole binding interface closing. Yet, the profiles P(X) in Fig. 4d reveal more discrete-like (two-state) dynamics of interface remodeling. Hence, quantitative analysis shows the evidence that hole 'a' is fluctuating bottleneck with tension-dependent width of the A:a knob-hole interface, i.e. X = X( f ). We were not able to discriminate between the discrete and continuous modes of evolution of X due to limited number of simulation runs, which took almost two years to complete. 4/29 Fluctuating bottleneck theory Bottleneck model of knob 'a' binding pocket: The main results from in vitro and in silico assays are the following: i) the A:a knob-hole bond first becomes stronger (increased affinity) and then becomes weaker (decreased affinity) with increasing force; ii) the force application promotes structural rearrangements resulting in binding interface remodeling; and iii) there is a competition between the dynamics of hole 'a' closure and kinetics of knob 'a' escape. To capture these observations, we propose a model which treats the binding pocket (hole 'a') as a fluctuating bottleneck.40–44 Pulling force affects both the dynamics of binding interface remodeling and kinetics of knob 'a' escape, and so the A:a bond lifetime τ and interface width X are coupled (Fig. 4e). Larger/smaller X facilitates faster/slower unbinding with shorter/longer bond lifetime τ. The A:a knob-hole bound state population P(X,t) is described by the kinetic equation: dP(X,t) dt = −K(X)P(X,t) + L(X,t) In equation (1) above, the first term describes the kinetics of knob 'a' escape from the bottleneck of size X with rate K(X) = kXα (1) (2) depending on shape parameter α (bottleneck geometry), and escape rate constant k. The rate constant is expected to increase with force, and here we use the Bell model32, 45 for k( f ) = k0 exp[σy f /kBT ], where k0 is the attempt frequency and σy is the transition distance for dissociation. Simulations show that i) X is determined by the cross-sectional area, and so we set α = 2 in equation (2); and that ii) characteristic timescale η (milliseconds) is much shorter than bond lifetimes (seconds; Fig. 2), and so we set X = (cid:104)X(cid:105). In equation (1), operator L describes the dynamics of X: By substituting equations (2) and (3) in equation (3), we arrive at the Smoluchowsky equation for P(X,t): ∂ P(X,t) ∂t ∂ 2 ∂ X 2 + κ ζ (X −(cid:104)X(cid:105)) ∂ ∂ X + κ ζ P(X,t)− k(cid:104)X(cid:105)2P(X,t) (cid:21) which can be solved as described in the SI to obtain the Green's function solution given by G(X,X0;t). To obtain the distribution of bond lifetimes P(t), we average over the initial values (X0) and sum over the final values (X), 0 dX P(t) = dX0G(X,X0;t)P(X0) (5) From simulations, the initial values of X are sharply peaked at a fixed value x0, and so P(X0) = δ (X0 − x0). By substituting P(X0) and expression for G(X,X0;t) into equation (5) and performing the integration, we obtain: exp(cid:2)−k(cid:104)X(cid:105)2t(cid:3)(cid:20) (cid:18) 1 + Er f (cid:104)X(cid:105) [2πkBT (1− e−2t/η )]1/2 (cid:19)(cid:21) L = ∂ ∂ X ∂ ∂ X + (cid:21) κ ζ (X −(cid:104)X(cid:105)) (cid:20)kBT (cid:20)kBT ζ = ζ (cid:90) ∞ (cid:90) ∞ 0 P(t) = k(cid:104)X(cid:105)2 2 The average bond lifetime is given by (cid:90) ∞ 0 (cid:104)τ(cid:105) = tP(t)dt (3) (4) (6) (7) In the continuous version, the binding pocket size X changes continuously (Fig. 4c), and the average size is given by (cid:104)X(t)(cid:105) = [X0e−t/η + (x0 − f /κ)(1− e−t)/η )]Θ( fc − f ) + x1Θ( f − fc), where Θ(x) is the Heaviside step function, X0 is the initial value and η = ζ /κ is the characteristic time for conformational fluctuations of the bottleneck (see SI). At a critical force f ≈ fc, (cid:104)X(cid:105) should reaches the minimum x1 = x0 − fc/κ (interface is closed). In the discrete version, X is a discrete random variable (Fig. 4d) interconverting between the open state X = x0 and closed state X = x1 < x0 (Fig. 4d) with the populations s0 = 1/(1 + exp[−ε0/kBT ]) and s1 = 1− s0 (ε0 is the energy difference). Pulling force favors the closed conformation by changing the state populations, and so (cid:104)X( f )(cid:105) = x0s0( f ) + x1s1( f ) and s0( f ) = [1 + exp[−(ε0 − σx f )/kBT )]−1, where σx is the transition distance. The critical force is defined as force f = fc at which s1( f ) (cid:29) s0( f ) (see SI). A:a knob-hole bond lifetimes: insights into molecular dimensions, and nanomechanics: We performed a fit of theo- retical curves of (cid:104)τ(cid:105) vs. f calculated using equations (S2), (S3), (6) and (7) to experimental data points (Fig. 2b). The friction coefficient was set to ζ = 10−3 pN nm−1s (diffusion of amino acids in water). The performance of continuous and discrete 5/29 versions of the model compared in Fig. 5 shows excellent agreement between theory and experiment. Model parameters are in Table 1. The continuous model has five parameters: interface stiffness κ, force-free interface width x0, minimal interface width (hole 'a' is closed) x1, escape rate constant k0, and transition distance for dissociation σy (i.e. distance from the bound state to the transition state along dissociation path). The discrete model has additional parameters: transition distance from the low- to high-affinity bound state σx and energy difference between these states ε0 (Table 1). Parameters κ, x0, and x1 can be directly accessed in the simulations, which revealed the following ranges: 10-30 pN/nm for κ, 2.0-2.7 nm for x0, and 0.5-0.9 nm for x1. Hence, the continuous model provides better agreement with simulations (Table 1). Discussion We employed the optical trap to measure the dynamic strength of single A:a knob-hole bonds – the strongest non-covalent interactions in fibrin polymerization (Fig. 1). We used the force-clamp to profile the average bond lifetime (cid:104)τ(cid:105) as a function of tensile force f applied to dissociate the fibrin-fibrinogen complex. We found that the bond lifetimes show the biphasic catch-slip behavior, namely the average bond lifetime (cid:104)τ(cid:105) first increases and then decreases with f (Fig. 2). Physiological importance of catch-slip behavior can be exemplified with shear-enhanced platelet adhesion on vWF-coated surfaces where GP1bα-vWF catch bonds promote primary hemostasis at the sites of vessel wall injury.15, 22, 46 Theoretical models have been proposed to explain the dynamic transition from catch-bonds to slip-bonds.24, 31, 32, 47–50 Yet, the structural origins of this counterintuitive behavior are not yet understood. We showed that force signals measured in the pulling experiments represent the strength of individual knobs 'a' coupled to holes 'a'. First, we adjusted the surface densities of the reacting molecules, so that the incidence of interactions lasting >0.5s (i.e. representing A:a bonding) did not exceed 10% of the total number of surface-to-surface contacts (binding attempts). This implies that forced dissociation events corresponding to multiple A:a knob-hole bonds were highly unlikely due to sufficiently low surface density of reacting molecules (300 nm2/molecule on a bead) and a small contact area (450 nm2). Second, when the fibrin-fibrinogen interactions were measured with the force-ramp, we only observed a sharp single peak in the histograms of bond rupture forces with very rare jagged signals.7 The unimodal nature of distributions of bond rupture forces implies that unbinding signals represents dissociation of single knob-hole bonds. Multiple interactions might have occurred in some of the measurements, but infrequently, resulting in strong bead attachments lasting >60 s. These were discarded and not used in subsequent data analysis. Experiments with proteolytic fibrin(ogen) fragments bearing both knobs 'a' and 'b' (fragment desAB-E) or only knob 'a' (fragment desA-E) showed catch-slip transition in both fragments, but in the absence of knob 'b' the bond lifetimes were shorter and the catch-slip transition peak was less pronounced (Fig. 2b). Hence, the catch-slip behavior was enhanced when both knobs 'a' and 'b' were exposed, which suggests positive cooperativity between knobs 'a' and 'b'. We employed MD simulations of atomic structural models of the A:a knob-hole complex (Fig. 1g) to resolve the structural basis and to illuminate the molecular mechanism(s) of dynamic transition from catch-bonds to slip-bonds in fibrin (Methods). We used low damping coefficient γ = 3.0 ps1 for more efficient sampling of the conformational space51 and to minimize the effect of fast ν f = 103 − 104 µm/s pulling speeds we had used due to very long computational time (it took ∼24 months to complete the computational tasks on 4 GPUs GeForce GTX780). Using dynamic force-ramp simulations, we were able to establish kinetic partitioning of the A:a knob-hole bond disassembly into the rapid/slow dissociation paths (Pathway 1/2) corresponding to lower/higher rupture forces (Figs.3 and S7). To show that observed variation of rupture forces was not due to statistical fluctuations, we probed the dynamics of binding contacts between residues in knob 'a' and hole 'a' defining the binding affinity Q. For example, in Pathway 1, Q decreased to zero (with some fluctuations) starting from 20 contacts (low-affinity bound state); in Pathway 2, Q increased initially to 30-35 contacts (high-affinity bound state) and then decreased to zero at a higher rupture force (Fig. S8a; see also Fig. S7). A 1.5-fold enhancement in binding affinity points to tension-induced bond stabilization through recruitment of additional binding contacts that become available for interaction with knob 'a' at higher tensile forces. We analyzed entire maps of residue-residue binding contacts (Fig. S8c,d), which revealed the important role played by the movable flap. In the low-affinity bound state, the movable flap – one of the three binding determinants in hole 'a' – is far away from knob 'a' (snapshots 2a,3a in Fig. 3); therefore, the low-affinity bound state facilitates rapid detachment of knob 'a' which slips easily from hole 'a'. In the high-affinity bound state, the movable flap translocates toward and catches knob 'a', forming additional binding contacts between the flap and knob 'a' (snapshots 2b,3b in Fig. 3). Importantly, this transition also triggers the loop I and interior region straightening, which results in formation of additional binding contacts between knob 'a' and the interior region (Fig. S8b). Therefore, in the high-affinity bound state hole 'a', comprised by the movable flap, loop I, and interior region (Fig. 1g), shrinks and the A:a binding interface narrows, which results in labored knob 'a' detachment. The increase in binding affinity was found to be positively correlated with the displacement of movable flap (Fig. S8b). Hence, a tension-induced increase in A:a knob-hole bond strength is entirely due to spatial rearrangement of the binding interface in hole 'a'. To better understand the biphasic catch-slip dynamic behavior of A:a knob-hole bonds, we developed new theory inspired by fluctuating bottleneck model.40–44 The model treats the binding interface size X as a Gaussian random variable (fluctuating 6/29 bottleneck) and accounts for the tension-dependent decrease in X (see equations (S2) and (S3)). Simulations showed that X is roughly equal the dimension (radius) of binding interface. For this reason we used a quadratic sink K(X) ∼ kX 2 with the Bell-type dependence of k on f (equations (2), (S5), (S6)). Because we were unable to determine from simulations whether the interface interconverted between several or many conformational states, we considered a discrete (two-state) version and continuous version of the model (Fig. 4). We obtained the distribution of A:a knob-hole bond lifetimes P(t) (equation (6)), which was then used to calculate (cid:104)τ(cid:105) as a function of f . Excellent agreement between theoretical curves of (cid:104)τ(cid:105) and experimental data points was achieved (Fig. 5), which enabled us to estimate parameters of the model accumulated in Table 1. Both continuous and discrete versions of the model give similar values of knob 'a' escape rate k0 = 0.11-0.12 nm2s−1 and transition distance σy = 0.25-0.27 nm, but the continuous version shows better agreement with estimates of other model parameters from simulations: the interface stiffness κ = 15.7 pN/nm (theory) vs. 10-30 pN/nm (simulations), initial (force-free) interface size x0 = 2.7 nm (theory) vs. 2.0-2.7 nm (simulations), and minimal interface size x1 = 0.74 nm (theory) vs. 0.5-0.9 nm (simulations). Hence, our theory provides evidence for a manifold of high-affinity bound states with continuous dependence of binding affinity on f . We also calculated the profiles of (cid:104)τ(cid:105) vs. f but for a linear sink K(X) ∼ X in equation (2), yet, the values of model parameters showed worse agreement with simulations (see Table S2). We used values of model parameters for the continuous fluctuating bottleneck model with quadratic sink (Table 1), which demonstrated the best agreement with the simulations, to calculate the distributions of bond lifetimes P(t). Theoretical curves of P(t) are compared with experimental histograms in Fig. S9, which shows that experiment and theory agree very well. Using the values of model parameters (Tables 1 and S2) we compared the prediction of all four models (continuous vs. discrete version with quadratic vs. linear sink) for the critical force fc of transition from the catch-to-slip regime of dissociation of A:a knob-hole bond. We found that for the continuous version fc = 30.7 pN (quadratic sink) and 30 pN (linear sink), whereas for the discrete version fc = 44.6 pN (quadratic sink) and 40.1 pN (linear sink). Hence, the continuous model provides a better prediction for the 30-35 pN critical force, which corresponds to the 4-5 s maximum bond lifetime (Fig. 5). To conclude, we demonstrated experimentally, resolved computationally, and modeled theoretically the catch-slip dynamic transition in A:a knob-hole bonds in fibrin. The movable flap plays an important role of a tension-dependent molecular switch, which triggers the crossover from the catch regime to the slip regime of bond dissociation. In the catch regime, the binding affinity of hole 'a' progressively grows with force owing to mechanical interface remodeling and formation of additional binding contacts, which results in bond strengthening. This trend continues until the size of binding pocket becomes comparable with the molecular dimension of knob 'a' at a critical force f = fc, at which point the slip regime sets in. In the slip regime, the binding affinity of hole 'a' gradually decreases with the increasing force f > fc due to disruption of binding contacts, which weakens the bond. The fluctuating bottleneck theory can be used to model biomolecular complexes displaying rich complex dynamics of interface remodeling. (Patho)physiologically the catch-bond behavior is equivalent to shear-enhanced strengthening of A:a knob-hole bonds that might favor fibrin polymerization in blood flow and might prevent breakup and damage of stressed clots in vasculature. This effect is especially important at the early stages of fibrin formation in blood flow when the incipient clot is small and the hydrodynamic shear stress is low, so that the tensile forces are in the range that strengthens the knob-hole interactions. This is an entirely new aspect of fibrin nanomechanics that needs further investigation. The novel and unexplored mechano-chemical aspects of fibrin polymerization addressed in this study for the first time will advance our understanding of blood clotting and will provide a firm foundation for the development of new approaches to control and modulate this process. Online Methods Optical trap-based model system: Our model system for probing the bimolecular interactions is based on an optical trap that uses a focused laser beam to generate the pico-Newton mechanical force to hold and move microscopic particles, such as micron-size polystyrene beads.7, 52–56 A custom-built optical trap previously described in detail54 was used to measure the mechanical strength of individual bi-molecular protein-protein complexes under a constant tensile force. The core of the laser tweezers system is a AxioObserver Z1 inverted microscope and a 100x 1.3NA Fluor lens combined with a FCBar Nd:YAG laser (=1,064 nm) with 4W power in continuous TEM-00 mode. A computer-operated two-dimensional acousto-optical deflector (AOD) was used to control the trap position. The force exerted by the trap on the bead displaced by an amount ∆x was measured with a quadrant detector and the trap position was corrected with an electronic feedback loop to keep the force constant. This system enabled control of the duration of compressive contact between interacting surfaces T , the magnitude of compressive force fc and the magnitude of the tensile force f = kopt∆x (kopt is the optical trap stiffness). The measured quantity was the time needed to separate the interacting surface-attached proteins (bond lifetime) τ. All experiments were conducted with the average trap stiffness of kopt = 0.10±0.02 pN/nm. Force calibration and trap stiffness were routinely confirmed by the Stokes' force method. LabVIEW(cid:114) software was used to control and record laser beam deflection, to move the piezoelectric stage, and to analyze data off-line. Surfaces and proteins: A single fibrinogen- or fragment D-coated bead was trapped and repeatedly brought into contact 7/29 with a fibrin- or fragment E-coated pedestal. When fibrinogen or fragment D (with holes 'a' and 'b') on the bead attached to monomeric fibrin or activated fragment E (with both knobs 'a' and 'b' or knobs 'a' only) on the pedestal, the trap exerted a constant force f = kopt∆x to trigger the complex dissociation.7 Purified human fibrinogen or its fragment E (both from HYPHEN BioMed, France) were bound covalently to spherical silica pedestals 5 m in diameter anchored to the bottom of a chamber. Pedestals coated with a thin layer of polyacrylamide were activated with 10% glutaraldehyde, after which the proteins were immobilized overnight at 4 ◦C from 1 mg/ml solution in 20 mM HEPES pH 7.4 containing 150 mM NaCl and 3 mM CaCl2. After washing off the non-covalently adsorbed protein, 2 mg/ml bovine serum albumin (BSA) in 0.055 M borate buffer pH 8.5 was added as a blocker. To form fibrin-coated pedestals, the immobilized fibrinogen was treated with human α-thrombin (Enzyme Research Laboratories, South Bend, IN) (1 U/ml, 37 ◦C, 1 hr) followed by washing of the chambers with 20 volumes of 100 mM HEPES pH 7.4 containing 150 mM NaCl, 3 mM CaCl2, 2 mg/ml BSA, and 0.1% (v/v) Triton X-100 before the measurements. To form pedestals coated with fibrin fragment E bearing knobs 'a' (fragment desA-E) or knobs 'a' and 'b' (fragment desAB-E), the immobilized fibrinogen fragment E with uncleaved fibrinopeptides was treated with batroxobin (Batroxobin moojeni, CenterChem, Stamford, CT) (1 BU/ml, 37 ◦C, 1 hr) or human α-thrombin (1 U/ml, 37 ◦C, 1 hr), respectively, followed by washing of the chambers. 1 unit of batroxobin activity (BU) was calibrated to be equal to 1 unit of thrombin activity (U) in terms of the rate of fibrinopeptide A release. Fibrinogen or fragment D (HYPHEN BioMed, France) were bound covalently to carboxylate-modified 1.75-m latex beads (Bangs Laboratories, Carmel, IN) activated by N-(3-dimethylaminopropyl)-N'-ethylcarbodiimide hydrochloride. The immobilization step lasted 15 min at 4 ◦C in 0.055 M borate buffer pH 8.5 containing 150 mM NaCl and 3 mM CaCl2. BSA was used as a blocker. When immobilized from 20 g/ml solution containing 100% of fibrinogen labeled with I125, the surface density of I125-fibrinogen was determined to be about (11±2)×10−9 g/m2, which approaches the point of surface binding saturation. Both fibrinogen and fragment D were used with the same solution concentration (20 g/ml). Measurements of protein-protein interactions: Experiments were performed at room temperature in 100 mM HEPES pH 7.4 containing 150 mM NaCl, 3 mM CaCl2 with 2 mg/ml BSA and 0.1% (v/v) Triton X-100 added to reduce non-specific interactions. 1 l of the fibrinogen- or fragment D-coated bead suspension (107 beads/ml) was added to 50 l of the working buffer and flowed into a chamber containing pedestals with immobilized fibrin or fragment E on their surface. After the chamber was placed on the microscope stage, a single bead was trapped and the stage moved manually to bring a pedestal within 1-2 microns of the trapped bead. After starting the bead oscillation, the separation of the pedestal and the bead was then reduced until they touched each other repeatedly with a compressive force fc = 20-30 pN and contact duration T = 0.5 s. The constant pulling force was varied from f = 5 to 60 pN. Trap displacement signals were recorded at 2000 scans per second and a bond lifetime was measured for each pedestal-bead touching event. Several tens of pedestal-bead pairs were analyzed for each set of conditions. The binding-unbinding events from individual files were summarized; the total number of bond lifetime values recorded for each set of experimental conditions varied from ∼3,000 to ∼4,500. The bond lifetimes <0.5 s represented non-specific interactions and were not susceptible for specific inhibition. These short bond lifetimes were not included into data analysis and modeling. All-atomic structural model of A:a knob-hole complex: The N-terminal motif Gly-Pro-Arg (GPR) in α chain is the main functional sequence of the knob 'a', and it is complementary to the hole 'a' in γ-nodule of adjacent fibrin molecule. The N-terminal β chain motif Gly-His-Arg-Pro (GHRP) is a major part of the knob 'b' that binds to the hole 'b' located in β -nodule of adjacent fibrin.57 The available atomic structures (PDB entry 1FZA, 1FZB, 1FZC)58 contain only the D:D interface, i.e. the β - and γ-nodules of two cross-linked fibrin monomers with bound knob-mimetic peptides GPRP and GHRP, which occupy the corresponding binding pockets in γ- and β -nodules, respectively (Fig. 1). We used the PDB data to reconstruct a physiologically relevant model of the A:a complex comprising the γ-nodule bound with the N-terminal end of the α-chain of adjacent fibrin monomer. We utilized the complete atomic structure of a short fibrin oligomer.39 The model contains the N-terminal ends of α-chains (with knobs 'a') and β -chains (with knobs 'b') bound to the complementary sites in the γ-nodule (with holes 'a') and β -nodules (with holes 'b') as shown in Fig. 1e. All in silico models were constructed using CHARMM.59 This structure was equilibrated for ∼100 ns using the MD simulations in implicit solvent. The equilibrated structure of the fibrin trimer was truncated to separate the central nodule of one trimer strand with two exposed knobs 'a' bound to the complementary holes 'a' of the γ-nodules in the other strand (Fig. 1f). To mimic experimental conditions of tensile force application, we constrained the Cα-atoms of γLys159 of two γ-nodules and pulled at the entire central domain. This resulted in a partial extension of α chains. We separated the γ-nodule (residues γLys140–γVal411) with attached N-terminal part of α chain (residues αGly17–αCys36). The obtained structural model of A:a knob-hole complex is shown in Fig. 1g. The γ-nodule (with hole 'a') consists of three globular domains called A-domain (γVal143–γTrp191), B-domain (γThr192–γAla286 and γLys380–γLeu392), and P-domain (γGly287–γMet379). The latter is a knob 'a'-binding domain,60 while the B-domain has a 3-stranded β -sheet stack (regulatory domain; residues γIle242–γGly283) located near the P-domain (binding domain). The γ-nodule has three binding determinants: loop I (region I; γTrp315–Trp330), interior region (region II; γTrp335–Asn365), and moveable flap (region III; γPhe295–Thr305) displayed in Fig. 1g,4, 61 which define the strength of A:a knob-hole bond.62 8/29 Dynamic force measurement of the A:a knob-hole interactions: We employed the all-atom MD simulations using Solvent Accessible Surface Area (SASA) model of implicit solvation with CHARMM19 unified hydrogen force-field63–65 implemented on a GPU. We used a lower damping coefficient γ = 3.0 ps−1 (vs. γ = 50 ps−1 for ambient water at 300K) for more efficient sampling of the conformational space.51 In the force-ramp measurements, we used time-dependent force f (t) = kopt (ν f t − ∆x), where ν f is the virtual optical trap velocity, kopt is the spring constant, and ∆x is the displacement of a pulled residue. We constrained the Cα-atoms of γLys159 and pulled the Cα-atom of αCys36 in the direction perpendicular to the A:a binding interface (Fig. 1g). We generated a total of 20 simulation runs with ν f = 103, 104 µm/s and kopt = 100 pN/nm. In the force-clamp measurements, we used constant tensile force f = f · n with force magnitude f = 30, 35, 40, 50, 80 pN (Fig. 1g). We performed 3 simulation runs (a total of 5 µs) for each force value. Acknowledgements The authors thank Andrey Mekler for technical assistance. This work was supported by NSF (grant DMR1505662 to JWW and VB), American Heart Association (grant-in-aid 13GRNT16960013 to VB and JWW), Russian Foundation for Basic Research (grant 15-37-21027, 15-01-06721 to AZ and grant 14-04-32066 to OK) and the Program for Competitive Growth at Kazan Federal University. Author contributions statement R.I.L. – designed and performed experiment, wrote the manuscript; O.K. – performed simulations and analyzed the results, developed the model, wrote the manuscript; F.M. – performed model fitting; A.Z. – analyzed results of simulations; K.A.M. – analyzed the results of simulations, wrote the manuscript; J.W.W. – designed experiment, wrote manuscript; V.B. – designed simulations and model, wrote the manuscript. Competing financial interests The authors declare no competing financial interests. References 1. Litvinov, R. I. & Weisel, J. W. What is the biological and clinical relevance of fibrin? Semin. Thromb. Hemost. 42, 333–343 (2016). 2. Weisel, J. W. & Litvinov, R. I. Fibrin formation, structure and properties. Subcell Biochem. 82, 405–456 (2017). 3. Standeven, K. F., Arieons, R. A. S. & Grant, P. J. The molecular physiology and pathology of fibrin structure/function. Blood Rev. 19, 275–288 (2005). 4. Laudano, A. P. & Doolittle, R. F. Synthetic peptide derivatives that bind to fibrinogen and prevent the polymerization of fibrin monomers. Proc. Natl. Acad. Sci. USA 75, 3085–3089 (1978). 5. Laudano, A. P. & Doolittle, R. F. Studies on synthetic peptides that bind to fibrinogen and prevent fibrin polymerization. Structural requirements, number of binding sites, and species differences. Biochem. 19, 1013–1019 (1980). 6. Spraggon, G., Everse, S. J. & Doolittle, R. F. Crystal structures of fragment D from human fibrinogen and its crosslinked counterpart from fibrin. Nat. 389, 455–462 (1997). 7. Litvinov, R. I., Gorkun, O. V., Owen, S. F., Shuman, H. & Weisel, J. W. Polymerization of fibrin: specificity, strength, and stability of knob-hole interactions studied at the single-molecule level. Blood 106, 2944–2951 (2005). 8. Litvinov, R. I. & Weisel, J. W. Shear strengthens fibrin: the knob–hole interactions display 'catch-slip' kinetics. J. Thromb. Haemost. 11, 1933–1935 (2013). 9. Hertig, S. & Vogel, V. Catch bonds. Curr. Biol. 22, R823 (2012). 10. Marshall, B. T. et al. Direct observation of catch bonds involving cell-adhesion molecules. Nat. 423, 190–193 (2003). 11. Rakshita, S., Zhang, Y., Manibog, K., Shafraza, O. & Sivasankar, S. Ideal, catch, and slip bonds in cadherin adhesion. Proc. Natl. Acad. Sci. USA 106, 18815–18820 (2012). 12. Kong, F., Garcia, A. J., Mould, A. P., Humphries, M. J. & Zhu, C. Demonstration of catch bonds between an integrin and its ligand. J. Cell Biol. 185, 1275–1284 (2009). 13. Thomas, W. E., Trintchina, E., Forero, M., Vogel, V. & Sokurenko, E. V. Bacterial adhesion to target cells enhanced by shear force. Cell 109, 913–923 (2002). 9/29 14. Sauer, M. M. et al. Catch-bond mechanism of the bacterial adhesin FimH. Nat. Comm. 7, 10738 (2016). 15. Yago, T. et al. Platelet glycoprotein Ibα forms catch bonds with human WT vWF but not with type 2B von Willebrand disease vWF. J. Clin. Invest. 118, 3195–3207 (2006). 16. Feghhi, S. et al. Glycoprotein Ib-IX-V complex transmits cytoskeletal forces that enhance platelet adhesion. Biophys. J. 111, 601–608 (2016). 17. Guo, B. & Guilford, W. H. Mechanics of actomyosin bonds in different nucleotide states are tuned to muscle contraction. Proc. Natl. Acad. Sci. USA 103, 9844–9849 (2006). 18. Akiyoshi, B. et al. Tension directly stabilizes reconstituted kinetochore-microtubule attachments. Nat. 468, 576–579 (2010). 19. Rai, A. K., Rai, A., Ramaiya, A. J., Jha, R. & Mallik, R. Molecular adaptations allow dynein to generate large collective forces inside cells. Cell 152, 172–182 (2013). 20. Nair, A., Chandel, S., Mitra, M. K., Muhuri, S. & Chaudhuri, A. Effect of catch bonding on transport of cellular cargo by dynein motors. Phys. Rev. E 94, 032403 (2016). 21. Chen, W., Lou, J. & Zhu, C. Forcing switch from short- to intermediate- and long-lived states of the αA domain generates LFA-1/ICAM-1 catch bonds. J. Biol. Chem. 285, 35967–35978 (2010). 22. McEver, R. P. & Zhu, C. Rolling cell adhesion. Annu. Rev. Cell Dev. Biol. 26, 363–396 (2010). 23. Waldron, T. T. & Springer, T. A. Transmission of allostery through the lectin domain in selectin-mediated cell adhesion. Proc. Natl. Acad. Sci. USA 106, 85–90 (2008). 24. Sarangapani, K. K. et al. Regulation of catch bonds by rate of force application. J. Biol. Chem. 286, 32749–32761 (2011). 25. Helms, G., Dasanna, A. K., Schwarz, U. S. & Lanzer, M. Modeling cytoadhesion of plasmodium falciparum-infected erythrocytes and leukocytes-common principles and distinctive features. FEBS Lett. 590, 1955–1971 (2016). 26. Manibog, K., Li, H., Rakshit, S. & Sivasankar, S. Resolving the molecular mechanism of cadherin catch bond formation. Nat. Comm. 5, 3941 (2014). 27. Gunnerson, K. N., Pereverzev, Y. V. & Prezhdo, O. V. Atomistic simulation combined with analytic theory to study the response of the P-selectin/PSGL-1 complex to an external force. J. Phys. Chem. B 113, 2090–2100 (2009). 28. Chakrabarti, S., Hinczewski, M. & Thirumalai, D. Phenomenological and microscopic theories for catch bonds. J. Struct. Biol. 197, 50–56 (2017). 29. Vernerey, F. J. & Akalp, U. Role of catch bonds in actomyosin mechanics and cell mechanosensitivity. Phys. Rev. E 94, 012403 (2016). 30. Bullerjahn, J. T. & Kroy, K. Analytical catch-slip bond model for arbitrary forces and loading rates. Phys. Rev. E 93, 012404 (2016). 31. Barsegov, V. & Thirumalai, D. Dynamics of unbinding of cell adhesion molecules: Transition from catch to slip bonds. Proc. Natl. Acad. Sci. U. S. A. 102, 1835–1839 (2005). 32. Barsegov, V. & Thirumalai, D. Dynamic competition between catch and slip bonds in selectins bound to ligands. J. Phys. Chem. B 110, 26403–26412 (2006). 33. Pereverzev, Y. V., Prezhdo, O. V., Forero, M., Sokurenko, E. V. & Thomas, W. E. The two-pathway model for the catch-slip transition in biological adhesion. Biophys. J. 89, 91446–91454 (2005). 34. Lou, J. & Zhu, C. A structure-based sliding-rebinding mechanism for catch bonds. Biophys. J. 92, 1471–1485 (2007). 35. Chakrabarti, S., Hinczewski, M. & Thirumalai, D. Plasticity of hydrogen bond networks regulates mechanochemistry of cell adhesion complexes. Proc. Natl. Acad. Sci. USA 111, 9048–9053 (2014). 36. Liu, F. & Ou-Yang, Z. Force modulating dynamic disorder: A physical model of catch-slip bond transitions in receptor- ligand forced dissociation experiments. Phys. Rev. E 74, 051904 (2006). 37. Wei, Y. Entropic-elasticity-controlled dissociation and energetic-elasticity-controlled rupture induce catch-to-slip bonds in cell-adhesion molecules. Phys. Rev. E 77, 031910 (2008). 38. Litvinov, R. I. et al. Polymerization of fibrin: direct observation and quantification of individual B:b knob-hole interactions. Blood 109, 130–138 (2007). 39. Zhmurov, A. et al. Structural basis of interfacial flexibility in fibrin oligomers. Struct. 24, 1907–1917 (2016). 10/29 40. Zwanzig, R. Dynamical disorder: Passage through a fluctuating bottleneck. J. Chem. Phys. 97, 3587 (1992). 41. Barsegov, V., Chernyak, V. & Mukamel, S. Multitime correlation functions for single molecule kinetics with fluctuating bottlenecks. J. Chem. Phys. 116, 4240–4251 (2002). 42. Hyeon, C., Hinczewski, M. & Thirumalai, D. Evidence of disorder in biological molecules from single molecule pulling experiments. Phys. Rev. Lett. 112, 138101 (2014). 43. Bicout, D. J. & Szabo, A. Escape through a bottleneck undergoing non-markovian fluctuations. J. Chem. Phys. 108, 5491 (1998). 44. Wang, J. & Wolynes, P. Passage through fluctuating geometrical bottlenecks. The general gaussian fluctuating case. Chem. Phys. Lett. 212, 427–433 (1993). 45. Bell, G. L. Models for the specific adhesion of cells to cells. Sci. 200, 618–627 (1978). 46. Savage, B., Saldivar, E. & Ruggeri, Z. M. Initiation of platelet adhesion by arrest onto fibrinogen or translocation on von Willebrand factor. Cell 84, 289–297 (1996). 47. Prezhdo, O. V. & Pereverzev, Y. V. Theoretical aspects of the biological catch bond. Acc. Chem. Res. 142, 693–703 (2009). 48. Thomas, W. E. Mechanochemistry of receptor-ligand bonds. Curr. Opin. Struct. Biol. 19, 50–55 (2009). 49. Zhu, C., Lou, J. & McEver, R. P. Catch bonds: physical models, structural bases, biological function and rheological relevance. Biorheol. 42, 443–462 (2005). 50. Chen, H. & Alexander-Katz, A. Polymer-based catch-bonds. Biophys. J. 100, 174–182 (2011). 51. Falkovich, S., Neelov, I. & Darinskii, A. Mechanism of shear deformation of a coiled myosin coil: Computer simulation. Polym. Sci. Ser. A+ 62, 662 (2010). 52. Litvinov, R. I., Shuman, H., Bennett, J. S. & Weisel, J. W. Binding strength and activation state of single fibrinogen-integrin pairs on living cells. Proc. Natl. Acad. Sci. USA 99, 7426–7431 (2002). 53. Litvinov, R. I., Bennett, J. S., Weisel, J. W. & Shuman, H. Multi-step fibrinogen binding to the integrin αIIbβ 3 detected using force spectroscopy. Biophys. J. 89, 2824–2834 (2005). 54. Litvinov, R. I. et al. Dissociation of bimolecular αIIbβ 3-fibrinogen complex under a constant tensile force. Biophys. J. 100, 165–173 (2011). 55. Litvinov, R. I. et al. Resolving two-dimensional kinetics of receptor-ligand interactions using binding-unbinding correlation spectroscopy. J. Biol. Chem. 287, 35272 (2012). 56. Litvinov, R. I., Farrell, D. H., Weisel, J. W. & Bennett, J. S. The platelet integrin αIIbβ 3 differentially interacts with fibrin versus fibrinogen. J. Biol. Chem. 291, 7858–7867 (2016). 57. Medved, L. & Weisel, J. W. Recommendations for nomenclature on fibrinogen and fibrin. J. Thromb. Haemost. 7, 355–359 (2009). 58. Everse, S. J., Spraggon, G., Veerapandian, L., Riley, M. & Doolittle, R. F. Crystal structure of fragment double-d from human fibrin with two different bound ligands. Biochem. 37, 8637–8642 (1998). 59. Brooks, B. R. et al. CHARMM: The biomolecular simulation program. J. Comput. Chem. 30, 1545–1614 (2009). 60. Yee, V. C. et al. Crystal structure of a 30 kDa C-terminal fragment from the γ-chain of human fibrinogen. Struct. 15, 125–138 (1997). 61. Kostelansky, M. S., Betts, L., Gorkun, O. V. & Lord, S. T. 2.8 A crystal structures of recombinant fibrinogen fragment D with and without two peptide ligands: GHRP binding to the "b" site disrupts its nearby calcium-binding site. Biochem. 41, 12124–12132 (2002). 62. Kononova, O. et al. Molecular mechanisms, thermodynamics, and dissociation kinetics of knob-hole interactions in fibrin. J. Biol. Chem. 288, 22681–22692 (2013). 63. Ferrara, P., Apostolakis, J. & Caflisch, A. Evaluation of a fast implicit solvent model for molecular dynamics simulations. Proteins 46, 24–33 (2002). 64. Zhmurov, A. et al. Mechanical transition from α-helical coiled coils to β -sheets in fibrin(ogen). J. Am. Chem. Soc. 134, 20396–20402 (2012). 65. Kononova, O. et al. Mechanistic basis for the binding of RGD-and AGDV-peptides to the platelet integrin αIIbβ 3. Biochem. 56, 1932–1942 (2017). 11/29 Table 1. Model parameters for the A:a knob-hole bottleneck: interface stiffness κ, initial (maximal) and minimal interface width x0 and x1, escape rate constant k0, transition distance for dissociation σy, and transition distance for conformational transition state σx and energy difference between two states ε0 (for discrete two-state version). The parameter values were obtained by fitting theoretical curves of (cid:104)τ( f )(cid:105) (equation (7) to the experimental bond lifetimes (Fig. 5). κ, pN/nm x0, nm x1, nm k0, nm2s−1 Model Continuous Discrete 15.7 100.0 2.7 4.3 0.74 1.6 0.11 0.12 σy, nm σx, nm ε0, kcal/mol 0.25 0.27 - 0.8 - 0.55 12/29 Figure 1. Major steps and main structural determinants in fibrin polymerization. Cleavage by thrombin of the N-terminal parts of the α-chain - fibrinopeptides A and B (FpA and FpB) in fibrinogen (panel a) converts it into fibrin with exposed knobs 'A' and 'B' (panel b), which bind to the complemental hole 'a' in the γ-nodules (blue) and 'b' in the β -nodule (green), facilitating fibrin oligomerization (panel c). Panel d shows the fibrin fragments utilized in dynamic force spectroscopy experiments in vitro and in silico. Panel e: all-atom model of the short fibrin oligomer39 with α-, β - and γ-chains colored in orange, green and red, respectively. The β - and γ-nodules are highlighted in light blue and light grey, respectively. Panel f: truncated portion of fibrin oligomer (with γ-nodules and central nodule) used in simulations. The pulling force is applied through a virtual plane, crossing the central nodule in the middle (Cα-atoms of γLys159 in the γ-nodule were constrained). Panel g: all-atom model of A:a knob-hole complex containing γ-nodule with hole 'a' (grey blue) and the N-terminal part of α-chain with knob 'A' (orange). Hole 'a' contains domains A, B, and P; three-stranded β -sheets stack in B-domain is shown in purple. Also shown are the major binding determinants: loop I (blue), interior region (green), and moveable flap (red) all interacting with GPR-motif-containing knob 'A'.62 In pulling simulations, tensile force was applied to the Cα-atom of αCys36 (Cα-atom of γLys159 was constrained). 13/29 Figure 2. Kinetics of forced dissociation of the A:a knob-hole bonds. Panel a: Plots of the average bond lifetime (cid:104)τ(cid:105) as a function of constant tensile force f for fibrin-fibrinogen (Fn:Fg) interactions in comparison with the data for control protein pairs both lacking knobs 'a' (Fg:Fg) or one of the interacting proteins (BSA), or lacking both knobs 'a' and holes 'a' (Fg:BSA and Fn:BSA). Panel b: Plots of (cid:104)τ(cid:105) > 0.5 s as a function of f for fibrinogen fragment D (with holes 'a' and 'b') interacting with fragments desAB-E (bearing knobs 'a' and 'b') or desA-E (with knobs 'a' only), or E (with no knobs). 14/29 Figure 3. Dynamic interface remodeling and A:a knob-hole bond dissociation from the low-affinity (Pathway 1) and high-affinity (Pathway 2) bound states. Shown are profiles of unbinding force F and binding affinity Q as functions of time t from force-ramp simulations (ν f = 104 µm/s and kopt = 100 pN/nm). Pathway 1/2 is characterized by lower/higher values of F and monotonic/non-monotonic evolution of Q. Structural snapshots numbered 1-4 correspond to the equally numbered regions in the curves of F vs. t (red) and Q vs. t (green) displaying the progress of dissociation from the initial bound state (structure 1) to the fully dissociated state (structure 4) along each pathways. Low-affinity bound state: movable flap interacts with the β -sheet stack of B-domain (red arrow and dashed circle in structures 2a and 3a). High-affinity bound state: movable flap moves from the periphery toward knob 'a' offering additional binding contacts; loop I and interior region straighten up, which results in the interface closing (red arrow and dashed circle in structures 2b and 3b). 15/29 Figure 4. Fluctuating bottleneck model of forced dissociation of A:a knob-hole bond. Panel a: structural snapshots I-III, showing gradual interface closing. Panel b: evolution of the interface width X for two representative simulations resulted in bond dissociation from the low-affinity bound state (blue curve) and high-affinity bound state (red curve). Panel c: The probability distributions P(X) of interface width X for two simulations (from panel b) showing continuous evolution of X. Panel d: P(X) for simulations showing discrete (two-state) evolution of X from the low- to high-affinity bound states centered around x0 to x1, respectively (the inset shows P(X) averaged over all simulation runs showing discrete evolution of X). Panel e: binding packet (hole 'a') modeled as a fluctuating bottleneck of width X decreasing from the initial value x0 to final value x1. 16/29 Figure 5. Catch-slip dual force response of A:a knob-hole bond. Shown is the non-monotonic behavior of the average bond lifetime (cid:104)τ(cid:105) with the tensile force f . The diamonds are experimental data points from Fig. 2b. The curves correspond to the fits of continuous version (equations (S2), (6), (7)) and discrete two-state version (equations (S3), (6), (7)) of fluctuating bottleneck model. The initial growth of (cid:104)τ(cid:105) at low forces f < fc ≈ 30-35 pN is followed by the decay to zero at higher forces f > fc, which marks the transition from the catch-regime to the slip-regime of the A:a knob-hole bond dissociation. 17/29 Supplementary Information Fluctuating bottleneck model: In the continuous version of the model, the binding pocket size X is a continuous random variable (Fig. 4e). Evolution of X is governed by the Langevin equation: ζ X(t) = −∂U/∂ X + R(t), where ζ is the friction coefficient, and R(t) is Gaussian random force with the average (cid:104)R(t)(cid:105) = 0 and standard deviation (cid:104)R(t)R(t(cid:48))(cid:105) = 2ζ kBTδ (t −t(cid:48)) (kBT – temperature). The distributions of X are Gaussian-like (Fig. 4c). Hence, for f = 0, X evolves on a harmonic potential U = κ(X − x0)2 with stiffness κ and equilibrium width x0. When 0 < f < fc ( fc is the critical force (see below)), X decreases (Fig. 4b) approaching new equilibrium x0 − f /κ, and the deterministic force is ∂U/∂ X = κ(X − x0) + f = κ[X − (x0 − f /κ)]. By substituting this expression into Langevin equation and solving this equation, we obtain the average size (cid:104)X(cid:105) and fluctuations (cid:104)∆X(cid:105)2: (cid:104)X(t)(cid:105) = X0e−t/η + (x0 − f /κ)(1− e−t)/η ) and (cid:104)∆X(cid:105)2 = kBT ζ (1− e−t)/η ) (S1) where X0 is the initial value and η = ζ /κ is the characteristic time for conformational fluctuations of the binding pocket (bottleneck). At a critical force f ≈ fc, (cid:104)X(t)(cid:105) should reaches the minimum x1 = x0 − fc/κ, which can be expressed as (cid:104)X(t)(cid:105) = [X0e−t/η + (x0 − f /κ)(1− e−t)/η )]Θ( fc − f ) + x1Θ( f − fc) (S2) where Θ(x) is the Heaviside step function. In the discrete version of the model, X is a discrete random variable interconverting between the open state X = x0 and closed state X = x1 < x0 (Fig. 4d) with the populations s0 = [1 + exp(−ε0/kBT )]−1 and s1 = 1− s0 (ε0 is the energy difference). Pulling force application favors the closed conformation by changing the state populations, in which case (cid:104)X(cid:105) and s0,1 become force-dependent, (cid:104)X( f )(cid:105) = x0s0( f ) + x1s1( f ) and s0 = (1 + exp[−(ε0 − σx)/kBT ])−1 (S3) where σx is the transition distance. The critical force is defined as force f = fc at which s1( f ) (cid:29) s0( f ). Pulling force affects both the dynamics of binding interface remodeling and kinetics of knob 'A' escape, and so the A:a bond lifetime τ and interface width size X are coupled (Fig. 4e). Larger/smaller X facilitates faster/slower unbinding with shorter/longer bond lifetime τ. The A:a knob-hole bound state population P(X,t) is described by the kinetic equation: dP(X,t) dt = −K(X)P(X,t) + LP(X,t) In equation (S4) above, the first term describes the kinetics of knob 'A' escape from the bottleneck of size X with rate K(X) = kXα (S4) (S5) depending on shape parameter α (bottleneck geometry), and escape rate constant k (when K = 0; knob 'A' is tightly locked in hole 'a'). The rate constant is expected to increase with force, and here we use the Bell model:31, 32, 45 k( f ) = k0 exp[−σy f /kBT ] (S6) where k0 is the attempt frequency and σy is the transition distance for dissociation. Simulations show that i) X is determined by the cross-sectional area, and so we set α = 2 in equation (S5); and that ii) characteristic timescale η (milliseconds) is much shorter than bond lifetimes (seconds; Fig. 2), and so we set X = (cid:104)X(cid:105) in equation (S5). In equation (S4), operator L describing the dynamics of X: By substituting equations (S6) and (S7) into equation (S4), we arrive at the Smoluchowsky equation for P(X,t): which can be solved as described below. The Green's function solution is given by L = ∂ ∂ X (cid:21) κ ζ (X −(cid:104)X(cid:105)) ζ ∂ ∂ X + (cid:20)kBT (cid:20)kBT (cid:20)2πkBT = κ ∂ P(X,t) ∂t ∂ 2 ∂ X 2 + ζ ∂ ∂ X + κ ζ κ ζ (X −(cid:104)X(cid:105)) (cid:21)−1/2 (1− e−2t/η ) exp G(X,X0;t) = (cid:21) P(X,t)− k(cid:104)X(cid:105)2P(X,t) (cid:20) − κ(X −(cid:104)X(cid:105))2 2kBT (1− e−2t/η ) (cid:21) k(cid:104)X(cid:105)2 exp(cid:2)−k(cid:104)X(cid:105)2t(cid:3) (S7) (S8) (S9) 18/29 To obtain the expression for P(t), we need average over the initial values (X0) and sum over the final values (X), P(t) = dX dX0G(X,X0;t)P(X0) (S10) (cid:90) ∞ 0 (cid:90) ∞ 0 From simulations, the initial values of X are sharply peaked fixed value x0, and so we use the Dirac delta function, P(X0) = δ (X0 − x0). By substituting P(X0) and equation (S9) into equation (S10) and performing the integration, we obtain the distribution of bond lifetimes: (cid:20) (cid:18) (cid:19)(cid:21) P(t) = k(cid:104)X(cid:105)2 2 exp[−k(cid:104)X(cid:105)2t]· 1 + Er f (cid:104)X(cid:105) [2πkBT /κ(1− e−2t/η )]1/2 In equation (S11), the average size (cid:104)X(cid:105) is given by equations (S2) and (S3) for the continuous and discrete versions, respectively. The average bond lifetime is given by Derivation of equation (S9): Equation (S8) in the text above can be solved by using the Zwanzig's ansatz: (S11) (S12) (S13) (S14) Upon the substitution of this ansatz into equation (S8), we obtain the following system of the ordinary differential equations for functions µ(t) and ν(t): The first equation (S14) can be readily integrated to obtain the solution for µ(t). The obtained solution can be substituted into the second equation (S14), which can then be solved for ν(t). We obtain: (cid:104)τ(cid:105) = tP(t)dt (cid:90) ∞ (cid:18)2π 0 κ (cid:19)−1/2 (cid:20) − (X −(cid:104)X(cid:105))2 2 P(t) = exp (cid:21) µ(t)− ν(t) (cid:26) µ = −2kBT /ζ µ2 + 2κ/ζ µ ν = kBT /ζ µ + k(cid:104)X(cid:105)2 − κ/ζ (cid:40) µ(t) = ν(t) = 1 P(X,X0;t) = 1 C1e−2t/η +kBT /κ 2 log(cid:2)kBT +C1κe−2t/η(cid:3) + k(cid:104)X(cid:105)2 +C2 (cid:20)2πkBT (cid:20) − κ(X −(cid:104)X(cid:105))2 2kBT (1− e−2t/η ) (1− e−2t/η ) (cid:21)−1/2 exp κ When substitute into equation (S13), arrive at the expression: which is equation (S9) above. (cid:21) k(cid:104)X(cid:105)2 exp(cid:2)−k(cid:104)X(cid:105)2t(cid:3) (S15) (S16) 19/29 Table S1. Statistical analysis of experimentally determined bond lifetimes for different interacting protein-coated surfaces (specific versus non-specific protein-protein interactions) collected at a constant pulling force f = 30 pN and contact duration T = 0.5 s. Shown are the percentages of the bond lifetimes for the following time ranges: τ < 0.03 s, 0.04 s < τ < 0.5 s and τ > 0.5 s. Interacting proteins Fibrinogen/BSA <0.03 s 96.9% 98.1% 96.7% 93.0% Fibrin/Fibrinogen in the presence of 2 mM GPRPam 92.6% Fibrinogen/Fibrinogen Fibrin/Fibrinogen Fibrin/BSA 0.04 - 0.5 s >0.5 s 0.6% 0.8% 1.4% 6.5% 1.3% 2.5% 1.1% 2.0% 0.6% 6.1% No. of touching events 4,136 3,170 3,481 4,125 4,023 Bond lifetime ranges Table S2. The model parameters for the A:a knob-hole bond dissociation kinetics: the binding interface stiffness κ, the initial (maximal) and minimal interface size (diameter) x0 and x1, the force-free rate of disruption of binding (receptor-ligand) residue-residue contacts per unit area k0, transition distance for dissociation σy, and transition distance for conformational transition state σx and energy difference between two states ε0 (for discrete two-state version). The values were obtained by fitting the theoretical profiles of the average bond lifetime (cid:104)τ( f )(cid:105) as a function of applied constant force f (curves) (equation (7)) to the experimental bond lifetime data (points; see Fig. 5). We used both continuous version (equation (S2)) and discrete version (equation (S3)) of the fluctuating bottleneck model of the receptor-ligand binding interface with a linear 'sink' K ∼ X (α = 1; see equation (2)). Model Continuous Discrete κ, pN/nm x0, nm x1, nm k0, nm2s−1 10.0 100.0 3.7 5.0 0.7 0.65 0.25 0.095 σy, nm σx, nm ε0, kcal/mol 0.175 0.195 1.242 17.0 - - 20/29 Figure S1. The log-linear plot of the distributions of bond lifetimes P(t) collected at a constant pulling force f = 50 pN for the fibrin-fibrinogen complexes in the absence and presence of 1 mM GPRP, a competitive inhibitor of the A:a knob:hole interactions. The vertical dashed line shows the bond lifetime threshold at 0.5 s, above which the A:a knob:hole interactions were found to be sensitive to the presence of GPRP peptide. Therefore, these data were considered to reflect the fibrin-fibrinogen specific interactions. The inset compares the probability of interactions lasting longer than 0.5 s in the absence and presence of GPRP peptide. 21/29 Figure S2. The binding probability for different interacting protein pairs showing that fibrin-fibrinogen complexes with lifetimes τ > 0.5 s collected at f = 50 pN formed much more frequently and that the formation of these complexes is prevented by GPRP. 22/29 Figure S3. The average bond lifetimes (cid:104)τ(cid:105) vs. applied constant force f for specific fibrin-fibrinogen interactions calculated with the cutoff set at 0.03 s and 0.5 s. 23/29 Figure S4. The binding probability as a function of the constant force f for fibrin-fibrinogen interactions calculated using the average lifetimes (cid:104)τ(cid:105) > 0.03 s and (cid:104)τ(cid:105) > 0.5 s. 24/29 Figure S5. The average bond lifetime (cid:104)τ(cid:105) > 0.5 s as a function of the constant tensile forces f for interactions of fibrinogen fragment D (with holes 'a' and 'b') with fragment desAB-E (bearing knobs 'A' and 'B') in the absence and presence of 1 mM GPRP. 25/29 Figure S6. The average bond lifetime (cid:104)τ(cid:105) > 0.5 s as a function of the constant tensile forces f for interactions of fibrinogen fragment D (with holes 'a' and 'b') with fragment desAB-E (bearing knobs 'A' and 'B') obtained using 10-fold different concentrations of fragment D (1 mg/ml vs. 0.1 mg/ml). A 10-fold difference in fragment D concentration translates to different surface densities achieved during fragment D immobilization. 26/29 Figure S7. Dynamics of the A:a knob-hole interaction probed by varying mechanical tension. Shown are four representative profiles of the molecular (unbinding) force F (panels a and b) and the number of residue-residue binding contacts reinforcing the bound state Q (panels c and d) as functions of time t from two sets of the force-ramp simulations with pulling speed ν f = 103 µm/s (panels a and c) and ν f = 104 µm/s (panels b and d). The text notations designate the most representative trajectories displaying the complex dissociation from the low-affinity bound state (pathway 1; blue curves) and high-affinity bound state (pathway 2; red curves). 27/29 Figure S8. Binding affinity of the A:a knob-hole complex. Panel a: Time evolution of the total number of binding contacts Q from four representative force-ramp trajectories in silico (the inset magnifies initial portion of the curves, showing an increase or decrease in Q). Panel b: Time evolution of Q (purple solid), number of binding contacts between moveable flap and knob 'A' QMT (purple dashed), and movable flap displacement relative to the B-domain DMT (magenta) from force-clamp simulations with f = 30 pN (vertical dashed lines emphasize correlated alterations in Q and DMT ). Panel c-d: Maps of residue-residue contacts stabilizing the complex interface (residues γSer240–γLys380) in the native state (bottom triangle) and intermediate state before dissociation (top triangle) for two simulations (blue and red curves in panel c) showing dissociation from the low-affinity bound state (panel c) and high-affinity state (panel d). A colored pixel corresponds to a contact, formed by amino acids in the i-th raw and j-th column, if the centers-of-mass of their side chains are in 5.5 A proximity. The color opacity is proportional to the contacts persistence. Different colors denote contacts formed by residues in knob 'A' (orange) and residues in one of the binding determinants in γ-nodule: loop I (blue), interior region (green), and movable flap (red). The β -sheets stack of B-domain is shown in purple. The solid ovals circle the areas of new contacts formed between the movable flap and β -sheet stack in B-domain (left) and between the movable flap and knob 'A' (right), which the native state lacks (dashed oval). 28/29 Figure S9. Probability distributions of bond lifetimes obtained for different values of tensile force f = 10, 20, 30, 40 pN. Comparison of experimental (red histograms) vs. modeling (blue curves) results (equation 6). 29/29
1512.01005
2
1512
2016-02-29T18:09:11
Growth, collapse, and stalling in a mechanical model for neurite motility
[ "physics.bio-ph" ]
Neurites, the long cellular protrusions that form the routes of the neuronal network are capable to actively extend during early morphogenesis or to regenerate after trauma. To perform this task, they rely on their cytoskeleton for mechanical support. In this paper, we present a three-component active gel model that describes neurites in the three robust mechanical states observed experimentally: collapsed, static, and motile. These states arise from an interplay between the physical forces driven by growth of the microtubule-rich inner core of the neurite and the acto-myosin contractility of its surrounding cortical membrane. In particular, static states appear as a mechanical traction/compression balance of these two parallel structures. The model predicts how the response of a neurite to a towing force depends on the force magnitude and recovers the response of neurites to several drug treatments that modulate the cytoskeleton active and passive properties.
physics.bio-ph
physics
Growth, collapse, and stalling in a mechanical model for neurite motility Pierre Recho1,∗ Antoine Jerusalem2,† and Alain Goriely1‡ 1 Mathematical Institute, University of Oxford, Oxford OX26GG, United Kingdom and 2 Department of Engineering Science, University of Oxford, Oxford OX13PJ, United Kingdom (Dated: March 1, 2016) Neurites, the long cellular protrusions that form the routes of the neuronal network are capable to actively extend during early morphogenesis or to regenerate after trauma. To perform this task, they rely on their cytoskeleton for mechanical support. In this paper, we present a three- component active gel model that describes neurites in the three robust mechanical states observed experimentally: collapsed, static, and motile. These states arise from an interplay between the physical forces driven by growth of the microtubule-rich inner core of the neurite and the acto- myosin contractility of its surrounding cortical membrane. In particular, static states appear as a mechanical traction/compression balance of these two parallel structures. The model predicts how the response of a neurite to a towing force depends on the force magnitude and recovers the response of neurites to several drug treatments that modulate the cytoskeleton active and passive properties. I. INTRODUCTION Neurons are cells with long and thin (∼ 1µm in di- ameter) quasi one-dimensional processes called neurites, a term which comprises the axon which emits electric signals and dendrites which are generally shorter and re- ceive signals. These processes emerge from the cell body (the soma) and, during embryonic development or re- generation after a trauma, are able to crawl over large distances to reach targets from other neurons thus form- ing a complex nervous network essential for both per- ception and motion. Understanding how neurites can establish these long distance connections is a problem which was pioneered more then a century ago [1] and is of paramount therapeutic importance. For instance, injuries of the spinal cord are often characterised by an irreversible and debilitating loss of motor and sensory functions of the lower body (paraplegia, tetraplegia) be- cause disrupted neurites cannot overcome the inflamma- tion and are incapable to initiate extensions that would rebuilt the broken connections [2–4]. The cytoskeleton of neurites is the mechanical scaf- fold that maintains their morphology and motility [5–8]. Extrinsic and intrinsic guidance clues may be viewed as agents that influence the physical state of the cytoskele- ton via biochemical pathways [9–12]. For example, the concentration of calcium is known to influence the Rho pathway which in turn modulates the neurite contractil- ity and can lead to a reversible collapse that shortens the axon [13]. The neuronal cytoskeleton (see Ref. [8] for an ex- tensive review and Fig. 1 for a simplified scheme) is a meshwork of three main types of biological polymers: F- actin, microtubules, and neurofilaments which all con- tribute mechanically [14]. While neurofilaments are pas- sive and apolar, both F-actin and microtubules are ca- ∗ [email protected][email protected][email protected] pable to polymerise at one end (addition of G-actin and tubulin subunits, respectively) and depolymerise at the other (by removal of subunits) with potentially low (∼ 1 min) turnover duration. They can also both be cross- linked by molecular motors (myosin II for F-actin, dynein or kinesin for microtubules) that are able to exert ac- tive stresses inside the meshwork [15–17]. Following Ref. [5], we define two different compartments of the neurite where the cytoskeleton is organised in a different way: the kinetoplasm, or growth cone (GC), at the proximal end of the neurite and the axoplasm which connects the GC to the soma. The axoplasm contains a core array of para-axial microtubules connected by passive cross- linkers (MAP: Microtubules Associated Proteins) that generate a network with a quasi-lattice structure [18]. These microtubules are highly stable (turnover duration of hours) possibly due to the presence of MAPs. In the axoplasm, F-actin is mostly organised into a corti- cal mesh around the microtubule's inner core and the presence of myosin II motors in this cortex leads to the presence of contractile stresses [19]. These two mesh- works are physically connected by different types of spe- cial proteins (such as +TIP, see Ref. [8]) which mediate force transmission between them. In continuity with this cortex, the GC is almost free of microtubules, apart from those engaging into filopodia while F-actin is organised in a lamellipodium similar to the ones found in cells spe- cialised in crawling (such as keratocytes [20]). Filaments polymerising at the leading edge (the P-domain) pro- trude the membrane and are then advected backward by a retrograde flow powered by myosin II motors that con- centrate at the trailing edge of the GC (the T-domain) [21]. Actin then accumulates into thick bundles in the T-domain that constitutes the main obstacle preventing microtubules from entering the GC. The cytoskeleton is connected to the external substrate/cellular matrix by special proteins specialised in adhesion such as integrins and cadherins [22, 23]. Numerous authors have proposed physical models to explain how the neurite cytoskeleton drives its motility [7, 25–27]. These models can be divided into two main 6 1 0 2 b e F 9 2 ] h p - o i b . s c i s y h p [ 2 v 5 0 0 1 0 . 2 1 5 1 : v i X r a 2 towards the soma, the velocity decays exponentially, sug- gesting a fluid-like behaviour of the neurite. However, if the applied force is not large enough, the neurite may undergo a finite deformation instead of acquiring a finite velocity [47] and both rheological models of Ref. [44] and [46] indicate a long term stiffness of the neurite two orders of magnitude smaller than the short time one. Further- more, the loading rate is also known to play an important role in the possible action potential impairment of the axon or in the transport properties alteration resulting from a loss of connectivity of the microtubules network [18, 48–50]. Dynamic loading is not studied in this article and the loading is assumed to be quasi-static. Axoplasm active growth and contractility have been proposed as the potential drivers for retraction or elongation of the neu- rite in presence of an applied force [44], and contractility is explicitly incorporated as a force opposing elongation in Ref. [33, 46]. The GC has been characterised as a Maxwell viscoelas- tic fluid with a relaxation time of a few seconds and an active contractile pre-stress stemming from the motor ac- tivity at the rear of the cone [51]. The F-actin poly- merisation driven formation of filopodia extension and retraction has been physically described in Ref. [52, 53]. In the present paper, we follow the suggestion of Ref. [33] that the theory of active gel may be used to unify these aforementioned models in order to obtain a global picture of neurite motility. Our one-dimensional model is based on the particular geometry of the neurite cy- toskeleton and fundamental balance laws. An analysis of its solutions reveals that, depending on the neurite pas- sive and active rheology, the neurite can collapse to the soma, remain static, or grow at finite velocity. The in- terchangeability of these three states is consistent with experiments that modify the state of the cytoskeleton and its substrate adhesivity with drugs. In particular, we recover within a common framework the following general trends emerging from different sets of experiments probing the mechanical and structural en- vironment of growing neurites: • Growth under axial force: As mentioned above and already observed 30 years ago [30], a steady axial force applied by a cantilever at the proximal tip of a neurite elongates it [31–33, 44, 47, 54, 55]. This elongation is elastic if the force is below a cer- tain threshold. Above that threshold, the neurite grows with finite velocity [56]. • Retraction under microtubules depletion: As shown in Ref. [57], neurites retract in response to microtubles depletion and elongate even in vivo fol- lowing stimulation of microtubules polymerisation [58]. • Retraction with reduction of adhesivity: By culturing neurites on different substrates, it was shown that retraction is promoted when the sub- strate adhesivity is reduced [33]. FIG. 1. neurite extending from the soma. Adapted from Ref. Colors online. Schematic representation of the cytoskeleton of a [24]. classes depending on whether they imply that the GC pulls the trailing axoplasm thanks to F-actin polymerisa- tion pushing the membrane in the P-domain [28, 29] and myosin II contractility pulling from the T-domain [30– 33], or whether it is rather microtubules which, from the axoplasm, polymerise against the T-domain and propel the GC [34, 35], or both [36]. As the rate of polymerisa- tion of microtubules depends on the force applied at their tip [37], both effects have been considered in a single sim- ple model [38]. Experimentally, the physical forces aris- ing from the F-actin or the microtubule meshworks are important since drug treatments lowering myosin II con- tractility (blebbistatin), preventing actin polymerisation (cytochalasin B), depolymerisating microtubules (noco- dazole) or on the contrary stabilising them (taxol) each influence the tip velocity of a crawling neurite in a con- centration dependent manner [15, 24, 39–42]. Interest- ingly, different levels of cytochalasin B can lead to either an increase [40] (high level) or a decrease [43] (low level) of the speed. This dependence suggests an antagonistic role of the acto-myosin meshwork in the propulsion. Mechanical models require a rheological characterisa- tion of the axoplasm and the GC. The axoplasm has been described as a Burgers viscoelastic material based on its fast elastic response (seconds to minutes) to a force ap- plied laterally or at the tip [44–46] while it elongates at a constant rate on longer time scales (a few hours) in re- sponse to a constant high force [32, 33]. In Ref. [32, 33], experiments tracking the mitochondria docked on the mi- crotubule array have shown that, close to the T-domain, this network flows forward with a velocity comparable to the velocity of the neurite. Away from the T-domain where the source term Sµ = kµ p − kµ d ρµ, 3 (2) FIG. 2. online Schematic of the neurite model geometry. Colors • Motility is related to contractility: Finally, neurites initiate their motility in a robust way when exposed to drugs that impair their active contrac- tility [42, 59]. The last two effects [60] are particularly important as possible therapeutic targets to promote axon regenera- tion after trauma [59]. The paper is organised as follows: In Section II, we de- velop a mechanical model for the axoplasm (acto-myosin and microtubules phases) alone and study its motility properties under an applied proximal traction force. In Section III, we model the GC (acto-myosin phase only) motile properties in response to a traction force applied at the trailing edge. In Section IV, we combine both models by assuming stress continuity at the T-domain to obtain a complete model of a growing neurite and show that it compares well with experiments. II. AXOPLASM PROPULSION Following Ref. [61], we model the microtubule network core of the axoplasm as a one-dimensional morphoelastic rod whose material points are indexed by the coordinate x ∈ [0, ln(t)], 0 denoting the connection with the soma and ln(t) the moving boundary between the GC and the axoplasm (T-domain) as shown in Fig. 2. Notice our model does not separate the contribution of the neurofil- amants from the one of the microtubules. They are thus viewed as a passive reinforcing structure contributing to the overall network elasticity [14]. This rod can only de- form along the axis and is in frictional contact with a vis- cous contractile "sleeve" (the cortex) which is supported by a static background. The tip of the neurite is sub- jected to a given traction force exerted either by the GC in normal growth conditions or by a micropipette, in ex- periments where the GC is lifted from the substrate [33]. A. Balance of mass Microtubule network Let ρµ denote the mass density of microtubules and vµ their velocity in the lab reference frame. The mass balance equation reads ∂tρµ + ∂x(ρµvµ) = Sµ, (1) follows a first-order kinetic with a polymerisation rate p and a depolymerisation term kµ kµ d ρµ, proportional to the density [62]. We assume here that the tubulin (mi- crotubule subunits) concentration is homogeneous [63] along the neurite because its motor driven transport is much faster (∼ 1µm.s−1) than the crawling velocity −1). We can rewrite Eq.(2) as Sµ = (¯ρ−ρµ)/τ (∼ 10µm.h p /kµ where ¯ρ = kµ d is the density at chemical equilibrium and τ = 1/kµ d is the turnover timescale associated with microtubule renewal. Here we have adopted a mean field description of the network and we do not consider the microtubule polarity. Indeed, while this information is likely to be important for transport properties along the neurite, microtubules have a clear forward polarity in the axon and a mixed one in dendrites [8]. Yet, these two structures can equivalently move, suggesting that polar- ity may not be a fundamental variable in this physical process. Also note that while we do not account for the influence of a potential loading on the kinetic rates kµ p and kµ d , our mean field description captures the load- dependant dynamic of the whole microtubule array (see Section II E). Assuming that no microtubule comes from the growth cone, Eq. (1) is equipped with a no-flux boundary condi- tion at the tip of the neurite ln = vµ(ln(t), t). (3) For simplicity, we also impose a no-flux boundary condi- tion at the connection with the soma, so that vµ(0, t) = 0. Notice, that at both ends, there is no assumption on the flux of tubulin which adjusts to maintain a constant con- centration as hypothesised in (2). Cortical network Denoting ρc the mass density of actin in the cortex, we can write an conservation equation similar to Eq. (1): ∂tρc + ∂x(ρcvc) = Sc, (4) where vc is the velocity of the actin network in the cortex in the lab reference frame. However, we assume that the actin network is highly compressible compared to the microtubule network. Therefore this equation decouples from the rest of the system and actin density can be found post-factum when the velocity field vc is known using the method of characteristics [64]. This point is not tackled in the present paper and Sc is thus left unspecified. We again assume that there is no filamentous actin flux from the soma to the cortex: vc(0, t) = 0. Unlike the microtubule network, the cortical actin is not stopped at the T-domain and can flow freely in the growth cone. Thus, there is no condition such as Eq. (3) for the cortical flow. B. Balance of linear momentum Microtubule network The microtubule network is in contact with the cortical actin network through differ- ent types of cross-linking proteins that can actively bind and unbind (see Ref. [8] for a review). Assuming a suf- ficiently fast binding/unbinding dynamic [65], we model this contact as a viscous friction. Neglecting inertia, the balance of linear momentum reads ∂xσµ = ζµ(vµ − vc), (5) where ζµ is a friction coefficient and σµ is the internal axial stress rescaled by the microtubule network width, that is, if hc is the width of the cortex and hµ the width of the microtubule network both assumed to be constant, then σµ = (1−w)Σµ, where Σµ is the axial Cauchy stress (axial force per unit area). We denote w = hc hµ + hc ∈ [0, 1] as the ratio of the cortical over total width of the axon. At the leading edge, the axoplasm is subjected to a traction stress Q. Thus, the boundary condition associ- ated to Eq. (5) reads σµ(ln(t), t) = (1 − w)Q. Cortical network Similarly, the linear momentum bal- ance in the cortical layer reads ∂xσc = −ζµ(vµ − vc) + ζcvc, (6) where ζc is a friction coefficient of the cortex with respect to the substrate through adhesive proteins [22]. Here, σc = wΣc is the rescaled axial stress, so that the bound- ary condition at the leading edge is, σc(ln(t), t) = wQ. C. Constitutive relations To close our system of equations we posit two assump- tions about the rheology of the microtubules and cortical networks. Microtubule network Given the long turnover time of microtubules inside the axoplasm which are highly sta- bilised [5, 8] and their high stiffness compared to F-actin filaments, we consider this network to be elastic for the timescale of interest (hours). We assume a logarithmic elastic stress-strain dependence σµ = −(1 − w)E log , (7) (cid:18) ρµ (cid:19) ρ0 which has the advantage to prevent both infinite contrac- tion and dilution. The natural density of microtubules at 4 which no stress is created is ρ0. In principle, this density can be modulated by the presence of molecular motors in the microtubule array [33]. It is worth mentioning that our results are robust with respect to the choice of other increasing concave function than log. In fact, realistic parameters show that ρµ is rather close to ρ0 thus imply- ing that in the range of strain considered in physiological conditions, a linear relation between the stress and the local density (representing the strain in 1D) could poten- tially be acceptable too. Cortical network The turnover duration of an actin fibre in the cortex is fast (few seconds [8, 51]) and we use a linear viscous law for this phase to relate the stress to the strain rate. In addition, we assume that there is an active contractile stress created by the myosin II motor activity [19]: σc = w(η∂xvc + χc). (8) The bulk viscosity of F-actin is η, χ > 0 is the contrac- tility coefficient and c the concentration of motors. The conservation equation for c is (see Ref. [66, 67] for further details), ∂tc + ∂x(cvc) − D∂xxc = ¯c − c τc . (9) The motors are advected with the actin filament that they bind but can also thermally diffuse with a diffusion coefficient D. The linear reaction term accounts for the attachment/detachment of motors with a cycle time τc. The concentration of motors at chemical equilibrium is ¯c. We can supplement this equation with a no-flux bound- ary condition at the soma, ∂xc(0, t) = 0 and assume that the motor concentration in the T-domain is a constant c(ln(t), t) = c0. Note that we do not resolve the radial component of the stress in our model (similarly to a neurite constrained in a channel [68]). Radial stress will be important for further investigation on the shape and turning of neurites which is outside the scope of this paper. D. Final system Denoting σ = σc + σµ the total stress, v = vc the ve- locity of actin and ρ = ρµ the density of microtubules, combining our model equations, we obtain the final sys- tem − wη ζc ∂xxσ + σ = wχc − E(1 − w) log ∂tρ + ∂x(ρv) − E(1−w) ∂tc + ∂x(cv) − D∂xxc = ¯c−c ζµ ∂xxρ = ¯ρ−ρ , τ τc (cid:16) ρ (cid:17) ρ0 (10) where the velocity field is related to the stress by v = ∂xσ/ζc. The boundary conditions are: (cid:26) ∂xσ0 = 0, ∂xρ0 = 0 and ∂xc0 = 0 σln = Q, ρln = ρ0e−Q/E and cln = c0. The last no-flux boundary condition, ln = vln − E(1 − w) ζµ ln(t), ∂xρ ρ (11) (12) is a Stefan condition needed to compute the unknown time dependence of the free boundary ln(t). In general, initial conditions should also be given but here, we focus on steady states only. Equation (10)2 is obtained by expressing the velocity vµ using Eq.(5) and combining it with Eq.(1). We comment on the structure of Eq. (10): the stress is created non-locally over the so-called hydrodynamic length lc = (cid:112)η/ζc by two active agents. Motors from the cortex are pullers creating a contractile stress and microtubules, provided their density is larger than ρ0, are pushers creating a tensile stress due to the addition of tubulin subunits in the network (growth). If tubulin sub- units are removed (shrinking), the microtubules are also pullers. In the context of cell motility, such structures with pushers and pullers have been investigated in Ref. [69, 70] where it was shown that growing and contracting agents can conspire to achieve robust motile properties of an active segment. Here, we supplement the picture with the two simple and similar dynamic Eq. (10)2,3 ruling the distribution of pushers and pullers which are relevant in the case of axonal motility. Having already investigated the pullers dominated case in Ref. [71] for motility properties and for the formation of periodic F-actin rings [67] which are actually observed in axons [8, 72], we turn our attention to the pushers dominated case. To understand this regime, here and subsequently, we restrict our attention to the case where the concentration of motors is homogeneous in the cortex, i.e., c ≡ ¯c. The resulting system can then be rewritten in the following minimal form, (cid:40) wη∂xxv − ζcv = E(1 − w) ∂xρ ∂tρ + ∂x(ρv) − E(1−w) ∂xxρ = ¯ρ−ρ ρ τ ζµ with boundary conditions, (cid:26) v0 = 0 and ∂xρ0 = 0 η∂xvln = Q − χ¯c and ρln = ρ0e−Q/E, (13) (14) along with the Stefan condition in Eq. (12). Substituting the non-dimensional quantities, σ = σ E , x = , t = t η/E , ρ = ρ ρ , and Q = Q/E, in Eq. (15) (13-14) and dropping the tildes for clarity, we x(cid:112)η/ζµ obtain the system (cid:26) w∂xxv − av = (1 − w) ∂xρ (cid:26) v0 = 0 and ∂xρ0 = 0 with boundary conditions, ∂tρ + ∂x(ρv) − (1 − w)∂xxρ = (1 − ρ) ρ ∂xvln = Q − Qc and ρln = eQµ−Q and the free boundary condition, ln = vln − (1 − w) ln(t). ∂xρ ρ 5 (16) (17) (18) We have now six non-dimensional parameters: • w, the relative width of the cortex with respect to the microtubule network; • a = ζc/ζµ, the ratio of friction coefficients; •  = η/(τ E), the ratio of the acto-myosin over mi- crotubule network viscosities; • Q, the applied load at ln scaled by E; • Qc = χ¯c/E > 0, the scaled contractile load; • Qµ = − log(¯ρ/ρ0) (see below). Note that these six non-dimensional parameters could be reduced to five by defining Q = Q− Qµ and ∆Qc µ = Qc− Qµ. However, to keep the treatment of the microtubule and acto-myosin meshwork parallel, we keep the three distinct loads. The system of Eq. (16-18) cannot be explicitly solved but some asymptotic cases give insight on how the physics of such medium works. E. "Solid" and "fluid" asymptotic cases The no-microtubule case, w → 1 In the absence of microtubules, the velocity field can be solved directly from Eq. (16)1 and we obtain, Q − Qc√ √ √ sinh( v(x, t) = ax) . a cosh( aln(t)) Plugging this expression in Eq. (18) leads to ln(t) = Q − Qc√ a √ tanh( aln(t)). This case was investigated in the absence of contraction (Qc = 0) in Ref. [32] and successfully compared to ex- periments where the GC was lifted and the axon was mechanically pulled with a cantilever. Notice however that in Ref. [32], v is the microtubule velocity while, here, v is the velocity of F-actin. √ The importance of axoplasmic contraction (Qc (cid:54)= 0) was recently demonstrated in Ref.[33]. In this case, there are two possible steady states. Either the load- ing is larger than the contractile stress, Q ≥ Qc, the finite velocity and the axon then extends at Vn = ln = (Q − Qc)/ a, or the loading is weaker than the contractile stress, Q < Qc, and the neurite then collapses back to the soma. We sketch the force velocity Vn(Q) relation in Fig. 3 (middle panel). This case can be referred as "fluid-like" growth given that the axoplasm is effectively modelled as a contractile viscous fluid. Alternatively, it was shown in Ref. [73] that this case can also be described as a morphoelastic rod by combining an elastic response with a fast evolution of the reference configuration. The no-cortex case, w → 0 In the absence of a cortex, we combine Eq. (16)1 and Eq. (16)2, to obtain the linear equation ∂tρ − (1 + a−1)∂xxρ = (1 − ρ). Its long-time asymptotics, on a semi-infinite domain x ≤ ln(t), can be found by considering the traveling wave reduction y = x − ln(t) in the domain y < 0, with ln(t) = Vnt. Denoting ( )(cid:48) the derivative with respect to y, we obtain −Vnρ − (1 + a−1)ρ(cid:48)(cid:48) = (1 − ρ), with boundary conditions, ρ0 = eQ−Qµ , ∂xρ−∞ = 0, and the front velocity given by, Vn = −(1 + a−1)eQµ−Q∂xρ0. The solution of this linear problem is given by ρ(y) = 1 − p(Q) 1 + p(Q) ey/l(Q). This expression depends on l(Q), which can be in- terpreted as the typical size of a boundary layer, over which the reaction of polymerisa- tion/depolymerisation of microtubules is maintained out of equilibrium at the tip of the neurite: chemical (cid:114) l(Q) = (1 + a−1)(1 + p(Q))  . The parameter p(Q) = eQ−Qµ − 1 > −1 represents the driving force leading to expansion or retraction. Indeed, the tip velocity can be expressed as p(cid:112)(1 + a−1) √ . 1 + p Vn = 6 FIG. 3. (a): Velocity-force relation in the absence of a cortical acto-myosin network. (b): Velocity-force relation in the absence of a microtubule network. (c): Two thresholds Velocity-force in the general case. If p < 0 (Q < Qµ), then the combined effect of external force and shrinking of microtubules leads to a collapse back to the soma while if p > 0 (Q > Qµ) the stress provided from microtubules growth is large enough to overcome an external load and a steady expansion is pre- dicted. In the absence of molecular motors, that is, with- out contraction, Qµ < 0 as microtubules are able to push [37]. But, in the presence of molecular motors, the sign of Qµ cannot be readily established [33]. We sketched the force velocity Vn(Q) relation on Fig. 3 (left panel). In this case the neurite is effectively modelled as a growing elastic solid. Notice that the velocity of the microtubules p(1 + a−1)ey/l l(1 + p − pey/l) vµ = (1 + a−1) ∂yρ ρ = also displays an exponential decay away from the tip of the neurite. This behavior is consistent with the experi- ments of Ref. [32]. From the two limiting cases discussed above, we can obtain a general picture of the dynamics when 0 < w < 1, for small and large values of the applied load Q: • If Q < Qµ, the axoplasm will collapse back to the soma. Indeed, both the cortex and the microtubule networks are in a collapse mode; • if Q > Qc, the axon length will increase at a finite velocity given that both the cortex and the micro- tubules network are in extension. Next, we consider the interval Q ∈ [Qµ, Qc] by studying possible static states where a finite load does not lead to motion. F. Static states Static states are the solutions of the following problem: (cid:26) w∂xxv − av = (1 − w) ∂xρ ∂x (ρv − (1 − w)∂xρ) = (1 − ρ), ρ (19) with the boundary conditions of Eq. (17) and where ln is a constant given by the condition vln = (1 − w) ln(t). ∂xρ ρ While there is no obvious solution to this nonlinear prob- lem, the following two limiting cases shed light on the general case. 7 Large microtubule network viscosity,  → 0 In the limit where the microtubule network viscosity is much larger than the acto-myosin network viscosity, we can take the limit  → 0. In this case Eq. (19)2 can be solved exactly: ρ(x) = exp , (20) Qµ − Q + σ(x) − Q a(1 − w) (cid:18) (cid:19) FIG. 4. Length-force relations for the static solutions given by Eq. (21) for different values of a (Fig 4 (a)) and w (Fig 4 (b)). Parameter ∆Qc µ = 5. Colors online which can then be substituted into Eq. (19)1 to obtain, w∂xxv − (1 + a)v = 0. for Q ∈ [Qµ, Q0 s] with The solution of this last equation is simply x√ w ln√ w v(x) = (Q − Qc) (cid:114) w 1 + a cosh sinh 1+a 1+a The stress is obtained by integrating v: σ(x) = wa 1 + a (Q − Qc) cosh (cid:18) (cid:18) (cid:19) (cid:18) (cid:19) (cid:19) . (cid:18) (cid:19) (cid:18) ln√ w 1+a cosh − cosh ln√ w 1+a (cid:19) x√ w 1+a , which can be substituted back into Eq. (20) to obtain a closed expression for the density. The last constraint is provided by integrating Eq. (16)2 and requiring that for steady states the average density of microtubules is conserved, i.e., (cid:90) ln 0 1 ln ρ(x)dx = 1. This constraint is now an integral equation for the static length ln: lneQ−Qµ = (cid:90) ln 0 exp −f0(Q − Qc) 1 − (cid:18) (cid:18) cosh cosh x√ w ln√ w 1+a 1+a (cid:19) (cid:19)  dx  Q0 s = Qµ + f0Qc 1 + f0 . (22) The parameter Q0 s is an average of Qµ and Qc weighted by the cortical width and the friction coefficients. The numerical solution of Eq. (21) between the two thresh- old loads is given in Fig. 4. The static length increases monotonically from 0 to ∞ between Qµ and Q0 s. Large cortex viscosity,  → ∞ In this limit, it is clear from the right hand-side of Eq. (16)2 that ρ converges to 1 almost everywhere in the layer, except in a boundary layer close to the tip where it has to satisfy the constraint ρln = eQµ−Q. To obtain the dynamics of the moving front ln we use the piecewise linear ansatz: (cid:26) 1 if x < ln − 1 (cid:0)eQµ−Q − 1(cid:1) (x + 1  ρ = Using this ansatz, we solve Eq. asymptotically in 1/. To leading order, we have, v(ln) = (1 − w)(1 − eQµ−Q) + w(Q − Qc) √ eQµ−Q wa  − ln) + 1 if x > ln − 1  . (cid:19) (cid:18)√ (16)1 to obtain v(ln) th aln√ w . This value is finite because the left handside of Eq. (16)1 is a regularising elliptic operator. To leading order, the front dynamic is then given by √ V = ln ∼ (1 − w) a 1 − eQµ−Q eQµ−Q . (21) where the constant f0 reads f0 = w/[(1 + a)(1 − w)]. In Appendix A, we show that there exists a steady solution We conclude that in the large  regime, there is no static front unless Q = Qµ. For Q > Qµ the axon increases indefinitely and collapses for Q < Qµ. wwww(a)(b) 8 This double force thresholds system is in agreement with experiments [56]. Physically, our model reveals that the applied tip stress Q (positive when the neurite is pulled and negative if it pushes against an obstacle) must be larger than the stress created by the growing micro- tubule network Qµ to avoid collapse but it must also be larger than the effective stalling stress Qs of the entire structure to lead to steady elongation. See Fig. 3 (c). Be- tween these thresholds, the neurite effectively behaves as a neutral solid in the sense that an increase of force leads to a global strain of the neurite which acquires a new rest length. We further speculate that the oscillations in the loading at the T-domain coming from oscillating filopo- dia [52] can lead to the small scale stop and go motion [22] experimentally observed during elongation. H. Further simplifications Analytical estimates can be obtained if we further simplify the model by using the fact that microtubule growths are localised at the tip of the axoplasm and pro- vides an effective advection velocity of the free boundary [19, 74, 75]. We assume that the axoplasm is a mixture of con- tractile acto-myosin network (with fraction w) and mi- crotubule network (with fraction 1 − w). In the non- dimensional notations used previously, the mechanics of the contractile phase is then given by Similarly, the growing microtubule phase is given by where ζeff = (1+a)n and ηeff = (1+)−m with parameters n > 0 and m > 0 chosen below. The front dynamic is given by the no-flux boundary condition, ln = wvcln + (1 − w)vµln . (23) and (24) and assuming that ηeff is Solving Eq. small enough (localised tip growth assumption) com- pared to ln, the moving front dynamic is given by ln = w(Q − Qc) tanh((cid:112)ζeffln) + (1 − w) Q − Qµ√ ηeff (25) √ which has a similar behaviour as the full model. Namely, • For Q < Qµ, ln → 0 and the neurite collapses to the soma. • For Qµ < Q < Qs, we have ln → ls n and the neurite ∂xσc = ζeffvc ∂xσc0 = 0 and σcln = Q.  σc = ∂xvc + Qc  σµ = ηeff∂xvµ + Qµ ∂xσµ = vµ ∂xσµ0 = 0 and σµln = Q, (23) (24) (a): Numerically constructed phase diagram of the FIG. 5. axoplasm. Parameters are a = 0.1, w = 0.1 and ∆Qc µ = 0.3. The dashed line is the analytic approximation from the meta- model presented in Section II G. (b): Typical steady state profiles of the microtubules velocity and density in the three phases for  = 0.001. The numerical method is presented in Appendix B. Colors online. G. General behavior Rather than tackling the difficult questions of unique- ness and stability (both local and global) of the solutions that we have given in the previous sections, we use nu- merical integration of problem (16) to build the phase di- agram shown on Fig. 5 (see Appendix B for the method). For a given set of parameters w, a, ∆Qc µ, we show in the (Q, )-plane the domain of existence of the three observed behaviours: Collapse, Static, and Motile. Essentially, the overall behaviour at finite  is as fol- lows: • For Q < Qµ, we observe a collapse of the neurite back to the soma. This collapse is associated with a backward flow of microtubules along the entire axoplasm (see Fig. 5) as observed experimentally [33]. We find numerically that the time to collapse decreases with increasing  since the effective re- sistance of the microtubule network to contraction decays. • For Qµ ≤ Q ≤ Qs, where Qµ < Qs < Q0 s: we ob- serve a stabilisation of the neurite in a static state stemming from an interplay between the growing core and the surrounding contractile sleeve. There- fore these static states may be interpreted as a ten- sile tightening of two parallel active networks. • For Q > Qs, we observe that the neurite tip moves with a finite velocity which increases with . The microtubules flow increases towards the tip and de- velops a boundary layer at the junction with the GC as observed experimentally [32, 33]. Accordingly, velocity-force relation is given in Fig. 3 (c). the general qualitative picture for the reaches a finite length given by (cid:18) Q − Qµ (cid:19) 1√ ζeff ls n = arctanh where feff = w/(1 − w)(cid:112)ηeff/ζeff. The threshold feff(Qc − Q) (26) load is given explicitly by Qs = Qµ + feffQc . 1 + feff To recover the value Qs → Q0 s (see Eq. (22)) in the limit  → 0, we choose n = 2. Then, we also have Qs → Qµ in the limit  → ∞. Finally, the parame- ter m ∼ 1/5 is chosen heuristically to approximate the value of Qs given by the general phase diagram (see Fig. 5). • For Q > Qs, we have ln → ∞ and the neurite reaches a finite velocity Vn given by Vn = (1 − w)√ ηeff (1 + feff) (Q − Qs) . (27) III. GROWTH CONE PROPULSION Ahead of the axoplasm, we do not distinguish the lamellipodial and filopodial phases of the GC in the fol- lowing 1D model. The GC is continuous with the acto- myosin cortex and we shall therefore model it as an visco- contractile material. We use the index gc to denote vari- ables related to the GC. Balance of mass The mass balance for actin reads ∂tρgc + ∂x(ρgcvgc) = −kdρgc where ρgc is the density of F-actin and vgc, its velocity in the lab reference frame. Here, kd is the bulk depoly- merisation rate [66]. This equation is supplemented with the kinetic boundary condition lgc = v(lgc(t), t) + vp, where vp is the localised polymerisation (G-actin to F- actin) velocity at the tip of the GC [19, 52, 74]. More gen- erally, vp may also depend on the microtubules extending into the filopodium [35] and on the external loading [19] at the proximal tip of the GC, but we shall not consider this dependence. Therefore, we assume here a stress free leading edge. As in the previous discussion, the high compressibility of F-actin leads to the decoupling of the actin density with the front dynamic. Balance of momentum. In the viscous regimes, the balance of linear momentum reads ∂xσgc = ζcvgc, (28) where ζc is the friction coefficient with respect to the sub- strate. This equation is supplemented with stress bound- ary conditions, σgcln(t) = Q and σgclgc(t) = 0 9 (29) where, as before, Q denotes the common traction force in the T-domain at the axoplasm/GC interface. The lead- ing edge of the cone is assumed to be stress free. Constitutive relation The constitutive relation in- cludes both a viscous and a contractile term: σgc = η∂xvgc + χ¯c − p. (30) The pressure p is defined numerically as a constant La- grange multiplier associated with the conservation of the one-dimensional volume of the GC, lgc(t) − ln(t) = L. (31) This constraint follows from both osmotic effects [76, 77] and the fact that few compressible microtubules are en- gaged into filopodia [78]. More general models taking into account global com- pressibility of the GC may be required to access variation of L when some rheological parameters such as the tip growth velocity or the contractility for instance are af- fected by drug treatments [79]. However, our goal here is to describe the entire neurite and we will not discuss these finer effects further. Crawling velocity of the cone Combining Eq. (28) and (30), we obtain, − l2 c ∂xxσgc + σgc = χ¯c − p. (32) (32) with boundary conditions (see Eq. Solving Eq. (29)) and satisfying the constraint of fixed length (see Eq. (31)), we obtain a closed expression for σgc and vgc which we use to compute the fronts dynamic (see Ref. [71, 75] for further details): lgc = ln = Vgc = 1 2 vp − √ Q ηζc tanh(L/(2lc)) the √ Introducing Vgc/(E/ this last expression becomes √ = velocities ηζu), Q = Q/E, ηζu) and vp = vp/(E/ dimensionless Vgc (cid:18) (cid:32) (cid:19) . (cid:33) . Vgc = 1 2 vp − √ Q a tanh(L/(2lc)) From simple physical parameters estimates (see Table I), we have L (cid:29) 2lc, so that tanh(L/(2lc)) ∼ 1 and we obtain, after dropping the tildes, (cid:18) (cid:19) Vgc = 1 2 vp − Q√ a . (33) The GC is propelled by the polymerisation of the actin network at the leading edge which pushes the membrane forward. But, the GC is also pulled by the traction force name F-actin viscosity Elasticity of microtubules symbol η E Microtubules viscosity Viscous friction coefficient Eτ ζc Contractility χ¯c F-actin polymerisation velocity vp GC Length L 10 typical value 103 Pa.s [51] 200 − 400 Pa [44–46] 1 − 5 × 106 Pa.s [32, 33] 1014 − 1015 Pa.m−2.s [23, 32] 10 − 102 Pa [33, 45, 46, 51] 2 × 10−8 m.s−1 [19, 52] 1 − 2 × 10−5 m [33, 51] 0.1 [72] 10ζc (estimated) Cortex to axon width Friction cortex/microtubules hydrodynamic length lc characteristic length characteristic time characteristic velocity characteristic stress E w ζµ (cid:112)η/ζc ∼ 1.5 × 10−6 m (cid:112)η/ζµ ∼ 4.4 × 10−7 m E/(cid:112)ηζµ ∼ 1.5 × 10−7 m.s−1 ∼ 3 s η/E ∼ 300 Pa TABLE I. Estimates of material coefficients. effect Drug blebbistatin inhibit myosin II contractility Qc ↓ inhibit Myosin II contractility Qc ↓ BDM parameters trend cytochalasin inhibit actin polymerisation latrunculin destroys the actin meshwork nocodazole depolymerises microtubules epothilone B polymerises microtubules taxol trypsin stabilises microtubules detaches the neurite vp ↓ (Qc ↓ high concentration) vp ↓, Qc ↓ Qµ ↑ Qµ ↓  ↓ (Qµ ↓ low concentration) a ↓ at the interface with the axoplasm. This traction force decreases the velocity of migration if Q > 0. The GC stops moving when Q reaches the stall force √ Qgc = avp. IV. FULL NEURITE CRAWLING A. Overall behaviour We can now combine the models for the motion of the GC and the axoplasm parts to obtain a full picture of the neurite dynamics. We use the analytic relations derived in Sections II G and III as they capture the main effects. We can distinguish three cases depending on the acto- myosin of the GC Qgc: • If Qgc < Qµ, then the axon collapses to the soma in finite time. • If Qµ < Qgc < Qs, then the axon has a finite static length. Using the approximation given by Eq. (26), we have (cid:18) Qgc − Qµ feff(Qc − Qgc) (cid:19) ls = 1√ ζeff arctanh . (34) TABLE II. Effects of classical drugs. • If Qs < Qgc, the axon acquires a finite steady state velocity. Using the approximation given by Eq. (27) and (33), we obtain 9 Pa by setting the neurite velocity to about 10µm.h leading to −1 V = √ 2 Qgc − Qs a + (1−w)√ ηeff (1 + feff) . (35) Qµ ∼ 0.03. We conclude that the main behaviour of the system is captured by the relative magnitude of the three stall forces of the different neurite phases: the microtubules network (Qµ), the entire axoplasm (Qs) and the acto- myosin of the GC (Qgc). B. Parameter estimation The three loads Qµ, Qs and Qgc depend on six non- dimensional parameters directly related to measurable material coefficients. Based on Table I, we have a ∼ 0.1,  ∼ 0.001, w ∼ 0.1, vp ∼ 0.14 and Qc ∼ 0.3. It is more difficult to asses the value of the microtubule network stall force Qµ as the presence of molecular mo- tors may induce contraction [33]. In Ref. [35], it is shown that the growth of microtubules engaging in filopodia can lead to a pushing stress of −90 Pa at the tip. However in agreement with Ref. [33], we assume that the axonal mi- crotubules exert a small pulling stress. Here, we choose C. Comparison with experiments Now that our model has been validated against clas- sical pulling experiments, it is interesting to see how its predictions compare with various pharmacological tests affecting the F-actin and microtubules meshworks. The effects of some classical drug treatments on the model parameters are collected in Table II. Retraction under microtubules depletion Experi- ments have shown that the depolymerisation of micro- tubules with nocodazole stops or even leads to the col- lapse of neurites depending on the concentration [40, 57, 79]. This treatment can be interpreted in our model as an increase of kµ d and thus an increase of Qµ and is qual- itatively captured in Fig. 6. Physically, the depolymerisation of microtubules low- ers the resistance of this growing network to contractile acto-myosin stress. As we show in the inset of Fig. 6 and as experimentally confirmed in Ref. [57], nocodazole in- duced collapse can be rescued by a latrunculin (destroys the acto-myosin cortex) or BDM (inhibits myosin II con- tractile activity) which effectively reduces Qc. Conversely, initiation of motility due to blebbistatin 11 FIG. 6. Perturbation of the microtubules network properties by nocodazole, taxol (inset: latrunculin and BDM reducing Qc): phase diagrams of the whole neurite state using expres- sions of the three driving loads Qµ, Qs and Qgc (default pa- rameters are a ∼ 0.1,  ∼ 0.001, w ∼ 0.1, vp ∼ 0.14, Qc ∼ 0.3 and Qµ ∼ 0.03). Colors online treatment (contractility inhibitor) is abolished if it is fol- lowed by a nocodazole treatment [42]. Additionally, an increase of microtubules polymerisation using epothilone B points towards a very promising therapeutic route to promote in vivo axonal outgrowth after injury of the spinal cord through the inhibitory environment due to the tissue scar [58]. Treatment with high concentration of taxol stabilises microtubules and slows down elonga- tion [80–82]. This can be interpreted in the model as a decrease of  and is also correctly captured as seen in Fig. 6. However, the effect of low concentration of taxol does not block the microtubules dynamic completely [82] but primarily lowers Qµ (by lowering kµ d ), thus leading to an increase of axonal outgrowth [82, 83]. Treatment of the acto-myosin meshwork We now turn to the treatments affecting the acto-myosin meshwork (Fig. 7) which has two antagonistic roles [25, 39]. On the one hand, it is pulling the axoplasm thanks to F-actin front polymerisation (vp) but the contractile acto-myosin cortex is also pulling the neurite backward. Remarkably, treatment with a low concentration of cytochalasin [43] reduces only the front F-actin protrusion (no filopodia) and can be interpreted as lowering vp effectively reducing the neurite velocity. Larger concentrations, on the con- trary, destroy the whole F-actin meshwork which strongly impacts the cortical contractility (Qc) and leads to an in- crease of neurite velocity as captured by the model [40]. More focussed experiments inhibiting contractility with blebbistatin [42, 59] confirm that contractility impair- ment robustly initiates neurite motility. Note again that cytochalasin (vp decrease) abolishes this blebbistatin in- FIG. 7. Perturbations of the acto-myosin meshwork by cy- tochalasin and blebbistatin: phase diagram of the whole neu- rite state using expressions of the three driving loads Qµ, Qs and Qgc (default parameters are a ∼ 0.1,  ∼ 0.001, w ∼ 0.1, vp ∼ 0.14, Qc ∼ 0.3 and Qµ ∼ 0.03). Colors online duced motility [42] as the model also predicts, see Fig. 7. Treatment of the substrate to modify adhesions Adhe- sion of the neurite with the substrate can also be strongly reduced with trypsin, which leads to a collapse or a stall of the neurite [33]. Such treatment can be modelled by lowering a. In Fig. 8 (a), we show that this effect is cor- rectly captured by our model. It is also known that the motility promoting effect of myosin II inhibition is ad- hesiveness dependent [42]. While blebbistatin promotes motility on polylysine substrates, it lowers motility on less adherent laminin substrates [60]. We can speculate that this is due to myosin II being strongly involved in the creation of focal adhesions for laminin substrates [15, 60]. As a result, in this case, a blebbsitatin treatment also considerably lowers adhesion (a) thus potentially leading to arrest (see Fig. 8 (a)). Finally, we also show in Fig. 8 (b) the effect of a cytochalasin treatment depending on the substrate adhesivity. While low level of cyotchalasin reduces the neurite velocity, we expect this effect to be attenuated on more adhesive substrates. V. CONCLUSION Starting from basic conservation laws, we have devel- oped and analysed a one-dimensional mechanical model of neurite motility based on a three-compartment cy- toskeletal structure. The model supports three robust states: Collapse, Static and Motile. Collapse arises when the growth of the microtubles and the GC induced trac- tion cannot overcome the cortical acto-myosin contrac- tility. On the contrary, extension at a finite velocity is provoked by the GC F-actin frontal polymerisation which generates a tension promoting growth of the microtubule network and overcoming cortical contractility. Interest- Appendix A: Existence of solutions to Eq. (21) 12 Eq. (21) can be rewritten in the following form, e Q = ψ(ln) where ψ(ln) = with e[(1+f0)Q0 s−f0 Q]g(u,ln)du, (cid:90) 1 0 (cid:16) uln (cid:17) (cid:17) (cid:16) ln lw lw FIG. 8. Perturbations of the substrate adhesion by trypsin, blebbistatin and cytochalasin: phase diagrams of the whole neurite state using expressions of the three driving loads Qµ, Qs and Qgc (default parameters are a ∼ 0.1,  ∼ 0.001, w ∼ 0.1, vp ∼ 0.14, Qc ∼ 0.3 and Qµ ∼ 0.03). Colors online ingly, between these two states, the neurite can also re- main static as a result of a tensile tightening between the microtubules growing network and the contractile acto- myosin sleeve operating in parallel. The respective position of the three stall forces of the microtubules, the axoplasm and the GC can be used to predict the state of the neurite and we explicitly re- late these loads to measurable material parameters. This framework allows for a number of model predictions in re- markable agreement with experimental drug treatments. It is our hope that the model will be used as a guide- line to design focussed experiments to discriminate the respective role of active (contractility, growth) and pas- sive (elasticity, viscosity, substrate adhesiveness) effects impacting neurite motility and leading to a better under- standing of the neuronal regeneration after a trauma. We did not investigate the shape of neurites which is also known to be an important signature of trauma [18, 84] as neurons swell or bead in response to fast pulling. To deal with this complex problem, a two- dimensional model must be used and the osmotic pres- sure regulation between the inside and the outside of the neurite must be taken into account [84]. More gener- ally, coupling of the cytoskeletal mechanics with the ions trafficking through channels and pumps at the plasmic membrane is an important challenge that will lead to a better insight on neurite guidance by chemical gradients as well as swelling of neurons during injury. authors Acknowledgements The thank Kristian Franze for helpful discussions during the international winter school on the physics of the brain held in Les Houches. P.R. acknowledges funding from OCCAM. P.R. and A.J. also acknowledge funding from the Eu- ropean Research Council under the European Union's Seventh Framework Programme (FP7 2007-2013)/ERC Grant Agreement No. 306587. g(u, ln) = 1 − cosh cosh (cid:113) wa and lw = orem, it follows that 1+a . Using the dominated convergence the- ψ(0) = 1 and ψ(∞) = e[(1+f0)Q0 s−f0 Q]. We also have (1 + f0)Q0 lw cosh = (cid:18) uln (cid:19) lw cosh sinh dψ dln (cid:20) (cid:124) (cid:90) 1 0 (cid:17)2 (cid:16) ln s − f0 Q (cid:18) ln (cid:19) lw (cid:123)(cid:122) − u cosh lw ≥sinh( ln(1−u) lw )≥0 du e[(1+f0)Q0 s−f0 Q]g(u,ln) (cid:18) ln (cid:19) lw sinh (cid:18) uln lw (cid:19)(cid:21) (cid:125) . As a result, if Q < Qs, then ψ(ln) is an increasing func- tion. If we additionally have Q > 0 then e Q is strictly between ψ(0) and ψ(∞) leading to the existence of a sin- gle solution of Eq. (21). Appendix B: Numerical method To solve the Cauchy problem of Eq. (16)-(17), we use the scaled space coordinate to deal with the mov- ing boundary y = x ln(t) , (B1) (cid:18) and denote the new unknown functions v(y, t) = v[ln(t)y, t] and ρ(y, t) = ln(t) ρ[ln(t)y, t]. Eq. (16)2 be- comes ∂y 1 ln ∂t ρ + ρ(v − y ln) − 1 − w 1 − ρ ln (B2) where the velocity field can be expressed through Eq. (16)1 as = ln ∂y ρ ln (cid:18) (cid:19) (cid:19) ∂yy v − av = w l2 n 1 − w ln ∂y ρ ρ . (B3) Accordingly, the boundary conditions of Eq. (17) become v0 = 0 and ∂y v1 = ln(Q − Qc), ln = v1 − (1 − w)∂y ρ ∂y ρ0 = 0 and ρ1 = lneQµ−Q, 1. ln ρ (B4) (B5) (B6) To fully specify the system, we impose the initial condi- tions ln(0) = l0 n and ρ(y, 0) = ρ0(y). Numerically, we did not find that the steady state phase reported in Fig. 5 was sensitive to the choice of initial conditions. The numerical scheme used to solve the Cauchy prob- lem Eq. (B2-B6) is based on the finite volume method [85] 13 which allows to conserve mass while handling localised states without spurious oscillations. Two regularly- spaced grids on the same interval [0, 1], denoted Z and Zd for its dual, are considered in parallel. An initial con- dition on ρ being given on Z, Eq. (B3) is solved using the boundary conditions of Eq. (B4) and the effective drift term v − y ln is computed on Zd using Eq. (B3). We then apply an upwind finite volume scheme to Eq. (B2) using the no-flux boundary conditions of Eq. (B5). This allows the computation of the updated concentra- tion profile ρ on Z which gives in turn the new initial data used for the next time step and the front dynamic through Eq. (B6). The same procedure is then repeated. The time interval for each time step is adapted so that the Courant-Friedrichs-Lewy condition is uniformly sat- isfied on Zd [85]. [1] S. R. Cajal and S. Cajal, Histologie du Syst`eme Nerveux cell biology 8, 216 (2006). de l'Homme et des Vert´ebr´es Maloine (1911). [2] J. Silver and J. H. Miller, Nature Reviews Neuroscience 5, 146 (2004). [3] L. C. Case and M. Tessier-Lavigne, Current biology 15, R749 (2005). [4] R. Deumens, A. Bozkurt, M. F. Meek, M. A. Marcus, E. A. Joosten, J. Weis, and G. A. Brook, Progress in neurobiology 92, 245 (2010). [5] T. Mitchison and M. Kirschner, Neuron 1, 761 (1988). [6] E. W. Dent and F. B. Gertler, Neuron 40, 209 (2003). [7] K. Franze and J. Guck, Reports on Progress in Physics [22] L. Bard, C. Boscher, M. Lambert, R.-M. M`ege, D. Cho- quet, and O. Thoumine, The Journal of Neuroscience 28, 5879 (2008). [23] C. E. Chan and D. J. Odde, Science 322, 1687 (2008). [24] R. B. Vallee, G. E. Seale, and J.-W. Tsai, Trends in cell biology 19, 347 (2009). [25] M. Aeschlimann, Biophysical Models of Axonal Path Finding (2000). [26] G. Kiddie, D. McLean, A. Van Ooyen, and B. Graham, Progress in brain research 147, 67 (2005). [27] D. M. Suter and K. E. Miller, Progress in neurobiology 73, 094601 (2010). 94, 91 (2011). [8] C. H. Coles and F. Bradke, Current Biology 25, R677 [28] C. S. Peskin, G. M. Odell, and G. F. Oster, Biophysical (2015). journal 65, 316 (1993). [9] M. Aeschlimann and L. Tettoni, Neurocomputing 38, 87 [29] A. Mogilner and G. Oster, Biophysical journal 84, 1591 (2001). (2003). [10] B. J. Dickson, Science 298, 1959 (2002). [11] L. A. Lowery and D. Van Vactor, Nature reviews Molec- [30] D. Bray, Journal of cell science 37, 391 (1979). [31] P. Lamoureux, R. E. Buxbaum, and S. R. Heidemann, ular cell biology 10, 332 (2009). Nature 340, 159 (1989). [12] J. Reingruber and D. Holcman, in Seminars in cell & developmental biology, Vol. 35 (Elsevier, 2014) pp. 189– 202. [13] K. Franze, J. Gerdelmann, M. Weick, T. Betz, S. Pawl- izak, M. Lakadamyali, J. Bayer, K. Rillich, M. Gogler, Y.-B. Lu, et al., Biophysical journal 97, 1883 (2009). [32] M. O'Toole, P. Lamoureux, and K. E. Miller, Biophysical journal 94, 2610 (2008). [33] M. O'Toole, P. Lamoureux, and K. E. Miller, Biophysical journal 108, 1027 (2015). [34] M. P. Van Veen and J. Van Pelt, Bulletin of mathematical biology 56, 249 (1994). [14] H. Ouyang, E. Nauman, and R. Shi, J Biol Eng 7, 21 [35] P. Rauch, P. Heine, B. Goettgens, and J. A. Kas, New (2013). [15] J. Brown and P. C. Bridgman, Journal of Histochemistry & Cytochemistry 51, 421 (2003). [16] D. H. Roossien, P. Lamoureux, and K. E. Miller, Journal of cell science 127, 3593 (2014). [17] W. Lu, P. Fox, M. Lakonishok, M. W. Davidson, and Journal of Physics 15, 015007 (2013). [36] J. Garc´ıa, J. Pena, S. Mchugh, and A. J´erusalem, Computer modeling in engineering and sciences 87, 411 (2012). [37] M. Dogterom and B. Yurke, Science 278, 856 (1997). [38] R. Buxbaum and S. Heidemann, Journal of theoretical V. I. Gelfand, Current Biology 23, 1018 (2013). biology 155, 409 (1992). [18] H. Ahmadzadeh, D. H. Smith, and V. B. Shenoy, Bio- [39] P. C. Letourneau, T. A. Shattuck, and A. H. Ressler, physical journal 106, 1123 (2014). Cell motility and the cytoskeleton 8, 193 (1987). [19] F. Julicher, K. Kruse, J. Prost, and J.-F. Joanny, Physics [40] T. Dennerll, H. Joshi, V. Steel, R. Buxbaum, and S. Hei- Reports 449, 3 (2007). demann, The Journal of cell biology 107, 665 (1988). [20] A. B. Verkhovsky, T. M. Svitkina, and G. G. Borisy, [41] K. Kollins, J. Hu, P. Bridgman, Y.-Q. Huang, and Current Biology 9, 11 (1999). G. Gallo, Developmental neurobiology 69, 279 (2009). [21] N. A. Medeiros, D. T. Burnette, and P. Forscher, Nature [42] E.-M. Hur, I. H. Yang, D.-H. Kim, J. Byun, W.-L. Xu, 14 P. R. Nicovich, R. Cheong, A. Levchenko, N. Thakor, F.- Q. Zhou, et al., Proceedings of the National Academy of Sciences 108, 5057 (2011). [64] P. Recho, T. Putelat, and L. Truskinovsky, Physical re- view letters 111, 108102 (2013). [65] K. Tawada and K. Sekimoto, Journal of theoretical biol- [43] J. Q. Zheng, J. Wan, and M. Poo, The Journal of neu- ogy 150, 193 (1991). roscience 16, 1140 (1996). [44] T. J. Dennerll, P. Lamoureux, R. E. Buxbaum, and S. R. Heidemann, The journal of cell biology 109, 3073 (1989). [45] R. Bernal, P. A. Pullarkat, and F. Melo, Physical review letters 99, 018301 (2007). [46] R. Bernal, F. Melo, and P. A. Pullarkat, Biophysical journal 98, 515 (2010). [47] J. Zheng, P. Lamoureux, V. Santiago, T. Dennerll, R. E. Buxbaum, and S. R. Heidemann, The Journal of neuro- science 11, 1117 (1991). [48] A. J´erusalem, J. A. Garc´ıa-Grajales, A. Merch´an- P´erez, and J. M. Pena, Biomechanics and modeling in mechanobiology 13, 883 (2014). [66] K. Kruse, J.-F. Joanny, F. Julicher, J. Prost, and K. Sekimoto, The European Physical Journal E 16, 5 (2005). [67] E. Hannezo, B. Dong, P. Recho, J.-F. Joanny, and S. Hayashi, Proceedings of the National Academy of Sci- ences 112, 8620 (2015). [68] C. Tomba, C. Braini, B. Wu, N. S. Gov, and C. Villard, Soft Matter 10, 2381 (2014). [69] A. Carlsson, New journal of physics 13, 073009 (2011). [70] P. Recho and L. Truskinovsky, Mathematics and Mechan- ics of Solids (2015), 10.1177/1081286515588675. [71] P. Recho, T. Putelat, and L. Truskinovsky, Journal of the Mechanics and Physics of Solids 84, 469 (2015). [49] W. W. Ahmed and T. A. Saif, Scientific reports 4, 4481 [72] K. Xu, G. Zhong, and X. Zhuang, Science 339, 452 (2014). (2013). [50] A. Shamloo, F. Manuchehrfar, and H. Rafii-Tabar, Jour- [73] D. E. Moulton, T. Lessinnes, and A. Goriely, Journal of nal of biomechanics 48, 1241 (2015). the Mechanics and Physics of Solids 61, 398 (2012). [51] T. Betz, D. Koch, Y.-B. Lu, K. Franze, and J. A. Kas, Proceedings of the National Academy of Sciences 108, 13420 (2011). [74] B. Rubinstein, M. F. Fournier, K. Jacobson, A. B. Verkhovsky, and A. Mogilner, Biophysical journal 97, 1853 (2009). [52] T. Betz, D. Lim, and J. A. Kas, Physical review letters [75] P. Recho and L. Truskinovsky, Physical Review E 87, 96, 098103 (2006). 022720 (2013). [53] E. M. Craig, D. Van Goor, P. Forscher, and A. Mogilner, [76] H. Jiang and S. X. Sun, Biophysical journal 105, 609 Biophysical journal 102, 1503 (2012). (2013). [54] P. Lamoureux, G. Ruthel, R. E. Buxbaum, and S. R. Heidemann, The Journal of cell biology 159, 499 (2002). [55] T. D. Nguyen, I. B. Hogue, K. Cung, P. K. Purohit, and [77] T. Hui, Z. Zhou, J. Qian, Y. Lin, A. Ngan, and H. Gao, Physical review letters 113, 118101 (2014). [78] P. Recho, J.-F. Joanny, and L. Truskinovsky, Physical M. C. McAlpine, Lab on a Chip 13, 3735 (2013). Review Letters 112, 218101 (2014). [56] S. R. Heidemann and R. E. Buxbaum, Neurotoxicology [79] W. A. Sayyad, L. Amin, P. Fabris, E. Ercolini, and 15, 95 (1993). [57] F. J. Ahmad, J. Hughey, T. Wittmann, A. Hyman, M. Greaser, and P. W. Baas, Nature cell biology 2, 276 (2000). [58] J. Ruschel, F. Hellal, K. C. Flynn, S. Dupraz, D. A. Elliott, A. Tedeschi, M. Bates, C. Sliwinski, G. Brook, K. Dobrindt, et al., Science 348, 347 (2015). [59] P. Yu, L. Y. Santiago, Y. Katagiri, and H. M. Geller, Journal of neurochemistry 120, 1117 (2012). [60] A. R. Ketschek, S. L. Jones, and G. Gallo, Developmen- tal neurobiology 67, 1305 (2007). [61] M. A. Holland, K. E. Miller, and E. Kuhl, Annals of biomedical engineering 43, 1640 (2015). V. Torre, Scientific reports 5, 7842 (2015). [80] P. C. Letourneau and A. H. Ressler, The Journal of cell biology 98, 1355 (1984). [81] J. Bamburg, D. Bray, and K. Chapman, Nature 321, 788 (1986). [82] H. Witte, D. Neukirchen, and F. Bradke, The Journal of cell biology 180, 619 (2008). [83] F. Hellal, A. Hurtado, J. Ruschel, K. C. Flynn, C. J. Laskowski, M. Umlauf, L. C. Kapitein, D. Strikis, V. Lemmon, J. Bixby, et al., Science 331, 928 (2011). [84] P. A. Pullarkat, P. Dommersnes, P. Fern´andez, J.-F. Joanny, and A. Ott, Physical review letters 96, 048104 (2006). [62] J. Prost, F. Julicher, and J. Joanny, Nature Physics 11, [85] R. LeVeque, Finite volume methods for hyperbolic prob- 111 (2015). lems (Cambridge University Press, Cambridge, 2002). [63] P. C. Bressloff and E. Levien, Physical review letters 114, 168101 (2015).
1603.02995
2
1603
2016-03-25T17:22:07
Transient superdiffusion and long-range correlations in the motility patterns of trypanosomatid flagellate protozoa
[ "physics.bio-ph", "q-bio.QM" ]
We report on a diffusive analysis of the motion of flagellate protozoa species. These parasites are the etiological agents of neglected tropical diseases: leishmaniasis caused by Leishmania amazonensis and Leishmania braziliensis, African sleeping sickness caused by Trypanosoma brucei, and Chagas disease caused by Trypanosoma cruzi. By tracking the positions of these parasites and evaluating the variance related to the radial positions, we find that their motions are characterized by a short-time transient superdiffusive behavior. Also, the probability distributions of the radial positions are self-similar and can be approximated by a stretched Gaussian distribution. We further investigate the probability distributions of the radial velocities of individual trajectories. Among several candidates, we find that the generalized gamma distribution shows a good agreement with these distributions. The velocity time series have long-range correlations, displaying a strong persistent behavior (Hurst exponents close to one). The prevalence of "universal" patterns across all analyzed species indicates that similar mechanisms may be ruling the motion of these parasites, despite their differences in morphological traits. In addition, further analysis of these patterns could become a useful tool for investigating the activity of new candidate drugs against these and others neglected tropical diseases.
physics.bio-ph
physics
Transient superdiffusion and long-range correlations in the motility patterns of trypanosomatid flagellate protozoa Luiz G. A. Alves1,2,3,*, D´ebora B. Scariot4, Renato R. Guimaraes1,3, Celso V. Nakamura4, Renio S. Mendes1,3, Haroldo V. Ribeiro1 1 Departamento de F´ısica, Universidade Estadual de Maring´a, Maring´a, PR, 87020-900, Brazil 2 Department of Chemical and Biological Engineering, Northwestern University, Evanston, IL, 60208, United States of America 3 National Institute of Science and Technology for Complex Systems, CNPq, Rio de Janeiro, RJ, 22290-180, Brazil 4 Departamento de Ciencias B´asicas da Sa´ude, Universidade Estadual de Maring´a, Maring´a, PR, 87020-900, Brazil * [email protected] Abstract We report on a diffusive analysis of the motion of flagellate protozoa species. These parasites are the etiological agents of neglected tropical diseases: leishmaniasis caused by Leishmania amazonensis and Leishmania braziliensis, African sleeping sickness caused by Trypanosoma brucei, and Chagas disease caused by Trypanosoma cruzi. By tracking the positions of these parasites and evaluating the variance related to the radial positions, we find that their motions are characterized by a short-time transient superdiffusive behavior. Also, the probability distributions of the radial positions are self-similar and can be approximated by a stretched Gaussian distribution. We further investigate the probability distributions of the radial velocities of individual trajectories. Among several candidates, we find that the generalized gamma distribution shows a good agreement with these distributions. The velocity time series have long-range correlations, displaying a strong persistent behavior (Hurst exponents close to one). The prevalence of "universal" patterns across all analyzed species indicates that similar mechanisms may be ruling the motion of these parasites, despite their differences in morphological traits. In addition, further analysis of these patterns could become a useful tool for investigating the activity of new candidate drugs against these and others neglected tropical diseases. Introduction Kinetoplastids protists are responsible for numerous diseases in humans and animals. Many of these protozoa are the etiological agents of neglected tropical diseases [1]. These diseases affect the lives of approximately one billion people around the world [2] and are considered a serious public health problem in several countries. The main regions affected are developing countries located in tropical areas, where the parasites have appropriate natural conditions for life-cycle and insects vectors are abound [3]. Leishmaniasis, Chagas disease, and African sleeping sickness are examples of neglected PLOS 1/17 tropical diseases caused by kinetoplastids parasites -- Leishmania spp, Trypanosoma cruzi, and Trypanosoma brucei, respectively [4]. These protozoa have a complex life-cycle, alternating between invertebrate (vector) and vertebrate hosts. These parasites have a flagellum at least during one of the evolutionary forms of their life-cycle [5 -- 7]. The flagellum is a multifunctional organelle, responsible for cell propulsion and associated with control cell morphogenesis, chemotaxis, and cytokinesis process during last stage of the cell division cycle [8]. The parasites motility is a key to host-cell attachment invasion and colonization of host tissues [8 -- 13]. Diffusive motion is ubiquitous in nature and plays a fundamental role in the motility of swimming microorganisms [14 -- 20]. Researchers in statistical mechanics have focused on phenomena where anomalous diffusion are present [19]. Also, the field of non-extensive statistical mechanics has made an advance in the understanding of several systems where Boltzmann thermodynamic fails to explain the results [21 -- 23]. For instance, anomalous diffusion and non-Gaussian velocity distribution have been observed in the context of Hydra cells [23], where maximum entropy densities associated with non-standard entropic measures were used to describe the motion of these cells. These densities are related to nonlinear diffusive process such as the generalized Fokker-Planck equation proposed in the context of the non-extensive statistical mechanics [22]. Motility and diffusive patterns have also been investigated in protozoa [9, 10, 12, 24 -- 28]. For instance, T. brucei studies have focused on the movement of propulsion [9], flexibility and directionality [24], and body adaptations to the environment [25]. Cell-host interaction [12] for Leishmania spp and the flagellar beating [26] for T. cruzi have also been studied. Motility is strongly related to cell viability in all flagellate kinetoplastid species and it is widely used as a proxy measurement for viability [10, 27, 28]. Each protozoan specie has unique adaptations depending on the different living conditions. The investigation of the dynamics, diffusion and motion behavior of these microorganisms is an advance in the understanding of the microbial pathogenesis mechanism and in the field of diffusive patterns. However, a more complete and general understanding of motility patterns of these parasites is still lacking. To overcome this gap, we study the diffusive dynamics of causative agents of the neglected tropical diseases. By tracking the positions of these parasites, we present a complete characterization of their motility patterns. Specifically, we show that the spread of the trajectories is superdiffusive for short-times and that probability distributions related to the radial positions differs from the predictions of the usual diffusion equation. We further verify that the velocity time series of individual trajectories have long-range correlations and are well approximated by a generalized gamma distribution. Our results reveal some "universal" parasite motility patterns that could facilitate the identification of novel targets for therapeutic intervention. Furthermore, it could be expanded to screen aspects of cell viability. Materials and methods Parasites maintenance Leishmania amazonensis (MHOM/BR/75/Josefa strain) and Leishmania braziliensis promastigote forms were cultivated inside cell culture flasks containing Warren's medium (brain heart infusion plus bovine hemin and folic acid, pH 7.2) supplemented with 10% of fetal bovine serum (FBS). The parasites were incubated at 25◦C for 48 h. Trypanosoma cruzi epimastigote forms (Y strain) were maintained in LIT medium (Liver Infusion Tryptose, pH 7.2), supplemented with 10% of FBS and incubated at 28◦C during 96 h. Trypomastigote forms of Trypanosoma brucei brucei (EATRO-427 PLOS 2/17 strain) were cultivated in HMI-9 medium, supplemented with 10% of FBS, incubated at 37◦C and 5% CO2 tension for 24 h. These incubation periods are essential to harvest the protozoa in the exponential growth phase. Experimental setup, image acquisition and tracking After the incubation periods, we have prepared a suspension containing about 6 × 106 parasites for Leishmania species, T. cruzi, and T. brucei in Warren, LIT, and DMEM (Dulbecco's Modified Eagle Medium, supplemented with 2 mM L-glutamine, pH 7.4), respectively. The mediums were not supplemented with FBS. Next, 10 µL of the protozoa suspensions were placed between a glass slide and coverslip to start the image acquisition. The thickness between the glass slide and coverslip is comparable to the size of the protozoa (∼ 5 µm) [29 -- 31], reducing it to a two-dimensional problem. We used the Motic's BA410E microscope equipped with a 5.0 Megapixel CMOS camera at a resolution of 800 × 600 pixels, acquisition rate of 10 frames/second and magnification of 20×. The area covered by the microscope at this configuration is 285.12 × 213.84 µm2. The length of the acquired videos was 10 minutes for each sample. We repeated this procedure three times for each protozoan. In order to extract the trajectories from the videos, we used a Matlab algorithm of motion tracking in image sequences [32]. After, we excluded the trajectories with less than 500 time-steps to ensure that we have long enough trajectories for statistical analysis. For these trajectories, we removed the first and last 50 steps due to imprecision of the algorithm in track the protozoa positions. By visual inspection, we further removed the trajectories for which the algorithm mistook two or more microorganisms at some step along the path. All trajectories were smoothed by applying a moving average filter of length 10 and are available in S1 Dataset. The viscosity of the culture medium was measured in a viscometer (Visco Star Plus) at 25 ◦C and 50 rpm. The viscosity values and the number of trajectories analyzed for each protozoan are shown in Table 1. Table 1. Dataset summary Protozoan L. amazonensis L. braziliensis T. brucei T. cruzi # Tracks Medium Viscosity [mPa/s] 105 135 131 153 Warren Warren DMEM LIT 0.38 0.38 0.71 0.39 In Fig. 1A, we show typical trajectories of the L. amazonensis, that is, (cid:113) v2 xi (t), where (t) + v2 yi (cid:126)ri(t) = [xi(t), yi(t)], where xi(t) and yi(t) are the horizontal and vertical components of the position vector (cid:126)ri(t) in the time t for the i-th track. In Fig. 1B, we plot the corresponding radial velocity time series, vi(t) = vxi(t) = [xi(t + ∆t) − xi(t)]/∆t and vyi(t) = [yi(t + ∆t) − yi(t)]/∆t with ∆t = 1/10 s. Examples of trajectories and velocity time series of the other protozoa are shown in S1 Fig. From these figures, we can observe circular patterns that could be related to possible hydrodynamic interactions with the walls, as previously observed in the motion of the bacteria Escherichia coli [33]. Additional experiments considering different boundaries and thickness between the glass slide and the coverslip could further evaluate the effects of the walls on the motility of these protozoa. We have also calculated the average (Fig. 1C) and standard deviation (Fig. 1D) of the velocities for each protozoan specie. It is worth noting that the velocities of the swimming microorganisms depend on the viscosity of the culture medium [25] and other values can be found in the literature due to different viscosity [9, 24 -- 26]. For instance, it was found that the wildtype bloodstream form of the T. brucei can reach much higher velocities in the blood PLOS 3/17 (around 30 µm/s) [25]. Trypanosoma spp use the motility as a tool to evade the immune cells and remove the bind between their surface and antibodies molecules. In the blood vessels, where the protozoa is adapted to survive, red and white blood cells behave as support for the flagellum to propel the cell body. The same behavior is observed in more viscous liquids, justifying the increase in speed of the parasite in these environments [25]. Other factors that affect the velocities include chemical cues, oxygen content, pressure, flow and confinement [25]. Our results about the motility of Leishmania spp promastigote and T. cruzi epimastigotes suggest a very similar behavior profile. Further details about these similarities are given in the results and discussion section. Figure 1. Illustration of trajectories and velocities. A) Typical swimming trajectories of the L. amazonensis and B) the corresponding time series of the radial velocities v(t) are shown. See Fig. S1 Fig for other protozoa. C) Average velocities and D) standard deviation over all trajectories for each protozoan are represented in the bar plots. The error bars are 95% confidence intervals calculated via bootstrapping. Results and Discussion We start by characterizing the spreading of the trajectories of all protozoa. In order to do so, we have considered the time series of the magnitude of the position vector (cid:126)ri(t) after subtracting its initial position (cid:126)ri(0) = [xi(0), yi(0)]. We thus measure the spreading by evaluating the time dependence of the variance of the centered radial position σ2(t) = (cid:104)[(cid:126)ri(t) − (cid:104)(cid:126)ri(t)(cid:105)]2(cid:105) Nk(t)(cid:88) i=1 (cid:80)Nk(t) = 1 Nk(t) − 1 ((cid:126)ri(t) − (cid:104)(cid:126)ri(t)(cid:105))2 , (1) Nk(t) where (cid:104)(cid:126)ri(t)(cid:105) = 1 i=1 (cid:126)ri(t) is the average radial position and Nk(t) is the number of available trajectories for the k-th specie longer than t steps (see Table 1). For usual diffusive processes the variance is expected to increase linearly on time, that is, σ2(t) ∼ t. This usual behavior (or the Brownian motion) is related to absence of memory along the particle trajectory as well as indicates a finite characteristic scale for the position increments and for the waiting times between flights. However, more complex diffusive processes often display deviations of this linear behavior. When this PLOS 4/17 y [µm] x [µm]050100150200050100150200250v [µm/s]t [s]0.00.51.01.5010203040ABCD0.01.02.03.04.0Standard deviation [µm/s]L. amazonensis L. braziliensis T. bruceiT. cruzi 1.02.03.04.0Average velocity [µm/s]L. amazonensis L. braziliensis T. bruceiT. cruzi L. amazonensis L. amazonensis happens, a typical behavior for the variance is a power-law dependence on time [34], σ2(t) ∼ tλ , (2) where 0 < λ < 1 corresponds to subdiffusion and λ > 1 to superdiffusion. In our case, the evolution of the variances are shown in Fig. 2A, where is evident that the spreading of the protozoan trajectories occurs much faster than the expected by a Brownian motion. We further observe an approximate power-law behavior over two decades of the temporal scale (t < 10 seconds). By least square fitting the log-log relationships (log σ2(t) versus log t), we find that values of λ are actually much larger than one. As shown in Fig. 2B, λ ranges from 1.69 for the L. braziliensis to 1.93 for the T. brucei ; therefore, the four flagellate protozoa studied here display a strongly superdiffusive behavior for short-times. A similar exponent was observed for the motion of intracellular particles with λ ≈ 1.8 [35]. It is worth note that, biological swimmers can present some persistence related to their drive mechanism for short-times. We expect that the diffusion may eventually approach the usual regime for longer trajectories. Thus, the values of λ obtained here represent a initial short-time behavior, and other regimes could be observed for longer trajectories. In fact, we note from Fig. 2A that the curves start to bend downwards for t > 10s. In Fig. 2C, we show the exponent λ calculated within a window of size 30 seconds centralized in tw as a function of tw, where we note that the values of λ decrease with tw and approach the value expected by the usual diffusion. Although all studied species showed superdiffusive spreading for short-times, we noticed a intriguing aspect of movement related to velocities and diffusion analysis for T. brucei. This protozoan shows the smallest velocity and the greatest diffusion exponent in culture medium, suggesting that the T. brucei motility is more directional than the other protozoa species considered in this work. Directional motility probably occurs because T. brucei is a free extracellular parasite and spreads across several tissues. Trypomastigote forms of T. brucei are highly adapted to live in intercellular spaces [36 -- 38]. During spreading in the host T. brucei parasites penetrate between cells in the tissue where there are collagen fibers that can facilitate or hinder motion. For trypanosomes to reach the maximum forward velocity a specific density of cells is required. If the density of obstacles resembles collagen networks the protozoa swim backwards in order to avoid getting trapped [25]. These parasites have an entire antigen surface called variant surface glycoprotein (VSG) [39]. The VSG is different among individuals in a population of T. brucei and prevents specific binding with antibodies that could kill the parasites [39]. We hypothesize that the presence of surface molecules such as VSG's, which reduces the difficulty of the cell motion in its surroundings media, allows T. brucei parasite to diffuse faster. Epimastigote forms of T. cruzi and promastigote forms of Leishmania spp exhibit similar velocities and diffusion exponent. These forms are faster than the T. brucei trypomastigote forms, have smaller diffusion exponent and have different strategies in this stage of the life-cycle. In the natural environment, T. cruzi epimastigote and Leishmania spp promastigote do not need to travel long distances. Specifically, epimastigote forms of T. cruzi must migrate to the midgut of the insect vector during a specific period of the life-cycle for proliferation [40]. In the case of promastigote forms of Leishmania spp, they must be phagocytized by mammalian cells to complete their life-cycle. At this stage, the Leishmania parasites secrete a substance called promastigote secretory gel (PSG). The PSG is a mucin-like gel and has been shown to be an important factor for the amastigote growth in the intracellular environment [41, 42]. PSG production seems to be responsible for the formation of agglomerates (rosettes) and could play a role as constraint for the diffusive motion [43, 44]. Overall, we suggest that changes in motility parameters and surface PLOS 5/17 Figure 2. The transient superdiffusive spreading of the protozoan trajectories. A) Experimental values of the variance of the radial positions σ2(t) (red dots) for all species studied here (as indicated in the plots) are shown in log-log scale. The dashed lines represent power-law relationships σ2(t) ∼ tλ, where the values of λ were obtained by least square fitting a linear model to these log-log relationships (considering t < 10 seconds). The values of λ (and their 95% bootstrap confidence intervals) are shown in plots, and in panel B) they are presented in bar plots, where error bars correspond to the 95% bootstrap confidence intervals. C) Exponents λ calculated within a window of size 30 seconds centralized in tw as a function of tw. The different colors represent the four protozoa according to the ones used in Fig. 2B. The small shaded regions represent 95% bootstrap confidence intervals and the dashed gray line represents the usual regime (λ = 1). molecules that affect mechanical or physical constraints restricting the ability of cells to freely diffuse could significantly contribute to the virulence of these parasites. Another striking feature of diffusive processes is related to the probability distributions of the positions. For the usual diffusion with radial symmetry, this distribution can be obtained by solving the following differential equation where D is the diffusion coefficient and ∇2 is the two-dimensional Laplacian. By considering P (r → ∞, t) → 0 and P ((cid:126)r, 0) = δ2((cid:126)r ), we can show that probability distribution of the radial position is ∂P ((cid:126)r, t) = D∇2P ((cid:126)r, t) ∂t (cid:18) −r2 (cid:19) 4 D t . P (r, t) = r 2 D t exp It is worth noting that this solution leads to a linear behavior for the variance over time, that is, σ2(t) = 4Dt. Furthermore, the distributions given by the Eq. 4 are self-similar in time and collapse into a single curve, P (ξ) = 2 ξ exp(cid:0)−ξ2(cid:1) , PLOS (3) (4) (5) 6/17 A1.41.61.82.0λT. cruziT. bruceiL. braziliensisL. amazonensisBCL. braziliensisT. bruceiT. cruziλ=1.933 [1.927,1.938]L. amazonensisλ=1.820 [1.814, 1.826]λ=1.69 [1.67, 1.71]λ=1.73 [1.71, 1.75]Variance,σ2(t)Time, t [s]10−1310−1210−1110−1010−910−810−710−1100101102Variance,σ2(t)Time, t [s]10−1210−1110−1010−910−810−710−1100101102Variance,σ2(t)Time, t [s]10−1310−1210−1110−1010−910−810−710−1100101102Variance,σ2(t)Time, t [s]10−1310−1210−1110−1010−910−810−710−1100101102λtw[s]0.81.01.21.41.61.82.0151821242730 for the rescaled position ξ(t) = r(t)/σ(t). The usual diffusion equation (Eq. 3) and its solution (Eq. 4) can be understood as a null model for the distributions of the radial positions of the protozoa if we assume that they behave as Brownian particles, and deviations from this prediction is another indication of anomalous diffusion. We have calculated the time evolution of the empirical distributions of the radial positions. As shown in Fig. 3A, we note that these distributions shift toward positive values of r while also become broaden over time t. After considering the rescaled position ξ(t) = r(t)/σ(t), we observe that all distributions collapse into a single curve (Fig. 3B). This result demonstrates the empirical self-similar nature of the protozoa trajectories, but also reveals remarkable deviations between the empirical distributions and the expected by the usual diffusion equation (gray dashed lines in Fig. 3B). In an attempt to find a better description for these empirical distributions, we propose to replace the Gaussian term in Eq. 4 by a stretched Gaussian (characterized by another parameter δ > 0). Specifically, we have considered the following probability distribution for the radial positions P (r, t) = δ r Γ(2/δ) (2 D t)2/δ exp , (6) (cid:18)−rδ (cid:19) 2 D t where Γ(x) =(cid:82) ∞ 0 yx−1e−ydy is the gamma function. We observe that δ = 2 recovers the usual Gaussian term; however, for 0 < δ < 2 the tail of this distribution goes to zero slower than the usual case, whereas for δ > 2 it decays faster. A similar generalization was proposed by Richardson [45] in the context of atmospheric diffusion by considering a spatial-dependent diffusion coefficient, and is considered one of the first anomalous diffusion equation. The distribution given by the Eq. 6 is also self-similar and for the rescaled radial position ξ(t) = r(t)/σ(t) it can be written as P (ξ) = δ Γ (2/δ) ξ e−ξδ . (7) We have thus adjusted Eq. 7 to the window average values of the empirical distributions via least squares fitting. The continuous lines in Fig. 3B show that agreement is far from being perfect, but this generalization is a better description when compared with the distribution emerging from the usual diffusion equation (Eq. 5). Therefore, the diffusive motion of the protozoa studied here displays simultaneously an anomalous (enhanced) scaling of variance σ2(t) as well as radial position distributions with much longer tails than the expected by Brownian swimmers. The values of the best fitting parameters δ are shown in Fig. 3C, where we observe that the L. amazonensis and the T. brucei display larger tails (δ ≈ 1) when compared with the L. braziliensis and the T. cruzi (δ ≈ 1.25). It is also worth noting that the values of δ should be related to values of λ. In fact, the time dependence of the variance evaluated from Eq. 6 is σ2(t) ∝ t2/δ, leading to δ = 2/λ, a relationship that is roughly valid for the values δ obtained via the fitting procedure (Fig. 3C). In order to further characterize the diffusive motion of the protozoa, we investigate the radial velocity time series vi(t) related to the individual trajectories (Fig. 1B). We first ask whether these time series have short or long-range correlations. To answer this question, we have applied the detrended fluctuation analysis (DFA) [46, 47] to these time series. This technique consists of four steps: i) first, we define the integrated vi(t) with tmax being the profile Y (t) =(cid:80)t k=1[vi(k) − ¯vi], where ¯vi = 1/tmax length of the i-th time series; ii) next, we split Y (t) into Nn = tmax/n non-overlapping segments of size n; iii) for each segment, a local polynomial trend (here we have used a linear function, but higher orders do not change our results) is calculated and subtracted from Y (t), defining Yn(t) = Y (t) − pν(t), where pν(t) represents the local trend in the ν-th segment; iv) finally, we calculate the root-mean-square fluctuation (cid:80)tmax t PLOS 7/17 Figure 3. The self-similar nature of the protozoan positions. A) Time evolution of the cumulative distribution functions (CDF) of the radial position r(t) for each protozoan. The color code indicates the time t, ranging from blue (t = 2s) to red (t = 10s). B) The same distributions after considering the rescaled position ξ(t) = r(t)/σ(t). The black dots represent window averages of the CDFs and the error bars stand for one standard deviation. The gray dashed lines represent the distribution given by the normalized Gaussian distribution (Eq. 5). The continuous lines are the cumulative version of the distribution of Eq. 7 and the best values for the fitting parameter δ are shown in the plots. C) Comparison between the values of δ obtained via least squares fitting the window averages of the CDFs and the ones expected by the time dependence of the variances, that is, δ = 2/λ. PLOS 8/17 δ = 1.03 [1.01, 1.05]Bδ = 1.27 [1.23, 1.30]δ = 1.24 [1.18, 1.30]δ = 0.99 [0.96, 1.02]L. amazonensisT. bruceiT. cruziL. braziliensisCT. bruceiT. cruziL. amazonensisL. braziliensisL. amazonensisT. bruceiT. cruziL. braziliensisA0.00.20.40.60.81.00306090120CDFPosition, r0.00.20.40.60.81.0020406080CDFPosition, r0.00.20.40.60.81.00306090120150CDFPosition, r0.00.20.40.60.81.00306090120CDFPosition, rCDFξ0.00.20.40.60.81.00.01.02.03.04.0CDFξ0.00.20.40.60.81.00.01.02.03.04.0CDFξ0.00.20.40.60.81.00.01.02.03.04.0CDFξ0.00.20.40.60.81.00.01.02.03.04.00.00.51.01.5δ, 2/λ (cid:80)Nn ν=1(cid:104)Yn(t)2(cid:105)ν]1/2 , where (cid:104)Yn(t)2(cid:105)ν is the mean square value of function F (n) = [ 1 Nn Yn(t) over the data in the ν-th segment. If the velocity time series is self-similar, the fluctuation function F (n) presents a power-law dependence on the time scale n, that is, F (n) ∼ nh, where h is the Hurst exponent. If h = 1/2 the velocities are either uncorrelated or short-range correlated, whereas h (cid:54)= 1/2 indicates that the time series is longe-range correlated. We have applied the above procedure to all velocity time series and a typical behavior for the fluctuation function F (n) is depicted in Fig. 4A. In this log-log plot, we adjust a linear model for obtaining the Hurst exponent h, which is h = 1.12 for the original time series and is close to 1/2 for random shuffled versions of the time series. We calcule the Hurst exponent for all time series and the average values for each protozoan is shown in Fig. 4B. These averages are practically indistinguishable from each other and indicate that the velocities of the protozoa have long-range correlations. Furthermore, these velocities display a strong persistent behavior (since h ≈ 1), that is, positive increments in the velocities are followed by positive increments and negative increments are followed by negative increments much more frequently than by chance. We have also evaluated the distributions of h for each protozoan specie for all individuals (Fig. 4C), where we note that these distributions peak around h ≈ 1 and that they are quite overlapped. Figure 4. Long-range and persistent correlations in the protozoan velocities. A) Detrended fluctuation analysis (DFA) of the velocity time series v(t). A typical behavior for the fluctuation function F (n) versus the scale n (in a log-log plot) for a velocity time series of the protozoan T. brucei (See S2 Fig for other examples). The red dots are the values of F (n) obtained for the original time series, while the gray ones were obtained for a random shuffled version of this time series. The black dashed line is a least squares fit to the relationship log10[F (n)] versus log10[n] for the original time series and the gray continuous line is the same for the random shuffled version. The value of Hurst exponent is h = 1.12 for the original data and h ≈ 0.5 for the shuffled data (a similar behavior is observed for all time series). B) Average values of the Hurst exponent over all trajectories for each protozoan. The error bars are 95% confidence intervals calculated via bootstrapping. C) Probability distribution function (PDF) of the Hurst exponents h for each protozoan obtained via kernel density estimation method. Another intriguing question is whether the velocity distributions of the protozoa exhibit a particular functional form (Fig. 5A). In this case, the two-dimensional Maxwell-Boltzmann distribution (or Rayleigh distribution) P (v) = e−v2/T 2v T (v > 0), (8) is a natural null model for the protozoan velocity (T is a parameter). This distribution PLOS 9/17 h=1.12 [1.10, 1.14] h=0.53 [0.52, 0.55] BPDFHurst, h01230.30.60.91.21.51.8CT. bruceiA10Log (n)Log (F(n))101.001.051.101.151.201.25Hurst, hT. bruceiT. cruziL. amazonensisL. braziliensis−2.0−1.00.01.00.51.01.52.02.53.0 represents the velocity of two-dimensional gas particles in thermodynamic equilibrium (at a temperature T ) and also emerges when evaluating the distribution of the magnitude of velocity vectors whose the components are uncorrelated and normally distributed (with zero mean) [48]. This functional form has been found to describe quite well the velocities of humans in a very peculiar situation (a mosh pit) [49] and it has also been used in the attempt of modeling the velocity distributions of Hydra cells in a two-dimensional setup [23]. Deviations from this model give clues about weather the velocities are correlated or not. In our case, we have tested the (two-dimensional) Maxwell-Boltzmann hypothesis by adjusting this distribution for each velocity time series and verifying the goodness of the fit via Kolmogorov-Smirnov test. This hypothesis was rejected for almost all velocity time series (≈ 99%), a result that somehow agrees with the more complex behavior observed in the correlation analysis of these time series. Aiming to find a better description for these empirical distributions, we have considered the generalized gamma distribution P (v) = γ β Γ(α) (v/β)α γ−1 e−(v/β)γ , (v > 0) (9) where α and γ are the shape parameters and β is a scale parameter. Despite of being an ad hoc generalization, it is worth noting that this distribution recovers the two-dimensional Maxwell-Boltzmann (for α = 1, γ = 2 and β2 = T ), and it has been employed by several authors as a wind speed model [50 -- 53]. For our data, the Kolmogorov-Smirnov test cannot reject the generalized gamma hypothesis in about 50% of trajectories (see S3 Fig), an improved description when compared with the two-dimensional Maxwell-Boltzmann distribution. Furthermore, we have tested for the usual gamma (Eq. 9 with γ = 1), log-normal, Weibull and q-exponential distributions, finding that they do not outperform the generalized gamma description. In Fig. 5B, we show the average values of the best fitting parameters α, γ and β. These values are practically indistinguishable among the four protozoa. An exception occurs for the T. brucei, which is characterized by a significantly smaller value of β. This fact is a direct consequence of the small standard deviation observed for the velocities of T. brucei (see Fig. 1D), since the standard deviation of v calculated from Eq. 9 is proportional to β. Conclusions We presented a description of the motility patterns of four trypanosomatid flagellate protozoa. By analyzing the time evolution of the positions of these protozoa, we identified that the spreading of their trajectories are strongly superdiffusive for short-times and characterized by self-similar probability distributions with longer tails than the expected by the usual Brownian motion. We also investigated the velocities of these protozoa, finding out that they have long-range correlations and present a strong persistent behavior. We further observed that the velocity distributions cannot be described by a two-dimensional Maxwell-Boltzmann distribution (the natural candidate for a random process). Instead, a generalized gamma distribution showed to be a better description. Thus, our results show that the motility patterns of these protozoa are anomalous in several ways and also reveal that these four protozoa exhibit similar behaviors, despite their morphological differences. Deviations from the behaviors reported here can be employed as an indicator of drug activity in drug tests. Authors contribution Conceived and designed the experiments: DBS, LGAA and RSM. Performed the experiments: DBS and LGAA. Analyzed the data: HVR, LGAA, RRG and RSM. PLOS 10/17 Figure 5. The velocities are not described by a two-dimensional Maxwell-Boltzmann distribution. A) Typical examples of cumulative distribution functions (CDF) of the velocities v(t) from a single trajectory of the protozoa. The gray continuous lines are the best fits obtained for the two-dimensional Maxwell-Boltzmann distribution (Eq. 8), whereas the dashed black lines are the fits for the generalized gamma distribution (Eq. 9). The values of parameters α, β and γ were obtained via maximum-likelihood method and are shown in the plots. We also present the p-values of the Kolmogorov-Smirnov test showing that we cannot reject the gamma hypothesis for these particular trajectories. We tested all trajectories and the KS test cannot reject this hypothesis for about 50% of the trajectories (all p-values are shown in S3 Fig). B) Average values of the best fitting parameters over all trajectories. The error bars are 95% confidence intervals calculated via bootstrapping. Contributed reagents/materials/analysis tools: CVN, DBS, HVR, LGAA, RRG and RSM. Wrote the paper: CVN, DBS, HVR, LGAA, RRG and RSM. Prepared the figures: LGAA. Funding This work has been supported by the agencies Conselho Nacional de Desenvolvimento Cient´ıfico e Tecnol´ogico (CNPq), Coordena¸cao de Aperfei¸coamento de Pessoal de N´ıvel Superior (CAPES), and Funda¸cao Arauc´aria (FA). HVR thanks the financial support of CNPq (grant 440650/2014-3), LGAA thanks for the financial support of CAPES (grant 99999.006842/2015-01) and RSM thanks FA (grant 263/2014) for partial financial support. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. PLOS 11/17 AT. bruceip-value = 0.96 CDFVelocity, v [μm/s]α=0.38β=0.34γ=2.41L. amazonensisp-value = 0.77 CDFVelocity, v [μm/s]α=0.28β=0.61γ=3.84L. braziliensisp-value = 0.17 CDFVelocity, v [μm/s]α=0.33β=0.32γ=5.78T. cruzip-value = 0.18 CDFVelocity, v [μm/s]α=0.19β=0.55γ=5.050.01.53.04.56.07.5γ0.00.30.60.9β0.00.51.01.52.02.5αT. bruceiT. cruziL. amazonensisL. braziliensisB10−11000.00.10.20.30.40.50.610−11000.00.10.20.30.40.510−11000.00.10.20.30.410−310−210−11000.00.20.40.60.81.0 Supporting Information S1 Dataset Dataset containing the trajectories employed in this study. The xls file contains one sheet for each protozoan specie. In each sheet, the columns represent the trajectories and the lines represent the time (in 1/10 seconds scale). The cell value is the position {x, y} for a given trajectory and time. S1 Fig. Figure S1. Trajectories and velocities. The left panel shows typical trajectories of the four protozoa. The right panel shows the corresponding velocity time series v(t). PLOS 12/17 v [µm/s]t [s]0.00.51.01.5y [µm]x [µm]050100150200050100150200250y [µm]x [µm]050100150200050100150200250v [µm/s]t [s]0.00.20.40.6010203040v [µm/s]t [s]0.00.40.81.2y [µm]x [µm]050100150200050100150200250v [µm/s]t [s]0.00.51.01.5010203040L. braziliensisT. bruceiT. cruziL. amazonensisL. braziliensis T. bruceiT. cruziL. amazonensisy [µm]x [µm]050100150200050100150200250010203040010203040 S2 Fig. Figure S2. Fluctuation function. Another typical examples of detrended fluctuation analysis (DFA) for the velocity time series. S3 Fig. Figure S3. Kolmogorov-Smirnov test. The red lines show the p-values of Kolmogorov-Smirnov test of the generalized gamma hypothesis for all velocities time series. About 50% of the velocity time series have p > 0.05 (horizontal line). PLOS 13/17 L. braziliensisT. bruceiT. cruzi−2.0−1.5−1.0−0.50.00.51.01.52.02.5−2.0−1.00.01.00.51.01.52.02.53.0−1.5−1.0−0.50.00.51.01.52.02.5h=1.12 L. amazonensish=1.15 h=1.04 h=1.06 −1.5−1.0−0.50.01.01.52.0Log (n)Log (F(n))Log (n)Log (F(n))Log (n)Log (F(n))1010Log (n)Log (F(n))101010101010p−valueTrack #0.00.20.40.60.81.00255075100L. amazonensis51% p-value > 0.05 p−valueTrack #0.00.20.40.60.81.00255075100125L. braziliensis49% p-value > 0.05 p−valueTrack #0.00.20.40.60.81.00255075100125p−valueTrack #0.00.20.40.60.81.00255075100125150T. bruceiT. cruzi58% p-value > 0.05 56% p-value > 0.05 Acknowledgments We thank Peter B. Winter, Meagan Bechel, and Nicol´as Pel´aez for helping with the manuscript revision. References 1. Nussbaum K, Honek J, Cadmus CM, Efferth T. Trypanosomatid parasites causing neglected diseases. Curr Med Chem. 2010; 17: 1594. doi: 10.2174/092986710790979953 2. Manderson L, Aagaard-Hansen J, Allotey P, Gyapong M, Sommerfeld J. Social research on neglected diseases of poverty: continuing and emerging themes. PLoS Negl Trop Dis. 2009; 3: e332. doi: 10.1371/journal.pntd.0000332 3. Anonymous. Parasites - leishmaniasis. Epidemiology & risk Factors. Centers for disease control and prevention. 2015; Available: http://www.cdc.gov/parasites/leishmaniasis/epi.html. 4. Anonymous. Neglected tropical diseases. World Health Organization. 2015; Available: http://www.who.int/neglected_diseases/diseases/en/. 5. Baron S. Medical Microbiology 4th ed. In Milton RJ, Kim K-S editors. Galveston: University of Texas Medical Branch at Galveston; 1996. pp. 1-18. 6. Kennedy PG. Human african trypanosomiasis of the cns: current issues and challenges. J Clin Invest. 2004; 113: 496. doi: 10.1172/JCI200421052 7. Grab DJ, Kennedy PG. Traversal of human and animal trypanosomes across the blood-brain barrier. J Neurovirol. 2008; 14: 344. doi: 10.1080/13550280802282934 8. Langousis G, Hill KL. Motility and more: the flagellum of Trypanosoma brucei. Nat Rev Microbiol. 2014; 12: 505. doi: 10.1038/nrmicro3274 9. Rodr´ıguez JA, Lopez MA, Thayer MC, Zhao Y, Oberholzer M, Chang DD, et al. Propulsion of African trypanosomes is driven by bihelical waves with alternating chirality separated by kinks. Proc Natl Acad Sci USA. 2009; 106: 19322. doi: 10.1073/pnas.0907001106 10. Hill KL. Biology and mechanism of trypanosome cell motility. Eukaryot Cell 2003; 2: 200. doi: 10.1128/EC.2.2.200-208.2003 11. Ginger ML, Portman N, McKean PG. Swimming with protists: perception, motility and flagellum assembly. Nat Rev Microbiol. 2008; 6: 838. doi: 10.1038/nrmicro2009 12. Forestier C-L, Machu C, Loussert C, Pescher P, Spath GF. Imaging host cell-leishmania interaction dynamics implicates parasite motility, lysosome recruitment, and host cell wounding in the infection process. Cell Host Microbe 2011; 9: 319. doi: 10.1016/j.chom.2011.03.011 13. Alizadehrad D, Kruger T, Engstler M, Stark H. Simulating the complex cell design of Trypanosoma brucei and its motility. PLoS Comput Biol. 2015; 11: e1003967. doi: 10.1371/journal.pcbi.1003967 PLOS 14/17 14. Thurner S, Wick N, Hanel R, Sedivy R, Huber LA. Anomalous diffusion on dynamical networks: a model for epithelial cell migration. Physica A 2003; 320: 475. doi: 10.1016/S0378-4371(02)01598-4 15. Song L, Nadkarni SM, Bodeker HU, Beta C, Bae A, Franck C, et al. Dictyostelium discoideum chemotaxis: Threshold for directed motion. Eur J Cell Biol. 2006; 85: 981. doi: 10.1016/j.ejcb.2006.01.012 16. Amselem G, Theves M, Bae A, Bodenschatz E, Beta C. A stochastic description of Dictyostelium chemotaxis. PLoS ONE 2012; 7: e37213. 17. Theves M, Taktikos J, Zaburdaev V, Stark H, Beta C. A bacterial swimmer with two alternating speeds of propagation. Biophys J. 2013; 105: 1915. doi: 10.1016/j.bpj.2013.08.047 18. Makarava N, Menz S, Theves M, Huisinga W, Beta C, Holschneider M. Quantifying the degree of persistence in random amoeboid motion based on the Hurst exponent of fractional Brownian motion. Phys Rev E 2014; 90: 042703. 19. Metzler R, Jeon J-H, Cherstvy AG, Barkai E. Anomalous diffusion models and their properties: non-stationarity, non-ergodicity, and ageing at the centenary of single particle tracking. Phys Chem Chem Phys. 2014; 16: 24128. doi: 10.1039/C4CP03465A 20. Theves M, Taktikos J, Zaburdaev V, Stark H, Beta C. 2015. Random walk patterns of a soil bacterium in open and confined environments. EPL 109: 28007. doi: 10.1209/0295-5075/109/28007 21. Tsallis C. Possible generalization of Boltzmann -- Gibbs statistic J Stat Phys. 1988; 52: 479. 22. Plastino AR, Plastino A. Non-extensive statistical mechanics and generalized Fokker-Planck equation. Physica A. 1995; 222: 347. doi: 10.1016/0378-4371(95)00211-1 23. Upadhyaya A, Rieub J-P, Glaziera JA, Sawadac Y. Anomalous diffusion and non-Gaussian velocity distribution of Hydra cells in cellular aggregates. Physica A. 2001; 293: 549. doi: 10.1016/S0378-4371(01)00009-7 24. Uppaluri S, Nagler J, Stellamanns E, Heddergott N, Herminghaus S, Engstler M, et al. Impact of microscopic motility on the swimming behavior of parasites: straighter trypanosomes are more directional. PLoS Comput Biol. 2011; 7: e1002058. doi: 10.1371/journal.pcbi.1002058 25. Heddergott N, Kruger T, Babu SB, Wei A, Stellamanns E, Uppaluri S, et al. Trypanosoma motion represents an adaptation to the crowded environment of the vertebrate bloodstream. PLoS Pathog. 2012; 8: e1003023. doi: 10.1371/journal.ppat.1003023 26. Ballesteros-Rodea G, Santill´an M, Mart´ınez-Calvillo S, Manning-Cela R. Flagellar motility of Trypanosoma cruzi epimastigotes. J Biomed Biotechnol. 2012; 2012: 520380. doi: 10.1155/2012/520380 27. Zarley JH, Britigan BE, Wilson ME. Hydrogen peroxide-mediated toxicity for Leishmania donovani chagasi promastigotes. Role of hydroxyl radical and protection by heat shock. J Clin Invest. 1991; 88: doi: 1511. 10.1172/JCI115461 PLOS 15/17 28. Broadhead R, Dawe HR, Farr H, Griffiths S, Hart SR, Portman N, et al. Flagellar motility is required for the viability of the bloodstream trypanosome. Nature 2006; 440: 224. doi 10.1038/nature04541 29. Rozenberg G. Microscopic Haematology: A Practical Guide for the Laboratory, 3nd ed. London: Churchill Livingstone; 2011. 30. Farrar J, Hotez P, Junghanss T, Kang G, Lalloo D, White NJ. Manson's tropical diseases 23th ed. Oxford: Saunders; 2013. 31. de Noya BA, Gonz´alez ON, Robertson LJ. Trypanosoma cruzi as a foodborne pathogen. Springer International Publishing; 2015. 32. Wauthier, F. Motion tracking in image sequences. Department of Statistics, University of Oxford. 2015; Available: http://www.stats.ox.ac.uk/~wauthier/tracker/. 33. Lauga E, DiLuzio WR, Whitesides GM, Stone HA. Swimming in circles: motion of bacteria near solid boundaries. Biophys J. 2006; 90: 400. doi: 10.1529/biophysj.105.069401 34. Metzler R, Klafter J. The random walk's guide to anomalous diffusion: a fractional dynamics approach. Phys Rep. 2000; 339: 1. doi: 10.1016/S0370-1573(00)00070-3 35. Reverey JF, Jeon J-H, Bao H, Leippe M, Metzler R, Selhuber-Unkel C. Superdiffusion dominates intracellular particle motion in the supercrowded cytoplasm of pathogenic Acanthamoeba castellanii. Sci Rep. 2015; 5: 11690. doi: 10.1038/srep11690 36. Barry JD, McCulloch R. Antigenic variation in trypanosomes: enhanced phenotypic variation in a eukaryotic parasite. Adv Parasitol. 2001; 49: 1. doi: 10.1016/S0065-308X(01)49037-3 37. Donelson JE. Antigenic variation and the African trypanosome genome. Acta Trop. 2003; 85: 391. doi: 10.1016/S0001-706X(02)00237-1 38. Lythgoe KA, Morrison LJ, Read AF, Barry JD. Parasite-intrinsic factors can explain ordered progression of trypanosome antigenic variation. Proc Natl Acad Sci USA. 2007; 104: 8095. doi: 10.1073/pnas.0606206104 39. Pays E, Vanhamme L, P´erez-Morga D. Antigenic variation in Trypanosoma brucei: facts, challenges and mysteries. Curr Opin Microbiol. 2004; 7: 369. doi: 10.1016/j.mib.2004.05.001 40. Souza W. Basic cell biology of Trypanosoma cruzi. Curr Pharm Des. 2002; 8: 269. doi: 10.2174/1381612023396276 41. Bates PA. Transmission of Leishmania metacyclic promastigotes by phlebotomine sand flies. Int J Parasitol. 2007; 37: 1097. doi: 10.1016/j.ijpara.2007.04.003 42. Rogers M, Kropf P, Choi B-S, Dillon R. Podinovskaia M, Bates P, et al. Proteophosophoglycans regurgitated by Leishmania-infected sand flies target the L-arginine metabolism of host macrophages to promote parasite survival. PLoS Pathog. 2009; 5: e1000555. doi: 10.1371/journal.ppat.1000555 43. Rogers ME. The role of Leishmania proteophosphoglycans in sand fly transmission and infection of the mammalian host. Front Microbiol. 2012; 3: 223. doi: 10.3389/fmicb.2012.00223 PLOS 16/17 44. Rogers ME, Chance ML, Bates PA. The role of promastigote secretory gel in the origin and transmission of the infective stage of Leishmania mexicana by the sandfly Lutzomyia longipalpis. Parasitol. 2002; 124: 498. doi: 10.1017/S0031182002001439 45. Richardson LF. Atmospheric diffusion shown on a distance-neighbour graph. Proc R Soc London A. 1926; 110: 709. doi: 10.1098/rspa.1926.0043 46. Peng CK, Buldyrev SV, Havlin S, Simons M, Stanley HE, Goldberger AL. Mosaic organization of DNA nucleotides. Phys Rev E. 1994; 49: 1685. doi: 10.1103/PhysRevE.49.1685 47. Kantelhardt JW, Koscielny-Bunde E, Rego HHA, Havlin S, Bunde A. Detecting long-range correlations with detrended fluctuation analysis. Physica A 2001; 295: 441. doi: 10.1016/S0378-4371(01)00144-3 48. Atkins P, de Paula J. Physical Chemistry: Thermodynamics, structure, and change 8th ed. Oxford: Oxford University Press 2006. 49. Silverberg JL, Bierbaum M, Sethna JP, Cohen I. Collective motion of humans in mosh and circle pits at heavy metal concerts. Phys Rev Lett. 2013; 110: 228701. doi: 10.1103/PhysRevLett.110.228701 50. Auwera L, Meyer F, Malet L. The use of the Weibull three-parameter model for estimating mean wind power densities. J Appl Meteorol 1980; 19: 819. doi: 10.1175/1520-0450(1980)019<0819:TUOTWT>2.0.CO;2 51. Kiss P, Janosi IM. Comprehensive empirical analysis of ERA-40 surface wind speed distribution over Europe. Energy Convers Manage. 2008; 49: 2142. doi: 10.1016/j.enconman.2008.02.003 52. Carta JA, Ramirez P, Velazquez S. A review of wind speed probability distributions used in wind energy analysis Case studies in the Canary Islands. Renew Sustain Energy Rev. 2009; 13: 933. DOI: 10.1016/j.rser.2008.05.005 53. Morgan EC, Lackner M, Vogel RM, Baise LG. Probability distributions for offshore wind speeds. Energy Convers Manage. 2011; 52: 15. doi: 10.1016/j.enconman.2010.06.015 PLOS 17/17
1012.0031
1
1012
2010-11-30T21:41:13
Enhanced DNA sequencing performance through edge-hydrogenation of graphene electrodes
[ "physics.bio-ph", "cond-mat.mes-hall" ]
We propose using graphene electrodes with hydrogenated edges for solid-state nanopore-based DNA sequencing, and perform molecular dynamics simulations in conjunction with electronic transport calculations to explore the potential merits of this idea. The results of our investigation show that, compared to the unhydrogenated system, edge-hydrogenated graphene electrodes facilitate the temporary formation of H-bonds with suitable atomic sites in the translocating DNA molecule. As a consequence, the average conductivity is drastically raised by about 3 orders of magnitude while exhibiting significantly reduced statistical variance. We have furthermore investigated how these results are affected when the distance between opposing electrodes is varied and have identified two regimes: for narrow electrode separation, the mere hindrance due to the presence of protruding hydrogen atoms in the nanopore is deemed more important, while for wider electrode separation, the formation of H-bonds becomes the dominant effect. Based on these findings, we conclude that hydrogenation of graphene electrode edges represents a promising approach to reduce the translocation speed of DNA through the nanopore and substantially improve the accuracy of the measurement process for whole-genome sequencing.
physics.bio-ph
physics
Enhanced DNA sequencing performance through edge-hydrogenation of graphene electrodes Yuhui He,† Ralph H. Scheicher,∗,‡ Anton Grigoriev,‡ Rajeev Ahuja,‡,¶ Shibing Long,† ZongLiang Huo,† and Ming Liu∗,† Laboratory of Nano-Fabrication and Novel Devices Integrated Technology, Institute of Microelectronics, Chinese Academy of Sciences, Beijing 100029, China, Condensed Matter Theory Group, Department of Physics and Astronomy, Box 516, Uppsala University, SE-751 20 Uppsala, Sweden, and Applied Materials Physics, Department of Materials Science and Engineering, Royal Institute of Technology (KTH), SE-100 44 Stockholm, Sweden E-mail: [email protected]; [email protected] Abstract We propose using graphene electrodes with hydrogenated edges for solid-state nanopore- based DNA sequencing, and perform molecular dynamics simulations in conjunction with electronic transport calculations to explore the potential merits of this idea. The results of our investigation show that, compared to the unhydrogenated system, edge-hydrogenated graphene electrodes facilitate the temporary formation of H-bonds with suitable atomic sites in the translocating DNA molecule. As a consequence, the average conductivity is drastically raised by about 3 orders of magnitude while exhibiting significantly reduced statistical variance. We have furthermore investigated how these results are affected when the distance between oppos- ing electrodes is varied and have identified two regimes: for narrow electrode separation, the ∗To whom correspondence should be addressed †Chinese Academy of Sciences ‡Uppsala University ¶Royal Institute of Technology 1 mere hindrance due to the presence of protruding hydrogen atoms in the nanopore is deemed more important, while for wider electrode separation, the formation of H-bonds becomes the dominant effect. Based on these findings, we conclude that hydrogenation of graphene elec- trode edges represents a promising approach to reduce the translocation speed of DNA through the nanopore and substantially improve the accuracy of the measurement process for whole- genome sequencing. Tremendous recent advances have been made in the fabrication of solid-state nanopores1,2 and in their envisioned application for rapid whole-genome sequencing.3 -- 7 The basic concept centers around the idea that the four types of nucleobases occurring in DNA (adenine, thymine, cytosine, guanine; in the following abbreviated as A, T, C, G) possess different local electronic densities of states, which are electrically distinguishable and could thus in principle be used to differentiate be- tween them. Very recently, the electrical detection of single isolated nucleotides residing between nanoelectrodes has been realized, identifying three of the four nucleotides based on a statistical distribution of electrical conductivity curves.8 However, the development of solid-state nanopore- based DNA sequencing continues to struggle with a series of extremely challenging requisites, in particular single-base resolution during the polynucleotide translocation through the nanopore, optimized contrast in the electrical signals between the four different types of nucleotides, and a general improvement of signal-to-noise ratio.9,10 Among these requisites, one of the biggest challenges is the realization of single-base resolu- tion. Sufficiently thin nanoelectrodes are required, in order to have no more than one nucleotide within a close interaction range to the transverse electrodes at any given time when the nucleotides on the target DNA are passing one by one through the nanopore ([figure][1][]1a). Considering that a nucleotide possesses dimensions of roughly 1 nm, it can be comprehended why it is extremely challenging to prepare and electrically connect sufficiently thin nanopore-embedded electrodes to achieve single-base resolution. To potentially solve this issue, an intriguing proposal was recently made, namely to prepare a nanogap in graphene and use its edges as electrodes for DNA sequencing.11 Being a one-atom- 2 Figure 1: (a) Schematic view of single-stranded DNA translocating through a nanopore under the application of a longitudinal electrical field Ex, while the transverse tunneling current is recorded for the purpose of sequencing. Wx is the width of the transverse nanoelectrodes. Dy is the inner diameter of the nanopore which characterizes the gap distance between opposing transverse na- noelectrodes. (b) Cross-section visualization (cut along the x − z plane) of the complete atomistic setup employed in the present simulation work, showing the silicon-nitride membrane/nanopore (blue and yellow), a translocating single-stranded DNA molecule, as well as water molecules (red and white) and counter ions (ochre and cyan). (c) Cross-section visualization (cut along the y − z plane) of the setup, showing only the edge-hydrogenated graphene electrodes (cyan and white) and a cartoon-version of the translocating single-stranded DNA molecule with the sugar-phosphate backbone represented as a ribbon and the nucleobases as protruding sticks (colors are used to dis- tinguish between different nucleotides). thick planar sheet of carbon atoms,12,13 graphene represents the ultimate limit of how thin a na- noelectrode could possibly be, and hence, the associated prospects for single-base resolution are expected to be optimal. Taking several other advantages of graphene into account, such as its ability to be tailored by nanolithography,14 graphene definitely shows great promise for use as nanoelec- trodes in DNA sequencing. An experimental setup related to that proposed in Ref. 11, however not consisting of a nanogap, but rather a nanopore, drilled with an electron beam into graphene, was actually very recently realized independently by three research groups15 -- 17 and DNA translo- cation through these fabricated graphene nanopores was successfully demonstrated. In addition, a theoretical study based on density functional theory has been performed to explore the potential detection capabilities of nucleotides inside a graphene nanopore.18 These achievements represent an important milestone on the road towards realization of solid-state nanopore-based DNA se- quencing, and it can be expected that further advances will allow the processing of graphene to fabricate electrodes for the performance of transverse conductance measurements (as opposed to 3 ionic blockage current measurements). A drawback of graphene electrodes is that their conductance is much reduced compared to that of gold nanoelectrodes. This can be ascribed to several physical mechanisms: first, as a semi-metal or zero-gap semiconductor, the intrinsic electronic conductibility of graphene is smaller than that of the metal gold. Second, the coupling between graphene electrodes and DNA is smaller than that between gold nanoelectrodes and DNA due to the much smaller space extension of carbon outer or- bitals compared to those of gold. Since the coupling strength drops exponentially with decreasing overlap, the transverse tunneling conductance using graphene electrodes will be reduced dramat- ically relative to that using gold nanoelectrodes. Such a view is also supported directly from our studies when evaluating the coupling elements which are on an average found to be larger for gold than for carbon at a given nanopore diameter. This small conductance, in turn, will lead to a deteri- orated signal-to-noise ratio, i.e., the noise caused by ionic currents and by structure fluctuations of DNA during the translocation process19 will smoothen out any characteristic electrical signatures of the nucleotides, thus making it very difficult or even impossible to accurately determine the DNA sequence. Given that the introduction of a hydrogen bond (H-bond) can enhance electron tunneling rates over vacuum tunneling,20 we propose here hydrogenation of the edges of graphene electrodes in order to improve their sensitivity for the DNA sequencing purpose. A somewhat weakened form of H-bonds is expected to form between the hydrogen atoms at the graphene electrode edges and those atoms carrying a partial negative charge on the DNA nucleobases, since graphene is only slightly electronegative and as a result the hydrogen atoms on the graphene edge will only carry a relatively small positive charge (calculated by us from density functional theory to be around +0.16 e) compared to the typical positive charge in an H-bond. For simplicity, we will continue to refer in the following to these bonds as H-bonds, but it should be understood that they are generally weaker than common H-bonds. Several advantages can be expected from using edge-hydrogenation: 1. H-bonds formed between graphene electrodes and the translocating DNA bases can enhance 4 the coupling between them, and thus substantially increase the magnitudes of transverse tunneling currents. As a consequence, the current measurability is greatly improved and the speed with which the DNA sequence can be read is raised since the increased electrical currents no longer require a time-consuming femto-ampere amplification setup.9,10 2. The nucleotides of the DNA molecule possess many internal degrees of freedom and can assume a large number of conformations when passing through the nanopore. These atomic conformations are overall very similar but exhibit atomic-scale differences. Considering that tunneling currents are exponentially sensitive even to atomic-scale changes of orientation and distance, severe variance of the measured conductance can be expected and hence the signal-to-noise ratio is deteriorating. The introduction of H-bonds can favorably affect the orientation and position of nucleotides when DNA is passing through the nanopore, and thus help to reduce the conductance variance. 3. The H-bonds will cause an attractive force which is stronger than that from van der Waals interaction, so that they can slow down the translocation of DNA through the nanopore, providing more time for the transverse conductance measurement of each nucleotide located within the nanogap between transverse electrodes, thus sampling over inevitable noise and molecular motion. There is no risk that the DNA molecule would get stuck in the nanopore due to the H-bonds because they are sufficiently weak so that they can be broken easily by the longitudinally oriented electric driving field. [figure][1][]1a gives a schematic view of solid-state nanopore-based DNA sequencing: a single- stranded DNA molecule is driven through the nanopore electrophoretically by a longitudinal elec- trical field while the transverse tunneling conductance is recorded for the purpose of identifying the type of the individual nucleotides. Here, the thickness of nanoelectrodes (Wx) should be no greater than a critical value to achieve single-base resolution of the target DNA. At the same time, it should also provide adequate coupling to the target DNA in order to obtain a sufficiently large transverse tunneling conductance. [figure][1][]1b and [figure][1][]1c provide cross-section views 5 of the nanopore setup with edge-hydrogenated graphene electrodes employed in our simulations. Our theoretical calculations and simulations were implemented as follows: first, the electrical static potential charge distribution on the graphene electrodes is determined self-consistently with density functional theory method (here, the generalized gradient approximations is employed as implemented in the BLYP exchange correlation functional). Next, taking into account the previ- ously determined charge distribution on the graphene electrodes, translocation of single-stranded DNA through the nanopore is simulated with molecular dynamics for which the software codes NAMD221 and VMD22 have been used following the procedure described in Ref. 23. The real- time electronic structure of a translocating DNA molecule is then obtained within the extended Hückel model. Finally, the transverse tunneling conductivity is calculated using the Landauer- Büttiker formula and non-equilibrium Green's function method. (For further details about the molecular dynamics simulation settings and electrical property calculations, we refer the reader to our previous work.24) In [figure][2][]2a we plot the transverse differential conductance distribution curves (Gd) of a poly(dA)30 chain translocating through a nanopore with a distance of Dy = 1.1 nm between op- posing graphene electrodes for three different scenarios: edge-hydrogenated graphene electrodes (black line), unhydrogenated electrodes (red line), and pseudo-hydrogenated electrodes (blue line). The definition of the pseudo-hydrogenated electrodes setup will be provided below. In this work, transverse differential conductance is employed for gathering sequencing data, because it can directly exhibit the characteristic local electronic densities of states of the target molecules and thus optimize the contrast between the different nucleotides.24 Here, the conductance distribution curves reveal an expansive distribution that extends over more than 3 orders of magnitude. This spread is attributed to the variation of molecule-electrode contact distances associated with the di- verse molecular conformations during the translocation process through the nanopore. However, a well-defined single maximum is discernible in each conductance distribution. The position of these respective maxima indicates the conductivity of the most probable set of resembling nucleotide conformations when passing through the nanoelectrode gaps under the influence of a driving field. 6 Figure 2: (a) Transverse differential conductance (Gd) distribution curves of poly(dA)30 translocat- ing through a nanopore using edge-hydrogenated graphene electrodes (black line), unhydrogenated electrodes (red line), and pseudo-hydrogenated electrodes (blue line). The nanopore inner diameter (Dy) amounts to 1.1 nm, the longitudinal driving field has a strength of Ex = 5 kcal/(mol Å), and the transverse bias voltage V0 is set at 3.2 V, which is near the position of a characteristic eigen-level of adenine. (b) A snapshot extracted from the molecular dynamics simulation of the DNA transloca- tion through the nanopore, showing a moment when two H-bonds (dotted yellow lines) are formed simultaneously between the nitrogen atom of a a DNA nucleobase and two H atoms attached to the graphene-edge. For the sake of clarity, only relevant atoms from the edge-hydrogenated graphene electrodes and the DNA molecule have been visualized, omitting water molecules, counter ions, and the silicon-nitride membrane. The width of the peak characterizes the degree of variation in orientation and position of the DNA bases when they are located between the nanoelectrodes: the sharper the peaks, the less variation in orientation and position. An extremely idealized case could be imagined where DNA translocates through the nanopore like a stiff rod with no internal degrees of freedom at all. This hypothet- ical scenario would correspond to a very sharp peak in the conductance distribution curve, i.e., the overlap between conductance distribution curves for different types of nucleotides would be virtually zero, and as a consequence, the corresponding electrical signatures would become fully 7 distinguishable and the identification of individual nucleotides would be straightforward. In real- ity, the best-case scenario is a sufficiently narrow distribution of the measured conductance with minimal overlap between the distribution curves of different nucleotides. In order to see the qualitative and quantitative advantages in the performance of hydrogenated graphene electrodes more directly, we have also plotted for comparison the conductance distribu- tion curve for graphene electrodes with bare edges (i.e., unhydrogenated) in [figure][2][]2a (red line). It can be clearly seen that hydrogenation of the edges substantially increases the transverse tunneling conductance by about 3 orders of magnitude and leads to a much more narrow distribu- tion. The following question naturally arises then: since, taken by itself, the introduction of hydrogen atoms shortens the atomic-scale distance between graphene electrodes and the translocating DNA molecule, resulting in an increase of the molecule-contact coupling strength and the associated transverse tunneling current, how much contribution to the increased conductance originates from the actual formation of H-bonds (as shown in [figure][2][]2b), and how much is due to the mere presence of hydrogen atoms at the contact? In order to quantitatively answer this question we eliminate the contribution of the H-bonds while preserving the orbital overlap by replacing the hydrogen atoms on the graphene-edge by a set of dummy hydrogen-like atomic orbitals in our simulations. In this artificial scenario, which we refer to as pseudo-hydrogenated, no H-bonds can be formed at the contact and any increase of the transverse tunneling conductance originates only from the shortened molecule-contact distance. The obtained conductance distribution curve is plotted in [figure][2][]2a (blue line). By comparing the peak positions between the three cases, it becomes apparent that the H-bond formed at the contact is responsible for the most substantial contribution (about 2 orders of magnitude) to the increase in the transverse tunneling conductance. Furthermore, by comparing the peak widths of the three cases, we find that both the hydrogenation and the pseudo-hydrogenation of the graphene electrode edges lead to a significant reduction in the conductance variation. It thus appears that both confining effect from the hydrogen atoms at the graphene edge, and orientation effects caused by the formation of H-bonds play a role. Quantitative 8 evaluation of the confining effect is achieved by comparing the peak width of the conductance distribution curve obtained using pseudo-hydrogenated graphene and that using unhydrogenated graphene, while an evaluation of orientation effects can be achieved by comparing the cases of edge-hydrogenated graphene and of pseudo-hydrogenated graphene. This analysis shows that for nanopores with very small diameters, confining effects play the major role for the reduction of the conductance variation. Figure 3: Transverse differential conductance (Gd) distribution curves of poly(dA)30 translocating through a nanopore using edge-hydrogenated graphene electrodes (black line), unhydrogenated electrodes (red line), and pseudo-hydrogenated electrodes (blue line). The nanopore inner diameter (Dy) is 1.3 nm wide, the longitudinal driving field has a strength of Ex = 5 kcal/(mol Å), and the transverse bias voltage V0 is set at 3.2 V. The inset shows the transverse transmission spectra at a random snapshot during the translocation process for edge-hydrogenated graphene electrodes (black line) and for unhydrogenated graphene electrodes (red line). In order to maximize the chance for the formation of H-bonds between edge-hydrogen atoms on the graphene electrodes and the translocating DNA bases, the gap between the nanoelectrodes (Dy) should be made as narrow as possible. To explore the dependence of H-bond formation on the nanopore inner diameter, we simulated the translocation of poly(dA)30 through a nanopore with Dy = 1.3 nm, which is just about one row of carbon atoms in graphene wider than the 1.1 nm of the previously considered case presented in [figure][2][]2. The calculated conductance distribu- tion curves are plotted in [figure][3][]3, where the data for the system using edge-hydrogenated graphene electrodes is drawn in black, that using unhydrogenated electrodes in red, and that using pseudo-hydrogenated electrodes in blue. 9 [figure][3][]3 clearly demonstrates that although the overall conductivity is significantly re- duced when compared to the system with Dy = 1.1 nm (an unavoidable side effect of the larger na- noelectrode gap), edge-hydrogenation of graphene leads to an even more prominent improvement of conductance measurability and reduction of conductance variance: when using unhydrogenated graphene electrodes, the transverse tunneling conductance is too small (about 10−8 nS) and pos- sesses a too broad distribution (about 5 orders of magnitude) to be of any use for DNA nucleotide detection, while hydrogenation of graphene electrodes can significantly improve the conductivity (by about 3 orders of magnitude) and reduce the associated variance (amounting to merely about 2 orders of magnitude). Upon comparison with the artificial scenario of using pseudo-hydrogenated electrodes, it be- comes apparent that the confining effect and the orientation effect play nearly equal roles for the reduction of conductance variance in case of a larger nanopore diameter. The physical mechanism of these substantial changes can be identified from the inset of [figure][3][]3, where the transverse transmission spectra at a random snapshot are plotted for poly(dA)30 translocation with edge- hydrogenated graphene electrodes (black line) and with unhydrogenated graphene electrodes (red line): edge-hydrogenation causes the transmission peaks to become higher and wider, indicating much better coupling of the DNA molecule with the transverse electrodes. Another striking observation from the plots in ?? is that the sum of counts under the distribution curves is significantly increased when considering edge-hydrogenation on the graphene electrodes. Our analysis shows that two factors contribute to this increase: one is that for transverse conduc- tance measurements using unhydrogenated graphene electrodes, quite a few of the results are be- low the threshold of 10−10 nS to be realistically measurable in any experiments, and are therefore dropped from the distribution curve; another is that by using edge-hydrogenated graphene elec- trodes the DNA translocation speed is reduced. Reducing the DNA translocation speed during the detection process is a crucial requisite for DNA base identification in nanopores, since each nucleotide should remain between the transverse nanoelectrodes sufficiently long to sample over any unavoidable noise background. This slower translocation speed can also be directly observed 10 in our molecular dynamics simulations, which exhibit a reduction of about 20% in speed when H-bonds are formed. For a weaker electric driving field, this percentage-wise reduction could be even higher. But it can also be noted in [figure][2][]2 that the number of counts for the 1.1 nm wide nanopore is about the same for both hydrogenated and pseudo-hydrogenated case. Therefore, the formation of H-bonds alone cannot fully explain our results. Rather, the hindrance of DNA translocation through the nanopore for smaller diameters by repulsive interaction due to presence of graphene- edge hydrogen atoms is also responsible for a slower translocation speed (as observed in our molecular dynamics simulations). For the larger nanopore diameter of 1.3 nm ([figure][2][]2), this hindrance becomes less of an issue, as the DNA molecule has sufficient space to evade the protruding hydrogen atoms at the graphene edges. In that case, the formation of H-bonds becomes the main effect for the slowing-down of the translocation process, as can be seen from the larger number of counts for the hydrogenated over the pseudo-hydrogenated case. Figure 4: Transverse differential conductance (Gd) distribution curves of poly(dX)30 (X = A, T, C, G) translocating through a nanopore using edge-hydrogenated graphene electrodes. The gap between the transverse electrodes (Dy) is 1.1 nm, the longitudinal driving field Ex = 5 kcal/(mol Å), and the transverse bias voltage V0 is set at 4.8 V, near the position of a LDOS maximum of guanine. The inset shows the corresponding results obtained using bare-edge (unhydrogenated) graphene electrodes. [figure][4][]4 plots Gd distribution curves of poly(dX)30 (where X stands for A, T, C, and G, respectively) translocating through a nanopore with Dy = 1.1 nm. Data for edge-hydrogenated 11 graphene electrodes is shown in the main panel of the figure, while data for the unhydrogenated system is plotted in the inset. Here, the transverse bias voltage V0 is set at 4.2 V where the local electronic density of states (LDOS) for guanine has a maximum. As a result, Gd of G is orders of magnitude larger than that of the other three nucleotides, making it possible to easily identify G. Although it appears from [figure][4][]4 as if the bases A, C, and T could not be distinguished from another, one should keep in mind that a proper adjustment of V0 to the positions of the respective maxima in the LDOS of other nucleotides can actually resolve these differences, as we have shown recently on the example of a setup based on gold nanoelectrodes.24 Finally, it should be emphasized that, according to our calculations, the likelihood for H-bond formations becomes dramatically enhanced when the nanopore inner diameter Dy is about 1.3 nm or less. This requirement for the nanopore diameter originates from the comparatively small partial positive charges on the graphene-edge hydrogen atoms: considering that the hydrogen atoms carry a charge of about +0.16 e (calculated within the framework of density functional theory) and not the approximately +0.35 to +0.45 e typical for hydrogen atoms forming usual H-bonds, the dis- tances at which DNA nucleobases are effectively attracted towards the graphene edge to form these weakened H-bonds should be rather small. Experimentally, the preparation of such tiny nanopores is expected to be extremely challenging. However, we would like to point out that our analysis of the effect of H-bonds on the transverse tunneling conductance is in principle not limited to the hydrogenation of graphene edges. In fact, H-bonds could be introduced in other ways, such as, e.g., by attaching a functional group to the nanoelectrodes. Previous theoretical calculations25 in- dicate that a nanoelectrode chemically functionalized with a probing base can form H-bonds with the DNA nucleotides as well. Our analysis and conclusions still apply, and such H-bond-assisted nucleotide recognition has in fact been verified experimentally.26 -- 28 In summary, through edge-hydrogenation of the graphene electrodes in a nanopore-based DNA sequencing setup, the transverse tunneling conductance can be drastically raised by about 3 orders of magnitude, thus improving the conductance measurability substantially. At the same time, the variation in the conductance will be significantly reduced, leading to a faster and more reliable 12 identification of the four nucleotide types. The pico-siemens tunneling conductance facilitates reading the nucleotide sequence at a much greater speed than what is possible with only femto- ampere tunneling currents. The analysis of our simulation shows that the formation of H-bonds between the hydrogenated graphene edges and the DNA nucleobases plays a crucial role in the described improvements. Acknowledgement We gratefully acknowledge financial support from the National Science Fund for Distinguished Young Scholars (Grant No. 60825403), China Ministry of Science and Technology (Contract No. 2010CB934200), the Swedish Foundation for International Cooperation in Research and Higher Education (STINT), the Swedish Research Council (VR, Grant No. 621-2009-3628), Wenner- Gren Foundations, Carl Tryggers Stiftelse för Vetenskaplig Forskning, and the Uppsala University UniMolecular Electronics Center (U3MEC). References (1) Wu, M. Y.; Smeets, R. M. M.; Zandbergen, M.; Ziese, U.; Krapf, D.; Batson, P. E.; Dekker, N. H.; Dekker, C.; Zandbergen, H. W. Nano Lett. 2009, 9, 479-484. (2) Taniguchi, M.; Tsutsui, M.; Yokota, K.; Kawai, T. Appl. Phys. Lett. 2009, 95, 123701. (3) Storm, A. J.; Storm, C.; Chen, J. H.; Zandbergen, H.; Joanny, J. F.; Dekker, C. Nano Lett. 2005, 5, 1193-1197. (4) Zwolak, M.; Di Ventra, M. Nano Lett. 2005, 5, 421-424. (5) Lagerqvist, J.; Zwolak, M.; Di Ventra, M. Nano Lett. 2006, 6, 779-782. (6) Iqbal, S. M.; Akin, D.; Bashir, R. Nat. Nanotechnol. 2007, 2, 243-248. (7) Dekker, C. Nat. Nanotechnol. 2007, 2, 209-215. 13 (8) Tsutsui, M.; Taniguchi, M.; Yokota, K.; Kawai, T. Nat. Nanotechnol. 2010, 5, 286-290. (9) Branton, D.; Deamer, D.; Marziali, A.; Bayley, H.; Benner, S. A.; Butler, T.; Di Ventra, M.; Garaj, S.; Hibbs, A.; Huang, X.; Jovanovich, S. B.; Krstic, P. S.; Lindsay, S.; Ling, X. S.; Mastrangelo, C. H.; Meller, A.; Oliver, J. S.; Pershin, Y. V.; Ramsey, J. M.; Riehn, R.; Soni, G. V.; Tabard-Cossa, V.;Wanunu, M.; Wiggin, M.; Schloss, J. A. Nat. Biotechnol. 2008, 26, 1146-1153. (10) Zwolak, M.; Di Ventra, M. Rev. Mod. Phys. 2008, 80, 141-165. (11) Postma, H. W. Ch. Nano Lett. 2010, 10, 420-425. (12) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.; Grig- orieva, I. V.; Firsov, A. A. Science 2004, 306, 666-669. (13) Castro Neto, A. H.; Guinea, F.; Peres, N. M. R.; Novoselov, K. S.; Geim, A. K. Rev. Mod. Phys. 2009, 81, 109-162. (14) Tapasztó, L.; Dobrik, G.; Lambin, P.; Biró, L. P. Nat. Nanotechnol. 2008, 3, 397-401. (15) Schneider, G. F.; Kowalczyk, S. W.; Calado, V. E.; Pandraud, G.; Zandbergen, H. W.; Van- dersypen, L. M. K.; Dekker, C. Nano Lett., 2010, 10, 3163-3167. (16) Merchant, C. A.; Healy, K.; Wanunu, M.; Ray, V.; Peterman, N.; Bartel, J.; Fischbein, M. D.; Venta, K.; Luo, Z.; Johnson, A. T. C.; Drndi´c, M. Nano Lett., 2010, 10, 2915-2921. (17) Garaj, S.; Hubbard, W.; Reina, A.; Kong, J.; Branton, D.; Golovchenko, J. A. Nature 2010, 467, 190-193. (18) Nelson, T.; Zhang, B.; Prezhdo, O. V. Nano Lett., 2010, 10, 3237-3242. (19) Krems, M.; Zwolak, M.; Pershin, Y. V.; Di Ventra, M. Biophys. J. 2009 97, 1990-1996. (20) Wuttke, D. S.; Bjerrum, M. J.; Winkler, J. R.; Gray, H. B. Science 1992, 256, 1007-1009. 14 (21) Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kale, L.; Schulten, K. J. Comp. Chem. 2005, 26, 1781-1802. (22) Humphrey, W.; Dalke, A.; Schulten, K. J. Molec. Graphics 1996, 14, 33. (23) Comer, J; Dimitrov, V.; Zhao, Q; Timp, G; Aksimentiev, A. Biophys. J. 2009, 96, 593-608. (24) He, Y.; Shao, L.; Scheicher, R. H.; Grigoriev, A.; Ahuja, R.; Long, S.; Ji, Z.; Yu, Z.; Liu, M. Appl. Phys. Lett. 2010, 97, 043701. (25) He, H.; Scheicher, R. H.; Pandey, R.; Rocha, A. R.; Sanvito, S.; Grigoriev, A.; Ahuja, R.; Karna, S. P. J. Phys. Chem. C 2008, 112, 3456-3459. (26) Ohshiro, T.; Umezawa, Y. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 10-14. (27) Chang, S.; He, J.; Kibel, A.; Lee, M.; Sankey, O.; Zhang, P.; Lindsay, S. Nat. Nanotechnol. 2009, 4, 297-301. (28) Chang, S.; Huang, S; He, J., Liang, F., Zhang, P.; Li, S.; Chen, X.; Sankey, O.; Lindsay, S. Nano Lett. 2010, 10, 1070-1075. 15
1807.06378
1
1807
2018-07-17T12:26:04
Loss of vimentin intermediate filaments decreases peri-nuclear stiffness and enhances cell motility through confined spaces
[ "physics.bio-ph", "q-bio.CB" ]
The migration of cells through tight constricting spaces or along fibrous tracks in tissues is important for biological processes, such as embryogenesis, wound healing, and cancer metastasis, and depends on the mechanical properties of the cytoskeleton. Migratory cells often express and upregulate the intermediate filament protein vimentin. The viscoelasticity of vimentin networks in shear deformation has been documented, but its role in motility is largely unexplored. Here, we studied the effects of vimentin on cell motility and stiffness using mouse embryo fibroblasts derived from wild-type and vimentin-null mice. We find that loss of vimentin increases motility through small pores and along thin capillaries. Atomic force microscopy measurements reveal that the presence of vimentin enhances the perinuclear stiffness of the cell, to an extent that depends on surface ligand presentation and therefore signaling from extracellular matrix receptors. Together, our results indicate that vimentin hinders three-dimensional motility by providing mechanical resistance against large strains and may thereby protect the structural integrity of cells.
physics.bio-ph
physics
Loss of vimentin intermediate filaments decreases peri-nuclear stiffness and enhances cell motility through confined spaces Alison E. Patteson1,2*, Katarzyna Pogoda1,3, Fitzroy J. Byfield1, Elisabeth E. Charrier1, Peter A. Galie1,4, Piotr Deptuła5, Robert Bucki5, and Paul A. Janmey1* 1 Institute for Medicine and Engineering, University of Pennsylvania, Philadelphia, PA 19104 2 Physics Department, Syracuse University, Syracuse, NY 13244 3 Institute of Nuclear Physics, Polish Academy of Sciences, PL-31342 Krakow, Poland 4 Department of Biomedical Engineering, Rowan University, Glassboro, NJ 08028 5 Department of Microbiological and Nanobiomedical Engineering, Medical University of Białystok, Mickiewicza 2C, Białystok, Poland *Corresponding author. Email: [email protected] (A.E.P.); [email protected], (P.A.J.) The migration of cells through tight constricting spaces or along fibrous tracks in tissues is important for biological processes, such as embryogenesis, wound healing, and cancer metastasis, and depends on the mechanical properties of the cytoskeleton. Migratory cells often express and upregulate the intermediate filament protein vimentin. The viscoelasticity of vimentin networks in shear deformation has been documented, but its role in motility is largely unexplored. Here, we studied the effects of vimentin on cell motility and stiffness using mouse embryo fibroblasts derived from wild-type and vimentin-null mice. We find that loss of vimentin increases motility through small pores and along thin capillaries. Atomic force microscopy measurements reveal that the presence of vimentin enhances the perinuclear stiffness of the cell, to an extent that depends on surface ligand presentation and therefore signaling from extracellular matrix receptors. Together, our results indicate that vimentin hinders three- dimensional motility by providing mechanical resistance against large strains and may thereby protect the structural integrity of cells. Introduction Many important biological processes, such as embryogenesis 1,2 and wound healing 3,4 , depend on the ability of cells to move through the tight constricting spaces in tissues. This same ability - under disruptive conditions - can lead to invasive cell migration and cancer metastasis 5,6. A key event that triggers cell migration is the epithelial-mesenchymal transition (EMT), in which non- migratory epithelial cells lose cell-cell adhesions and transition to migratory polarized mesenchymal cells 1,2. During this transition, cells express and upregulate the intermediate filament protein vimentin (Fig. 1a), a wide-spread marker of EMT and mesenchymal cells 1,7. Vimentin's role in cell motility has been implicated in aggressive tumors throughout the body, including the prostate, breast, and lung, and high levels of vimentin expression in cancer cells correlate with accelerated tumor growth and poor prognosis 8. In vivo, cells move by either degrading the surrounding extracellular matrix (ECM) or deforming their bodies in order to squeeze through narrow openings of the ECM or conform to highly 1 curved fibrous tracks such as nerves or small blood vessels 9,10. The interstitial spaces of tissues are complex environments that contain pores, 0.1 to 30 μm in diameter 11,12, and pre-existing tracks, 100-600 μm in length 13, that act as paths for neutrophil 9 and cancer cell 9,12 migration. For small pores (2-3 μm), nuclear size and rigidity significantly limit migration 9,14,15, and when cells do squeeze through small pores, the large strain imposed on the cell and nucleus can lead to nuclear membrane rupture 16,17, accumulated DNA damage 16-18, and enhanced genome variation 18. A robust feature of vimentin intermediate filaments is their formation of a cage-like network that encircles the cell nucleus (Fig. 1a) 19 and contributes to the mechanical integrity of the cell 20-22. Even under conditions in which the peripheral vimentin networks is destabilized, such as growth on soft substrates, the perinuclear vimentin cage remains intact 23. Reconstituted vimentin networks are viscoelastic materials that significantly stiffen at large strains and are capable of withstanding extreme deformations that cause other cytoskeletal polymers (actin and microtubule) to fail 24,25. Vimentin's unique mechanical properties are hypothesized to confer enhanced resistance to cellular deformations especially at large strains 22,24. However, experimental demonstrations remain sparse, and little is known about how vimentin affects cell motility through constricting spaces. Results Loss of vimentin increases cell motility through constricting spaces To model cell migration in 3D environments, we designed micro-fluidic devices with confining channels. The channels were large enough to allow the nucleus to pass through yet small enough to constrict the vimentin perinuclear cage (Fig. 1, SI Fig 1-3). We observed that both wild-type (vim +/+) and vimentin-null (vim -/-) mEF exhibited spontaneous migration into collagen I- coated micro-channels (Fig. 1b) at migration rates similar to in vivo speeds (50-300 um/hr) 26 observed for amoeboid cancer cells. These persistent motions occurred in the absence of both chemical gradients and applied pressure gradients across the channels, consistent with previous studies using 3T3 fibroblasts 27. Here, we found that the ability to pass through the constrictions differs strikingly for cells with and without vimentin. Surprisingly, the loss of vimentin enhanced migration, enabling cells to enter the channels faster and increasing the probability of crossing the channel (Fig. 1b iii, SI Movie 1 & 2). Cells that do not pass through either change direction or get stuck, events that occur more frequently in wild-type than vimentin-null mEF (c.f. Fig. 2, SI Movie 3). Similar behaviors were observed in 3T3 fibroblasts, which also showed a marked decrease in motility compared to fibroblasts lacking vimentin (SI Fig. 4). To test the effects of vimentin on motility in fibrillar biopolymer matrices, we embedded the wild-type and vim -/- mEF in collagen I gels and tracked their motion over 17 hr (Methods). Bright-field images of the cells in the gel and sample trajectories are shown in Fig. 1c. In the collagen gels, the cells are highly confined (expected average 1-2 μm pore size 28), and their motility is limited. Under these conditions, the vimentin-null cell behavior is more varied than that of wild-type mEFs, yet a subset of vim -/- mEFs squeeze through the narrow pores and travel much greater distances than wild-type mEFs. 2 The micro-fluidic channels and collagen gel experiments highlight an unexpected trend: vimentin hinders motility through 3D confining spaces. This result contrasts with previous studies that show vimentin enhances motility on unconfined rigid substrates 20,29. What effect does cell confinement have on vimentin's role in motility? To address this question, we compared the instantaneous speeds of cells measured in environments with varying degrees of confinement, which include unconfined 2D glass substrates, micro-channels of various width, and highly- confined collagen gels (Fig. 1d, Methods). On the glass slides, the instantaneous speeds are 3 Figure 1: Vimentin decreases cell motility through constricting spaces (a) (i) Migratory cells often express the intermediate filament vimentin during important events such as the epithelial-mesenchymal transition, wound healing, and aggressive tumor growth. Vimentin forms a peri-nuclear cage, which could hinder migration through constricting interstitial spaces. (ii) Immunofluorescence image showing vimentin in a wild-type mouse embryo fibroblast (mEF). (iii) Unlike the rigid flat surfaces common to cell culture methods, interstitial spaces are three-dimensional environments containing small pores and highly curved tracks, such as along nerve fibers. (b) Cells are seeded in PDMS mico-fluidic devices in the absence of applied pressure and chemical gradients across the channels. (i) Brightfield images showing wild type (vim +/+) and vimentin null (vim -/-) mEF migrating through collagen I-coated micro-channels (SI Movies 1&2). Cells circled for visualization. (ii) Immunofluorescence image showing co-localization of vimentin and nucleus for wild-type mEF entering channel (Fibronectin-coated, 12 hr). (iii) Vim +/+ mEF take more time entering and are less likely to pass through micro-channels (collagen-coated) than vim -/- mEF. (c) (i) Bright field images of vim +/+ and vim -/- mEFs cultured in 3D collagen gels (24 hr, 2 mg/mL). (ii) Cell trajectories (over 17 hr) show that vim -/- mEF migrate more through gel than vim +/+ mEF. (d) Cell speed depends on vimentin and confinement. On unconfined glass slides, cell speed is similar between the two cell lines (N = 45+ cells each). In the micro-channels, the speed of vim -/- mEF increases with decreasing channel width at a significantly larger rate than vim +/+ mEF (N = 35-60 cells each). In collagen gels where confinement is strongest, vim -/- are faster than vim +/+ mEFs. Error bars denote standard error. Denotation ***, p < 0.001; **, p < 0.01. statistically similar (Fig. 1d), although wild-type mEF motion is more persistent than vim -/- (SI Fig. 5). In the micro-channels, cells experience increasing levels of confinement as the width decreases from 20 to 10 μm. For the vimentin-null cells, the confinement increases cell speed by a factor of 2.25 (p < 0.001), whereas wild-type speed increases by only 24% (p = 0.058). In the porous gels, cell speeds are significantly lower compared to the larger micro-channels. Unlike the glass slides, however, vimentin decreases cell speed in the 3D gels that confine the cells. Taken together, these experiments (Fig. 1) demonstrate that vimentin can significantly hinder migration and that this effect is stronger in more constricting environments. Increased cell speed with confinement has been observed previously 27,30 and is implicated with transitions to faster migration modes, such as amoeboid migration 30. Moderate confinement induces changes in the cytoskeleton, such as an accumulation of actin in the cortex and an alignment of microtubules that is important for migration along 3D tracks 31,32. In our micro-channels, application of microtubule and actin inhibitors causes a rapid decrease in motility for vim +/+ and vim -/- mEFs, indicating that the 3D motility observed here depends on actin and tubulin (SI Fig. 6 & 7). The results shown here (Fig. 1) also highlight that vimentin plays an important role in controlling 3D motility. 3D fibroblast motility depends on vimentin and surface integrin-coating We observe differences in cell behavior between surfaces coated with collagen I and fibronectin (Fig. 2). For both cell lines, the projected spread area was greater in fibronectin-coated micro- channels compared to collagen I (Fig. 2a, SI Fig. 8), similarly to other stiff surfaces 22. For both ligands, the average spread area was approximately the same between the two cell lines, suggesting similar cell volumes despite having different cell speeds (Fig. 2a). Cells lacking vimentin were faster than wild-type mEF for both coatings: yet showed a 22% decrease (p = 0.011) on fibronectin compared to collagen I; whereas normal cell speeds were constant. Ligand coating impacted emergent behaviors in the micro-channels, such as twirling and persistency (Fig. 2b). When seeded on fibronectin, both cell types were more likely to exhibit 4 twirling (SI Movie 4), appearing to move in three-dimensional helices, compared to their movement on collagen. Interestingly, twirling was more frequent in normal cells than cells without vimentin. In addition, persistent migration – characterized here by the percentage of cells that pass through the channel – depended on the surface ligand coating and vimentin. Wild-type mEF became more persistent in the presence of fibronectin compared to collagen, whereas vimentin-null mEF were persistent on both. Cellular response to extracellular ligands is important for healthy tissue maintenance, for example fibronectin stimulates fibroblast migration in damaged tissues and collagen helps maintain homeostasis 33,34. The different effects of collagen-I and fibronectin on wild-type mEF Figure 2: Vimentin and surface-ligand coatings control 3D cell migration (a) Left: Immunofluorescence images of vim +/+ and vim -/- mEF in micro-channels coated with fibronectin. Right: Both cell types are more spread in channels coated with fibronectin than collagen I. Cells lacking vimentin slow down on fibronectin compared to collagen-coated channels. (b) In the microchannels, cells display different behaviors, such as twirling, getting stuck, and switching directions (SI Movies 3&4). Twirling appears more often in vim +/+ than vim -/- mEF and is more frequent on fibronectin-coated channels than collagen. Normal mEFs are more likely to cross channel in the presence of fibronectin. Error bars denote standard error. Denotation ***, p < 0.001; **, p < 0.01; *, p < 0.05. 5 motility suggest distinct biochemical signals from different integrins that engage these ligands. Overall, the results in Fig. 2 indicate a role of vimentin in regulating cell migration in response to different ligands, which includes promoting persistent migration in the presence of fibronectin, a result that may be important for understanding wound healing models. Vimentin hinders migration along highly-curved capillaries Similar to the behavior in micro-channels, vim -/- cells migrate faster than vim +/+ when cultured on glass capillaries with diameters ranging from 10-20 μm (Fig. 3a). Migration of vim - /- cells was also sensitive to the ligand coating of the substrate, with a higher migratory speed observed on collagen-coated capillaries compared to those coated with fibronectin (Fig. 3b). Both cell lines tended to have a high aspect ratio and were elongated in the axial direction rather than the radial direction of the capillaries, possibly due to the high bending energy of F-actin 35 (Fig. 3c). Nuclei of vim +/+ cells, cultured on capillaries, were observed to be off-center relative to the distribution of vimentin within the cell (Fig. 3d). Vimentin stiffens cells on substrates coated with collagen but not fibronectin To understand vimentin's role in the ability of cells to move and deform, we used atomic force microscopy (AFM) experiments to measure the stiffness of normal and vimentin-null mEFs (Fig. Figure 3. Loss of vimentin increases cell migration on highly curved capillaries (a) Bright field time-lapse images of vim +/+ and vim -/- mEF cultured on collagen-coated capillaries, taken over approximately 3 hr. (b) Vim -/- cells migrate faster than vim +/+ cells on both collagen- and fibronectin-coated capillaries (N=17-27 cells per condition). While the speed of vim +/+ cells is similar on both collagen- and fibronectin-coated substrates, the speed of vim -/- cells is higher on collagen. (c) Fluorescence images showing the typical distribution of vimentin and F-actin in a vim +/+ mEf and F-actin in a vim -/- mEF, cultured on collagen coated capillaries (21 hr). Dotted lines indicate the boundaries of the capillary. (d) Line profiles of a fluorescent image showing the typical position of the nucleus (stained with DAPI) relative to the distribution of vimentin (stained with anti-vimentin ab) in a vim +/+ cell on a collagen-coated capillary. Dotted lines indicate boundary of vimentin staining. A.U: Arbitrary units. Error bars denote standard error. Denotation ***, p < 0.001. 6 3). Notably, the loss of vimentin does not significantly change the expression levels of either F- actin and microtubules between these two cell lines 21, and we observe no obvious differences in the organization of these networks from fluorescence staining on collagen- (Fig. 3a) or fibronectin coated glass surfaces (SI Fig. 9). We measured cell stiffness in the perinuclear region of the cell, sampling multiple positions just outside the periphery of the nucleus, where filamentous vimentin is abundant (Fig. 3b). Cells were cultured on glass coverslips and then subjected to periodic indentation by an AFM tip, which was applied at a constant force amplitude of 3 nN and frequency 0.4 Hz. We found that cell's perinuclear stiffness depended strongly on the presence of vimentin and the type of adhesive ligand used for coating. On collagen-coated surfaces, the stiffness of normal mEFs was approximately 13 kPa, which was 2.5x times greater than for vimentin-null mEFs (p<0.001). On fibronectin, the stiffness of normal mEFs was lower (7.5 kPa) and more equal to the stiffness of vim -/- mEFs (5 kPa, p = 0.09). These trends are striking compared to previous Figure 4: Vimentin increases cell stiffness on collagen coated surfaces and impedes 3D migration (a) (i) Immunofluorescence images of vim +/+ and vim -/- mEF showing the cytoskeletal networks: actin, microtubule, and vimentin (as shown on collagen-coated glass slides). Vim -/- mEFs maintain robust actin and microtubule networks. (ii) To measure peri-nuclear stiffness, an AFM tip was used to probe the region juxtapose to the nucleus, sampling at multiple positions (as labeled) to account for cytoskeletal variations. (iii) Perinuclear stiffness depends on the presence of vimentin and the type of adhesive ligand used for coating: vim +/+ mEFs are stiffer than vim -/- mEF on glass slides coated with collagen but not fibronectin. (N = 30+ cells per condition). (b) (i) For the glass capillaries and micro-channel experiments, cell speed decreases with increasing cell stiffness: dotted lines connect points to guide the eye. Ligand coating for each condition, from left to right, collagen (Col), fibronectin (Fn), Fn, Col. (ii) Percentage of cells crossing the micro-channels versus cell stiffness E. Dashed line is fit of the data to activated model 𝑘"exp (−𝛽𝐸), which suggests that stiffness limits migration by increasing the total work 𝛽𝐸 for cells to cross channels. Error bars indicate 7 standard error. Denotation ***, p < 0.001. AFM experiments 22, which probed the endoplasmic region of these two cell lines, and found that vim +/+ mEFs were stiffer than vim -/- only when cells with maximal spread areas (spread area > 14, 000 μm2) were compared. The experiments shown here (Fig. 3) highlight an important role of vimentin in mediating stiffness in the perinuclear region of the cell, a role that also depends on the surface ligand coating. Cell speed in 3D environments correlates with cell stiffness Consistent with the hypothesis that cytoskeletal stiffness hinders cell migration through constricted spaces, we find that 3D cell speeds monotonically decrease with cell stiffness. Here, cell stiffness varies depending on surface ligand coating and the presence of vimentin. A scatter plot of population-averaged cell speed versus cell stiffness indicates a general trend that cell stiffness suppresses cell speeds in 3D environments, as shown in Fig. 4b for both the glass capillaries and micro-channels. Interestingly, on the 2D glass coverslips, vim +/+ and vim -/- cell speeds do not depend on cell stiffness. Our results (Fig. 4bi) suggest that softer cells move faster than stiff ones in 3D and vimentin's contribution to stiffness thereby decreases migration. The ability of cells to cross the micro-channels also decreases with cell stiffness (Fig. 4bii). To interpret this result, we suggest a minimal model that treats crossing the channel as an activated process with rates controlled by effective energy barriers needed to be overcome for cells to migrate across the channel. Assuming that peri-nuclear stiffness 𝐸 presents a significant energy barrier for the cell to cross the channel, the flux of cells is given by 𝑘"exp(−𝛽𝐸), where 𝑘" is the flux in the absence of cell deformation and 𝛽 is a constant determined by details of the required to cross the channel increases and the flux decreases by a factor exp(−𝛽𝐸), consistent coupling between cell deformation and channel geometry. As cell elasticity increases, the work with the observed decrease in cell flux (Fig. 4bii). Vimentin's contribution to peri-nuclear stiffness limits cell migration through small pores To test the limits of vimentin's effects on migration, we performed confined motility experiments using polycarbonate Transwell membranes with pore diameters varying from 3 to 8 μm, significantly smaller than the micro-channels. Cells were seeded on top of the membrane and allowed to adhere and migrate for 15 hr before being fixed and counted (Fig. 5a, Methods). The migration rates through the small pores depended on whether the membranes were coated with collagen or fibronectin (Fig. 5b). For collagen, vimentin-null mEF crossed at higher rates than wild-type mEF; even for the 3 μm pores – where the nucleus is expected to be limiting 9,14,15 – the presence of vimentin significantly impacted motility, decreasing speed by a factor of 4.2 (p = 0.004). For fibronectin, motility rates were equalized between the cell lines. Variations in pore migration (Fig. 5b) may be attributed to differences in cell stiffness: to cross smaller pores, cells must undergo larger strains and expend more energy, which increases with cell stiffness. Here, we estimate that the cell compressional strain varies from 0.47 to 0.80 based on the geometry of the pores (Fig. 5c, Methods). To interpret how cell migration depends on cell stiffness and cell strain, we extend the activated model of constricted migration, assuming that the energy barrier to cross the pore is proportional to the cell stiffness 𝐸 and the parameter 𝜖"𝜖, where 𝜖 is the compressional strain needed to cross the membrane and 𝜖" is a pre-strain based upon internal stresses that develop as cells adhere and spread on the surface. The flux of cells 8 Figure 5: Vimentin intermediate filaments impede migration through nucleus-limiting pores (a) (i) The role of vimentin in constricted cell migration is probed using Transwell membranes with pore diameters ranging from 3 to 8 μm. Vim +/+ and vim -/- mEF are seeded on top of membranes and allowed to migrate for 15 hr, then fixed and stained with crystal violet. (ii) Fluorescence images of cells localized to the membrane top and bottom show that more vimentin-null mEF translocate pores than wild type mEF, as shown for 8 μm collagen-coated pores. (b) The percentage % of cells that cross the filters depends on pore size and surface ligand coating. Fibronectin equalizes the crossing rates of the two cell types. (c) Smaller pore sizes require higher cell strain and more work to passage through. The % cross versus the compressional strain is fit to the activated model 𝑘"exp(−𝛽𝜖"𝜖𝐸 ), where 𝐸 is the AFM-measured cell stiffness; the fitting parameters 𝑘"= 16% and 𝛽𝜖" = 0.18 kPa-1 are each a single constant across all four data sets. Error bars indicate standard error. then decreases by a factor exp(−𝛽𝜖"𝜖𝐸 ) for increasing cell stiffness 𝐸 (a function of cell type and ligand coating) and strain 𝜖 (based on pore size alone). By fixing cell stiffness 𝐸 to values from the AFM measurements (Fig. 4), we fit the four curves in Fig. 5c to 𝑘"exp(−𝛽𝜖"𝜖𝐸 ) and obtain a single value for the constants 𝛽𝜖" = 0.18 kPa-1 and 𝑘" = 16%. The model seems to Denotation **, p < 0.01; *, p < 0.05. capture the main features of the experimental data and further supports the idea that constricted cell migration depends strongly on vimentin peri-nuclear stiffness. Discussion Intermediate filaments are important in development, tissue maintenance, metastasis, and disease. Using a knock-out model for vimentin, we have found that vimentin hinders cell motility through small pores and along thin capillaries, in contrast to previous studies, which have shown 9 that vimentin enhances motility on 2d plastic cell culture substrates. The strong effect of vimentin on 3D motility may be surprising, given that intermediate filaments are much softer than F-actin or microtubules and thus easiest to deform to large strains. Our results here indicate that strain-stiffening and resistance to breaking 24 may be the important mechanical features through which vimentin contributes to migration. It is possible that vimentin's role in regulating actin stress fiber assembly through rho-A 36 and possible other signaling mechanisms contribute to differences in motility for cells with and without vimentin, although no difference in actin expression levels have been reported for these two cell lines 21. The physiological effects of intermediate filaments in 3D cell migration remain unclear. Vimentin expression reflects a phenotypic characteristic that contributes to normal cell migration during EMT but also aggressive behavior of tumor cells. The use of vimentin-disrupting drugs such as withaferin A have been proposed as effective therapy for cancer treatment 8. Recent studies have shown that modification of keratin intermediate filament organization in pancreatic cancer cells by the phospholipid sphingosylphosphorylcholine (SPC) enhanced cell deformability and migration in micro-channels 37, which is suggested to correlate with increased metastatic potential. Here, we find that the loss of vimentin in a knock-out mouse model similarly increases cell deformability and enhances constricted cell migration. These results might seem counter-intuitive, considering that as epithelial cells undergo EMT and become more migratory they upregulate vimentin to replace cytokeratin rather than downregulating intermediate filaments generally. However, the greater localization of vimentin around the nucleus compared to other cytoskeletal filaments could help provide a more detailed understanding of why migratory cells express vimentin. One possible explanation is that the perinuclear vimentin network is required to cushion the nucleus or the cortical actin network during extreme strains associated with movement through constrictions that would otherwise cause damage to the cell. Indeed, our current work Patteson, et al. 38 reveals that the loss of vimentin increases nuclear damage, such as nuclear membrane rupture, during migration through small pores. The effects of vimentin and cell stiffness on 3D migration and the relation between cell stiffness and migration speed indicates that vimentin is important for resisting compressive stresses and maintaining the structural integrity of the cell during migration. Vimentin's role in cell deformability and migration may therefore have wide ranging implications for nuclear damage and repair, genome expression, cell cycle, and cell fate, which are important to the maintenance of tissues and progression of diseases, including cancer. Overall, our findings demonstrate that vimentin's perinuclear stiffness controls 3d motility and provides new insight into how cells might alter their cytoskeleton to maximize invasion in vivo without compromising cell integrity. Methods Cell Culture Wild-type mouse embryo fibroblasts and vimentin-null mEF were kindly provided by J. Ericsson (Abo Akademi University, Turku, Finland) and maintained in DMEM with 25 mM HEPES and sodium pyruvate (Life Technologies; Grand Island, NY) supplemented by 10% fetal bovine serum, 1% penicillin streptomycin (Gibco), and nonessential amino acids (Life Technologies). For NIH-3T3 fibroblasts (American Type Culture Collection, Manassas, VA), 10 cells were maintained in DMEM (Gibco) with 10% fetal bovine serum (ATCC) and 1% penicillin streptomycin, and 25 mM HEPES was added to media for micro-channel experiments. All cell cultures were maintained at 37°C and 5% CO2. Immunofluorescence Cells were fixed for immunofluorescence using 4% paraformaldehyde for 30 min at 37°C, permeabilized with 0.05% Triton X-100 solution in PBS (15 min., room temperature RT), and saturated with 1% serum albumin bovine (30 min, RT). For vimentin visualization, cells were incubated with primary anti-vimentin monoclonal rabbit antibody (1:200, RT, Abcam ab92547) or primary anti-vimentin polyclonal chicken antibody (1:200, RT, Novus NB300-223); secondary antibodies were anti-Rabbit Alexa Fluor 488 (1:1000, RT, Invitrogen A-11008) or anti-chicken Alexa Fluor 488 (1:1000, RT, Invitrogen A-11039). For visualizing microtubules, we use primary anti-tubulin monoclonal rat antibody (1:200, Serotec MCA77G) and secondary anti-rat AlexaFluor647 (Invitrogen A-21247). For immunostaining cells in micro-channels and on capillaries, primary antibodies were diluted to 1:1000 and kept overnight at 4°C. Cells were washed and stained with Rhodamine phalloidin 565 (Life Tech. r415) or Texas Red phalloidin and Hoechst 33342 (Molecular probes H-1399) or DAPI (Molecular probes) for 1 hr according to manufacturer's instructions. Cells were imaged with a Leica DMIRE2 inverted microscope with either a 40x (0.55 NA) air objective lens or 63x (0.70 NA) air objective lens. Microfluidic device fabrication and operation Fabrication - The microfluidic devices were built using standard soft lithography techniques and designed to prevent pressure gradients across the channels as described in Irimia, et al. 39. The positive silicon master was generated by spinning KMPR 2010 onto silicon wafers (Wafer World Inc., West Palm Beach FL) to create a 10 μm thick layer. The photoresist was soft baked for 5 min at 95°C and exposed to UV light through a chrome mask (CAD/Art Services, Inc., SI Fig. 2) with a mask aligner (ABM-USE, Inc., ABM3000HR). Unexposed KMPR2010 was developed with SU-8 developer and rinsed with isopropanol. This process was repeated 2 times in order to stack the three layers of the device that were aligned with a mask aligner. The positive wafer obtained was silanized (Sigma 448931) in a vacuum chamber overnight. A PDMS (Sylgard 184) solution at a 1:10 ratio (curing:elastomer) was mixed and degassed. This solution was then poured over the silanized positive silicon wafer and baked 90 min at 80°C to generate a negative mold. The negative mold was silanized overnight in a vacuum chamber. This protocol was repeated to obtain the silanized positive mold that was used to build the microfluidic device. Finally, A PDMS mixture was degassed, poured over the PDMS positive molds, baked for 90 min at 80°C, and removed from the mold. The device was punched with 0.5 mm access holes for tubing inserts. Channels were sealed with a glass microscope slide using an oxygen plasma chamber (Femto-Diener Electronic). Operation- To sterilize and clean the device, it was flushed with a solution of 70% (vol/vol) ethanol in deionized water (diH20), followed by rinsing with sterile diH20. Next, the device was submerged in sterile diH20r and degassed to remove bubbles. Channels were then rinsed with phosphate buffered saline (PBS) and coated with surface ligands by pumping in a 50 ug/mL solution of either collagen 1 (BD Biosciences, Franklin Lakes, NJ) or fibronectin (purified from Salmon plasma) in PBS and incubating for 1 hr at 37°C. Finally, channels were then washed 3x 11 with PBS, filled with cell culture media, and incubated at 37 °C for at least 30 min before seeding cells. Cell culturing and seeding- Cells were trypsinized using 0.5% trypsin (GIBCO) at 37°C, centrifuged to remove trypsin, and re-suspended in cell media at densities of 10 million cells/mL. Using a hand-held syringe (Hamilton Company, 81320) and tubing (Hamilton Company, 90622), cells were gently pumped into the device inlet. Cells were preferentially placed near the opening of the channel constrictions by manually tilting the device for 2 to 4 min and allowing gravity to pull cells in suspension toward the constrictions (SI Fig. 2&3). Fluid reservoirs (barrel-less syringes (BD, Ref 309657) containing cell media) were connected to the channel outlets with tubing and arranged to ensure no pressure driven flow through the channels. The device was kept in a Tokai-Hit Imaging Chamber (Tokai Hit, Shizuoka-ken, Japan) and maintained at 37°C and 5% CO2. Cells were allowed to adhere to channel surfaces for approximately 20-40 min before time-lapse imaging began. Time-lapse imaging was performed with a Leica DMIRE2 inverted microscope in bright-field with a 10x (0.3 NA) air lens. Images were taken every 4 min for 12-21 hr using an ASI x/y/z stage (MS – 2000, Applied Scientific Instrumentation) to capture multiple positions in the device. Capillary fabrication and operation Fabrication - Capillaries having diameters ranging from 10-20 um were pulled from larger borosilicate glass capillaries with a diameter of 1.6 mm (Richland Glass, Richland, NJ) using a Narishige PB-7 pipette puller. Cell culture chambers (length, 60 mm, width 240mm, height, 12mm) were printed using either ABS or PLA plastic (Biomedical library, University of Pennsylvania). To allow for visualization of the cells, windows were designed into the upper and lower sections of the chambers. Capillaries were affixed to the inner surface of the chambers using UV-curable glue (NOA 68, Norland Products). Operation- Capillaries were cleaned by rinsing once with ethanol, then rinsing with deionized H20, and then air dried. Chambers containing capillaries were then placed in a plasma cleaner (Harrick, PDC-32G) and exposed to air plasma for 5 mins. Overnight incubations with either 0.1 mg/ml collagen or fibronectin were then made while shaking on an orbital shaker. Capillaries were then rinsed 3X with PBS, sterilized under UV for 1 hr, and then incubated with culture media for at least 10 mins before cell seeding. Cell culturing and seeding - Cells were trypsinized using 0.5% trypsin (GIBCO) at 37°C then centrifuged to remove trypsin. Resuspended cells at a density of 2.5 x 105 cells/ml were added to the capillaries and allowed to attach for 60 mins at 37°C. Cell culture chambers were then transferred to a Tokai-Hit Imaging Chamber mounted on an ASI x/y/z stage and maintained at 37°C and 5% CO2. Time-lapse images were taken at multiple positions every 10 mins for 18-21 hrs using a 40x objective. Three-dimensional (3D) collagen gel preparation and imaging Gel preparation - Collagen gels (2 mg/mL) were prepared with final concentrations of 300 000 cells/mL by mixing together the following reagents in the order listed: pelleted and counted cells in media (10% v/v), 5x DMEM (20% v/v), FBS (10% v/v), 0.1 M NaOH (10% v/v), and 4 mg/mL collagen type 1 (Corning, REF 354236, 50 %v/v). Reagents are kept cold on ice while 12 Cell speed in micro-channels, glass slides, & capillaries - Cell trajectories 𝒓(𝑥,𝑦,𝑡) were determined over time 𝑡 as 𝒗(𝑡)=[𝐫(𝑡+∆𝑡)− 𝒓(𝑡)]/∆𝑡, where ∆𝑡 was approximately 30 min. with the more random motion in 2D (glass slides), we chose the maximum value of 𝒗(𝑡) as the cell speed was determined as the cell displacement over ∆𝑡 = 4 hr, a time step large enough to mixing. One mL of mixture was added to 20 mm dishes and maintained at 37°C and 5% CO2. Cells were imaged in bright field, 24 hr after seeding in gel (Fig. 1). Imaging nucleus for tracking in gels - To track cells in the 3D gels over time, the nuclei of vim +/+ and vim -/- mEF were fluorescently labeled with NLS-GFP. For these experiments, cells were transiently transfected with pEGF-C1-NLS. Forty-eight hr after transfection, cells were cultured in the collagen gel. Twenty-four hr later, cell nuclei were imaged at 10 min increments for 17 hr by using fluorescence microscopy and a 10x (0.3 NA) air lens. Cell area and motility measurements Cell area - The projected cell area changed over time as the cell entered and moved through the channel (SI Fig. 8). Thus, we measured cell area at three designated points: (i) one hr before cells entered the channel, (ii) at the middle of the channel length, and (iii) one hr after cell exited the channel. Cell area was determined by manually tracing the periphery of single cells (N = 25- 35 cell per condition) using ImageJ software (ImageJ Software, NIH, Bethesda, MD). determined by tracking the center of mass of cells at either 4 or 10 min. increments using ImageJ Software (NIH) and the Manual Tracking plugin (https://imagej.nih.gov/ij/). Here, cell speed was To compare the directed cell motion in the 3D environments (micro-channels and capillaries) measure of cell speed for each cell in each experimental condition. For average cell speed and persistence in the micro-channels and glass slides, see SI Fig. 5. Cells that divided or moved out of the frame of view were excluded. Cell speed in collagen gels – Because cell migration in the fibrillar collagen gels was limited, measure displacements (≈ 8 μm) but smaller than the cell persistence length: this method yielded approximately 50 measures of speed per cell by sampling over multiple times in a 17 hr video. Cells observed to move out of the field of view, divided, or died were excluded. Experiments were conducted three times for a total of 22-24 cells per condition. Cell stiffness measurements Cell peri-nuclear stiffness measurements were performed using atomic force microscopy (NanoWizard 4, JPK) equipped with a liquid cell and temperature control setup. Silicon nitride cantilevers (ORC8, Bruker) with nominal 0.1 N/m spring constant and tip half-opening angle of 36° ± 2° were used for cell nanoindentation. Quantitative characterization of nanomechanical properties of the cells was realized by recording of multiple force vs distance curves in the peri-nuclear region (Fig. 4ii) with the constant force of 3 nN and indentation rate equal 0.4 Hz. Modified Hertz model was fitted to the data and Young`s modulus of each point was calculated as described previously 40. Actin and microtubule inhibitors To determine the effects of microtubule and actin inhibitors on vim +/+ and vim -/- mEF migration in the micro-channels, the cells were treated with either the microtubule inhibitor nocodazole or actin inhibitor cytochalasin D (SI Fig. 6 & 7) while using time-lapse microscopy. Prior to administering inhibitors, cells were seeded in 10 μm collagen- 13 coated micro-channels for three to six hr. Then, cells were treated with either 1 μg/mL nocodazole (Sigma) or 1 μg/mL cytochalasin D (Sigma) by diluting the reagents in cell media and gently pumping the solution through the micro-fluidic device with a hand-held syringe and tubing connected to the device inlet. Cell migration was observed for the following 17 hr, and the resultant average speed of the cells was determined for 8-13 cells per condition (1 experiment each). Control experiments with DMSO only were performed for each cell line. To ensure drugs did not diffuse out of the channel, a test using 40 nm fluorescent particles (Invitrogen F8770) as tracers was used to confirm stable concentrations during the experiment (SI Fig. 7). We found that nocodazole and cytochalasin D treatment decreases cell speed for both vim +/+ and vim -/- mEF in the micro-channels (SI Fig. 6). Transwell migration assays Cells were seeded at sub-confluent concentrations (10-20 thousand cells/cm2) on polycarbonate Transwell membranes with pore diameters of 3 μm (Corning, CLS3414), 5 μm (Corning, CLS3421), and 8 μm (Corning, CLS3428). Membranes were pre- coated with either collagen 1 (50 ug/mL) or fibronectin (50 ug/mL). After seeding cells, the filters were maintained at 37°C and 5% CO2 for 15 hours. Cells were gently removed from either the top or bottom of the membrane with a cotton swab and immediately fixed with paraformaldehyde. To determine the number of cells per unit area, cells were stained with crystal violet and imaged at multiple locations across the membrane with a 10x objective. Cells were manually counted in 800x800 μm2 fields of view (12-30 locations per condition). The percentage of cells that cross the membrane (Fig. 5) was then determined by taking the ratio of the number of cells on the bottom of the membrane and the sum of cells on the filter top and bottom. Estimate of compressional cell strain – To estimate the cell strain through the Transwell filters, we assume that the cells maintain a constant volume of 1.76 pL (equivalent to a sphere of diameter 15 μm). This volume is less than the value determined by multiplying the average cell spread area by the channel height, which yields an overestimate of 2.1 +/- 0.2 pL for vim +/+ mEF and 2.4 +/- 0.2 pL for vim -/- mEF (SI Fig. 8). Compressional cell strain is then estimated as the magnitude of (L1-L)/L, where L is the cell size in the unstressed spherical state (15 μm) and L1 is the narrowest dimension of the cell while crossing the pore, the diameter of the pore (c.f. Fig. 5). Statistical Analysis Data presented as mean values ± standard errors (SE). Each experiment was performed a minimum of two times unless otherwise stated. The unpaired Student's t-test with two tails at the 95% confidence interval was used to determine statistical significance. Denotations: *, p <= 0.05; **, p < 0.01; ***, p < 0.001; ns, p > 0.05. The Fisher's exact test was used to confirm statistical significance between the proportion of cells that exit the channels or exhibit twirling behavior (Fig. 2c). SUPPLEMENTARY MOVIES SI Movie 1: Spontaneous migration of wild-type mEF crossing micro-channel. Wild-type mEF migrating through a 10x10 μm2 micro-fluidic constriction, coated with collagen-I. The length of the video is 9.3 hr. Images are taken at 4-minute increments with a 10x air objective. 14 SI Movie 2: Spontaneous migration of vimentin-null mEF crossing micro-channel. Vimentin-null mEF migrating through a 10x10 μm2 micro-fluidic constriction, coated with collagen-I. The length of the video is 6.3 hr. Images are taken at 4-minute increments with a 10x air objective. SI Movie 3: Wild-type mEF that does not cross channel. Wild-type mEF interacting with a micro-fluidic constriction. The cell does not enter the channel, but instead switches directions. Channel coated with collagen-I. The length of the video is 9.2 hr. Images are taken at 4-minute increments with a 10x air objective. SI Movie 4: Twirling motions of wild-type mEF in micro-channel. Wild-type mEF migrating through a 10x10 μm2 micro-fluidic constriction, coated with fibronectin. The cell appears to twirl around the channel, moving in a three-dimensional spiral. The length of the video is 6.1 hr. Images are taken at 4-minute increments with a 10x air objective. References 1. Yang, J. & Weinberg, R. A. Epithelial-Mesenchymal Transition: At the Crossroads of Development and Tumor Metastasis. Developmental Cell 14, 818–829 (2008). 2. Thiery, J. P., Acloque, H., Huang, R. Y. J. & Nieto, M. A. Epithelial-Mesenchymal Transitions in Development and Disease. Cell 139, 871–890 (2009). 3. Lamouille, S., Xu, J. & Derynck, R. Molecular mechanisms of epithelial– mesenchymal transition. Nat Rev Mol Cell Biol 15, 178–196 (2014). 4. Abreu-Blanco, M. T., Watts, J. J., Verboon, J. M. & Parkhurst, S. M. Cytoskeleton responses in wound repair. Cell. Mol. Life Sci. 69, 2469–2483 (2012). 5. Friedl, P. & Gilmour, D. Collective cell migration in morphogenesis, regeneration and cancer. Nat Rev Mol Cell Biol 10, 445–457 (2009). 6. Paul, C. D., Mistriotis, P. & Konstantopoulos, K. Cancer cell motility: lessons from migration in confined spaces. Nat Rev Cancer 17, 131–140 (2016). 7. Hay, E. D. The mesenchymal cell, its role in the embryo, and the remarkable signaling mechanisms that create it. Dev. Dyn. 233, 706–720 (2005). 8. Satelli, A. & Li, S. Vimentin in cancer and its potential as a molecular target for cancer therapy. Cell. Mol. Life Sci. 68, 3033–3046 (2011). 9. Friedl, P. & Bröker, E.-B. T Cell Migration in Three-dimensional Extracellular Matrix: Guidance by Polarity and Sensations. Developmental Immunology 7, 249–266 (2000). 10. Lugassy, C. & Barnhill, R. L. Angiotropic Melanoma and Extravascular Migratory Metastasis. Advances in anatomic pathology 14, 195–201 (2007). 11. Doerschuk, C. M., Beyers, N., Coxson, H. O., Wiggs, B. & Hogg, J. C. Comparison of neutrophil and capillary diameters and their relation to neutrophil sequestration in the lung. Journal of Applied Physiology 74, 3040–3045 (1993). of collective melanoma cell invasion. IntraVital 1, 32–43 (2014). 12. Weigelin, B., Bakker, G.-J. & Friedl, P. Intravital third harmonic generation microscopy 15 13. Wolf, K. et al. Collagen-based cell migration models in vitro and in vivo. Seminars in Cell 14. Davidson, P. M., Denais, C., Bakshi, M. C. & Lammerding, J. Nuclear Deformability & Developmental Biology 20, 931–941 (2009). Constitutes a Rate-Limiting Step During Cell Migration in 3-D Environments. Cel. Mol. Bioeng. 7, 293–306 (2014). 15. Harada, T. et al. Nuclear lamin stiffness is a barrier to 3D migration, but softness can limit survival. J Cell Biol 204, 669–682 (2014). 16. Denais, C. et al. Nuclear envelope rupture and repair during cancer cell migration. Science 352, 353–358 (2016). 17. Raab, M. et al. ESCRT III repairs nuclear envelope ruptures during cell migration to limit DNA damage and cell death. Science 352, 359–362 (2016). Irianto, J. et al. DNA Damage Follows Repair Factor Depletion and Portends Genome 18. Variation in Cancer Cells after Pore Migration. Current Biology 27, 210–223 (2017). 19. Goldman, R. D., Grin, B., Mendez, M. G. & Kuczmarski, E. R. Intermediate filaments: versatile building blocks of cell structure. Current Opinion in Cell Biology 20, 28–34 (2008). 20. Eckes, B. et al. Impaired mechanical stability, migration and contractile capacity in vimentin- deficient fibroblasts. J Cell Sci 111, 1897–1807 (1998). 21. Guo, M. et al. The Role of Vimentin Intermediate Filaments in Cortical and Cytoplasmic Mechanics. Biophysical Journal 105, 1562–1568 (2013). 22. Mendez, M. G., Restle, D. & Janmey, P. A. Vimentin Enhances Cell Elastic Behavior and Protects against Compressive Stress. Biophysical Journal 107, 314–323 (2014). 23. Murray, M. E., Mendez, M. G. & Janmey, P. A. Substrate stiffness regulates solubility of cellular vimentin. Molecular biology of the cell 25, 87–94 (2013). 24. Janmey, P. A., Euteneuer, U., Traub, P. & Schliwa, M. Viscoelastic Properties of Vimentin Compared with Other Filamentous Biopolymer Networks. Journal of Cell Biology 113, 155–160 (1991). 25. Kreplak, L., Bär, H., Leterrier, J. F., Herrmann, H. & Aebi, U. Exploring the Mechanical Behavior of Single Intermediate Filaments. Journal of Molecular Biology 354, 569–577 (2005). Pinner, S. & Sahai, E. Imaging amoeboid cancer cell motility in vivo. Journal of 26. Microscopy 231, 441–445 (2008). Irimia, D. & Toner, M. Spontaneous migration of cancer cells under conditions of 27. mechanical confinement. Integr. Biol. 1, 506–8 (2009). 28. Lang, N. R. et al. Estimating the 3D Pore Size Distribution of Biopolymer Networks from Directionally Biased Data. Biophysical Journal 105, 1967–1975 (2013). 29. Helfand, B. T. et al. Vimentin organization modulates the formation of lamellipodia. Molecular biology of the cell 22, 1274–1289 (2011). 30. Liu, Y.-J. et al. Confinement and Low Adhesion Induce Fast Amoeboid Migration of Slow Mesenchymal Cells. Cell 160, 659–672 (2015). 31. Carey, S. P. et al. Comparative mechanisms of cancer cell migration through 3D matrix and physiological microtracks. Am J Physiol Cell Physiol 308, C436–C447 (2015). 32. Balzer, E. M. et al. Physical confinement alters tumor cell adhesion and migration phenotypes. The FASEB Journal 26, 4045–4056 (2012). 16 33. Mutsaers, S. E., Bishop, J. E., McGrouther, G. & Laurent, G. J. Mechanisms of tissue repair: from wound healing to fibrosis. The International Journal of Biochemistry & Cell Biology 29, 5–17 (1997). 34. Humphrey, J. D., Dufresne, E. R. & Schwartz, M. A. Mechanotransduction and extracellular matrix homeostasis. Nat Rev Mol Cell Biol 15, 802–812 (2014). 35. Biton, Y. Y. & Safran, S. A. The cellular response to curvature-induced stress. Phys. Biol. 6, 046010–9 (2009). Jiu, Y. et al. Vimentin intermediate filaments control actin stress fiber assembly through 36. GEF-H1 and RhoA. Journal of Cell Science 130, 892–902 (2017). 37. Rolli, C. G., Seufferlein, T., Kemkemer, R. & Spatz, J. P. Impact of Tumor Cell Cytoskeleton Organization on Invasiveness and Migration: A Microchannel-Based Approach. PLoS ONE 5, e8726–8 (2010). Patteson, A. E. et. al. Forthcoming Irimia, D., Charras, G., Agrawal, N., Mitchison, T. & Toner, M. Polar stimulation and constrained cell migration in microfluidic channels. Lab Chip 7, 1783–8 (2007). Pogoda, K. et al. Depth-sensing analysis of cytoskeleton organization based on AFM data. Eur Biophys J 41, 79–87 (2011). 38. 39. 40. Acknowledgements This work was supported by the National Institutes of Health-National Institute of General Medical Sciences (P01 GM096971) and National Science Center, Poland under Grant: UMO-2015/17/B/NZ6/03473. We thank Robert Goldman for kindly sharing the pEGF-C1-NLS plasmid construct and Dennis Discher and Manu Tewari for help amplifying the plasmid. We also thank Eric Johnston and Nathan Bade for help developing the micro-fluidic device and Mateusz Cieśluk for his technical assistance during AFM experiments. Author Contributions A.E.P. designed, performed, and analyzed motility experiments using micro-fluidics devices, collagen gels, and Transwell membrane assays. F.B. designed and performed capillary experiments and analysis. K.P., R.B., and P.D designed and performed AFM measurements and analysis and K.P. assisted in collagen gel preparation. E.C. assisted with cell culture, and E.C. and P.G. developed micro-fluidic device. A.E.P., F.B., K.P., E.C., P.G., and P.A.J. contributed to project design and wrote the manuscript. 17
1112.1281
1
1112
2011-12-06T14:09:17
The stochastic evolution of a protocell. The Gillespie algorithm in a dynamically varying volume
[ "physics.bio-ph", "q-bio.CB" ]
In the present paper we propose an improvement of the Gillespie algorithm allowing us to study the time evolution of an ensemble of chemical reactions occurring in a varying volume, whose growth is directly related to the amount of some specific molecules, belonging to the reactions set. This allows us to study the stochastic evolution of a protocell, whose volume increases because of the production of container molecules. Several protocells models are considered and compared with the deterministic models.
physics.bio-ph
physics
The stochastic evolution of a protocell. The Gillespie algorithm in a dynamically varying volume. by T. Carletti and A. Filisetti Report naXys-22-2011 6 12 2012 Namur Center for Complex Systems University of Namur 8, rempart de la vierge, B5000 Namur (Belgium) http://www.naxys.be THE STOCHASTIC EVOLUTION OF A PROTOCELL. THE GILLESPIE ALGORITHM IN A DYNAMICALLY VARYING VOLUME. T. CARLETTI AND A. FILISETTI Abstract. In the present paper we propose an improvement of the Gillespie algorithm allowing us to study the time evolution of an ensemble of chemical reactions occurring in a varying volume, whose growth is directly related to the amount of some specific molecules, belonging to the reactions set. This allows us to study the stochastic evolution of a protocell, whose volume increases because of the production of container molecules. Several protocells models are considered and compared with the deterministic models. 1. Introduction All known life forms are composed of basic units called cells; this holds true from the single-cell prokaryote bacterium to the highly sophisticated eucaryotes, whose existence is the result of the coordination, in term of self-organization and emergence, of the behavior of each single basic unit. While present day cells are endowed with highly sophisticated regulatory mechanisms, that represent the outcome of almost four billion-years of evolution, it is generally believed that the first life-forms were much simpler. Such primordial life-bricks, the protocells, were most probably exhibiting only few simplified functionalities, that required a primitive embodiment structure, a protometabolism and a rudimentary genetics, so to guarantee that offsprings were "similar" to their parents [1, 15, 17]. Intense research programs are being established aiming at obtaining protocells capable of growth and duplication, endowed with some limited form of genetics [12, 13, 14, 17]. Despite all efforts, artificial protocells have not yet been reproduced in laboratory and it is thus extremely impor- tant to develop reference models [3, 10, 14, 16] that capture the essence of the first protocells appeared on Earth and enable to monitor their subsequent evolution. Due to the uncertainties about the details, high-level abstract models are particularly relevant. Quoting Kaneko [7] it is necessary to consider "simplified models able to capture universal behaviors, without carefully adding complicating details". Most of the models present in the literature are based on deterministic differential equations governing the evolution of the concentrations of the involved reacting molecules. Even if the results are worth discussing and provide important insights, it should be stressed that the former assumptions are rarely satisfied in a cell [5]. Firstly, the number of involved molecules is small and should be counted by integer numbers, hence the use of the concentrations can be questioned; secondly, the presence of the thermal noise introduces in the system a degree of stochasticity than cannot be trivially encoded by a differential equation, mostly because this makes the time evolution a stochastic process. One possible way to overcome such difficulties is to use the Chemical Master equation: given the present state of the system, namely the number of available molecules for each species, and the possible reactions among them, one can compute the transition probabilities to reach and leave the given state and thus get a partial differential equation describing the time evolution of the probability distribution of having a given number of molecules at any future times [5, 6]. Analytically solving the resulting equation is normally a very hard task, one should thus resort to use numerical methods. A particularly suitable one is the algorithm presented by Gillespie [5, 6], allowing to determine, as a function of the present state of the system, the most probable reaction and the most probable reaction time, i.e. the time at which such reaction will occur. Date: September 14, 2018. 1 2 T. CARLETTI AND A. FILISETTI Let us however observe that in the setting we are hereby interested in, the chemical reactions occur in a varying volume, because of the protocell growth; we thus need to adapt the Gillespie method to account for this factor. To the best of our knowledge, there are in the literature very few papers dealing with the Gillespie algorithm in a varying volume [8, 9]. Moreover in all these papers, the volume variation can be considered as an exogenous factor, not being directly related to the number of lipids forming the protocell membrane. So our main contribution is to improve the Gillespie algorithm taking into account the protocell varying volume which is moreover consistent with the increase of the number of lipids constituting the protocell membrane. The paper is organized as follow. In Section 2 we briefly recall the Surface Reaction Models of protocell, that would be used to compare our stochastic numerical scheme. Then in Sections 3 and 4 we will present our implementation of the Gillespie algorithm in a dynamically varying volume. Finally in Section 5 we will present some applications of our method. 2. Surface Reaction Models Among the available models for protocells, a particularly interesting one is the Surface Reactions Model [3, 16], SRM for short, and its applicability to the synchronization problem. Such model is roughly inspired by the Los Alamos bug hypothesis [14, 15] but which, due to its abstraction level, the SRM can be applied to a wider set of protocell hypotheses. The SRM is build on the assumption that a protocell should comprise at least one kind of "container" molecule (typically a lipid or amphiphile), hereby called C molecule, and one kind of replicator molecule - loosely speaking "genetic material", hereafter called Genetic Memory Molecule, GMM for short, and named with the letter X. There are therefore two kinds of reactions which are crucial for the working of the protocell: those which synthesize the container molecules Eq. (1) and those which synthesize the GMM replicators Eq. (2) (1) (2) Xi + Li αi GGGGGA Xi + C , Xi + Pj Mij GGGGGGGA Xi + Xj . In both cases Li and Pj are the buffered precursors, respectively of container molecules and of the j -- th GMM, while αi and Mij are the reaction kinetic constants. A second main assumption of the SRM, is that such reactions occur on the surface of the protocell, exposed to the external medium where precursors are free to move. Hence, as long as container molecules are produced, they are incorporated in the membrane that thus increases its size, until a critical point at which, due to physical instabilities, the membrane splits and two offsprings are obtained, each one getting half of the mother's GMMs and whose size is roughly half that of the mother just before the division. Under the previous assumptions and in the deterministic setting, one can prove [3, 16] that the number of membrane molecules and the number of GMMs evolve in time according to: where ⃗X = (X1, . . . , XN) represents the amount of each GMM, ⃗α = (α1, . . . , αN) is the vector of the reaction constants responsible for the production of C molecules from the X molecules plus some appropriate precursor. (Mij) denotes the reaction constant at which Xi is produced by Xj plus some precursor. β ∈ [2~3, 1] is a geometrical shape factor that relates the surface to the volume of the protocell and ρ is the lipid density (for more details the interested reader can consult [3, 16]). Let us observe that in this setting the precursors are assumed to be buffered and thus their amount to be constant, hence the latter can be incorporated into the constants α and M . the protocell will grow until some time t0 + ∆T1 at which the amount of C molecules has doubled So starting with a initial value of container molecules, C(t0) = C0, and of GMMs, ⃗X(t0) = ⃗X0, with respect to the initial value, C(t0 + ∆T1) = 2C0 and thus the protocell undergoes a division. (3) ⎧⎪⎪⎪⎨⎪⎪⎪⎩ ρ β−1 ⃗α ⋅ ⃗X C =  C ⃗X = C β−1M ⋅ ⃗X , STOCHASTIC EVOLUTION OF A PROTOCELL. 3 Each offspring will get half of the GMMs the mother protocell had just before the division, ⃗X (1) = ⃗X(t0 +∆T1)~2. And the protocell cycle starts once again. One can prove [3, 16] that under suitable conditions ⃗X (n) tends to a constant value once n goes to infinity, implying thus the emergence of synchronization of growth and information production. 3. The method the probability that a random combination of molecules from channel Rµ will react in the interval Let us observe that because of the protocell growth the volume is an increasing function of time. Actually one can relate the volume to the amount of container molecules via their density will hereby use the same symbol Xi to denote both the i -- th GMM and the integer number of molecules of type Xi in the system. Let us now improve the previous scheme by introducing a probabilistic setting `a la Gillespie. We thus consider a protocell made by a lipidic vesicle and containing a well stirred mixture of N GMMs, X1, . . . , XN , that may react through m elementary reaction channels Rµ, µ = 1, . . . , m, running within the volume V(t) of the protocell. V = C~ρ where C denotes the integer number of molecules forming the lipidic membrane. We For each reaction channel Rµ assume that there exists a scalar rate cµ such that cµdt + o(dt) is [t, t + dt) within the volume V(t). Let hµ(Y) be the total number of possible distinct combinations of molecules for a channel Rµ when the system is in state Y =(X1, . . . , XN , C), then we can define the propensity [9] of the reaction Rµ to be aµ(Y) = hµ(Y)cµ. One can prove [5] that for a binary reaction the rate cµ can be written in the form cµ = kµ~V , where kµ is a fixed constant. Similarly one can prove that for a reaction involving n different species, we get: cµ = kµ~V n−1. And thus for a single molecule reaction, i.e. a decay, we get cµ = kµ, namely independently from the volume. Let us now assume that among the m reactions, Q1 involve one single molecule, Q2 are binary reactions, Q3 are ternary reactions and so on. Of course Q1 + Q2 + ⋅ ⋅ ⋅ + QN +1 = m. We recall that we have N GMMs and the container type molecule C, hence N + 1 species. For short we will denote Q1 the set of indices µ for mono molecule reactions, and by Q the remaining ones. Let us observe that in this way some coefficient aµ, will depend both on the system state Y and on the More precisely to study the time evolution of the system we need to determine the probability Pµ(τY, t)dτ = Pnot(τY, t) × aµ(Y, t + τ)dτ , time via the volume V(t): aµ(Y, t) for µ ∈ Q. Pµ(τY, t)dτ , that given the system in the state Y = (X1, . . . , Xn, C) at time t, then the next reaction will occur in the infinitesimal time interval (t + τ, t + τ + dτ) and it will be the reaction where Pnot(τY, t) is the probability that no reaction occurs in (t, t + τ) given the state Y at time t, whereas the rightmost term denotes the probability to have a reaction Rµ in (t + τ, t + τ + dτ) To compute the first term Pnot, let us take s ∈[t, t + τ] and observe that: Pnot(s + dsY, t) = Pnot(sY, t)Pnot(dsY, t + s) = Pnot(sY, t)⎛⎝1 − Q aµ(Y, t + s)ds⎞⎠ , being 1 − ∑µ aµ(Y, t + s)ds the probability that no reaction will occur in (t + s, t + s + ds) once equation, passing to the limit ds → 0, and observing that Pnot(0Y, t) = 1, we get the solution: we are in state Y at time t + s. Thus rewriting the previous difference equation as a differential Rµ. This probability will be computed as given the state Y at time t + τ . (4) µ Pnot(τY, t) = exp−AQ1(Y)τ −S τ AQ1(Y) = Q aµ(Y) and AQ(Y, s + t) = Q AQ(Y, s + t) ds , aµ(Y, s + t) . µ∈Q1 µ∈Q 0 (5) where The apparent asymmetry in the exponential term in (5) is easily recovered by observing that AQ1(Y)τ = ∫ τ 0 AQ1(Y)ds. 4 T. CARLETTI AND A. FILISETTI We can thus conclude that (6) Pµ(τY, t)dτ = exp−AQ1(Y)τ −S t+τ t AQ(Y, s) ds aµ(Y, t + τ)dτ . Let us observe that the rightmost term is correctly aµ(Y, t + τ), namely the system is still in the state Y at time t + τ , because no reaction has been produced in (t, t + τ). Let us recall that the volume enters in the previous relation via the function AQ, more explicitely one has AQ(Y, s) = Q hµ(Y)kµ V(s) + Q hµ(Y)kµ (V(s))2 µ∈Q2 (7) that can be rewritten in terms of C molecules using the relation C = ρV . So our method applies to a different problem with respect to the one considered in [9], in fact in our case the volume growth is not imposed a priori but dynamically evolves accordingly to the reaction scheme, if C is produced then V increases otherwise it will keep a constant value, while in [9] the volume growth is an exogenous variable. µ∈QN +1 µ∈Q3 , hµ(Y)kµ (V(s))N + ⋅ ⋅ ⋅ + Q 4. The stochastic simulation algorithm in a growing volume the time evolution given by the model defined above. Given the system in some state Y at time t, we must determine the interval of time τ and Once we have the probability function Pµ(τY, t) we can build an algorithm that reproduces the reaction channel Rµ according to the probability distribution function Pµ(τY, t), and finally update the state Y → Y + νµ, where νµ is a stoichiometric vector representing the increase and decrease of molecular abundance due to the reaction Rµ. This will be accomplished following the standard approach by Gillespie [5] but taking care of the time dependence of the propensities. We will thus need to compute the cumulative probability distribution function and then make use of the inversion method [6], to determine the channel µ and the next reaction time τ , distributed according to Pµ(τY, t). (8) From (6) we can compute the cumulative distribution function F(τY, t) = S τ 0 Q µ Pµ(sY, t) ds , providing the probability that any reaction will occur in(t, t + τ) starting from the state Y at time t. The function F(τY, t) can be explicitely computed by F(τY, t) = 1 − exp−AQ1(Y)τ −S t+τ Proposition 4.1. Under the above assumptions we have AQ(Y, s) ds . (9) t Proof. The first step is to use (6) and perform a sum over all the channels µ to rewrite (8) as F(τY, t) = S τ Then we can observe that 0 (AQ1(Y) + AQ(Y, t + s)) exp−AQ1(Y)s − S t+s ∂sexp−AQ1(Y)s −S t+s AQ(Y, r) dr = ∂ t t AQ(Y, r) dr ds . and thus = −(AQ1(Y) + AQ(Y, t + s)) exp−AQ1(Y)s −S t+s F(τY, t) = −S τ ∂sexp−AQ1(Y)s − S t+s AQ(Y, r) dr , AQ(Y, r) dr ds ∂ 0 t t = 1 − exp−AQ1(Y)τ − S t+τ t AQ(Y, r) dr . (cid:3) STOCHASTIC EVOLUTION OF A PROTOCELL. 5 Once we have the cumulative distribution function we can obtain the value τ by drawing a radom number u1 from an uniform distribution in [0, 1] and then solve with respect to τ the implicit equation: u1 = 1 − exp−AQ1(Y)τ −S t+τ t AQ(Y, s) ds . Let us stress once again that this is not as straightforward as for the original Gillespie [5] scheme, or the simplified one presented in [9], because of the time dependence of AQ via the volume. One can nevertheless found suitable approximation for the integral, this will be the goal of the next sections. 4.1. The adiabatic assumption. Let assume that τ is very small, or which is equivalent, that the time scale of the chemical reactions involving the GMMs is much faster than the production of container molecules, hence the volume growth is very slow compared with the production of the chemicals Xi. Under this hypothesis one can assume that in the interval(t, t + τ) the volume doesn't vary and thus one can made the following approximation (10) (11) One can thus explicitely solve equation (10) to get: (12) τGill = − S t+τ t AQ(Y, s) ds ∼ AQ(Y, t)τ . AQ1(Y) + AQ(Y, t) log(1 − u1) , 1 that is the standard Gillespie result except now that AQ(Y, t) depends on time and as long the volume increases, then the contribution arising form AQ(Y, t) mights become smaller because AQ ∼ 1~V . 4.2. The next order correction. One can obtain a somehow better estimate valid in the case of comparable time scales for the reactions involving GMM and the container growth. The idea is to compute the integral in Eq. (10) using the following approximation: ∂AQ(Y, t) ∂t s + . . . ds S t+τ t (13) AQ(Y, s) ds = S τ 0 AQ(Y, t + s) ds = S τ ∂AQ(Y, t) ∂t 0 AQ(Y, t) + + O(τ 3) . τ 2 2 = AQ(Y, t)τ + C ⎛⎝ Q hµ(Y)kµ C(t) µ∈Q2 where ∂AQ(Y, t)~∂t can be obtained using the definition (7) and expressing the volume in terms of C = V(t)ρ, namely: ∂AQ(Y, t) = − + ⋅ ⋅ ⋅ + N Q + 2 Q C ∂t µ∈Q3 µ∈QN +1 To compute C~C we make the assumption that in a very short time interval, as the one we are interested in, the deterministic growth of the container is a good approximation for the stochastic underlying mechanism; this implies that we can use (3) Inserting the previous result into (13) and finally solving (10) with respect to τ , we can compute C C = C(t)ρ β−1 ⃗α ⋅ ⃗X(t) C(t) . (C(t))N ⎞⎠ . hµ(Y)kµ hµ(Y)kµ (C(t))2 the next reaction time up to correction of the order of τ 3, as: τGill = −(AQ1(Y) + AQ(Y, t)) +(AQ1(Y) + AQ(Y, t))2 − 2 log(1 − u1) AQ(Y, t) (14) , where we wrote for short AQ(Y, t) = ∂AQ(Y, t)~∂t and we selected the positive square root in such a way in the limit AQ(Y, t) → 0 we recover the previous solution (12). AQ(Y, t) 6 T. CARLETTI AND A. FILISETTI Remark 4.2 (On the existence of τGill). In the case of variable volume a new phenomenon can arise: the volume growth can be so fast that no reaction can occur in the interval (t, t + τ + dτ) for any τ . Mathematically this translates into a sign condition for the term under square root in (14), if: (15) then equation (10) has no real solution. log(1 − u1) <(AQ1(Y) + AQ(Y, t))2~(2 AQ(Y, t)) , This can be geometrically interpreted as follows. The relation (10) determines τGill as the intersection of the parabola −AQ1(Y) − AQ(Y, t)τ − AQ(Y, t)τ 2~2 with the horizontal line log(1 − u1), which is negative because u1 ∈ (0, 1). Such parabola intersect the y-axis at τ1 = 0 and τ2 = −2(AQ1(Y) + AQ(Y, t))~ AQ(Y, t) > 0 and it is concave. Then its absolute (negative) minimum is reached at the vertex τV = (t1 + t2)~2 and its value is (AQ1(Y) + AQ(Y, t))2~(2 AQ(Y, t)) and it is negative because AQ(Y, t) is negative. Hence if the horizontal line is below this value, i.e. condition (15) is verified, the parabola and the line do not have any real intersections (see Fig. 1). 0.06 0.04 0.02 0 −0.02 −0.04 −0.06 −0.08 −0.1 −0.12 −0.14 −0.2 τ1 τGill (τ1 + τ2)/2 τ2 log(1 −u1) 0 0.2 0.4 0.6 0.8 1 1.2 0.06 0.04 0.02 0 −0.02 −0.04 −0.06 −0.08 −0.1 −0.12 −0.14 −0.2 τ1 (τ1 + τ2)/2 τ2 0 0.2 0.4 0.6 0.8 1 1.2 log(1 −u1) Figure 1. Geometrical interpretation of the existence of the next reaction time τGill. Left panel : τGill is the smallest intersection between the parabola and the horizontal line log(1 − u1). Right panel τGill doesn't exist, the horizontal line is located below the minimum of the parabola. Let us also observe that, whenever it exists, τGill is always positive as it should be. In the case of a protocell the non existence of such next reaction time could be translated into the death by dilution of the protocell. 4.3. The next reaction channel. Whenever the next reaction time does exist, the next reaction channel is determined using the classical Gillespie method, namely by drawing a second uniformly (16) distributed random number u2 ∈[0, 1] and fix the channel µ such that: aν(Y, t + τ) , aν(Y, t + τ) ≤ u2a0(Y, t + τ) ≤ µQ ν=1 aν(Y, t + τ). where a0(Y, t + τ) = AQ1(Y) + AQ(Y, t + τ) = ∑m µ−1Q ν=1 ν=1 Remark 4.3. Let us observe that if all the reactions involve the same number of chemicals, then the determination of which reaction channel µ will be activated in the next reaction, doesn't depend on the volume which factorizes out from (16). In fact assuming all the reactions to involve p chemical, we obtain by definition and thus (16) rewrites: µ−1Q ν=1 aν(Y, t + τ) = hν(Y)kν hν(Y)kν [V(t + τ)]p ≤ u2 [V(t + τ)]p ∀ν ∈{1, . . . , m} , hν(Y)kν mQ [V(t + τ)]p , hν(Y)kν [V(t + τ)]p ≤ µQ ν=1 ν=1 which is clearly independent of the volume value V . STOCHASTIC EVOLUTION OF A PROTOCELL. 7 5. Some applications The aim of this section is to provide some applications of the previous algorithm to the study of the evolution of a protocell. 5.1. One single Genetic Memory Molecule. The simplest model is the one where only one GMM specie is present in the protocell [16] and thus only two chemical channels are active: channel 1, R1 (17) channel 2, R2 ∶ ∶ η X + P1 GGGGA 2X α X + L1 GGGGGA X + C , where P1 and L1 are, respectively, precursors of GMM, i.e. nucleotide, and precursors of am- phiphiles. One can thus compute the propensities in the state Y =(X, C) at time t: and a2(X, C, t) = h2(X, C) α a1(X, C, t) = h1(X, C) η P1X V(t) = η V(t) let us observe that we assume that precursors are buffered and thus they are constant. V(t) = α L1X V(t) ; (18) Because system (17) contains only bimolecular reactions, all the propensities are time depen- dent, hence AQ1 = 0 and AQ = a1(X, C, t) + a2(X, C, t) =(P1η + L1α)X~V(t), thus (10) simplifies into whose second order solution (14) is given by and ∂AQ(X, C, t) = − So we can finally obtain ∂t τGill = C L1αX  ρ provided u1 = 1 − exp−S t+τ t AQ(Y, s) ds , , L1αX AQ(Y, t) τGill = −AQ(Y, t) +(AQ(Y, t))2 − 2 log(1 − u1) AQ(Y, t) V(t)V(t) P1ηX V(t) V (t)=C(t)~ρ = − C V(t) + ¿ÁÁÀ C Cβ−1 Cβ−1 L1αX  ρ c2(β−1)(P1η + L1α) . 2α ρ log(1 − u1) ≥ − if u2(P1η + L1α)X namely 0 ≤ u2 ≤ + 2 C 2 P1η − ρ V 2 ρβ−1 ρL1αX 2 C 2 (P1η + L1α) . L1αρX 2(P1η + L1α) log(1 − u1) , Which reaction channel µ will active in the time interval[t, t + τ] can be obtained according to : if P1ηX V < u2(P1η + L1α)X V ≤ P1ηX ≤ (P1η + L1α)X V V P1η+L1α namely P1η P1η+L1α < u2 ≤ 1 then µ = 1 then µ = 2 . Let us observe that according to remark 4.3, the choice of µ doesn't depend on the volume, because only binary reactions are present. protocell splits into two offspring, almost halving the GMM amount. More precisely we assume that the first offspring will get a number of GMMs drawn according to a Binomial distribution Let C0 be the initial amount of container molecules, then we assume that once C(¯t) = 2C0 the with parameter p = 1~2 and n = X(¯t). From this step, for technical reason, only one randomly chosen offspring will be studied during each generation. In Fig. 2 we report a comparison between the deterministic (3) and the stochastic dynamics, under the adiabatic assumption for τGill, corresponding to the continuous growth phase of the container between two successive divisions. As one should expect, a system composed by a large 8 T. CARLETTI AND A. FILISETTI number of molecules exhibits small stochastic fluctuations whose average is not too far from the dynamics described by the deterministic model. Stochastic Integ ODE 1500 1000 1 X 500 0 0 2 4 time 6 8 x 10−3 2000 1800 1600 C 1400 1200 1000 0 Stochastic Integ ODE 2 4 time 6 8 x 10−3 b Figure 2. Stochastic vs ODE SRM protocell (3). Case of one GMM, left panel the time evolution of the amount of GMM, right panel the time evolution of the amount of C. Parameters are : η = 1, α = 1, L1 = 500, P 1 = 600, X1(0) = 100, C(0) = 1000, ρ = 200 and β = 2~3. In Fig. 3 we report the amount of GMM, X (k) (panel a), at the beginning of each protocell cycle and the duplication time (panel b), namely the interval of time needed to double the amount of C molecules, for both the stochastic and deterministic models. Once again one can clearly observe the small fluctuations of the stochastic system around the value obtained by the numerical integration of the deterministic description, Eq. (3). Let us observe that these fluctuations are due to the stochastic integrator scheme and also on the division mechanism. 1400 1300 1200 1100 ) k 1( X 1000 900 800 700 600 0 20 Stochastic Integ ODE 40 60 generation number (k) 3.5 x 10−3 3 k T ∆ 2.5 2 1.5 0 20 80 100 (a) Stochastic Integ ODE 40 60 generation number (k) 80 100 (b) Figure 3. Stochastic vs ODE SRM protocell (3). Case of one GMM, left panel the amount of GMM at the beginning of each division cycle, right panel the division time as a function of the number of elapsed divisions. Parameters are : η = 1, α = 1, L1 = 500, P 1 = 600, X1(0) = 100, C(0) = 1000, ρ = 200 and β = 2~3. We are now interested in studying the fluctuations dependence on the amount of molecules. We already know that for a sufficiently large number of molecules the stochastic dynamics follows closely the deterministic one and thus the fluctuations are small. On the other hand, one should expect that when the number of molecules decreases, then the fluctuation will rise and the system behavior could not be completely described by means of a deterministic approach. This is con- firmed by Fig. 4 and Fig. 5, where we can observe that a model composed by a small number of initial molecules, 20 times lesser than in the model presented in Fig. 2 exhibits larger stochastic fluctuations. STOCHASTIC EVOLUTION OF A PROTOCELL. 9 Stochastic Integ ODE 1 X 80 70 60 50 40 30 20 10 0 0 0.5 1 1.5 time 2 2.5 3 x 10−3 100 90 80 C 70 60 50 0 Stochastic Integ ODE 0.5 1 1.5 2 2.5 3 x 10−3 Figure 4. Stochastic vs ODE SRM protocell (3). Case of one GMM, left panel the time evolution of the amount of GMM, right panel the time evolution of the amount of C. Parameters are : η = 1, α = 1, L1 = 500, P 1 = 600, X1(0) = 5, C(0) = 50, ρ = 200 and β = 2~3. 100 ) k 1( X 90 80 70 60 50 40 30 20 0 20 Stochastic Integ ODE 14 x 10−4 12 10 k T ∆ 8 6 4 0 20 80 100 40 60 generation number (k) Stochastic Integ ODE 40 60 generation number (k) 80 100 Figure 5. Stochastic vs ODE SRM protocell (3). Case of one GMM, left panel the amount of GMM at the beginning of each division cycle, right panel the division time as a function of the number of elapsed divisions. Parameters are : η = 1, α = 1, L1 = 500, P 1 = 600, X1(0) = 5, C(0) = 50, ρ = 200 and β = 2~3. In Fig. 6 we summarize the results of several protocell models each one with a different amount of initial molecules, in order to appreciate the influence of the latter on the stochastic fluctuations. To compare with, we also report the case of the deterministic model. Because the kinetic constants are kept constant, the analytical theory for the deterministic model ensures that the division time doesn't vary [3]. Nevertheless the fewer is the initial amount of X0 and C0, the larger are the fluctuations present in the stochastic integration. To get a more complete understanding of the fluctuations dependence, we decided to measure them using the standard deviation of the protocell division time (after a sufficiently long transient phase). In Fig. 7 we report the standard deviation of the division time ∆T as a function of the initial amount of molecules. As expected the fluctuations strength decreases rapidly as soon as the number of molecules increases and the relation can be very well approximated by a power law distribution with exponent −0.54 ± 0.03 (linear best fit). 5.2. Two non -- interacting Genetic Memory Molecules. A slightly more sophisticated model can be obtained by considering two linear non interacting GMMs. The system can be described by the following chemical reactions: 10 T. CARLETTI AND A. FILISETTI k T ∆ 4.5 x 10−3 4 3.5 3 2.5 2 1.5 1 0.5 0 20 =10 =100 =1, C X 0 0 =10, C X 0 0 =50, C X 0 0 =100, C X =1000 0 0 =500 ODE 40 60 generation number (k) 80 100 Figure 6. Fluctuation dependence on the initial conditions. We report the divi- sion times as a function of the number of elapsed divisions, for 5 different protocells models. Protocell ○ : X1(0) = 5, C(0) = 10, protocell ◻: X1(0) = 10, C(0) = 100, protocell ▽: X1(0) = 50, C(0) = 500, protocell △: X1(0) = 100, C(0) = 1000. The black line denotes the deterministic protocell. All the remaining parameters have been fixed to: η = 1, α = 1, L1 = 500, P 1 = 600, ρ = 100 and β = 1. Fluctuations Power Law 10−3 ) T ∆ ( d t s 10−4 10−5 100 101 X 0 102 103 Figure 7. Fluctuation dependence on the initial conditions. We report the stan- dard deviation of the protocell division time as a function of the initial amount of molecules X0 (●) and a linear best fit, whose slope is = −0.54 ± 0.03. Parameters are: X(0) = 2n with n = 0, ..., 10, C(0) = 10X(0), η = 1, α = 1, L1 = 500, P 1 = 600, ρ = 100 and β = 1. channel 1, R1 channel 2, R2 channel 3, R3 channel 4, R4 ∶ ∶ ∶ ∶ η1 X1 + P1 GGGGGA 2X1 α1 X1 + L1 GGGGGGA X1 + C η2 X2 + P2 GGGGGA 2X2 α2 X2 + L2 GGGGGGA X2 + C , where Pi and Li are, respectively, precursors of the i -- th GMM, i.e. nucleotide, and precursors of amphiphiles used by the i -- th GMM to build a C molecule. As previously done, we compare the stochastic and the deterministic models. Results are reported in Figure 8 and one can still observe that in presence of a large number of molecules the STOCHASTIC EVOLUTION OF A PROTOCELL. 11 deterministic dynamics well approximates the stochastic model. On the other hand, the protocell division time exhibits large fluctuations around the deterministic value even in presence of a quite large number of molecules (see right panel Fig. 9). The parameters have been set in such a way only one GMM will survive according to the analytical theory for the deterministic model. One can observe that, despite the fluctuations, the same fate is obtained for the stochastic model (see right panel Fig. 9). Stochatic Integ (X 1 Stochatic Integ (X 2 ) ) EDO (X 1 EDO (X 2 ) ) 450 400 350 300 250 200 150 i X 100 0 0.5 1 1.5 2 2.5 time 3 3.5 4 4.5 x 10−3 2000 1900 1800 1700 1600 C 1500 1400 1300 1200 1100 1000 0 Stochastic Integ ODE 0.5 1 1.5 2 2.5 time 3 3.5 4 4.5 x 10−3 Figure 8. Stochastic vs ODE SRM protocell (3). Case of two GMMs, left panel the time evolution of the amount of GMM during a division cycle, right panel the time evolution of the amount of C molecules. Parameters are : η1 = η2 = 1, α1 = α2 = 2, L1 = 500, L2 = 600, P 1 = 600, P2 = 670, X1(0) = X2(0) = 100, C(0) = 1000, ρ = 200 and β = 2~3. 700 600 500 400 300 200 100 i X 0 0 20 x 10−3 3 2.8 2.6 k T ∆ 2.4 2.2 2 1.8 0 20 80 100 Stochastic Integ (X 1 Stochastic Integ (X 2 ) ) EDO (X 1 EDO (X 2 ) ) 40 60 generation number (k) Stochastic Integ ODE 40 60 generation number (k) 80 100 Figure 9. Stochastic vs ODE SRM protocell (3). Case of two GMMs, left panel the amount of GMM at the beginning of each division cycle, right panel the division time as a function of the number of elapsed divisions. Parameters are : η1 = η2 = 1, α1 = α2 = 2, L1 = 500, L2 = 600, P 1 = 600, P2 = 670, X1(0) = X2(0) = 100, C(0) = 1000, ρ = 200 and β = 2~3. Once we reduce the number of involved molecules, the stochastic fluctuations dramatically increase (see Fig. 10 and Fig. 11). As in the case of only one GMM, when two non interacting linear GMMs are present the size of the stochastic fluctuations as a function of the initial number of molecules follows a power law distribution with exponent −0.51 ± 0.05 (linear best fit), see Fig. 12: the fewer are the molecules in the system, the larger are the fluctuations around the deterministic dynamics. 12 T. CARLETTI AND A. FILISETTI 30 25 20 i X 15 10 5 0 0 Stochastic Integ (X 1 Stochastic Integ (X 2 ) ) EDO (X 1 EDO (X 2 ) ) 0.2 0.4 0.6 0.8 time 1 1.2 1.4 1.6 x 10−3 110 100 90 C 80 70 60 50 0 Stochastic Integ ODE 0.2 0.4 0.6 0.8 time 1 1.2 1.4 1.6 x 10−3 Figure 10. Stochastic vs ODE SRM protocell (3). Case of two GMMs, left panel the time evolution of the amount of GMM during a division cycle, right panel the time evolution of the amount of C molecules. Parameters are : η1 = η2 = 1, α1 = α2 = 2, L1 = 500, L2 = 600, P 1 = 450, P2 = 670, X1(0) = X2(0) = 5, C(0) = 50, ρ = 200 and β = 2~3. ) k i( X 45 40 35 30 25 20 15 10 5 0 0 20 x 10−4 14 Stochastic Integ ODE 12 10 k T ∆ 8 6 4 2 0 20 40 60 division number (k) 80 100 Stocastic Integ (X 1 Stocastic Integ (X 2 ) ) EDO (X 1 EDO (X 2 ) ) 40 60 generation number (k) 80 100 Figure 11. Stochastic vs ODE SRM protocell (3). Case of two GMMs, left panel the amount of GMM at the beginning of each division cycle, right panel the division time as a function of the number of elapsed divisions. Parameters are : η1 = η2 = 1, α1 = α2 = 2, L1 = 500, L2 = 600, P 1 = 450, P2 = 670, X1(0) = X2(0) = 5, C(0) = 50, ρ = 200 and β = 2~3. A new phenomenon arises in the case of two GMMs modeled by a stochastic process. There can be a breaking of the symmetry emerging in systems composed of two identical GMMs (i.e equal kinetic constants, equal initial amounts and availability of precursors) present with a few initial amount of each one. Although adopting a deterministic approach the dynamics of the two replicators would be perfectly the same, a small fluctuation in the very first instants of the protocell evolution entails the dilution of one of the two replicators and thus a different fate for the protocell. Let us observe that the probability to have a large fluctuation is never zero, thus waiting for a sufficiently long time, a specie can always disappear from the system and thus giving rise to the the breaking of the symmetry phenomenon. See Fig. 13 where we report, as a function of the initial amount of molecules Xi(0), i = 1, 2, the proportion of simulations where the symmetry breaking has been observed repeating 50 times each simulation with the same set of parameters and initial conditions during 100 generations. In this paper we presented a new stochastic integration algorithm based on the one introduced by Gillespie. Our contribution is devoted to the explicit introduction of the volume variation in 6. Conclusion STOCHASTIC EVOLUTION OF A PROTOCELL. 13 Fluctuations Power Low 10−3 ) T ∆ ( d t s 10−4 10−5 100 101 X 0 102 103 Figure 12. Fluctuation dependence on the initial conditions. We report the standard deviation of the protocell division time as a function of the initial amount of molecules Xi(0), i = 1, 2, (●) and a linear best fit, whose slope is = −0.51 ± 0.05. Parameters are: X1(0) = X2(0) = 2n with n = 0, ..., 10, C(0) = 10X1(0), η1 = η2 = 1, α1 = α2 = 2, L1 = 500, L2 = 500, P 1 = 500, P2 = 600, ρ = 100 and β = 1. 1 0.8 0.6 0.4 0.2 s n o i t a l u m i s y r t e m m y s n e k o r b f o o i t a R 0 0 10 20 X 0 30 40 50 Figure 13. Symmetry breaking phenomenon. Each point denotes the frac- tion of runs exhibiting the symmetry breaking phenomenon, during 100 gen- Parameters are: X1(0) = X2(0) = erations, over 50 identical replicas. [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 25, 50], C(0) = 10X, η1 = η2 = 1, α1 = α2 = 2, L1 = 500, L2 = 500, P 1 = 600, P2 = 600, ρ = 100 and β = 1. the algorithm, which moreover is directly related to the amount of contained molecules, and thus it evolves in a self-consistent way. This algorithm straightforwardly adapts to the study of the evolution of a protocell, simplified form of cells, where an ensemble of chemical reactions occurs in a varying volume, the volume of the protocell, that in turn increases because of the production of container molecules. We presented several protocell models and we compare them with the analogous deterministic protocell models, namely solved using the ODE. In this preliminary study, we emphasized the role of the fluctuations and their dependence on the initial amount of molecules. The dynamics is richer than the deterministic one and thus it is worth studying, in particular we deserve to future investigations the case where several molecules interact in a linear way but including cross catalysis, i.e. the interaction matrix is not diagonal, or they interact in a non-linear way. Also the study of the emergence of time-periodic patterns due to the fluctuations, will be analyzed. An analytical treatment of the latter case could be possible using some recent technics developed by [11, 4], see also [2] where the space is also taken into account. 14 T. CARLETTI AND A. FILISETTI Acknowledgment The work of TC has been partially supported by the FNRS grant "Mission Scientifique 2010-2011". TC would like also to thank the European Center for Living Technol- ogy in Venice (Italy) for the warm hospitality during a short visiting stay that originates this collaboration. References [1] Alberts B. et al., Molecular Biology of the Cell, (Garland, New York) 2002. [2] Anna P. et al., Spatial model of autocatalytic reactions, PRE, 81, (2010), pp. 056110 [3] T. Carletti et al, Sufficient conditions for emergent synchronization in protocell models, J. Theor. Biol., 254, (2008), pp. 741. [4] T. Dauxois et al, Enhanced stochastic oscillations in autocatalytic reactions, PRE, 79, (2009), pp. 036112 [5] D.T. Gillespie, Exact stochastic simulation of coupled chemical reactions, J. of Phys Chem, 81, 25, (1977), pp. 2340. [6] D.T. Gillespie, Markov Processes: An Introduction for Physical Scientists, Academic Press, Boston, (1992). [7] Kaneko, K., Life : An introduction to complex system biology, Springer, The Netherlands, (2006) [8] A.M. Kierzek, STOKS: STOChastic Kinetic Simulations of biochemical systems with Gillespie algorithm, 18, (3), Bioinformatics, (2002), pp. 470-481. [9] T. Lu, D. Volfson, L. Tsimring and J. Hasty, Cellular growth and division in the Gillespie algorithm, Syst. Biol., 1, 1, (2004), pp. 121. [10] Luisi P. L., Ferri F. and Stano P, Naturwissenschaften, 93, (2006), pp. 1. [11] A.J. McKane and T.J. Newman, Predator-Prey Cycles from Resonant Amplification of Demographic Stochas- ticity, PRL, 94, (2005), pp. 218102. [12] Mansy, S.S. et al., Template-directed synthesis of a genetic polymer in a model protocell, Nature Letters, (2008), doi:10.1038/nature07018 [13] Oberholzer, T. et al., Enzymatic RNA replication in selfreproducing vesicles: an approach to a minimal cell, Biochemical and Biophysical Research Communications, 207, (1995), pp. 250257. [14] Rasmussen S., Chen L., Stadler B. and Stadler P., Proto-Organisim kinetics: evolutionary dynamics of lipid aggregates with genes and metabolism, Origins Life Evol. Biosphere, 34, (2004), pp. 171. [15] Rasmussen S. et al., Transitions from non living to living matter, Science, 303, (2004), pp. 963-965. [16] R. Serra et al., , A. Life, 13, (2007), pp. 123. [17] Szostak D., Bartel P. B. and Luisi P. L., Synthesizing life, Nature, 409, (2001), pp. 387-390. (Timoteo Carletti) naXys, Namur Center for Complex Systems and University of Namur, Namur, Belgium 5000 (Alessandro Filisetti) CIRI, Energy and Environment Interdipartimental Center for Industrial Re- search, University of Bologna and European Centre for Living Technology, University C´a Foscari of Venice
1912.03003
1
1912
2019-12-06T07:10:11
Effect of disorder and polarization sequences on two-dimensional spectra of light harvesting complexes
[ "physics.bio-ph", "physics.chem-ph" ]
Two-dimensional electronic spectra (2DES) provide unique ways to track the energy transfer dynamics in light-harvesting complexes. The interpretation of the peaks and structures found in experimentally recorded 2DES is often not straightforward, since several processes are imaged simultaneously. The choice of specific pulse polarization sequences helps to disentangle the sometimes convoluted spectra, but brings along other disturbances. We show by detailed theoretical calculations how 2DES of the Fenna-Matthews-Olson complex are affected by rotational and conformational disorder of the chromophores.
physics.bio-ph
physics
Noname manuscript No. (will be inserted by the editor) Effect of disorder and polarization sequences on two-dimensional spectra of light harvesting complexes Tobias Kramer · Mirta Rodr´ıguez 9 1 0 2 c e D 6 ] h p - o i b . s c i s y h p [ 1 v 3 0 0 3 0 . 2 1 9 1 : v i X r a 10. October 2019 Abstract Two-dimensional electronic spectra (2DES) provide unique ways to track the energy transfer dy- namics in light-harvesting complexes. The interpreta- tion of the peaks and structures found in experimentally recorded 2DES is often not straightforward, since sev- eral processes are imaged simultaneously. The choice of specific pulse polarization sequences helps to disentan- gle the sometimes convoluted spectra, but brings along other disturbances. We show by detailed theoretical cal- culations how 2DES of the Fenna-Matthews-Olson com- plex are affected by rotational and conformational dis- order of the chromophores. Keywords Two-dimensional spectroscopy · Light- harvesting complex · FMO (Fenna -- Matthews -- Olson complex) 1 Introduction The investigation of energy transfer in light-harvesting complexes (LHCs) is naturally performed by optical spectroscopy. To study the time scales and interme- diate steps associated with the energy transfer from the antenna to the reaction center requires time- and frequency resolving methods (Blankenship 2014). One example are transient absorption spectra which probe the response of the system after an excitation (pump) pulse. Transient absorption spectra can be described as projected two-dimensional electronic spectra (2DES), which in turn rely on a sequence of four pulses to sepa- Tobias Kramer Zuse Institute Berlin, Germany E-mail: [email protected] (cor- responding author) Mirta Rodr´ıguez Zuse Institute Berlin, Germany E-mail: [email protected] rate the excitation and emission frequencies. The theo- retical modeling of 2DES within non-perturbative theo- ries of the exciton dynamics requires considerable com- putational effort due to the possibility of excited state absorption, giving rise to the presence of two excitons in the molecular complex (Mukamel 1995; Cho 2008; Hamm and Zanni 2011). The excitons are coupled to vibrations of the molecules, which induces decoherence and dissipation. The dissipation directs the excitons to lower lying states, where often the reaction center of LHCs is located (Blankenship 2014). The first-principle calculation of excitonic site-energies and inter-pigment couplings requires a mixed molecular- dynamics quantum-chemistry approach due to the mod- ulations of the excitonic energies by the solvent and the protein scaffolding, see Olbrich et al. (2011); Aghtar et al. (2013). For typical distances of chromophores in LHCs and at physiological temperatures, a commonly used model is the Frenkel-type Hamiltonian, which con- siders the excitons as "system" and the vibrational states of the molecule as environment ("bath") (May and Kuhn 2008). For LHCs the excitonic couplings, temperature, and reorganization energy can be of comparable magni- tude, which precludes the application of weak coupling approximations. The hierarchical equation of motion (HEOM) method introduced by Tanimura and Kubo (1989) solves the Frenkel exciton-model dynamics in an exact way and does not rely on small parameter as- sumption (Ishizaki and Fleming 2009; Kreisbeck et al. 2011). A comprehensive comparison of Redfield and HEOM spectra for the Fenna-Matthews-Olson (FMO) complex is given by Hein et al. (2012); Kramer et al. (2018b), for the Photosystem I supercomplex and Fo- erster theory by Kramer et al. (2018a), and for com- bined Foerster-Redfield theory applied to the light har- 2 Tobias Kramer, Mirta Rodr´ıguez (k, l, m, n) (1, 1, 2, 2), (1, 1, 3, 3), (1, 2, 1, 2), (1, 2, 2, 1), (1, 3, 1, 3), (1, 3, 3, 1) (2, 1, 1, 2), (2, 1, 2, 1), (2, 2, 1, 1), (2, 2, 3, 3), (2, 3, 2, 3), (2, 3, 3, 2) (3, 1, 1, 3), (3, 1, 3, 1), (3, 2, 2, 3), (3, 2, 3, 2), (3, 3, 1, 1), (3, 3, 2, 2) (1, 1, 1, 1), (2, 2, 2, 2), (3, 3, 3, 3) (k, l, m, n) (1, 2, 1, 2), (1, 2, 2, 1), (1, 3, 1, 3) (1, 3, 3, 1), (2, 1, 1, 2), (2, 1, 2, 1) (2, 3, 2, 3), (2, 3, 3, 2), (3, 1, 1, 3) (3, 1, 3, 1), (3, 2, 2, 3), (3, 2, 3, 2) Cklmn (cid:104)0◦, 0◦, 0◦, 0◦(cid:105) + 1 15 + 1 15 + 1 15 + 1 5 12 ,− 1 + 1 − 1 12 ,− 1 12 ,− 1 + 1 12 ,− 1 + 1 12 12 ,+ 1 12 ,+ 1 12 ,− 1 12 ,+ 1 12 12 12 Cklmn (cid:104)45◦, −45◦, 90◦, 0◦(cid:105) Table 1 Cklmn coefficients for isotropic averaging of the (cid:104)0◦, 0◦, 0◦, 0◦(cid:105) and (cid:104)45◦, −45◦, 90◦, 0◦(cid:105) polarization sequences vesting complex II (LHCII) by Kreisbeck et al. (2014); Novoderezhkin and van Grondelle (2017). Here, we discuss how the protein structures of LHCs are reflected in 2DES for different pulse polarizations and how rotational and static disorder affects the spec- tra. This analysis is required to interpret sequences of 2DES in terms of signature of coherent exciton dynam- ics. 2 Two-dimensional electronic spectra (2DES) Most commonly 2DES is applied in the "all parallel" setup, where all pulses are polarized along the same di- rection. In this case, 2DES provides direct insights in to the dynamics of the excitonic energy transfer, since the stimulated emission (SE) and excited state absorp- tion (ESA) signals trace the population dynamics of exciton states during the delay time T2 between the second and third pulse. For the FMO complex con- tained in green sulfur bacteria (Olson and a Romano 1962; Fenna and Matthews 1975; Blankenship 2014), the transfer dynamics mirrored in 2DES has been an- alyzed with HEOM by Chen et al. (2011); Hein et al. (2012); Kreisbeck and Kramer (2012). In contrast to the SE/ESA contributions, the ground state bleaching (GSB) pathway mirrors the linear absorption signal at long delay times and does not provide direct insights into the population dynamics (Kramer and Rodriguez 2017). However, GSB is particularly susceptible to vi- brational motion, which enters the 2DES (Tiwari et al. 2013; Kreisbeck et al. 2013) and produces beating sig- nals in the cross-peak amplitudes as function of the delay time T2. These oscillatory signals have been ob- served by Engel et al. (2007) in experimental 2DES and need to be separated from electronic coherences. Over the years various polarization sequences have been ex- plored to disentangle the overlapping and oscillating peaks composing a typical 2DES. 2.1 Polarization sequences The 2DES of light harvesting complexes often shows large and overlapping regions, making it difficult to di- rectly assign peaks to single excitonic energies. Vari- ous extension of 2DES have been proposed, see the re- views by (Nuernberger et al. 2015; Lambrev et al. 2019) and have been used to isolate specific features in the spectra. To confirm the structural information of LHCs obtained from crystals (Tronrud et al. 2009) and elec- tron microscopy (B´ına et al. 2016), circular and linear dichroism studies are important to probe the relative orientations of the transition dipole moments (Lindor- fer and Renger 2018). By choosing specific polariza- tion sequences, 2DES is also suitable to interrogate the molecular configuration (orientation and magnitude) of the transition-dipole moments. To see this, we consider exemplary the stimulated emission rephasing (SE,RP) pathway (see Kramer et al. (2018b), Eq. (65)): SSE,RP(T3, T2, T1p0, p1, p2, p3) = (t1)ρ0 µ− p0 (0)µ+ p2 (t3)µ+ p1 p3 (1) with time intervals T1 = t1, T2 = t2−t1, T3 = t3−t2 and four impulsive interactions of the excitonic states with the light pulses represented by the dipole operators µ with polarizations pi. The recorded electric field of the signal is given by the sum of all pathways and incurs an additional conjugation (see Hamm and Zanni (2011), Eq. (4.29)): Esig ∝ i (SGSB,RP + SSE,RP + SESA,RP) . (2) iTr(cid:2)µ− (t2)(cid:3), Within the HEOM method the time-propagation of the coupled exciton-vibrational system is performed numer- ically. The first T1 and last T3 interval of the time- dependent signal (1) are Fourier transformed to the fre- quency domain and represents the excitation (ω1) and emission (ω3) frequencies. To account for an ensemble of randomly oriented molecules, a rotational isotropic av- erage of the signal has to be performed. This leads to a tensorial expression of all possible directional Cartesian Effect of disorder and polarization sequences on two-dimensional spectra 3 Fig. 1 Calculated 2DES of a dimer. Upper row: Real part of the rephasing signal for (cid:104)0◦, 0◦, 0◦, 0◦(cid:105) polarization sequence at delay time T2 = 40 fs, lower row, same for the (cid:104)45◦, −45◦, 90◦, 0◦(cid:105) polarizations. From left to right the contributions of the different pathways are shown, the value above the color bar indicates the magnitude of the signal. The right panels show the combined, experimentally accessible signal. In the all-parallel sequence diagonal peaks are enhanced, while in the (cid:104)45◦, −45◦, 90◦, 0◦(cid:105) sequence the cross-peaks are visible with additional zero crossings. components, which is however reduced by symmetry to 21 terms at most (Gordon 1968; Yuen-Zhou et al. 2014; Gelin et al. 2017; Kramer et al. 2018b). Upon trans- forming the time trace to the frequency domain and isotropic rotational average we obtain (cid:90) ∞ (cid:90) ∞ × 3(cid:88) (cid:104)SRP(ω3, T2, ω1)(cid:105)rot = dT1 dT3e−iT1ω1+iT3ω3 0 0 CklmnSSE,RP(T3, T2, T1p0,k, p1,l, p2,m, p3,n). k,l,m,n=1 (3) The Cklmn coefficients are determined by Cklmn = δklδmn [(f0f1)(f2f3) − (f0f2)(f1f3) − (f0f3)(f1f2)] /30 +δkmδln [(f0f2)(f1f3) − (f0f1)(f2f3) − (f0f3)(f1f2)] /30 +δknδlm [(f0f3)(f1f2) − (f0f1)(f2f3) − (f0f2)(f1f3)] /30, where fi denotes the unit vector of the electric field of the ith pulse. The rotational average can be used to address specific dipole combinations. To simulate the experimental ensemble requires additionally to ac- count for slower conformational variations of the chro- mophores (for instance bending and twisting modes), which lead to fluctuations of the site energies, known as static disorder. Static disorder requires to run simu- lations for different site energies and leads to a further blurring of the 2DES, albeit with different effects at changing locations in the ω1-ω3 frequency-plane. The (cid:104)0◦, 0◦, 0◦, 0◦(cid:105) polarization sequence In the all parallel polarization sequence, GSB and the combined SE+ESA pathways contribute with similar magnitude to the total 2DES signal, see the analy- sis of the 2DES of C. tepidum by Kramer and Ro- driguez (2017), Fig. 3. The isotropic averaging coeffi- cients are listed in Tab. 1. Typical 2DES of FMO for the (cid:104)0◦, 0◦, 0◦, 0◦(cid:105) polarization sequence computed with DM-HEOM (Noack et al. 2018; Kramer et al. 2018b) are shown in Kramer et al. (2018b), Fig. 11. The Hamil- tonian and the dipole directions are listed in Kramer et al. (2018b), Tab. 1 and Eq. (77), taken from Adolphs and Renger (2006). The energy transfer is clearly visible in experiments performed by Brixner et al. (2005) and in theoretical computations (Hein et al. 2012; Kreis- beck and Kramer 2012) in form of the growing inten- sities of the lower cross peaks compared to the diag- onal peaks. On top of the energy decay Engel et al. (2007); Panitchayangkoon et al. (2010) reported oscil- latory amplitudes, which are interpreted as a combina- tion of ground-state bleach induced vibrational modes and electronic coherences. The electronic coherences are expected to decay on the time-scale of the combined dephasing and relaxation decoherence time (Kreisbeck 4 Tobias Kramer, Mirta Rodr´ıguez approximate expression for the 2DES of a dimer with dipole directions given by D1, D2 and Hamiltonian (cid:18)−E J (cid:19) √ J +E Hdimer = (5) with eigenvalues {−∆E/2, +∆E/2 = E2 + J 2}. An analytic expression for the 2DES is obtained by consid- ering a purely coherent dynamics for the density matrix ρ(t) = e−iH(t−t0)/¯hρ(t0)e+iH(t−t0)/¯h, and only introducing decoherence through the inverse of the decoherence time α = 1/τdecoherence. Inserting the time-evolution in Eq. (6) yields the time-representation of the rephasing signal SERP4490(T1, T2, T3) = 2(D1 × D2)2 (6) cos E2 + J 2(T1 + 2T2 + T3) e−α(T1+T2+T3). (7) (cid:16)(cid:112) (cid:17) The cross product of the two transition-dipole direc- tions in Eq. (7) leads to a vanishing contribution of any parallel dipole component. The Fourier transform to the frequency domain results in (cid:20) (2ω1 − ∆E+)(∆E+ + 2ω3)e−i∆ET2 SERP4490(ω1, T2, ω3) = (D1 × D2)2 (cid:21) × − (∆E+ + 2ω1)(∆E+ − 2ω3)ei∆ET2 Γ = 2π(∆E− − 2ω3)(−∆E+ − 2ω3) ×(cid:0)J 2 − (E− + ω1)(ω1 − E+)(cid:1) , (8) where we introduced ∆E± = ∆E ± i2α, E± = E ± iα. At the cross peak location this expression simplifies to SERP4490(ω1 = ∆E/2, T2, ω3 = −∆E/2) = Γ Γ , − (D1 × D2)2 2πα2 e−iT2∆E/¯h − α2eiT2∆E/¯h (∆E + iα)2 . (9) (cid:32) (cid:33) The ESA rephasing term has the same form (up to an overall sign), albeit with a slightly different decoher- ence parameter α(cid:48) due to the coupling to the vibrational states of two separated pigments. The observed signal is the real part of the difference of the SE and ESA path- ways and brings along an additional sign change in the ω1, ω3 plane, see Fig. 1 for a model dimer with energies E = 50 cm−1 and couplings J = −90 cm−1. This re- sults in alternating stripes of positive/negative contri- butions and shifts the highest/lowest intensities away from the cross-peak locations already in the individ- ual pathways (SE,ESA). This is in contrast to the all- parallel polarization sequence, where ESA has only neg- ative components and SE positive values, which largely stay in place. Fig. 2 The orientations of the transition dipoles in the Fenna- Matthews-Olson complex (FMO monomeric unit 3ENI) along the NB and ND nitrogens. and Kramer 2012) of the two eigenenergies at the loca- tion of cross-peak. The slope of the spectral density towards zero frequency determines the pure dephas- ing time, while the value of the spectral density at the eigenenergies sets the relaxation rate. Both contribute to the decay time, see Kreisbeck and Kramer (2012), SI, and Kramer and Kreisbeck (2014), Fig. 8. Another manifestation of the spectral density J(ω) of each pig- ment m in 2DES is the reorganization energy λm: λm = Jm(ω) πω dω. (4) (cid:90) ∞ 0 The reorganization energy leads to a downward shift of the diagonal and cross peaks with increasing delay (Dost´al et al. 2016; Kramer and Rodriguez 2017). The observed shift is consistent with the organization en- ergies around 40/cm assigned by (Adolphs and Renger 2006) to the bacteriochlorophylls contained in the FMO complex. The (cid:104)45◦,−45◦, 90◦, 0◦(cid:105) polarization sequence Different polarization directions have been used by Hochstrasser (2001); Zanni et al. (2001) to enhance var- ious features in 2D IR spectra and 2DES (Schlau-Cohen et al. 2012). For a polarization sequence with two or- thogonal pulse pairs, the isotropic rotational averag- ing suppresses the (1, 1, 1, 1), (2, 2, 2, 2), and (3, 3, 3, 3) components, and leads to alternating signs of the re- maining 12 contributions listed in in Tab. 1 for the (cid:104)45◦,−45◦, 90◦, 0◦(cid:105) case. This selection of pathways elim- inates the ground-state bleaching signal (Westenhoff et al. 2012) and emphasizes all dipole pairs with or- thogonal orientation. To demonstrate this, we derive an Effect of disorder and polarization sequences on two-dimensional spectra 5 Fig. 3 Calculated 2DES of FMO for delay times T2 = {40, 100, 500} fs. Imaginary part of the rephasing signal for (cid:104)45◦, −45◦, 90◦, 0◦(cid:105) polarization sequence for increasing values of static disorder (SD), {0, 30, 50} cm−1, temperature T = 100 K. See Fig. 5 by Rodr´ıguez and Kramer (2019) for the real part of the rephasing signal. Corresponding measured spectra by Thyrhaug et al. (2018) are shown in their Fig. 2c, SI Fig. 2a, SI Fig. 2b and favor the 30 − 50 cm−1 static disorder cases. This results in alternating stripes aligned along the diagonal. The selective excitation of orthogonal dipole orien- tations leads to a different 2DES of the FMO complex compared to the all-parallel polarization sequence. In the FMO complex, the transition dipoles are aligned within a few degrees with the nitrogen atoms NB-ND in the bacteriochlorophylls, see Fig. 2. The dipoles of bac- teriochlorophyll pairs (1, 3), (1, 5), (2, 7), (3, 4), (4, 5), (5, 7) are almost orthogonally arranged, but strong exci- tonic couplings in conjunction to the orthogonal direc- tions are only existing between the neighbouring pig- ments (3, 4), (4, 5) (Vulto et al. 1998, 1999; Adolphs et al. 2008). The (cid:104)45◦,−45◦, 90◦, 0◦(cid:105) sequence thus en- hances the contribution of these specific pairs of bac- teriochlorophylls in the complex, however at the added complexity of alternating signs in the signal, which are strongly affected by static disorder. Thyrhaug et al. (2016) measured 2DES with another polarization se- quence, taken to be (cid:104)90◦, 90◦, 0◦, 0◦(cid:105). Corresponding com- 6 Tobias Kramer, Mirta Rodr´ıguez Fig. 4 Calculated 2DES of FMO. Imaginary part of the rephasing signal for the (cid:104)45◦, −45◦, 90◦, 0◦(cid:105) polarization sequence at 500 fs delay time in the absence of static disorder. Shown is the decomposition of the experimentally accessible signal (sum of SE and ESA, resulting in stripes) into the SE and ESA components (GSB does not contribute). puted 2DES are discussed by Kramer et al. (2018b), Fig. 12. While the (cid:104)45◦,−45◦, 90◦, 0◦(cid:105) sequence has been sug- gested and used for studying cross-peak dynamics in FMO, it is more prone to disorder average than the all-parallel configuration. The alternating signs of the stripes in the real part of the rephasing signal of the (cid:104)45◦,−45◦, 90◦, 0◦(cid:105) sequence tend to diminish the sig- nal and only elongated stripes along the diagonal re- main. Numerically, we compute the disorder average from 5000 realizations of disorder added to the site en- ergies with standard deviation 30 cm−1 and 50 cm−1 . The resulting 2DES are efficiently encoded in a neural network representation following Rodr´ıguez and Kramer (2019). This encoding allows us to study various disor- der realizations, shown in Fig. 3 for delay times T2 = 40 fs, 100 fs, and 500 fs. The total 2DES is the addition of SE and ESA con- tributions (see Fig. 4) and develops a stripe-like pat- tern. Note, that in the experimental signals recorded by Thyrhaug et al. (2018) Supplementary Fig. 2 (corre- sponding to the lower two panels in Fig. 3) the stripes are almost perfectly parallel to the diagonal, already at shorter delay times (100 fs) compared to the simula- tion. A similar theoretical spectrum for the 40 fs delay time is shown by Thyrhaug et al. (2018), Fig. 2d, com- puted with a stochastic averaging approach instead of the tensorial method used here. 3 Conclusion The understanding of 2DES of LHCs requires to ensem- ble average individual molecular complexes over rota- tional and static disorder. This averaging process has been used experimentally in combination with specific polarization sequences with the goal to enhance cer- tain relative pigment orientations (see Thyrhaug et al. (2016, 2018)). Our calculations of FMO spectra show that the cross-peaks in these sequences are diminished and split by cancellation effects of positive/negative sig- nal contributions from the SE and ESA pathways. With increasing static disorder elongated structures arise and blur specific cross-peak contributions. The disorder sen- sitivity is different compared to the all-parallel 2DES polarization which causes less cancellations due to sign changes in the signal, but is directly affected by broad- ening of peaks due to static disorder. The 2DES of the FMO complex demonstrate that the theoretical models of the exciton energy transfer are in general agreement with the experimental obser- vations. This includes the values for the excitonic cou- plings, site energies, and vibrational density of states. The excitonic, vibrational and reorganization energies derived from theory and experiment bring along elec- tronic coherences, typically decaying within 300 fs at ambient temperatures (Kreisbeck and Kramer 2012). The functional role of these coherences is still debated since Avery et al. (1961). Kramer and Kreisbeck (2014) find that resonances between vibrational states and elec- tronic energy differences can either hinder or enhance transfer times. For LHCs with more pigments a simultaneous fitting of 2DES, circular and linear dichroism, and transient absorption spectra will be required to further interpret the experimental data and to provide systematic pre- dictions. A recent review by Lambrev et al. (2019) of the experimental and theoretical 2DES of LHCs comes to a similar conclusion, especially for the light harvest- ing complex II (LHC II). Conflict of Interest: The authors declare that they have no conflict of interest. Effect of disorder and polarization sequences on two-dimensional spectra 7 Acknowledgements The work was supported by the Ger- man Research Foundation (DFG) grants KR 2889 and RE 1389 ("Realistic Simulations of Photoactive Systems on HPC Clus- ters with Many-Core Processors"). We acknowledge compute time allocation by the North-German Supercomputing Al- liance (HLRN). M.R. has received funding from the European Union's Horizon 2020 research and innovation programme un- der the Marie Sklodowska-Curie grant agreement No. 707636. References Adolphs J, Renger T (2006) How Proteins Trigger Excita- tion Energy Transfer in the FMO Complex of Green Sul- fur Bacteria. Biophysical Journal 91(8):2778 -- 2797, DOI 10.1529/biophysj.105.079483 Adolphs J, Muh F, Madjet MEA, Renger T (2008) Calcu- lation of pigment transition energies in the FMO pro- tein. Photosynthesis Research 95(2-3):197 -- 209, DOI 10.1007/s11120-007-9248-z Aghtar M, Strumpfer J, Olbrich C, Schulten K, Kleinekathofer U (2013) The FMO Complex in a Glycerol -- Water Mixture. The Journal of Physical Chemistry B 117(24):7157 -- 7163, DOI 10.1021/jp311380k Avery J, Bay Z, Szent-Gyorgyi A (1961) On the Energy Transfer in Biological Systems. Proceedings of the Na- tional Academy of Sciences of the United States of Amer- ica 47(11):1742 -- 1744 B´ına D, Gardian Z, V´acha F, Litv´ın R (2016) Native FMO- reaction center supercomplex in green sulfur bacteria: An electron microscopy study. Photosynthesis Research 128(1):93 -- 102, DOI 10.1007/s11120-015-0205-y Blankenship RE (2014) Molecular Mechanisms of Photosyn- thesis, 2nd edn. Wiley, Oxford, UK Brixner T, Stenger J, Vaswani HM, Cho M, Blanken- ship RE, Fleming GR (2005) Two-dimensional spec- troscopy of electronic couplings in photosynthesis. Nature 434(7033):625 -- 628, DOI 10.1038/nature03429 Chen L, Zheng R, Jing Y, Shi Q (2011) Simulation of the two-dimensional electronic spectra of the Fenna- Matthews-Olson complex using the hierarchical equa- tions of motion method. The Journal of Chemical Physics 134(19):194508 -- 194508, DOI 10.1063/1.3589982 Cho M (2008) Coherent Two-Dimensional Optical Spec- troscopy. Chemical Reviews 108(4):1331 -- 1418, DOI 10.1021/cr078377b Dost´al J, Psenc´ık J, Zigmantas D (2016) In situ map- ping of the energy flow through the entire photosyn- thetic apparatus. Nature Chemistry 8(7):705 -- 710, DOI 10.1038/nchem.2525 Engel GS, Calhoun TR, Read EL, Ahn TK, Mancal T, Cheng YC, Blankenship RE, Fleming GR (2007) Evidence for wavelike energy transfer through quantum coherence in photosynthetic systems. Nature 446(7137):782 -- 786, DOI 10.1038/nature05678 Fenna RE, Matthews BW (1975) Chlorophyll arrangement in a bacteriochlorophyll protein from Chlorobium limicola. Nature 258(5536):573 -- 577, DOI 10.1038/258573a0 Gelin MF, Borrelli R, Domcke W (2017) Efficient orienta- tional averaging of nonlinear optical signals in multi- chromophore systems. The Journal of Chemical Physics 147(4):044114 -- 044114, DOI 10.1063/1.4996205 Gordon R (1968) Correlation Functions for Molecular Motion. In: Advances in Magnetic and Optical Resonance, vol 3, ACADEMIC PRESS INC., pp 1 -- 42, DOI 10.1016/B978- 1-4832-3116-7.50008-4 Hamm P, Zanni MT (2011) Concepts of 2D Spectroscopy. Cambridge University Press, Cambridge Hein B, Kreisbeck C, Kramer T, Rodr´ıguez M (2012) Mod- elling of oscillations in two-dimensional echo-spectra of the Fenna-Matthews-Olson complex. New Journal of Physics 14(2):023018 -- 023018, DOI 10.1088/1367- 2630/14/2/023018 Hochstrasser RM (2001) Two-dimensional IR-spectroscopy: Polarization anisotropy effects. Chemical Physics 266(2- 3):273 -- 284, DOI 10.1016/S0301-0104(01)00232-4 Ishizaki A, Fleming GR (2009) On the adequacy of the Redfield equation and related approaches to the study of quantum dynamics in electronic energy transfer. The Journal of Chemical Physics 130(23):234110 -- 234110, DOI 10.1063/1.3155214 Kramer T, Kreisbeck C (2014) Modelling excitonic-energy transfer in light-harvesting complexes. AIP Conference Proceedings 1575:111 -- 135, DOI 10.1063/1.4861701 Kramer T, Rodriguez M (2017) Two-dimensional electronic spectra of the photosynthetic apparatus of green sul- fur bacteria. Scientific Reports 7:45245 -- 45245, DOI 10.1038/srep45245 Kramer T, Noack M, Reimers JR, Reinefeld A, Rodr´ıguez M, Yin S (2018a) Energy flow in the Photosystem I supercomplex: Comparison of approximative theories with DM-HEOM. Chemical Physics 515:262 -- 271, DOI 10.1016/j.chemphys.2018.05.028 Kramer T, Noack M, Reinefeld A, Rodr´ıguez M, Zelinskyy Y (2018b) Efficient calculation of open quantum system dy- namics and time-resolved spectroscopy with distributed memory HEOM (DM-HEOM). Journal of Computational Chemistry 39(22):1779, DOI 10.1002/jcc.25354 Kreisbeck C, Kramer T (2012) Long-lived electronic coher- ence in dissipative exciton dynamics of light-harvesting complexes. Journal of Physical Chemistry Letters 3(19):2828 -- 2833, DOI 10.1021/jz3012029 Kreisbeck C, Kramer T, Rodr´ıguez M, Hein B (2011) High- performance solution of hierarchical equations of mo- tion for studying energy transfer in light-harvesting com- plexes. Journal of Chemical Theory and Computation 7(7):2166 -- 2174, DOI 10.1021/ct200126d Kreisbeck C, Kramer T, Aspuru-Guzik A (2013) Disen- tangling electronic and vibronic coherences in two- dimensional echo spectra. Journal of Physical Chemistry B 117(32):9380 -- 9385, DOI 10.1021/jp405421d Kreisbeck C, Kramer T, Aspuru-Guzik A (2014) Scalable High-Performance Algorithm for the Simulation of Exci- ton Dynamics. Application to the Light-Harvesting Com- plex II in the Presence of Resonant Vibrational Modes. Journal of Chemical Theory and Computation 10:4045 -- 4054, DOI 10.1021/ct500629s Lambrev PH, Akhtar P, Tan HS (2019) Insights into the mechanisms and dynamics of energy transfer in plant light-harvesting complexes from two-dimensional electronic spectroscopy. Biochimica et Biophysica Acta (BBA) - Bioenergetics DOI 10.1016/j.bbabio.2019.07.005 Lindorfer D, Renger T (2018) Theory of Anisotropic Circu- lar Dichroism of Excitonically Coupled Systems: Appli- cation to the Baseplate of Green Sulfur Bacteria. The Journal of Physical Chemistry B 122(10):2747 -- 2756, DOI 10.1021/acs.jpcb.7b12832 May V, Kuhn O (2008) Optical field control of charge trans- mission through a molecular wire. I. Generalized master equation description. Physical Review B 77(11):115439 -- 115439, DOI 10.1103/PhysRevB.77.115439 8 Tobias Kramer, Mirta Rodr´ıguez Simulations of Optical Spectra of the FMO Complex from the Green Sulfur Bacterium Chlorobium tepidum at 6 K. The Journal of Physical Chemistry B 102(47):9577 -- 9582, DOI 10.1021/jp982095l Vulto SIE, de Baat MA, Neerken S, Nowak FR, van Ameron- gen H, Amesz J, Aartsma TJ (1999) Excited State Dy- namics in FMO Antenna Complexes from Photosyn- thetic Green Sulfur Bacteria: A Kinetic Model. The Jour- nal of Physical Chemistry B 103(38):8153 -- 8161, DOI 10.1021/jp984702a Westenhoff S, Palecek D, Edlund P, Smith P, Zigman- tas D (2012) Coherent Picosecond Exciton Dynamics in a Photosynthetic Reaction Center. Journal of the American Chemical Society 134(40):16484 -- 16487, DOI 10.1021/ja3065478 Yuen-Zhou J, Krich JJ, Kassal I, Johnson A, Aspuru-Guzik A (2014) Ultrafast Spectroscopy. IOP Publishing, DOI 10.1088/978-0-750-31062-8 Zanni MT, Ge NH, Kim YS, Hochstrasser RM (2001) Two- dimensional IR spectroscopy can be designed to elimi- nate the diagonal peaks and expose only the crosspeaks needed for structure determination. Proceedings of the National Academy of Sciences 98(20):11265 -- 11270, DOI 10.1073/pnas.201412998 Mukamel S (1995) Principles of Nonlinear Optical Spec- troscopy. Oxford University Press, Oxford Noack M, Reinefeld A, Kramer T, Steinke T (2018) DM- HEOM: A Portable and Scalable Solver-Framework for the Hierarchical Equations of Motion. 2018 IEEE International Parallel and Distributed Processing Symposium Workshops 10.1109/IPDPSW.2018.00149 (IPDPSW) p 947, DOI Novoderezhkin VI, van Grondelle R (2017) Modeling of exci- tation dynamics in photosynthetic light-harvesting com- plexes: Exact versus perturbative approaches. Journal of Physics B: Atomic, Molecular and Optical Physics 50(12):124003 -- 124003, DOI 10.1088/1361-6455/aa6b87 Nuernberger P, Ruetzel S, Brixner T (2015) Multidimensional Electronic Spectroscopy of Photochemical Reactions. Angewandte Chemie International Edition 54(39):11368 -- 11386, DOI 10.1002/anie.201502974 Olbrich C, Jansen TLC, Liebers J, Aghtar M, Strumpfer J, Schulten K, Knoester J, Kleinekathofer U (2011) From atomistic modeling to excitation transfer and two- dimensional spectra of the FMO light-harvesting com- plex. Journal of Physical Chemistry B 115(26):8609 -- 8621, DOI 10.1021/jp202619a Olson JM, a Romano C (1962) A new chlorophyll from green bacteria. Biochimica et Biophysica Acta 59(3):726 -- 728, DOI 10.1016/0006-3002(62)90659-5 Panitchayangkoon G, Hayes D, a Fransted K, Caram JR, Harel E, Wen J, Blankenship RE, Engel GS (2010) Long- lived quantum coherence in photosynthetic complexes at physiological temperature. Proceedings of the National Academy of Sciences of the United States of America 107(29):12766 -- 70, DOI 10.1073/pnas.1005484107 Rodr´ıguez M, Kramer T (2019) Machine learning of two-dimensional spectroscopic data. Chemical Physics 520:52 -- 60, DOI 10.1016/j.chemphys.2019.01.002 Schlau-Cohen GS, Ishizaki A, Calhoun TR, Ginsberg NS, Ballottari M, Bassi R, Fleming GR (2012) Elucidation of the timescales and origins of quantum electronic co- herence in LHCII. Nature Chemistry 4(5):389 -- 395, DOI 10.1038/nchem.1303 Tanimura Y, Kubo R (1989) Time Evoultion of a Quantum System in Contact with a Nearly Gaussian-Markoffian Noise Bath. Journal of the Physics Society Japan 58(1):101 -- 114, DOI 10.1143/JPSJ.58.101 Thyrhaug E, Z´ıdek K, Dost´al J, B´ına D, Zigman- tas D (2016) Exciton Structure and Energy Transfer in the Fenna -- Matthews -- Olson Complex. The Journal of Physical Chemistry Letters 7(9):1653 -- 1660, DOI 10.1021/acs.jpclett.6b00534 Thyrhaug E, Tempelaar R, Alcocer MJP, Z´ıdek K, B´ına D, Knoester J, Jansen TLC, Zigmantas D (2018) Iden- tification and characterization of diverse coherences in the Fenna -- Matthews -- Olson complex. Nature Chemistry 10(7):780 -- 786, DOI 10.1038/s41557-018-0060-5 Tiwari V, Peters WK, Jonas DM (2013) Electronic reso- nance with anticorrelated pigment vibrations drives pho- tosynthetic energy transfer outside the adiabatic frame- work. Proceedings of the National Academy of Sciences 110(4):1203 -- 1208, DOI 10.1073/pnas.1211157110 Tronrud DE, Wen J, Gay L, Blankenship RE (2009) The structural basis for the difference in absorbance spectra for the FMO antenna protein from various green sul- fur bacteria. Photosynthesis Research 100(2):79 -- 87, DOI 10.1007/s11120-009-9430-6 Vulto SIE, de Baat MA, Louwe RJW, Permentier HP, Neef T, Miller M, van Amerongen H, Aartsma TJ (1998) Exciton
1111.0573
3
1111
2012-01-04T03:05:33
Emergent Behaviors from A Cellular Automaton Model for Invasive Tumor Growth in Heterogeneous Microenvironments
[ "physics.bio-ph", "physics.comp-ph" ]
Understanding tumor invasion and metastasis is of crucial importance for both fundamental cancer research and clinical practice. In vitro experiments have established that the invasive growth of malignant tumors is characterized by the dendritic invasive branches composed of chains of tumor cells emanating from the primary tumor mass. The preponderance of previous tumor simulations focused on non-invasive (or proliferative) growth. The formation of the invasive cell chains and their interactions with the primary tumor mass and host microenvironment are not well understood. Here, we present a novel cellular automaton (CA) model that enables one to efficiently simulate invasive tumor growth in a heterogeneous host microenvironment. By taking into account a variety of microscopic-scale tumor-host interactions, including the short-range mechanical interactions between tumor cells and tumor stroma, degradation of extracellular matrix by the invasive cells and oxygen/nutrient gradient driven cell motions, our CA model predicts a rich spectrum of growth dynamics and emergent behaviors of invasive tumors. Besides robustly reproducing the salient features of dendritic invasive growth, such as least resistance and intrabranch homotype attraction, we also predict nontrivial coupling of the growth dynamics of the primary tumor mass and the invasive cells. In addition, we show that the properties of the host microenvironment can significantly affect tumor morphology and growth dynamics, emphasizing the importance of understanding the tumor-host interaction. The capability of our CA model suggests that well-developed in silico tools could eventually be utilized in clinical situations to predict neoplastic progression and propose individualized optimal treatment strategies.
physics.bio-ph
physics
1 Emergent Behaviors from A Cellular Automaton Model for Invasive Tumor Growth in Heterogeneous Microenvironments Yang Jiao1∗, Salvatore Torquato1,2,∗∗ 1 Physical Science in Oncology Center, Princeton Institute for the Science and Technology of Materials, Princeton University, Princeton New Jersey 08544, USA 2 Department of Chemistry and Physics, Princeton Center for Theoretical Science, Program in Applied and Computational Mathematics, Princeton University, Princeton New Jersey 08544, USA ∗ E-mail: [email protected] ∗∗ E-mail: [email protected] Abstract Understanding tumor invasion and metastasis is of crucial importance for both fundamental cancer re- search and clinical practice. In vitro experiments have established that the invasive growth of malignant tumors is characterized by the dendritic invasive branches composed of chains of tumor cells emanating from the primary tumor mass. The preponderance of previous tumor simulations focused on non-invasive (or proliferative) growth. The formation of the invasive cell chains and their interactions with the pri- mary tumor mass and host microenvironment are not well understood. Here, we present a novel cellular automaton (CA) model that enables one to efficiently simulate invasive tumor growth in a heterogeneous host microenvironment. By taking into account a variety of microscopic-scale tumor-host interactions, including the short-range mechanical interactions between tumor cells and tumor stroma, degradation of extracellular matrix by the invasive cells and oxygen/nutrient gradient driven cell motions, our CA model predicts a rich spectrum of growth dynamics and emergent behaviors of invasive tumors. Besides robustly reproducing the salient features of dendritic invasive growth, such as least resistance and intra- branch homotype attraction, we also predict nontrivial coupling of the growth dynamics of the primary tumor mass and the invasive cells. In addition, we show that the properties of the host microenvironment can significantly affect tumor morphology and growth dynamics, emphasizing the importance of under- standing the tumor-host interaction. The capability of our CA model suggests that well-developed in silico tools could eventually be utilized in clinical situations to predict neoplastic progression and propose individualized optimal treatment strategies. 2 Author Summary The goal of the present work is to develop an efficient single-cell based cellular automaton (CA) model that enables one to investigate the growth dynamics and morphology of invasive solid tumors. Recent experiments have shown that highly malignant tumors develop dendritic branches composed of tumor cells that follow each other, which massively invade into the host microenvironment and ultimately lead to cancer metastasis. Previous theoretical/computational cancer modeling neither addressed the question of how such chain-like invasive branches form nor how they interact with the host microenvironment and the primary tumor. Our CA model, which incorporates a variety of microscopic-scale tumor-host interactions (e.g., the mechanical interactions between tumor cells and tumor stroma, degradation of extracellular matrix by the tumor cells and oxygen/nutrient gradient driven cell motions), can robustly reproduce experimentally observed invasive tumor evolution and predict a wide spectrum of invasive tumor growth dynamics and emergent behaviors in various different heterogeneous environments. Further refinement of our CA model could eventually lead to the development of a powerful simulation tool that can be utilized in the clinic to predict neoplastic progression and propose individualized optimal treatment strategies. Introduction Cancer is not a single disease, but rather a highly complex and heterogeneous set of diseases that can adapt in an opportunistic manner, even under a variety of stresses. It is now well accepted that genome level changes in cells, resulting in the gain of function of oncoproteins or the loss of function of tumor suppressor proteins, initiate the transformation of normal cells into malignant ones and neoplastic progression [1, 2]. In the most aggressive form, malignant cells can leave the primary tumor, invade into surrounding tissues, find their way into the circulatory system (through vascular network) and be deposited at certain organs in the body, leading to the development of secondary tumors (i.e., metastases) [3]. The emergence of invasive behavior in cancer is fatal. For example, the malignant cells that invade into the surrounding host tissues can quickly adapt to various environmental stresses and develop resistance to therapies. The invasive cells that are left behind after resection are responsible for tumor recurrence and thus an ultimately fatal outcome. Therefore, significant effort has been expended to understand the mechanisms evolved in the invasive growth of malignant tumors [2, 4 -- 7] and their treatment [8, 9]. It is generally accepted that the invasive behavior of cancer is the outcome of many complex interactions 3 occurring between the tumor cells, and between a tumor and the host microenvironment [3]. Tumor inva- sion itself is a complex multistep process involving homotype detachment, enzymatic matrix degradation, integrin-mediated heterotype adhesion, as well as active, directed and random motility [4]. In recent in vitro experiments involving glioblastoma multiforme (GBM), the most malignant brain cancer, it has been observed that dendritic invading branches composed of chains of tumor cells are emanating from the primary tumor mass; see Fig. 1. Such invasive behaviors are characterized by intrabranch homotype attraction and least resistance [4]. Although recently progress has been made in understanding certain aspects of the complex tumor- host interactions that may be responsible for invasive cancer behaviors [4, 10 -- 12], many mechanisms are either not fully understood or are unknown at the moment. Even if all of the mechanisms for cancer invasion could be identified, it is still not clear that progress in understanding neoplastic progression and proposing individualized optimal treatment strategies could be made without the knowledge of how these different mechanisms couple to one another and to the heterogeneous host microenvironment in which tumor grows [13]. Theoretical/computational cancer modeling that integrates apart mechanisms, when appropriately linked with experimental and clinical data, offers a promising avenue for a better understanding of tumor growth, invasion and metastasis. A successful model would enable one to broaden the conclusions drawn from existing medical data, suggest new experiments, test hypotheses, predict behavior in experimentally unobservable situations, and be employed for early detection and prognosis. Indeed, cancer modeling has been a very active area of research for the last two decades (see Refs. [14] and [13] for recent reviews). A variety of interactions between tumor cells and between tumor and its host microenvironment have been investigated [15 -- 27,29 -- 32], via continuum [25 -- 28,32], discrete [16,20,33] or hybrid [19, 21 -- 23] mathematical models. Very recently, multiscale mathematical models [22, 23, 28] have been employed to study the effects of the host microenvironment on the morphology and phenotypic evo- lution of invasive tumors and it has been shown that microenvironmental heterogeneity can dramatically affect the growth dynamics of invasive tumors. Although the simulated tumors showed certain invasive characteristics (e.g., developing protruding surfaces), no dendritic invasive branches emerged. In response to the challenge to develop an "Ising" model for cancer growth [13], we generalize a cell-based discrete cellular automaton (CA) model that we have developed [16 -- 18, 20, 21] to investigate the invasive growth of malignant tumors in heterogeneous host microenvironments. To the best of our knowledge, this generalized CA model is the first to investigate the formation of invasive cell chains and 4 their interactions with the primary tumor mass and the host microenvironment. Our cellular automaton model takes into account a variety of microscopic-scale tumor-host interactions, including the short-range mechanical interactions between tumor cells and tumor stroma, the degradation of extracellular matrix by the invasive cells and oxygen/nutrient gradient driven cell motions and thus, it can predict a wide range of growth dynamics and emergent behaviors of invasive tumors. In particular, our CA model robustly reproduces the salient features of dendritic invasive growth observed in experiments, which is characterized by least resistance and intrabranch homotype attraction. The model also predicts nontrivial coupling of the growth dynamics of the primary tumor mass and the invasive cells, e.g., the invasive cells can facilitate the growth of primary tumor in harsh microenvironment. Moreover, we show that the properties of the host microenvironment can significantly affect tumor growth dynamics and lead to a variety of tumor morphology. These emergent behaviors naturally arise due to various microscopic-scale tumor-host interactions, which emphasizes the importance of taking into account microenvironmental heterogeneity in understanding cancer. Further refinement of our model could eventually lead to the development of a powerful in silico tool that can be utilized in the clinic. As a demonstration of the capability and versatility of our CA model, we mainly consider invasive tumor growth in two dimensions, although the model is easily extended to three dimensions. Indeed, the algorithmic details of the model are given for any spatial dimension. Materials and Methods Biophysical Background of the CA Model Voronoi Tessellation: The Underlying Cellular Structure The underlying cellular structure is modeled using a Voronoi tessellation of the space into polyhedra [34], based on centers of spheres in a packing generated by a random sequential addition (RSA) process [16,21] (see Fig. 2). In particular, nonoverlapping spheres are randomly and sequentially placed in a prescribed region until there is no void space for additional spheres, i.e., saturation is achieved. Such a saturated RSA packing possesses relative small variations in its Voronoi polyhedra and thus, has served as models for many biological systems [35,36]. We refer to the polyhedra associated with the Voronoi tessellation as automaton cells. These automaton cells may correspond to real cells or tumor stroma (e.g., clusters of the 5 ECM macromolecules). In previous studies, such automaton cells have represented clusters of real cells of various sizes or have implicitly represented healthy tissues [16]. Thus, the Voronoi tessellation associated with RSA sphere centers provides a highly flexible model for real-cell aggregates with a high degree of shape isotropy. For example, one can use a variable automaton cell size to simulate avascular tumor growth from a few malignant cells to its macroscopic size [16]. In addition, such a Voronoi tessellation can reduce the undesired growth bias due to the anisotropy of ordered tessellations based on square and simple cubic lattices. Since our new CA model explicitly takes into account the interactions between a single cell and its neighbors and microenvironment, each automaton cell here represents either a single tumor cell or a region of tumor stroma. Thus, the linear size of a single automaton cell is approximately 15 − 20 µm and the linear size of the 2D simulation domain is approximately 5 mm, which contains ∼ 100 000 automaton cells. In the current model, we mainly focus on the effects of ECM macromolecule density and ECM degradation by malignant cells on tumor growth. Henceforth, we will refer to the host microenvironment (or tumor stroma) as "ECM" for simplicity. Each ECM associated automaton cell is assigned a particular density ρECM, representing the density of the ECM molecules within the automaton cell. A tumor cell can occupy an ECM associated automaton cell only if the density of this automaton cell ρECM = 0, which means that either the ECM is degraded or it is deformed (pushed away) by the proliferating tumor cells. Microenvironment Heterogeneity The microenvironment in which tumor grows is usually highly heterogeneous, composed of various types of stromal cells and ECM structures. The ECM is a complex mixture of macromolecules that provide mechanical supports for the tissue (such as collagens) and those that play an important role for cell adhe- sion and motility (such as laminin and fibronectin) [22, 37, 38]. For different individuals with tumors, the ECM in the host microenvironments generally possess distinct mechanical and transport properties. By explicitly representing the ECM using automaton cells with different macromolecule densities, the effects of microenvironment heterogeneity on tumor growth can be very well explored. For example, various distributions of the ECM densities (i.e., the densities of the ECM macromolecules) can be employed to mimic the actual heterogeneous host microenvironment of the tumor. Certain tumor stroma contains fibroblasts, which actively produce ECM macromolecules leading to a higher ECM density around these cells. The automaton cells representing the ECM with larger densities are considered to be more rigid 6 and more difficult to degrade. Since each automaton cell associated with the ECM has its own density, this allows a variation of ECM characteristics on the length scale comparable to that of a single tumor cell. In addition, the tumor in our model is only allowed to grow in a compact growth-permitting region. This is to mimic the physical confinement of the host microenvironment, such as the boundary of an organ. In other words, only automaton cells within this region can be occupied by the cells of the tumor as it grows. In general, the growth-permitting region can be of any shape that best models the organ shape. Here we simply choose a spherical region to study the effects of the heterogeneous ECM on tumor growth. More sophisticated growth-region shapes have been employed to investigate the effects of physical confinement on tumor growth [20, 21]. Furthermore, we assume a constant radially symmetric nutrient/oxygen gradient in the growth-permitting region with the highest nutrition concentration at the boundary of this region, i.e., it is a vascular boundary. However, this assumption can also be relaxed. Tumor Cell Phenotypes and Interactions with the Host Microenvironment For highly malignant tumors, we consider the cells to be of one of the two classes of phenotypes: either invasive or non-invasive. Following Ref. [16], the non-invasive cells remain in the primary tumor and can be proliferative, quiescent or necrotic, depending on the nutrition supply they get. For avascular tumor growth, as we focus here, the nutrition the tumor cells can get are essentially those diffuse into the tumor through its surface. As the tumor grows, the amount of nutrition supply, which is proportional to the surface area of the tumor, cannot meet the needs of all cells whose number increases with the tumor volume, leading to the development of necrotic and quiescent regions. Following Ref. [16], characteristic diffusion lengths are employed to determine the states of a non-invasive cell. For example, quiescent cells more than δn away from the tumor surface become necrotic (see details in the next section). The diffusion length δn (also the characteristic thickness of the rim of living tumor cells) depends on the size of the primary tumor. As a proliferative cell divides, its daughter cell effectively pushes away/degrades the surrounding ECM and occupies the automaton cell originally associated with the ECM [39 -- 41]. It is easier for a tumor cell to take up an ECM associated automaton cell with lower density (i.e., less rigid ECM regions) than that with higher density (i.e., more rigid ECM regions) and thus, the tumor growth is affected by the ECM heterogeneity through the local mechanical interaction between tumor cells and the ECM. If there is no 7 space available for the placement of a daughter cell within a distance δp from the proliferative cell, the proliferative cell turns quiescent. The invasive cells are considered to be mutant daughters of the proliferative cells [42], which can gain a variety of degrees of ECM degradation ability χ (i.e., the matrix-degradative enzymes) and motility µ that allow them to leave the primary tumor and invade into surrounding microenvironment [43]. We consider the invasive cells can move from one automaton cell to another only if the ECM in the target automaton cell is completely degraded (i.e., with ρECM = 0). Each trial move of an invasive cell involves the degradation of the ECM in its neighbor automaton cells, followed by a possible move to one of the automaton cells whose ECM is completely degraded; otherwise the invasive cell does not move. The number of trial moves of an invasive cell and to what extent it degrades the ECM are respectively determined by µ and χ (see the following section for details). The oxygen/nutrient gradient also drives the invasive cells to move as far as possible from the primary tumor [44], which takes up the majority of oxygen/nutrients. The motility µ is the maximum possible number of such trial moves. In addition, we assume that the invasive cells do not divide as they migrate. Algorithmic Details We now provide specific details for the CA model to study invasive tumor growth in confined hetero- geneous microenvironment. In what follows, we will simply refer to the primary tumor as "the tumor" and explicitly use "invasive" when considering invasive cells. After generating the automaton cells by Voronoi tessellation of RSA sphere centers, an ECM macromolecule density ρECM ∈ (0, 1) is assigned to each automaton cell within the growth-permitting region, which represents the heterogeneous host microenvironment. Then a tumor is introduced by designating any one or more of the automaton cells as proliferative cancer cells. Time is then discretized into units that represent one real day. At each time step: • Each automaton cell is checked for type: invasive, proliferative, quiescent, necrotic or ECM associ- ated. Invasive cells degrade and migrate into the ECM surrounding the tumor. Proliferative cells are actively dividing cancer cells, quiescent cancer cells are those that are alive, but do not have enough oxygen and nutrients to support cellular division and necrotic cells are dead cancer cells. • All ECM associated automaton cells and tumorous necrotic cells are inert (i.e., they do not change 8 type). • Quiescent cells more than a certain distance δn from the tumor's edge are turned necrotic. The tumor's edge, which is assumed to be the source of oxygen and nutrients, consists of all ECM associated automaton cells that border the neoplasm. The critical distance δn for quiescent cells to turn necrotic is computed as follows: δn = aL(d−1)/d t , where a is prescribed parameter (see Table 1), d is the spatial dimension and Lt is the distance between the geometric centroid xc of the tumor and the tumor edge cell that is closest to the quiescent cell under consideration. The position of the tumor centroid xc is given by xc = x1 + x2 + · · · + xN N , where N is the total number of noninvasive cells contained in the tumor, which is updated when a new noninvasive daughter cell is added to the tumor. • Each proliferative cell will attempt to divide with probability pdiv into the surrounding ECM (i.e., the automaton cells associated with the ECM) by degrading and pushing away the ECM in that automaton cell. We consider that pdiv depends on both the physical confinement imposed by the boundary of the growth-permitting region and the local mechanical interaction between the tumor cells and the ECM, i.e., pdiv =   p0 2 [(1 − r Lmax if any ECM associated automaton cell within ) + (1 − ρECM)] the predefined growth distance is in the growth- permitting microenvironment 0 the predefined growth distance is in the growth- if no ECM associated automaton cell within permitting microenvironment where p0 is the base probability of division (see Table 1), r is the distance of the dividing cell from the tumor centroid, Lmax is the distance between the closest growth-permitting boundary cell in 9 the direction of tumor growth and the tumor's geometric centroid xc and ρECM is the ECM density of the automaton cell to be taken by the new tumor cell. When a ECM associated automaton cell is taken by a tumor cell, it density is set to be zero. The predefined growth distance (δp) is described in a following bullet point. • If a proliferative cell divides, it can produce a mutant daughter cell possessing an invasive phenotype with a prescribed probability γ (i.e., the mutation rate). The invasive daughter cell gains ECM degradation ability χ and motility µ, which enables it to leave the tumor and invade into surrounding ECM. The rules for updating invasive cells are given in a following bullet point. If the daughter cell is noninvasive, it is designated as a new proliferative cell. • A proliferative cell turns quiescent if there is no space available for the placement of a daughter cell within a distance δp from the proliferative cell, which is given by δp = bL(d−1)/d t , where b a nutritional parameter (see Table 1), d is the spatial dimension and Lt is the distance between the geometric tumor centroid xc and the tumor edge cell that is closest to the proliferative cell under consideration. • An invasive cell degrades the surrounding ECM (i.e., those in the neighboring automaton cells of the invasive cell) and can move from one automaton cell to another if the associated ECM is completely degraded. For an invasive cell with motility µ and ECM degradation ability χ, it will make m attempts to degrade the ECM in the neighboring automaton cells and jump to these automaton cells, where m is an arbitrary integer in [0, µ]. For each attempt, the surrounding ECM density ρECM is decreased by δρ, where δρ is an arbitrary number in [0, χ]. Using random numbers for ECM degradation ability and cellular motility is to take into account tumor genome heterogeneity, which is manifested as heterogeneous phenotypes (such as different m and δρ). When the ECM in multiple neighboring automaton cells of the invasive cell are completely degraded (i.e., ρECM = 0), the invasive cell moves in a direction that maximizes the nutrients and oxygen supply. Here we assume that the migrating invasive cells do not divide. The degraded ECM shows the invasive path of the tumor. 10 The aforementioned automaton rules are briefly illustrated in Fig. 3. We note that non-invasive tumor growth can be studied by imposing a mutation rate γ = 0. This enables us to compare the growth dynamics of invasive and non-invasive tumors and in turn to investigate the effects of coupling growth of the primary tumor mass and the invasive cells. Although we only consider spherical-growth-permitting regions here, the CA rules given above allow growth-permitting regions with arbitrary shapes. The important parameters mentioned in the bullet points above are summarized in Table 1. In the following, we will employ our CA model to investigate the growth dynamics of malignant tumors with different degrees of invasiveness in variety of different heterogeneous microenvironments. Quantitative Metrics for Tumor Morphology To characterize quantitatively the morphology of simulated tumors, we present several scalar metrics that capture the salient geometric features of the primary tumor, dendritic invasive branches or the entire invasive pattern. These metrics include the ratio β of invasive area over tumor area (defined below), the specific surface s of the invasive pattern, the asphericity α of the primary tumor and the angular anisotropy metric ψ for the invasive branches. The metrics are computed for all simulated tumors and compared to available experimental data. We note that the invasive pattern associated with a neoplasm includes both the primary tumor and the invasive branches. Following Ref. [4], the tumor area AT is defined as the area of the circumcircle of the primary tumor (see Fig. 4(a)) and the invasive area AI is the area of the region between the effective circumcircle of the invasive pattern and the circumcircle of the primary tumor (see Fig. 4(a)). The radius of the effective circumcircle of the invasive pattern is defined to be the average distance from the invasive branch tip to the tumor center. The ratio β = AI /AT as a function of time t reflects the degree of coupling between the primary tumor and the invasive cells. If β(t) is linear in t, there is no coupling; otherwise the two are coupled. The specific surface s [34] for the invasive pattern is defined as the ratio of the total length of the perimeter of the invasive pattern over its total area. In general, s is inversely proportional to the size of the tumor and thus, large tumors have small s values. Moreover, given the tumor size, tumors with a large number of long dendritic invasive branches possess a large value of s. And s is minimized for perfectly circular tumors with s = 2/RT , where RT is the radius. Since s depends on the size of the tumor, which makes it difficult to compare tumors with different sizes, in the calculations that follow 11 we employ a normalized s with respect to 2/RT for an arbitrary-shaped tumor with effective radius RT (i.e., the average distance from tumor edge to tumor center). For simplicity, we will still refer to the normalized specific surface as "specific surface" and designate it with symbol s. The asphericity α of the primary tumor is defined as the ratio of the radius of circumcircle Rc of the primary tumor over its incircle radius Rin [45], i.e., α = Rc/Rin (see Fig. 4(b)). A large α value indicates a large deviation of the shape of primary tumor from that of a perfect circle, i.e., the tumor is more anisotropic. To quantify the degree of anisotropy of the invasive branches, we introduce the angular anisotropy metric ψ. In particular, the entire invasive pattern is evenly divided into na sectors with lines emanating from the tumor center (see Fig. 4(c)). The angular anisotropy metric ψ is defined as ψ = Pna na · ℓave i=1 ℓ(i) − ℓave , where ℓ(i) is the average length of the invasive branches within the ith sector and ℓave = Pna i=1 ℓ(i) na , (1) (2) is the average length of all invasive branches. For tumors with invasive branches of similar lengths that are uniformly angularly distributed, the metric ψ is small. Large fluctuations of both invasive branch length and angular distribution can lead to large ψ values. In the following, we use na = 16 to compute ψ for the simulated invasive tumors. Results Model Validation To verify the robustness and predictive capacity of our CA model, we first employ it to quantitatively reproduce the observed invasive growth of a GBM multicellular tumor spheroid (MTS) in vitro [4]. In particular, the boundary of the growth-permitting region is considered to be vascularized, i.e., a growing tumor can receive oxygen and nutrients from the growth-permitting region. A constant radially symmetric nutrient/oxygen gradient in the growth-permitting region with the highest nutrient/oxygen 12 concentration at the vascular boundary is used. Initially, approximately 250 proliferative tumor cells are introduced at the center of the growth-permitting region with homogeneous ECM and tumor growth is started. This corresponds to an initial MTS with diameter DMTS ≈ 310 µm which is consistent with the in vitro experiment set-up [4]. The following values of the growth and invasiveness parameters are used: p0 = 0.384, a = 0.58 mm1/2, b = 0.30 mm1/2, γ = 0.05, χ = 0.55, µ = 3. Note that the value of p0 corresponds to a cell doubling time of 40 hours, which is consistent with the reported experimental data [4]. A small value of ECM density ρECM = 0.15 is used, which corresponds to the soft DMEM medium used in the experiment [4]. In the visualizations of the tumor that follow, we use the following convention: the ECM in the growth-permitting region is white, and gray outside this region. The ECM degraded by the tumor cells is blue. In the primary tumor, necrotic cells are black, quiescent cells are yellow and proliferative cells are red. The invasive tumor cells are green. Figure 5(a) and (b) respectively show the morphology of simulated MTS and a magnification of its invasive branches with increasing branch width towards the proliferative core. Specifically, one can clearly see that within the branches, chains of cells are formed as observed in experiments [4] (see Fig. 1). The invasive cells tend to follow one another (which is termed "homotype attraction") since paths of degraded ECM are formed by pioneering invasive cells and it is easier for other cells to follow and enhance such paths than degrading ECM to create new paths by themselves. In other words, invasive cells tend to take paths with "least resistance". We note that no CA rules are imposed to force such cellular behaviors. Instead, they are emergent properties that arise from our simulations. The ratio of invasion area over primary tumor area β = AI /AT as a function of time for the simulated tumor is computed and compared to reported experimental data [4] (see Fig. 5(c)). One can clearly see that our simulation results agree with experimental data very well. Moreover, the deviation of β(t) from a linear function of t indicates that the growth of primary tumor and the invasive branches are strongly coupled [4] Other metrics for tumor morphology such as the specific surface s of the invasive pattern, the sphericity α of the primary tumor and the angular anisotropy metric ψ for the invasive branches are computed from our simulation results and from the image of invasive MTS in Fig. 1(a) at 24 hours after initialization. The values are given in Table 2, from which one can see again a good agreement. Thus, we have shown that our CA model is both robust and quantitatively accurate with properly selected parameters. 13 Simulated Invasive Growth in Heterogeneous Miroenvironments Having verified the robustness and predictive capacity of our CA model, we now consider three types of ECM density distributions, i.e., homogeneous, random and sine-like, to systematically study the effects of microenvironment heterogeneity on invasive tumor growth (see Fig. 6). These ECM density distributions represent real host microenvironments in which a tumor grows. (Details about these ECM distributions are given in the following sections.) Again, the boundary of the growth-permitting region is considered to be vascularized with a constant radially symmetric nutrient/oxygen gradient in the growth-permitting region pointing to the tumor center. We note that although generally the nutrient/oxygen concentration field in vivo is more complicated, previously numerical studies that considered the exact evolution of nutrient/oxygen concentrations have shown a decay of the concentrations toward the tumor center [22,23]. Since the directions of cell motions are determined only by the nutrient/oxygen gradient, our constant- gradient approximation is a very reasonable one. In the beginning, a proliferative tumor cell is introduced at the center of the growth-permitting region and tumor growth is initiated. The growth parameters for the primary tumors in all cases studied here are the same and are given in Table 1. The invasiveness parameters and ECM densities are variable and specified in each case separately. The values of the growth parameters for the CA model were chosen to be consistent with GBM data from the medical literature [16, 20]. Specifically, the value of the base probability of division is p0 = 0.192, which corresponds to a cell doubling time of 4 days [46, 47]. This value is used for all of the cases of invasive growth that follow. Since our CA model takes into account general microscopic tumor-host interactions, we expect that the general growth dynamics and emergent behaviors predicted by the model will qualitatively apply to other solid tumors. We note that all of the reported growth dynamics and emergent properties of the simulated tumors for any specific set of growth and invasiveness parameters are repeatedly observed in 25 independent simulations. Effects of Cellular Motility We first simulate the growth of malignant tumors with different degrees of invasiveness in a homogeneous ECM with ρECM = 0.45. In particular, we consider three invasive cases with the same mutation rate γ = 0.05 and ECM degradation ability χ = 0.9, but different cell motility µ = 1, 2, 3. A non-invasive growth case (i.e., γ = 0) in the same microenvironment (ρECM = 0.45) is also studied for comparison purposes. 14 Figure 7 shows the simulated growing tumors 100 days after initiation (plots showing the full growth history of the tumors are given in the Supplementary Information). The computed metrics for tumor morphology are given in Table 3. The primary tumors for both invasive [Figs. 7(b),(c) and (d)] and non-invasive [Fig. 7(a)] cases develop necrotic and quiescent regions. For invasive tumors, when the cell motility is small (i.e., µ = 1), the invasive cells do not form dendritic invasive branches but rather clump near the outer border of the proliferative rim [see Fig. 7(b)], forming bumpy invasive concentric-like shells with relatively small specific surface (e.g., s = 1.09 on day 100). Such invasive shells significantly enhance the growth the primary tumor, i.e., the size of the primary tumor in Fig. 7(b) is much larger than in Figs. 7(a), (c) and (d). A quantitative comparison of the tumor sizes is given in the Supplementary Information. By contrast, for larger cell motility, long dendritic invasive branches are developed as manifested by the large specific surface (e.g., s = 7.89 on day 100 and s = 9.73 on day 120 for µ = 3). In particular, one can clearly see that within the branches the cells tend to follow one another to form chains, as observed in experiments [4]. We emphasize that no rules are imposed to force the cells to follow on another in our CA model. This homotype attraction is purely due to the mechanical interaction between the invasive cells and the ECM, i.e., once a path of invasion is established by a leading invasive cell (by degrading the ECM), other invasive cells nearby turn to follow and enhance this path since the resistance for migration is minimized on a existing path. Furthermore, we can see that larger cell motility (i.e., high malignancy) leads to more invasive branches [see Figs. 7(c) and (d)] and thus, a larger specific area of the invasive pattern. Effects of the ECM Rigidity It is not very surprising that isotropic tumor shapes and and invasive patterns are developed in a ho- mogeneous ECM with relative low density (i.e.,the ECM is soft) compared to the ECM degradation ability of the invasive tumor cells. However, real tumors are rarely isotropic, primarily due to the host microenvironment in which they grow, which we now explore. Consider the invasive growth of a tumor in a much more rigid ECM than that in the previous section, i.e., ρECM = 0.85. The invasiveness parameters used are γ = 0.05, µ = 3 and χ = 0.9. The snap shots of the growing tumor are shown in Fig. 8 and the tumor morphology metrics are given in Table 3. It can be clearly seen that both the size of the primary tumor and the extent of its invasive branches are 15 much smaller than those of the tumors growing in a softer ECM (see Fig. 7). Importantly, although the ECM is still homogeneous, due to its high rigidity, the primary tumor develops an anisotropic shape with protrusions in the proliferation rim caused by the invasive branches (e.g., α = 1.40 and ψ = 1.02 on day 100). Since the invasive cells have degraded the ECM either completely or partially along the invasive branches, it is easier for the proliferative cells in the primary tumor to take these "weak spots" than to push against the rigid ECM themselves. Again, we emphasize that we do not force the cells to behave this way by imposing special CA rules; this behavior results purely from the mechanical interaction between the tumor and its host and the coupling dynamics of invasive and non-invasive tumor cells. Effects of the ECM Heterogeneity: Random Distribution of ECM Density The real host microenvironment for tumors are far from homogeneous in general. To investigate how heterogeneity of ECM affects the tumor growth dynamics, we use a random distribution of ECM density, i.e., for each ECM associated automaton cell, it density ρECM is a random number uniformly chosen from the interval [0, 1] (see Fig. 6(b)). The invasiveness parameters used are γ = 0.05, µ = 3 and χ = 0.9 and snap shots of the growing tumor are shown in Fig. 9. The tumor morphology metrics are given in Table 4. Note that the primary tumor develops a rough surface and slightly anisotropic shape in the early growth stages (e.g., α = 1.32 on day 50 and α = 1.34 on day 80), which reflects the ECM heterogeneity [Figs. 9(a) and (b)]. Since the characteristic heterogeneity length scale is comparable to a single cell, its effects are diminished (e.g., α = 1.18 on day 100 and α = 1.15 on day 120) as the tumor grows larger and larger [Fig. 9(c) and (d)]. (In other words, on large length scales, the ECM is still effectively homogeneous.) However, the anisotropy in the invasive pattern (i.e., the extents of invasive branches in different directions) still persists (e.g., ψ = 0.64 on day 100) even though the primary tumors almost resumes an isotropic shape. Effects of the ECM Heterogeneity: Sine-like Distribution of ECM Density To represent large-scale heterogeneities in the ECM, we use a sine-like distribution of ECM density, i.e., for an automaton cell with centroid (x1, x2, . . . , xd), the associated ECM density is given by ρECM(x1, . . . , xd) = 1 2d [sin( 2πx1 L ) + 1][sin( 2πx2 L ) + 1] · · · [sin( 2πxd L ) + 1], (3) 16 where d is the spatial dimension and L is the edge length of the d-dimensional cubic simulation box. A two-dimensional sine-like ECM distribution is shown in Fig. 6(c). The red spots correspond to large ρECM and high ECM rigidity; they can be considered as effective obstacles (e.g., brain ventricles) that hinder tumor growth. Figures 10(a),(b) and (c),(d) show the snap shots of invasive tumors growing in the aforementioned ECM on day 80 and day 120, with invasiveness parameters γ = 0.05, µ = 1 and χ = 0.9 and γ = 0.05, µ = 3 and χ = 0.9, respectively. The plots showing the full growth history is given in Supplementary Information and the tumor morphology metrics are given in Table 4. We can see that in the early growth stage, both the primary tumor and invasive pattern in the two cases are significantly affected by the ECM heterogeneity. In particular, the tumors are highly anisotropic in shape and the invasive branches clearly favor two orthogonal directions associated with low ECM densities (e.g., α = 1.61, ψ = 1.18 for µ = 1 on day 80; and α = 1.67, ψ = 1.33 for µ = 3 on day 80). For the case with large cellular motility, anisotropy effects are diminished in later growth stages (α = 1.21, ψ = 0.23 for µ = 3 on day 120). For small cellular motility, anisotropy in both primary tumor shape and the invasive pattern persists (α = 1.26, ψ = 0.98 for µ = 1 on day 120). Furthermore, one can see that again invasive cells with low motility significantly facilitate the growth of the primary tumor. However, instead of forming "bumpy" concentric-like shells as in homogeneous ECM, the invasive cells form large invasive cones, protruding into the ECM. These invasive cones are followed by weak protrusion of the proliferative rim, leading to bumpy surface of the primary tumor. The fact that such complex growth dynamics are only observed for tumors growing heterogeneous ECM emphasizes the crucial importance of understanding the effects of physical heterogeneity in cancer research. Discussion We have developed a novel cellular automaton (CA) model which, with just a few parameters, can produce a rich spectrum of growth dynamics for invasive tumors in heterogeneous host microenvironment. Besides robustly reproducing the salient features of branched invasive growth, such as least resistance and intrabranch homotype attraction observed in in vitro experiments, our model also enables us to systematically investigate the effects of microenvironment heterogeneity on tumor growth as well as the coupling growth of the primary tumor and the invasive cells. In particular, we have shown that 17 in homogeneous ECM with low densities (i.e., soft microenvironment), both the shape of the primary tumor and invasive pattern are isotropic. For high cellular motility cases, the invasive cells form extended dendritic invasive branches; while for low cellular motility cases, the invasive cells clump near the primary tumor surface and form a bumpy concentric-like shell that facilitates the growth of the primary tumor. Tumors growing in a highly rigid homogeneous ECM can develop anisotropic shapes, facilitated by the invasive cells that degrade the ECM; both the tumor size and the extent of invasive branches are much smaller. In heterogeneous ECM, both the primary tumor and invasive pattern are significantly affected during the early growth stages, i.e., anisotropic shapes and patterns are developed to avoid high density/rigid regions of the ECM. If the characteristic length scale of the heterogeneities is comparable to the macroscopic tumor size, such effects can persist in later growth stages. In addition, invasive cells with large motility can significant diminish the anisotropy effects by their ECM degradation activities. We emphasize that we did not manipulate the behavior of cells by imposing artificial CA rules to give rise to these complex and rich growth dynamics. Instead, these are emergent behaviors that naturally arise due to various microscopic-scale tumor-host interactions that are incorporated into our CA model, including the short-range mechanical interaction between tumor cells and tumor stroma, and the degradation of extracellular matrix by the invasive cells. Note that the growth dynamics of tumors growing in a heterogeneous microenvironment is distinctly different than those in a homogeneous microenvironment. This emphasizes the importance of under- standing the effects of physical heterogeneity of the host microenvironment in modeling tumor growth. Here we just make a first attempt to take into account a simple level of host heterogeneity, i.e., by consid- ering the ECM with variable density/rigidity. Currently, the invasion of the malignant cells into the host microenivronment is considered to be a consequence of invasive cell phenotype gained by mutation, and is not triggered by environmental stresses. However, the effects of environmental stresses can be easily taken into account. For example, a CA rule can be imposed that if the division probability of a malignant cell is significantly reduced by ECM rigidity, i.e., it is extremely difficult to push away/degrade ECM to make room for daughter cells, the malignant cell leaves the primary tumor and invades into soft regions of surrounding ECM. This would lead to reduced tumor invasion (i.e., development of the dendritic invasive branches) in soft microenvironments but enhanced invasion in rigid microenvironments [48]. Moreover, the spatial-temporal evolution of more complicated and realistic nutrient/ oxygen fields can be incorporated into our CA model. This can be achieved by solving the coupled nonlinear partial differ- 18 ential equations governing the evolution of the nutrient/oxygen concentrations as was done in Refs. [19] and [21]. Since the CA rules are given for any spatial dimension, our model is readily generalized to three dimensions. In addition, the model can be easily modified to incorporate other host heterogeneities, such as stromal cells, blood vessels and the shape anisotropy of the host organ [20, 21]. As currently imple- mented, a single 2D simulation takes less than 0.5 hours on a 32-bit 1.56Gb Memory 1.44GHz dual core Dell Workstation. We expect that a 3D simulation will take no longer than 24 hours on a supercomputer when a proper parallel implementation is used. Such an in silica tool not only enables one to investigate tumor growth in complex heterogeneous microenvironment that closely represents the real host microenvironments but also allows one to infer and even reconstruct individual host microenvironment given limited growth data of tumors (such as shape and size at various times). Such microstructural information of the individual host would be extremely valuable for developing individualized treatment strategies. For example, based on the host microstructure one can design special encapsulation and transport agents that maximize drug delivery efficiency [13]. In our current CA model, the microscopic parameters governing tumor invasion are variable and can be arbitrarily chosen within a feasible range as given in Table 1. Given sufficient and reliable experimental data of invasive tumor growth, the parameters in our CA model can be uniquely determined and thus, the model can produce robust predictions on the neoplastic progression. Although the current CA model is specifically implemented to reproduce and predict the growth dynamics of invasive solid tumors in vitro, further refinement of the model could eventually lead to the development of a powerful simulation tool that can be utilized clinically. For example, more complicated and realistic host heterogeneities such as the vascular structure, various stromal cells, the corresponding spatial-temporal evolution of the nutrient/oxygen concentrations as well as environmental stress-induced mutations should be incorporated as we described earlier. If the robustness of the refined model could be validated clinically, we would expect it to produce quantitative predictions for in vivo tumor growth, which are valuable for tumor prognosis and proposing individualized treatment strategies. Acknowledgments The authors are grateful to Robert Gatenby and Bob Austin for valuable comments on our manuscript. The authors are also grateful to the anonymous reviewers for their valuable comments. 19 References 1. Coffey DS (1998) Self organization, complexity and chaos: The new biology for medicine. Nat Med 4: 882-885. 2. Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell 100:57-70. 3. Fearon ER, Vogelstein B (1990) A genetic model for colorectal tumorigenesis. Cell 61:759-767. 4. Deisboeck TS, Berens ME, Kansal AR, Torquato S, Rachamimov A, Louis DN, Chiocca EA (2001) Patterns of self-organization in tumor systems: Complex growth dynamics in a novel brain tumor spheroid model. Cell Proliferat. 34:115-134. 5. Fidler IJ (2003) The pathogenesis of cancer metastasis: the "seed and soil" hypothesis revisited. Nat Rev Cancer 3:453-458. 6. Kerbel RS (1990) Growth dominance of the metastatic cancer cell: cellular and molecular aspects. Adv Cancer Res 55:87-132. 7. Liotta LA, Kohn EC (2003) Cancer's deadly signature. Nat Genet 33:10-11. 8. Chen ZJ, Gillies GT, Broaddus WC, Prabhu SS, Fillmore H, Mitchell RM, Corwin FD, Fatouros PP (2004) A realistic brain tissue phantom for intraparenchymal infusion studies. J. Neurosurgery 101:314-322. 9. Frieboes HB, Edgerton ME, Fruehauf JP, Rose FR, Worrall LK, Gatenby RA, Ferrari M, Cristini V (2009) Prediction of drug response in breast cancer using integrative experimental/computational modeling. Cancer Res. 69:4484-4492. 10. Crossa FR, Tinkelenberga AH (1991) A potential positive feedback loop controlling cln1 and cln2 gene expression at the start of the yeast cell cycle. Cell 65:875-883. 20 11. Brand U, Fletcher JC, Hobe M, Meyerowitz EM, Simon1dagger R (2000) Dependence of stem cell fate in arabidopsis on a feedback loop regulated by clv3 activity. Science 289:617-619. 12. Kitano H (2004) Cancer as a robust system: implications for anticancer therapy. Nat Rev Cancer 4:227-235. 13. Torquato S (2011) Toward an Ising model of cancer and beyond. Phys Biol 8:015017. 14. Byrne HM (2010) Dissecting cancer through mathematics: from the cell to the animal model. Nat Rev Cancer 10:221-230. 15. Anderson ARA, Chaplain MAJ (1998) Continuous and discrete mathematical models of tumor- induced angiogenesis. Bull Math Biol 60:857900. 16. Kansal AR, Torquato S, Harsh GR, Chiocca EA, Deisboeck TS (2000) Simulated brain tumor growth using a three-dimensional cellular automaton. J Theor Biol 203:367-382. 17. Kansal AR, Torquato S, Chiocca EA, Deisboeck TS (2000) Emergence of a subpopulation in a computational model of tumor growth. J Theor Biol 207:431-441. 18. Schmitz JE, Kansal AR, Torquato S (2002) A Cellular Automaton Model of Brain Tumor Treat- ment and Resistance. J Theor Med 4:223-239. 19. Gevertz JL, Torquato S (2006) Modeling the effects of vasculature evolution on early brain tumor growth. J Theor Biol 243:517-531. 20. Gevertz JL, Gillies GT, Torquato S (2008) Simulating tumor growth in confined heterogeneous environments. Phys Biol 5:036010. 21. Gevertz JL, Torquato S (2009) Growing heterogeneous tumors in silico. Phys Rev E 80:051910. 22. Anderson ARA (2005) A hybrid mathematical model of solid tumor invasion: the important of cell adhesion. Math Med Biol 22:163-186. 23. Anderson ARA, Weaver AM, Cummings PT, Quaranta V (2005) Tumor morphology and pheno- typic evolution driven by selective pressure from microenvironment. Cell 127:905-915. 24. Bankhead III A, Heckendorn RB (2007) Using evolvable genetic cellular automata to model breast cancer. Genet Program Evolvable Mach 8:381-393. 21 25. Gatenby RA (1996) Application of competition theory to tumour growth: implications for tumour biology and treatment. Eur J Cancer 32:722-726. 26. Gatenby RA, Gawlinski ET (1996) A reaction-diffusion model of cancer invasion. Cancer Res. 56:5745-5753. 27. Gatenby RA, Gawlinski ET, Gmitro AF, Kaylor B, Gillies RJ (2006) Acid-mediated tumor invasion: a multidisciplinary study. Cancer Res. 66:5216-5223. 28. Frieboes HB, Zheng XM, Sun CH, Tromberg B, Gatenby R, Gristini V (2006) An integrated computational/experimental model for tumor invasion. Cancer Res. 66:1597-1604. 29. Bellomo N, Preziosi L (2000) Modelling and mathematical problems related to tumor evolution and its interaction with the immune system. Math Comput Model 32:413-452. 30. Scalerandi M, Sansone BC, Condat CA (2001) Diffusion with evolving sources and competing sinks: Development of angiogenesis. Phys Rev E 65:011902. 31. Scalerandi M, Sansone BC (2002) Inhibition of vascularization in tumor growth. Phys Rev Lett 89:218101. 32. Kim Y, Friedman A (2010) Interaction of tumor with its micro-environment: A mathematical model. Bull Math Biol 72:1029-1068. 33. Stein AM, Demuth T, Mobley D, Berens M (2007) A mathematical model of glioblastoma tumor spheroid invasion in a three-dimensional in vitro experiment. Biophys J 92:356-365. 34. Torquato S (2002) Random Heterogeneous Materials: Microstructure and Macroscopic Properties. (Springer-Verlag, New York). 35. Torquato S, Stillinger FH (2010) Jammed hard-particle packings: From Keplerto Bernal and Be- yond. Rev Mod Phys 82:2633-2672. 36. Patel AB, Gibson WT, Gibson MC, Nagpal R (2009) Modeling and inferring cleavage patterns in proliferating epithelia. PLoS Compt Biol 5:e1000412. 37. Gevertz JL, Torquato S (2008) A novel three-phase model of brain tissue microstructure. PLoS Comput. Biol. 4:e1000152. 22 38. Burridge K, Chrzanowska-Wodnicka M (1996) Focal adhesions, contractability, and signalling. Annu. Rev. Cell Dev. Biol. 12:463-518. 39. Sarntinoranont M, Rooney F, Ferrari M (2003) Interstitial stress and fluid pressure within a growing tumor. Annals of Biomedical Engineering 31:327-335. 40. Gordon VD, Valentine MT, Gardel ML, Andor-Ardo D, Dennison S, Bogdanov AA, Weitz DA, Deisboeck TS (2003) Measuring the mechanical stress induced by expanding multicellular tumor system: a case stduy. Experimental Cell Res. 289:58-66. 41. Liotta LA, Rao CN, Barsky SH (1983) Tumor invasion and the extracellular matrix. Lab. Invest. 49:636-649. 42. Boyle JO, Hakim J, Koch W (1993) The incidence of p53 mutations increases with progression of head and neck cancer. Cancer Res. 53:4477-4480. 43. Stetler-Stevenson WG, Aznavoorian S, Liotta LA (1993) Tumor cell interactions with the extra- cellular matrix during invasion and metastasis. Annu. Rev. Cell Biol. 9:541-573. 44. Lawrence JA, Steeg PS (1996) Mechanisms of tumor invasion and metastasis. World J. Urol. 14:124-130. 45. Torquato S, Jiao Y (2009) Dense Packings of the Platonic and Archimedean Solids. Nature 460:876- 879. 46. Hoshino T, Wilson CB (1979) Cell kinetic analyses of human malignant brain tumors (gliomas). Cancer 44:956-962. 47. Pertuiset B, Dougherty D, Cromeryer C, Hoshino T, Berger M, Rosenblum ML (1985) Stem cell studies of human malignant brain tumors. Part 2: Proliferation kinetics of brain-tumor cells in vitro in early-passage cultures. J. Neurosurg. 63:426-432. 48. Guiot C, Pugno N, Delsanto PP, Deisboeck TS (2007) Physical aspects of cancer invasion. Phys. Biol. 4:P1-P6. Figure Legends 23 (a) (b) Figure 1. GBM multicelluar tumor spheroid (MTS) gel assay showing dendritic invasive branches. (a) The invasive branches centrifugal evolve from the central MTS. The linear size of central MTS is approximately 400 µm. (b) The invasive branches are composed of chains of invasive cells. The images are adapted from Ref. [4]. (a) (b) Figure 2. A 2D Voronoi tessellation and the associated point configuration. (a) A Voronoi tessellation of the 2D plane into polygons which are the automaton cells in our model. (b) The associated point configuration for the tessellation, generated by randomly placing nonoverlap circular disks in a prescribed region, i.e., the random sequential addition process [34]. 24 (a) (b) (c) (d) Figure 3. Illustration of cellular automaton rules. Necrotic cells are black, quiescent cells are yellow, proliferative cells are red and invasive tumor cells are green. The ECM associated automaton cells are white and the degraded ECM is blue. (a) A proliferative cell (dark red) is too far away from the tumor edge to get sufficient nutrients/oxygen and it will turn quiescent in panel (b). A quiescent cell (dark yellow) is too far away from the tumor edge and it will turn necrotic in panel (b). Another proliferative cell (light red) will produce a daughter proliferative cell in panel (b). (b) The dark red proliferative cell and the dark yellow quiescent cell in panel (a) turned quiescent and necrotic, respectively. The light red proliferative cell in panel (a) produced a daughter cell. Another proliferative cell (light red) will produce a mutant invasive daughter cell. (c) The light red proliferative cell in (b) produced an invasive cell. (d) The invasive cell degraded the surrounding ECM and moved to another automaton cell. (a) (b) (c) Figure 4. Schematic illustration of the quantities in the definitions of tumor morphology metrics. (a) Invasive area AI and tumor area AT associated with the invasive pattern. (b) Circumcircle with radius Rc and incircle with radius Rin associated with the primary tumor. (c) Evenly dividing the invasive pattern into na = 8 sectors for computing angular anisotropy metric ψ. 25 3 2.5 2 1.5 1 0.5 0 0 Experimental Data Simulation Results 100 50 Time t (hours) (c) 150 T A / I A = β (a) (b) Figure 5. Simulated invasive growth of MTS in vitro. (a) A snapshot of the simulated growing MTS at 24 hours after initialization. The region circled is magnified in panel (b). (b) A magnification of the circled region in panel (a). One can clearly see that the invasive cells (green) are following each other to form chains within the dendritic branches (blue), as observed in experiment [4]. (c) Comparison of β = AI /AT as a function of time associated with the simulated MTS and the in vitro experimental data [4]. (a) (b) (c) Figure 6. Different distributions of ECM densities. (a) Uniform distribution. (b) Random distribution, i.e., the value of ρECM is completely independent of ρECM values of other automaton cells. (c) Sine-like distribution defined by Eq. (3) to mimic the obstacles for a growing tumor. 26 (a) µ = 0 (b) µ = 1 (c) µ = 2 (d) µ = 3 Figure 7. Simulated growing tumors in homogeneous ECM with ρECM = 0.45 on day 100 after initiation. For the invasive growth, the mutation rate is γ = 0.05 and ECM degradation ability is χ = 0.9, (a) Tumor cells are noninvasive, i.e., γ = 0. (b) Invasive tumor with cellular motility µ = 1. (c) Invasive tumor with cellular motility µ = 2. (d) Invasive tumor with cellular motility µ = 3. Note that the size of the primary tumor whose growth is facilitated by the concentric-like shell formed by clumpped invasive cells (b) is much larger than the other cases. Invasive cells with a larger motility lead to more dendritic invasive branches. (a) (b) (c) (d) Figure 8. Evolution of simulated tumor in homogeneous ECM with ρECM = 0.85. The mutation rate is γ = 0.05, the cell motility is µ = 3 and ECM degradation ability is χ = 0.9. (a) Growing tumor on day 50. (b) Growing tumor on day 80. (c) Growing tumor on day 100. (d) Growing tumor on day 120. Note that although the ECM is homogeneous, due to its high rigidity, the primary tumor develops an anisotropic shape with protrusions in the proliferation rim caused by the invasive branches. Also note that the invasive cells clump at the tips of certain invasive branches due to the high ECM rigidity. 27 (a) (b) (c) (d) Figure 9. Evolution of simulated tumor in random ECM. The mutation rate is γ = 0.05, the cell motility is µ = 3 and ECM degradation ability is χ = 0.9. (a) Growing tumor on day 50. (b) Growing tumor on day 80. (c) Growing tumor on day 100. (d) Growing tumor on day 120. Note that both the primary tumor and invasive pattern are affected (i.e., becoming anisotropic) by the ECM heterogeniety in the early growth stages. Also note that unlike the case in Fig.6(d), the invasive cells clump at the tips of invasive branches since they have reached the boundary of the growth-permitting region. (a) µ = 1 (b) µ = 1 (c) µ = 3 (d) µ = 3 Figure 10. Simulated growing tumors in sine-like ECM with different motilities (µ = 1, 3). The mutation rate is γ = 0.05 and ECM degradation ability is χ = 0.9. (a) Growing tumor with cell motility µ = 1 on day 80. (b) Growing tumor with cell motility µ = 1 on day 120. (c) Growing tumor with cell motility µ = 3 on day 80. (d) Growing tumor with cell motility µ = 3 on day 120. Note that both the primary tumor and invasive pattern in the two cases are significantly affected by the ECM heterogeneity, i.e., the tumors are highly anisotropic in shape and the invasive branches clearly favor two orthogonal directions associated with low ECM densities in the early growth stages. 28 Table 1. Parameters and terms in the CA model Time dependent terms Local tumor radius (varies with cell positions) Lt Lmax Local maximum tumor extent (varies with cell positions) δp δn pdiv Characteristic proliferative rim thickness Characteristic living-cell rim thickness (determines necrotic fraction) Probability of division (varies with cell positions) Growth parameters Base probability of division, linked to cell-doubling time (0.192 and 0.384) Base necrotic thickness, controlled by nutritional needs (0.58 mm1/2) Base proliferative thickness, controlled by nutritional needs (0.30 mm1/2) Invasiveness parameters Mutation rate (determines the number of invasive cells, 0.05) ECM degradation ability (0.4 − 1.0) Cell motility (the number of "jumps" from one automaton cell to another, 1 − 3) p0 a b γ χ µ ρECM ECM density (determines the ECM rigidity and varies with positions, 0.0 − 1.0) Other terms Summarized here are definitions of the parameters for tumor growth and invasion, and all other (time-dependent) quantities used in the simulations. For each parameter, the number(s) listed in parentheses indicates the value or range of values assigned to the parameters during the simulations. The values of the parameters are chosen such that the CA model can reproduce reported growth dynamics of GBM from the medical literature [4, 16, 20]. Table 2. Comparison of tumor morphology metrics associated with simulated MTS and experimental data at 24 hours after tumor initialization. Metrics Simulated MTS Experimental data Specific surface s Asphericity α Angular anisotropy metric ψ 9.24 1.09 0.17 9.78 1.12 0.19 29 Table 3. Morphology metrics for simulated tumors growing in homogeneous ECM. Noninvasive tumor in ECM with ρECM = 0.45 Metrics Specific surface s Asphericity α Day 50 Day 80 Day 100 Day 120 1.23 1.21 1.13 1.18 1.09 1.08 1.04 1.12 Invasive tumor with µ = 1 in ECM with ρECM = 0.45 Metrics β = AI /AT Specific surface s Asphericity α Angular anisotropy metric ψ Day 50 Day 80 Day 100 Day 120 0.66 1.76 1.23 0.13 0.38 1.48 1.12 0.29 0.21 1.26 1.08 0.33 0.19 1.18 1.06 0.17 Invasive tumor with µ = 2 in ECM with ρECM = 0.45 Metrics β = AI /AT Specific surface s Asphericity α Angular anisotropy metric ψ Day 50 Day 80 Day 100 Day 120 1.28 1.94 1.42 0.86 2.12 3.92 1.38 0.67 2.67 3.67 1.16 0.64 2.08 3.28 1.23 0.45 Invasive tumor with µ = 3 in ECM with ρECM = 0.45 Metrics β = AI /AT Specific surface s Asphericity α Angular anisotropy metric ψ Day 50 Day 80 Day 100 Day 120 2.14 1.71 1.38 1.25 2.43 4.28 1.27 0.68 2.64 7.89 1.13 0.41 2.89 9.73 1.8 0.18 Invasive tumor with µ = 3 in ECM with ρECM = 0.85 Metrics β = AI /AT Specific surface s Asphericity α Angular anisotropy metric ψ Day 50 Day 80 Day 100 Day 120 - 1.21 1.33 - 5.17 3.40 1.36 1.32 3.89 3.91 1.40 1.02 2.63 5.79 1.56 0.65 30 Table 4. Morphology metrics for simulated tumors growing in heterogeneous ECM. Invasive tumor with µ = 3 in random ECM Metrics Day 50 Day 80 Day 100 Day 120 β = AI /AT Specific surface s Asphericity α Angular anisotropy metric ψ 2.54 2.47 1.32 0.63 4.14 3.97 1.34 0.87 2.13 4.65 1.18 0.64 2.78 8.98 1.15 0.19 Invasive tumor with µ = 1 in sine-like ECM Metrics β = AI /AT Specific surface s Asphericity α Angular anisotropy metric ψ Day 50 Day 80 Day 100 Day 120 2.78 1.89 1.42 1.23 2.46 2.95 1.61 1.18 1.28 2.73 1.49 1.09 0.86 1.92 1.26 0.98 Invasive tumor with µ = 3 in sine-like ECM Metrics β = AI /AT Specific surface s Asphericity α Angular anisotropy metric ψ Day 50 Day 80 Day 100 Day 120 5.24 2.51 1.19 1.36 3.86 4.12 1.67 1.33 3.13 6.13 1.48 0.88 2.96 8.76 1.21 0.23
1209.2217
1
1209
2012-09-11T04:20:28
Counting statistics for genetic switches based on effective interaction approximation
[ "physics.bio-ph", "cond-mat.stat-mech", "q-bio.QM" ]
Applicability of counting statistics for a system with an infinite number of states is investigated. The counting statistics has been studied a lot for a system with a finite number of states. While it is possible to use the scheme in order to count specific transitions in a system with an infinite number of states in principle, we have non-closed equations in general. A simple genetic switch can be described by a master equation with an infinite number of states, and we use the counting statistics in order to count the number of transitions from inactive to active states in the gene. To avoid to have the non-closed equations, an effective interaction approximation is employed. As a result, it is shown that the switching problem can be treated as a simple two-state model approximately, which immediately indicates that the switching obeys non-Poisson statistics.
physics.bio-ph
physics
Counting statistics for genetic switches based on effective interaction approximation Graduate School of Informatics, Kyoto University, Yoshida Hon-machi, Sakyo-ku, Kyoto-shi, Kyoto 606-8501, Japan Jun Ohkubo∗ Applicability of counting statistics for a system with an infinite number of states is investigated. The counting statistics has been studied a lot for a system with a finite number of states. While it is possible to use the scheme in order to count specific transitions in a system with an infinite number of states in principle, we have non-closed equations in general. A simple genetic switch can be described by a master equation with an infinite number of states, and we use the counting statistics in order to count the number of transitions from inactive to active states in the gene. To avoid to have the non-closed equations, an effective interaction approximation is employed. As a result, it is shown that the switching problem can be treated as a simple two-state model approximately, which immediately indicates that the switching obeys non-Poisson statistics. I. INTRODUCTION Counting statistics is a scheme to calculate all statis- tics related to specific transitions in a stochastic system. In the counting statistics, a master equation with discrete states is used to derive time-evolution equations for gen- erating functions related to the specific transitions. The scheme has been used to investigate Forster resonance energy transfer, and many successful results have been obtained [1–3]. Although the scheme is basically formu- lated for a system with a finite number of states, it is possible to use the scheme to investigate a system with an infinite number of states. However, as exemplified later, we have non-closed equations in general, so that it would be needed to develop approximation schemes suit- able for specific systems. As a first step, it is important to check whether an approximation scheme for the count- ing statistics is available for the system with an infinite number of states or not. In the present paper, we focus on dynamics in genetic It has been shown that stochastic behavior switches. plays an important role in gene regulatory systems [4–6], and there are many studies for the stochasticity in the gene regulatory systems from experimental points of view (e.g., see [7, 8]) and theoretical ones (e.g., see [9–16]). Not only studies by numerical simulations, but also those by analytical calculations have been performed. Some ana- lytical expressions for the static properties, i.e., station- ary distributions for the number of proteins or mRNAs, have already been obtained. In addition, in order to in- vestigate the role of the stochasticity in genetic switches, dynamical properties, i.e., switching behavior between active and inactive gene states, have also been studied. Basically, such dynamical properties have been investi- gated by numerical simulations (e.g., see [17]); only for a simple system, analytical expressions for the first-passage time distribution have been obtained [18]. The genetic switch is described by a master equation with an infinite ∗Email address: [email protected] number of states. Hence, if we can use the scheme of the counting statistics in order to investigate the dynamical properties in the genetic switches, it will be helpful to ob- tain deeper understanding and intuitive pictures for the genetic switches. The aim of the present paper is to seek the applicability of the counting statistics in order to investigate the dy- namical property in the genetic switches. It immediately becomes clear that a straightforward application of the counting statistics derives intractable non-closed equa- tions. In order to obtain simple closed forms, we here employ an effective interaction approximation [19]. As a result, we will show that the switching problem can be treated as a simple two-state model approximately. This result immediately gives us intuitive understanding for the switching behavior and the non-Poissonian property. The present paper is constructed as follows. In Sec. II, we give a brief explanation of a stochastic model for the genetic switch. In Sec. III, the counting statistics is em- ployed in order to count the number of transitions in the genetic switch, and, as a result, a simple two-state model is derived approximately. The derived approximated re- sults are compared with those of Monte Carlo simulations in Sec. IV. Section V gives concluding remarks. II. MODEL A gene regulatory system consists of many compo- nents, such as genes, RNAs, and proteins. Here, a simpli- fied model is used; mRNAs are neglected for simplicity, and an activated gene assumes to directly increase the number of proteins. In addition, in the simplified model, a repressed gene cannot produce any proteins. The above model has been used to investigate the switching behav- ior in previous works, and, for example, see [13] for details of the model. We summarize the model studied in the present paper in Fig. 1. The binding interaction is assumed to be a re- pressed one, and the gene is activated only when the reg- ulatory proteins are not binding the gene. The proteins are produced from the gene in the active state with rate active inactive degradation FIG. 1: A schematic illustration of the self-regulating gene with repressed binding interaction. When the regulatory pro- teins are combining the gene, the gene is repressed and there is no production of proteins. If the regulatory proteins are released from the gene, the gene becomes active and it can produce the proteins. We consider the transition between the active and inactive states as a switch. g, and proteins are degraded spontaneously with rate d. The regulatory proteins bind the gene with a rate func- tion H(n), where n is the number of free proteins. For example, H(n) = hn for a monomer interaction case, and H(n) = hn(n − 1)/2 for a dimer interaction case, where h is a rate constant for the binding. f is a rate constant with which the regulatory proteins are released from the repressor site of the gene. We here give short comments for the model from the viewpoint of experiments. Using this simplified model, we can discuss the connection among the model parame- ters, the number of proteins, and the switching behaviors. While the number of proteins n can be observed or esti- mated experimentally, as far as we know, there has not been an experimental technique to observe the attach- ment and detachment of the regulatory proteins directly. We hope that developments of single-molecule observa- tions in future would enable us to give information about the switching dynamics. III. COUNTING STATISTICS FOR THE NUMBER OF TRANSITIONS A. Master equation for the number of proteins Analytical treatments for the self-regulating gene sys- tem have been developed, and an exact solution is known for the monomer interaction case, i.e., H(n) = hn [11, 18]. In order to simplify the analytical treatments, an additional assumption has been used in some previous works [13, 19]; i.e., some of proteins are assumed to be inert when the gene state is active. The inert proteins cannot repress the gene, and it is not degraded. For the monomer interaction case, there is only one inert protein; the number of inert protein for the dimer interaction case is two, and so on. Note that the assumption of the in- ert proteins does not have physical meanings; this only simplify the analytical treatments (for details, see [13]). However, it has been shown that this assumption has lit- tle influence of the gene system, and then we employ the 2 assumption in the present paper. Let αn and βn be states in which there are n free pro- teins for the active and inactive states, respectively. The probabilities for αn and βn at time t satisfy the following master equations; dP (αn, t) dt dP (βn, t) dt =g[P (αn−1, t) − P (αn, t)] + d[(n + 1)P (αn+1, t) − nP (αn, t)] − hnP (αn, t) + f P (βn, t), (1) =d[(n + 1)P (βn+1, t) − nP (βn, t)] + hnP (αn, t) − f P (βn, t), (2) where P (αn, t) and P (βn, t) are probabilities for n free proteins for the active and inactive states, respectively. As stated in Sec. I, the exact solutions for stationary distributions of the number of proteins have been derived, and those are expressed using the Kummer confluent hy- pergeometric functions. For details, see [11, 13]. B. Counting statistics Using the concept of the counting statistics [1–3], it is possible to investigate dynamical properties, i.e., all statistics for the switching behavior between the active and inactive states. In the present paper, as an example, we calculate the number of transitions from the inactive state to the active state. The generating functions for the transitions are immediately obtained from the mas- ter equations (1) and (2). A brief explanation of the counting statistics is given in the Appendix, and we here give consequences of the counting statistics. A probability, with which there are k transitions from the inactive state to the active state during time t, is denoted by P (kt). The generating function for P (kt) is defined as F (λ, t) = ∞ Xk=0 P (kt)λk, (3) where λ is a counting variable. The generating func- tion gives all information related to "inactive → ac- tive" transitions. According to the scheme of counting statistics, we split F (λ, t) into restricted generating func- tions {φ(αn, λ, t)} and {φ(βn, λ, t)}, where φ(αn, λ, t) and φ(βn, λ, t) are the generating functions for the sys- tem in states αn and βn at time t, respectively. Using the scheme of the counting statistics, we obtain the follow- ing time-evolution equations for the restricted generating functions {φ(αn, λ, t)} and {φ(βn, λ, t)}: dφ(αn, λ, t) dt dφ(βn, λ, t) dt =g[φ(αn−1, λ, t) − φ(αn, λ, t)] + d[(n + 1)φ(αn+1, λ, t) − nφ(αn, λ, t)] − hnφ(αn, λ, t) + λf φ(βn, λ, t), (4) =d[(n + 1)φ(βn+1, λ, t) − nφ(βn, λ, t)] + hnφ(αn, λ, t) − f φ(βn, λ, t). (5) Although Eqs. (4) and (5) are similar to Eqs. (1) and (2), note that the final term in the right hand side of Eq. (4) has a factor λ. The factor λ is introduced in order to count the number of transitions, and we can count the number of transitions related to this term (for details, see Appendix). Using the above restricted generating functions, the generating function F (λ, t) is calculated as F (λ, t) = ∞ Xn=0 {φ(αn, λ, t) + φ(βn, λ, t)} . (6) Next, we introduce the following generating functions for φ(αn, λ, t) and φ(βn, λ, t): α(λ, z, t) ≡ β(λ, z, t) ≡ ∞ ∞ Xn=0 Xn=0 φ(αn, λ, t)zn, φ(βn, λ, t)zn. (7) (8) It is straightforward to derive the time-evolution equa- tions for the new generating functions α(λ, z, t) and β(λ, z, t) from Eqs. (4) and (5); dα(λ, z, t) dt =(z − 1)(cid:20)gα(λ, z, t) − d ∂α(λ, z, t) ∂z (cid:21) ∂α(λ, z, t) − hz ∂z + λf β(λ, z, t), (9) dβ(λ, z, t) dt = − (z − 1)d ∂β(λ, z, t) ∂z + hz ∂α(λ, z, t) ∂z − f β(λ, z, t). (10) Using the generating function α(λ, z, t) and β(λ, z, t), the generating function F (λ, t) is given by 3 where we define α(λ, t) ≡ α(λ, z = 1, t) and β(λ, t) ≡ β(λ, z = 1, t). Note that Eqs. (12) and (13) contain the derivative of α(λ, z, t) with respect to z. Hence, the equations are not closed. If these terms are expressed simply using α(λ, t), we will have simultaneous differential equations written only by the generating functions α(λ, t) and β(λ, t); i.e., we have closed equations and hence the obtained equa- tions may be solved analytically. In the following anal- ysis, an effective interaction approximation is employed, and we will show that the above statistics can be approx- imated by a simple two-state model. C. Approximation for the interaction In the effective interaction approximation, the interac- tion function H(n) is replaced as a constant value. As shown in [19], the dependence of H(n) on n makes it dif- ficult to obtain analytical results, and it has been shown that the approximation gives qualitatively good results. Replacing the interaction function H(n) as H(n) = h, (14) where h is a constant, we obtain the following equations instead of Eqs. (12) and (13): dα(λ, t) dt dβ(λ, t) dt = − hα(λ, t) + λf β(λ, t), =hα(λ, t) − f β(λ, t). (15) (16) Note that Eqs. (15) and (16) are written only by α(λ, t) and β(λ, t). It means that the switching problem can be approximated as a simple two-state model if the effective interaction h is chosen adequately. We here briefly explain the choice of the effective in- teraction h using a simple example, i.e., the monomer binding interaction case. For the monomer binding in- teraction, the interaction function is calculated as follows [19]. In this case, the interaction function is hn. In or- der to obtain the effective interaction h, the number of proteins n is replaced as the average number of proteins, i.e., F (λ, t) = α(λ, z = 1, t) + β(λ, z = 1, t), (11) h = hhniα, (17) and therefore it is enough to solve the following time- evolution equations in order to calculate the generating function F (λ, t): dα(λ, t) dt = − h dβ(λ, t) dt =h ∂α(λ, z, t) ∂z ∂α(λ, z, t) ∂z (cid:12)(cid:12)(cid:12)(cid:12)z=1 (cid:12)(cid:12)(cid:12)(cid:12)z=1 + λf β(λ, t), (12) − f β(λ, t), (13) where hniα is the expectation of the number of free regu- latory proteins under a condition that the gene is in the active state (conditional expectation). The conditional expectation can be calculated from the stationary distribution of the number of proteins. Note that the generating functions α(λ, z, t) and β(λ, z, t) are reduced to generating functions for the stationary distri- bution of the number of proteins when λ = 1. Hence, as shown in [19], they are written as follows. α(z) ≡ lim t→∞ β(z) ≡ lim t→∞ α(λ = 1, z, t) = AF [a, b, N (z − 1)], (18) β(λ = 1, z, t) = 1 + h f! AF [a − 1, b − 1, N (z − 1)] − α(z) (19) where A = f /(f + h) and N = g d , a = 1 + f d , b = 1 + f + h d . F (p, q, r) is the Kummer confluent hypergeometric func- tion, F (p, q, r) ≡ (p)n (q)n rn n! , ∞ Xn=0 (20) where (p)n = p(p+ 1)(p+ 2) · · · (p+ n− 1). We, therefore, obtain hniα ≡ 1 α(1) ∂ ∂z = g(d + f ) d(d + f + h) . (21) α(z)(cid:12)(cid:12)(cid:12)(cid:12)z=1 By inserting Eq. (21) into Eq. (17), the following self- consistent equation is derived: h = h g(d + f ) d(d + f + h) . (22) Solving Eq. (22), we obtain h = −(d2 + f d) +p(d2 + f d)2 + 3hgd(d + f ) 2d . (23) We finally comment on a solution of the simple two- state model (Eqs. (15) and (16)). The simple two-state model can be solved exactly [1, 3], and the probability distribution P (kt) for the number of "inactive → active" transitions during time t is explicitly written as follows: P (kt) =(cid:18) (1 − γ 2)T 2γ (cid:19)k e−T k!p8γT /π ×(cid:8)2γ(k + T )Ik−1/2(γT ) + (1 + γ 2)T Ik+1/2(γT )(cid:9) , (24) where T = (f + h)t/2, γ 2 = 1 − 4f β(1)/(f + h), and In(z) are modified Bessel functions of the first kind. This expression (24) immediately gives us the non-Poissonian picture of the phenomenon. IV. NUMERICAL RESULTS 4 t = 10 t = 100 0.6 (a) y t i l i b a b o r P 0.4 0.2 0.0 0 4 8 12 16 20 k t = 10 t = 100 0.6 (b) y t i l i b a b o r P 0.4 0.2 0.0 0 4 8 12 16 20 k FIG. 2: Probability distributions for the number of "inactive → active" transitions. (a) Monomer binding interaction case. (b) Dimer binding interaction case. In each figure, filled cir- cles and filled boxes are Monte Carlo results for time t = 10 and t = 100, respectively. Solid and dashed lines corresponds to approximated analytical results of Eq. (24) for time t = 10 and t = 100, respectively. analytical results with those of Monte Carlo simula- tions. The original genetic switch explained in Sec. II was simulated using a standard Gillespie algorithm [20]. The parameters used in the simulation are as follows: d = 1, f = 0.1. Note that these parameters were selected as one of the typical val- ues used in the previous works [13, 19]. g = 50.0, h = 0.004, Firstly, we consider the monomer binding interaction case. According to the discussions in Sec. 3.3, the value of the effective interaction h is calculated as h = 0.173. Fig- ure 2(a) shows the results of the analytical calculations (Eq. (24)) and those of the Monte Carlo simulations. Although there are quantitative differences, the results shows that the approximated two-state model captures the essential features of the phenomenon. Next, we consider a dimer binding interaction case, i.e., H(n) = hn(n−1)/2. In this case, the effective interaction is calculated as follows: h = h hn(n − 1)iα 2 . (25) In order to check the validity of the analytical treat- ments and the approximations, we here compare the As shown in [19], the effective interaction h is obtained by solving the following self-consistent equation: h = h 2 1 α(1) ∂ 2 ∂z2 α(z)(cid:12)(cid:12)(cid:12)(cid:12)z=0 . (26) We here numerically solved the self-consistent equation (Eq. (26)), and the calculated value of the effective inter- action is h = 1.358. Using the calculated value, we depict the analytical results and the corresponding Monte Carlo results in Fig. 2(b). From the comparison, we confirmed that the approximated two-state model is available even in the dimer binding interaction case. Although results are not shown, we performed numerical simulations for other some parameters, and checked the validity of the analytical treatments. For example, even for parameter regions in which the probability distribution of the num- ber of proteins has bistability, the approximation scheme works well. V. CONCLUSIONS In the present paper, we studied an analytical scheme to extract information related to the dynamical behavior in genetic switches. Using an effective interaction ap- proximation, a simple two-state model is obtained, and we confirmed that the two-state model captures the fea- tures of the phenomenon. Note that in the analytical treatments, we did not neglect the stochastic properties of the system (except for the effective interaction approxi- mation); i.e., we can calculate all statistics for transitions approximately, including higher order moments. It could be possible to apply the above effective expression for the transitions between the active and inactive states to more complicated gene regulatory networks without loss of the stochasticity; this would give us deeper understanding for the switching behavior of the gene regulatory sys- tems including static, dynamical, and stochastic behav- iors. In addition, the idea of the effective interaction may be similar to the mean-field approximation in statistical physics; the interaction is replaced with the average. It may be possible to develop higher-order approximations using the analogy with the conventional approximation schemes in statistical physics; this is an important future work. We discussed properties only in the stationary states, because the effective interaction approximation has been applied only for the stationary states at the moment; the average number of proteins (or higher moments) should be estimated adequately, and it was calculated by using the analytical solutions for the stationary distributions of the number of proteins. Recently, exact time-dependent solutions for a self-regulating gene have been derived [21]. Hence, it may be possible to extend the effective interac- tion approximation to non-stationary states. If so, the effective interaction h would be time-dependent, and, at least numerically, it is possible to calculate various moments for the counting statistics for time-dependent 5 systems [22]. We expect that the simple description developed in the present paper is available for various cases, such as complicated regulatory systems and time- dependent systems, and that the description gives new insights for the regulation mechanisms and stochastic be- haviors. ACKNOWLEDGMENTS This work was supported in part by grant-in-aid for scientific research (Nos. 20115009 and 21740283) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan. Appendix A: Generating function for counting statistics Here, we give a brief explanation for the counting statistics for readers' convenience (For details, see [1–3].) In the framework of counting statistics, the quantity of interest is the number of target transitions. It is needed to set multiple target transitions in the genetic switches, and the genetic switches have two states, i.e., active and inactive states. In the following explanations, a simple setting, in which there is only one transition matrix and only one target transition, will be discussed because it is straightforward to apply the following simple discussions to the genetic switches. Let {Knm} be a transition matrix. We here derive the generating function for counting the number of events of a specific target transition iA → jA. Denote the prob- ability, with which the system starts from state m and finishes in state n with k transitions from iA to jA during time t, as Pnm(kt). In order to calculate the probability Pnm(kt), we here define a probability G′ kl(t) with which the system evolves from state l to state k, provided no iA → jA transitions occur during time t. By using the probability G′ kl(t), the probability Pnm(kt) is calculated as Pnm(kt) = G′ njA (t) ∗ KjAiA (t)G′ iAjA (t) ∗ · · · ∗ KjAiA (t)G′ iAjA (t) ∗ KjAiA (t)G′ iAm(t), k−1 {z } (A1) where g1(t) ∗ g2(t) ≡ R t 0 g1(t − t′)g2(t′)dt′ denotes the convolution. This formulation means that an occurrence of the target transition iA → jA is sandwiched in between situations with no occurrence of the target transition, and it is repeated k times. Next, we construct the generating function φnm(χ, t) of the probability Pnm(kt): φnm(χ, t) = ∞ Xk=0 λkPnm(kt). (A2) That is, the generating function φnm(λ, t) gives the statistics of the number of transition iA → jA during time t under the condition that the system starts from state m and ends in state n. The generating function φnm(λ, t) satisfies the following integral equation φnm(λ, t) = G′ nm(t) +Z t 0 G′ njA (t − t′)λKjAiA (t′) φiAm(λ, t′)dt′, and obeys the following time-evolution equation d dt φnm(λ, t) + λG′ =Xi 0 (cid:18) d +Z t =Xi dt Kni(t)G′ im(t) − δn,jA KjAiA (t)G′ iAm(t) njA (0)KjAiA (t) φiA m(t) G′ njA (t − t′)(cid:19) λKjAiA (t′) φiAm(t′)dt′ Kni(t) φim(λ, t) − δn,jA (1 − λ)KjAiA (t) φiAm(λ, t), (A4) where φnm(λ, 0) = δn,m. In order to show (A4), we used the following two facts. Firstly, the probability of no target transitions, G′ nm(t), obeys d dt G′ nm(t) =Xi Kni(t)G′ im(t) − δn,jA KjAiA (t)G′ iAm(t), (A5) where G′ convolution is given by nm(0) = δn,m. Secondly, the derivative of the 6 g1(t − t′)g2(t′)dt′ 0 d dtZ t = g1(0)g2(t) +Z t 0 (cid:18) d dt g1(t − t′)(cid:19) g2(t′)dt′. (A6) (A3) Using the generating function φnm(λ, t), we construct restricted generating functions {φn(λ, t)} as follows: φn(λ, t) =Xm φnm(λ, t)pm(0), (A7) where pm(0) is a probability distribution at initial time t = 0. From (A4) and (A7), the restricted generating function satisfies φn(λ, t) d dt =Xi Kni(t)φi(λ, t) − δn,jA (1 − λ)KjAiA (t)φiA (λ, t), (A8) tions φn(λ, 0) =Pm and these equations should be solved with initial condi- φnm(λ, 0)pm(0) = pn(0). The sum- mation of {φn(λ, t)} for n gives the objective generating function for counting the number of events of the specific target transition. [1] I.V. Gopich and A. Szabo, J. Chem. Phys. 118, 454 (2003). [2] I.V. Gopich and A. Szabo, J. Chem. Phys. 122, 014707 (2005). A.M. Walczak, J.N. Onuchic, and P.G. Wolynes, Phys. Rev. E 72, 051907 (2005). [12] B.-L. Xu and Y. Tao, J. Theor. Biol. 243, 214 (2006). [13] D. Schultz, J.N. Onuchic, and P.G. Wolynes, J. Chem. [3] I.V. Gopich and A. Szabo, J. Chem. Phys. 124, 154712 Phys. 126, 245102 (2007). (2006). [14] V. Shahrezaei and P.S. Swain, Proc. Natl. Acad. Sci [4] M.B. Elowitz, A.J. Levine, E.D. Siggia, and P.S. Swain, U.S.A 105, 17256 (2008). Science 297, 1183 (2002). [15] A.M. Walczak and P.G. Wolynes, Biophy. J. 96, 4525 [5] C.V. Rao, D.M. Wolf, and A.P. Arkin, Nature 420, 231 (2009). (2002). [16] J. Venegas-Ortiz and M.R. Evans, J. Phys. A: Math. [6] M. Kaern, T.C. Elston, W.J. Blake, and J.J Collins, Na- Theor. 44, 355001 (2011). ture Rev. Genetics 6, 451 (2005). [17] H. Feng, B. Han, and J. Wang, J. Phys. Chem. B 115, [7] T.S. Gardner, C.R. Cantor, and J.J. Collins, Nature 403, 1254 (2011). 342 (2000). [18] P. Visco, R.J. Allen, and M.R. Evans, Phys. Rev. E 79, [8] H. Okano, T.J. Kobayashi, H. Tozaki, and H. Kimura, 031923 (2009). Biophys. J. 95, 1063 (2008). [9] J. Hasty, J. Pradines, M. Dolnik, and J.J Collins, Proc. Natl. Acad. Sci U.S.A 97, 2075 (2000). [19] J. Ohkubo, Phys. Rev. E 83, 041915 (2011). [20] D.T. Gillespie, J. Phys. Chem. 81, 2340 (1977). [21] A.F. Ramos, G.C.P Innocentini, and J.E.M. Hornos, [10] M. Sasai and P.G. Wolynes, Proc. Natl. Acad. Sci U.S.A Phys. Rev. E 83, 062902 (2011). 100, 2374 (2003). [22] J. Ohkubo and T. Eggel, J. Stat. Mech., P06013 (2010). [11] J.E.M. Hornos, D. Schultz, G.C.P. Innocentini, J. Wang,
1207.0524
2
1207
2012-11-14T19:57:19
Nonlinear diffusion effects on biological population spatial patterns
[ "physics.bio-ph", "cond-mat.stat-mech", "q-bio.PE" ]
Motivated by the observation that anomalous diffusion is a realistic feature in the dynamics of biological populations, we investigate its implications in a paradigmatic model for the evolution of a single species density $u(x,t)$. The standard model includes growth and competition in a logistic expression, and spreading is modeled through normal diffusion. Moreover, the competition term is nonlocal, which has been shown to give rise to spatial patterns. We generalize the diffusion term through the nonlinear form $\partial_t u(x,t) = D \partial_{xx} u(x,t)^\nu$ (with $D, \nu>0$), encompassing the cases where the state-dependent diffusion coefficient either increases ($\nu>1$) or decreases ($\nu<1$) with the density, yielding subdiffusion or superdiffusion, respectively. By means of numerical simulations and analytical considerations, we display how that nonlinearity alters the phase diagram. The type of diffusion imposes critical values of the model parameters for the onset of patterns and strongly influences their shape, inducing fragmentation in the subdiffusive case. The detection of the main persistent mode allows analytical prediction of the critical thresholds.
physics.bio-ph
physics
Nonlinear diffusion effects on biological population spatial patterns Eduardo H. Colombo1 and Celia Anteneodo1,2 1Department of Physics, PUC-Rio, Rio de Janeiro, Brazil 2National Institute of Science and Technology for Complex Systems, Rio de Janeiro, Brazil. Motivated by the observation that anomalous diffusion is a realistic feature in the dynamics of biological populations, we investigate its implications in a paradigmatic model for the evolution of a single species density u(x, t). The standard model includes growth and competition in a logistic expression, and spreading is modeled through normal diffusion. Moreover, the competition term is nonlocal, which has been shown to give rise to spatial patterns. We generalize the diffusion term through the nonlinear form ∂tu(x, t) = D∂xxu(x, t)ν (with D, ν > 0), encompassing the cases where the state-dependent diffusion coefficient either increases (ν > 1) or decreases (ν < 1) with the density, yielding subdiffusion or superdiffusion, respectively. By means of numerical simulations and analytical considerations, we display how that nonlinearity alters the phase diagram. The type of diffusion imposes critical values of the model parameters for the onset of patterns and strongly influences their shape, inducing fragmentation in the subdiffusive case. The detection of the main persistent mode allows analytical prediction of the critical thresholds. PACS numbers: 89.75.Fb, 89.75.Kd, 05.65.+b, 05.40.-a I. INTRODUCTION Pattern formation in population dynamics has been studied both experimentally and theoretically. In exper- iments, the dynamics of insects and bacterial colonies, amongst others, have been observed [1 -- 5]. From the the- oretical viewpoint, mean-field descriptions render macro- scopic or mesoscopic approximations to describe the be- havior of such complex systems. To take into account spatial inhomogeneities, one may construct a partial dif- ferential equation that rules the temporal evolution of the population density u(~x, t), a function of the spatial position ~x and time t. Within this family of models, a standard one was first introduced by Fisher [6]. It consists of a reaction-diffusion equation, taking into ac- count growth and competition in the usual logistic form. Recently, a generalization of Fisher equation (FE) was introduced [7 -- 9], namely, ∂ ∂t u(~x, t) = D∇2u(~x, t) + u(~x, t)(a − bJ[u(~x, t)]) , (1) where D, a, b are positive parameters and J is a func- tional of the density embodying nonlocality: J[u(~x, t)] = ZΩ f (~x, ~x′)u(~x′, t)d~x′ . (2) In the particular case f (~x, ~x′) = δ(~x − ~x′), the original (local) FE is recovered. The introduction of the non- local form of the competition term is motivated by the consideration that products released in the environment by the individuals may either harm or support neigh- bors' growth. Interestingly, this nonlocal component was shown to give rise to the formation of steady spatial patterns [8]. Diverse variants have been studied before. As influence functions f (~x, ~x′), square and smooth forms have been considered [8, 9]. Nonlocality in the reproduc- tion rate [2], dimensionality [10] and fluctuation effects [11] have been investigated too. In all those cases, how- ever, spatial spread was described by normal diffusion. Meanwhile, there are indications that the spreading of bi- ological populations is not due to purely random motion but influenced by the density, either to favor or to avoid crowding [12 -- 14]. Hence dispersal is guided by a state- dependent diffusion coefficient rather than by a constant one. An important class of generalized diffusion equations is constituted by the porous media equation ∂tu = ∂xxuν, originally defined for ν > 1 [15]. Although it was later extended to real ν > −1 [16], here we will restrict our analysis to ν > 0. Nonlinear diffusion is ruled by a state-dependent diffusion coefficient, proportional to uν−1, hence embracing the cases where the coefficient ei- ther grows [17] or decreases [18] with the density u. The generalization of Arrhenius law [19], the performance of thermal ratchets [20], and other properties such as ag- ing [21] that arise under this kind of diffusion have been studied before. Nonlinear diffusion equations in higher dimensions [22] or even with space-fractional derivatives [23] have been analytically solved. The nonlinearity leads to anomalous diffusion [16]: either superdiffusion for ν < 1 or subdiffusion for ν > 1, recovering normal dif- fusion when ν = 1. Microscopically, high density regions can slow down (ν < 1) or intensify (ν > 1) individual displacements, as a consequence of homophilic behaviors that rule the dynamics of self-diffusion favoring or not the mobility among other individuals. While ν < 1 reflects a reaction to sparseness (with high diffusion coefficient where the density is low), on the contrary, ν > 1 is asso- ciated to immobilization in poorly populated regions. We will analyze the effects of nonlinear diffusion on pattern formation by considering the one dimensional generalized FE with nonlocal competition ∂ ∂t u(x, t) = D ∂ 2 ∂x2 uν(x, t) + u(x, t)(a − bJ[u(x, t)]) . (3) Since alternative forms of the functional f (x, x′) do not yielded substantially different results [8], we will restrict our analysis to the case where f (x, x′) is constant for x−w ≤ x′ ≤ x+w, and zero otherwise, namely f (x, x′) = 1 2w Θ(w−x−x′), where Θ is the Heaviside step function. II. RESULTS Numerical integration of Eq. (3) was performed by means of a standard forward-time centered-space scheme with integration time step dt ≤ 10−3 and width of spatial grid cells dx ≤ 0.1. We set periodic boundary conditions, with periodic domain size L = 100. As initial condition we considered small amplitude random perturbations ei- ther above the null state or around the nontrivial homo- geneous solution. We also considered square pulses (with small random fluctuations) with zero values everywhere else. 2.0 1.5 1.0 0.5 0.0 100 10-4 10-8 ) t , x ( u (a) (b) 10-12 0 5 10 15 x 2 0 10 20 30 40 50 70 100 200 20 25 30 8 6 4 2 0 0 20 40 60 80 100 (a) n = 4.0 n = 1.0 n = 0.2 FIG. 2: Time evolution of the density profile obtained from numerical integration of Eq. (3) with a = b = 1, L = 100, D = 0.1, w = 10 and ν = 4, represented for different times t indicated on the figure, in (a) linear and (b) logarithmic scales. The thick line corresponds to t = 200. 8 ) , x ( u 10 5 0 101 100 10-1 10-2 10-3 10-4 10-5 rapidly increases for all x towards the level correspond- ing to the homogeneous solution, u0 = a/b (t < 10), while patterns develop. After t = 100 no substantial changes are detected at the crests. Between successive crests, the density tends zero (exponentially fast with time). This fragmentation or clusterization process [24] yields isolated population groups (clusters). Therefore fluxes between clusters are eliminated in the long time limit. This phenomenon is crucial in connection with "epidemic" spreads within a population. A. Stability analysis To determine the stability conditions we follow the standard procedure of considering a first-order pertur- bation around the homogeneous solution u0 = a/b: u(x, t) = u0 + ǫ exp(ikx + λkt) , (4) where ǫ is the perturbation initial amplitude, k the wave number and λk is the exponential rate of temporal behav- ior. Substituting Eq. (4) into Eq. (3) gives the dispersion relation λk = −νDuν−1 0 k2 − a sin(wk) wk , (5) that generalizes the one obtained by Fuentes et al. Defining the nondimensional rate Λk ≡ λk [9]. a , Eq. (5) can 0 10 20 30 x (b) 40 50 60 FIG. 1: Longtime patterns obtained from numerical integra- tion of Eq. (3), with a = b = 1, L = 100, D = 0.1, w = 10 and different values of ν indicated on the figure, in (a) linear and (b) logarithmic scales. In the inset, the longterm profiles in the full grid are shown. The profiles are plotted for t = 200, but they remain unchanged after t ≃ 100. Typical longtime patterns, robust under changes in the initial conditions here considered, are shown in Fig. 1. Notice that, while the number of peaks is not affected by changing ν, the form of the patterns becomes sub- stantially different. By increasing ν, the width (inverse concavity) of the crests increases and the density at the valleys decreases, such that for ν > 1 disconnected re- gions can arise. Figure 2 shows the time evolution for ν = 4, start- ing with small random values of the density u(x, 0). It 0.5 0.0 k -0.5 -1.0 -1.5 0 2 4 wk0/p 2 1 0 10-4 wk/p 10-3 b 10-2 10-1 6 8 10 FIG. 3: Dispersion relation Λk = λk/a vs scaled k (solid line), for β = 5 × 10−4. The dotted line corresponds to the term − sin(wk)/(wk) and the dashed line to the zero line, drawn In the inset, we show the position of the for comparison. first maximum k0 as a function of β ≡ νDuν−1 /(aw2). The vertical dotted line indicates the instability threshold. 0 be rewritten in a single parameter form as Λk = −β(wk)2 − sin(wk) wk , with β ≡ νD uν−1 0 a w2 . (6) Negative Λk means relaxation back to the uniform state. Figure 3 depicts the dispersion relation in a typical case where Λk can take positive values allowing instability growth. Even if the analysis at short times does not guaran- tee the later evolution towards a stationary state, the mode with largest growth rate k∗ (absolute maximum of the dispersion relation Λk vs k) could play a crucial role. This mode will excite other wavelengths though the coupling nonlinear term, however, if they are damped, k∗ will remain selected and its harmonics will shape the patterns. Substitution of the Fourier series expan- k=−∞ ck(t) exp(ikx) into Eq. (3), when ν = 1, leads to the following evolution equations for the Fourier coefficients ck [9] sion u(x, t) = P∞ dck dt = −Dk2ck + ack − bXm cm¯ck−m sin mw mw . (7) These equations are highly coupled through the last non- linear term. If ν 6= 1, there will be still an additional nonlinearity in the first term of the righthand side, any- way let us consider the case where the first term is very small, allowing the existence of unstable modes. The am- plitude of the mode corresponding to the uniform state, c0, grows with rate a until stabilization, as observed in numerical simulations, e.g., in the example of Fig. 2 the level u0 = a/b = 1 is attained at times of order 1/a. The mode with largest initial (positive) rate quickly de- velops and keeps dominating at intermediate timescales. Notice in Fig. 2 an almost perfect sinusoidal profile at time t ≃ 20. If a unique mode contributes to the sum in 3 Eq. (7), it grows with rate given by Eq. (5) until stabiliza- tion while the remaining modes will be dumped. Actu- ally a set of undamped harmonics, characteristic of each value of ν, also persists to shape the profiles. Typical Fourier spectra for the long-term patterns are shown in Fig. 4. Although we do not have a rigorous mathematical proof, we will see that numerical results indicate that the dominant persistent mode, defining pattern wavelength, is the fastest growing one at short times. k0 2.5 2.0 1.5 1.0 0.5 ck n = 0.5 n = 1.0 n = 4.0 0 1 2 k 3 4 FIG. 4: (Color online) Fourier spectra for the longtime pat- terns shown in Fig. 1. The vertical dotted line indicates the position of the dominant mode k0. Λk possesses infinite local maxima located at kn, n = 0, 1, . . .. Since the absolute maximum is the first one, then k∗ = k0 (see Fig. 3). In the inset of Fig. 3, k0 (numerically obtained) is plotted as a function of β. For sufficiently small β, Λk is dominated by the last term in Eq. (6), yielding k0 = θ0/w with θ0 ≃ 1.43 π. Fig. 4 that the dominant mode of long time patterns is in good accord with k0. Perturbations to the homogeneous solution vanish if Λk < 0. Since sin(wk) is bounded, then the instability condition Λk > 0 implies β < (wk)−3 . (8) For the mode with largest growth, considering the ap- proximation k0 = θ0/w ≃ 1.43π/w, one has the instabil- ity condition β ≡ νD uν−1 a w2 < θ−3 0 0 ≃ (1.43π)−3 . (9) This is equivalent to requiring the positivity of the first maximum. Notice in the inset of Fig. 3 that k0 ≃ 1.43π/w remains a good approximation in the whole in- stability region, below the threshold (vertical dotted line in the inset). Beyond this point the maximum becomes negative, hence the homogeneous solution recovers its stability for any wavelength. Equation (9) defines the critical value βc ≡ θ−3 for the onset of patterns. 0 As intuitively expected, on the basis of the homoge- nizing role of diffusion, the inequality in Eq. (9) indi- cates that the diffusion constant can not exceed a limiting L value for the perturbation to depart from the homoge- neous state. In accord with Eq. (9), in the limit D → 0, patterns are also observed. Then, diffusion is not a nec- essary ingredient for the onset of patterns but has a role in pattern shaping. Figure 5(a) shows the density pro- files that emerge for different values of D in the normal case ν = 1. For D = 0 patterns are noisy due to the lack of the smoothing effect of diffusion and the ampli- tude is less uniform but the wavelength ℓ is well defined. Moreover, between bumps, the density tends to zero as in the subdiffusive case of Fig. 2. The width of the bump at zero height, 2x0, is also well defined. Beyond fluctua- tions, the results are robust, at least under the types of initial conditions analyzed. In the absence of diffusion, the steady state must verify u(x)[a − bJ(x)] = 0, then either u(x) vanishes or its integral within the interval (x − w, x + w) must adopt the constant value a/b. The former case requires that the null solution becomes sta- ble in some regions. The latter requires that each cluster does not see the neighbors and any point in the cluster must be influenced by the full cluster. This means that 2x0 ≤ w ≤ ℓ − 2x0, which is verified in numerical exper- iments. For D = 10−5, patterns are still noisy at t = 100, but are expected to smooth out at much longer times. In this case also the density between crests goes to zero exponentially fast with time, as can be seen in Fig. 5(b). In the case of the figure, the clusterization occurs for D . 10−3 while for larger values of D not only the crests 0 20 40 60 80 100 (a) D = 10-1 D = 10-5 D = 0.0 4 but also the valleys stabilize in a finite value as time goes by. Then clusterization occurs for D below a threshold value only. It is noteworthy the resemblance of the noisy profiles with those observed in experiments with bacteria [4]. But here clusters arise naturally, without imposing absorbing nor zero flux boundaries. If D 6= 0, Eq. (9) predicts the existence of a minimal value of the interaction range w required for pattern for- mation, with all other parameters kept fixed. This crit- ical value does depend on the kind of diffusion through the factors ν and uν−1 . Notice also that for ν 6= 1 there is influence of u0 too, which is absent in the normal case (ν = 1). 0 According to the hypothesis that k0 is the characteris- tic wavenumber of the stationary pattern, and taking into account that boundary conditions are periodic (i.e., an integer number of wavelengths must accommodate into the size of the system L), the number of maxima m is given (in average) by m = k0L 2π = θ0 2π L w ≃ 0.715 L w . (10) Even in the cases when Eq. (10) gives an integer value, it is expected to furnish the number of peaks observed in the average. In practice, depending on the initial condi- tions, the crests come out and grow accommodating its number approximately to the rounded value of m. Fluc- tuations in the effective m are larger the larger m or the further is m from an integer value. For instance in the example of Fig. 1, instead of seven, eight crests are ob- served in some realizations, being m ≃ 7.15. These observations can be verified through numerical integration of the evolution equation. In Fig. 6, we show the number of maxima m of the patterns as a function of w together with the theoretical prediction given by Eq. (10). An excellent agreement between theoretical and ) 0 0 1 = t , x ( u 14 12 10 8 6 4 2 0 20 15 10 5 0 100 10-10 10-20 ) t , x ( u 10-30 20 0-20 30 40 50 60 100 m 10 0.5 1.0 1.5 2.0 100 (b) D = 10-5 35 40 25 30 x 1 0.1 1 10 100 w FIG. 5: (Color online) (a) Longtime patterns obtained from numerical integration of Eq. (3) up to time t = 100, with a = b = 1, ν = 1, w = 5, L = 100 and different values of D indicated on the figure. In the inset, the profiles in the full grid are shown. In (b) the time evolution for D = 10−5 is shown in logarithmic scale. Times are indicated on the figure. FIG. 6: Number of maxima m as a function of the nonlo- cality width w, for a = b = 1, L = 100 and D = 0.01, and different values of ν indicated on the figure. The solid line corresponds to the approximate theoretical relation Eq. (10) and the symbols to the outcomes of numerical simulations. 5 numerical outcomes can be observed. Then, in good ap- proximation, the dominant wavelength only depends on the relation L/w independently of the remaining parame- ters. However, these parameters participate in condition- ing pattern upraise through the critical value of β, given by Eq. (9), and may also influence pattern shape. In par- ticular, in the average, m does not depend on ν, as can be observed in the example of Fig. 1. The wavenumber is preserved in the limit D → 0, even if in approaching this limit more and more modes become unstable (i.e., more local maxima of Λk become positive), but the global max- imum remains approximately the same. Then, diffusion is not necessary for pattern formation, nor influences the characteristic wavelength. 2.0 1.5 w 1.0 0.5 0.0 0.0 u0 = 0.1 u0 = 0.5 u0 = 1.0 u0 = 2.0 PATTERNS NO PATTERNS 0.5 1.0 n 1.5 2.0 Notice also that condition (9) indicates that, although approximately the same number of maxima is expected independently of ν, there are critical thresholds νc be- yond which no patterns occur. This is illustrated in Fig. 7 for the case a = b = 1, for which νc = w2/(Dθ3 0). FIG. 8: Phase diagram of pattern formation in the ν w plane , for u0 = 1, following Eq. (9). Shadowed is the region where no patterns arise. The lines show the frontier of the phase diagram for other values of u0 = a/b. Patterns emerge for w > wc (above the critical line). Parameter values are L = 100, D = 0.01 and a = 1. 100 80 60 40 20 m w = 1 w = 2 0 0 1 2 3 4 5 n FIG. 7: Number of maxima m as a function of ν for two different values of w indicated on the figure, with a = b = 1, L = 100 and D = 0.01. Dashed lines correspond to the theoretical prediction given by Eq. (10), with the additional condition (9), defining a critical value νc beyond which no patterns occur (we set m = 0 in such case). Differently from normal diffusion, u0 = a/b is also de- terminant of pattern formation. How the phase diagram is altered by changes in u0 is illustrated in Fig. 8. The shadowed area represents the region where no patters emerge when u0 = 1. For other values of u0, only the frontier is shown. For u0 ≥ 1, the critical curve increases monotonically with ν, such that only small ν (superdif- fusion) allows pattern formation for a given interaction range w and remaining parameters fixed. However, the monotonic behavior of the critical curve is broken when u0 < 1. Thus, for low values of w there is also an upper critical value of ν for the onset of patterns, but for large w, patterns occur for any ν. B. Patterns shape Although the characteristic mode does not depend on ν, its amplitude does. This is shown in Fig. 9 where the amplitude, ∆u = umax − umin obtained from numerical simulations is represented as a function of w for different values of ν. For vanishing w we recover the local case f (x, x′) = δ(x − x′) in which no patterns emerge, as supported by numerical simulations. In agreement with Eq. (9), there is a critical value wc, at which the amplitude vanishes. Notice the abrupt decay of the amplitude at the criti- cal value. This threshold was not detected in previous works dealing with normal diffusion possibly because of n = 0.5 n = 1.0 n = 2.0 n = 4.0 D u 8 6 4 2 0 0.1 1 10 100 w FIG. 9: Pattern amplitude ∆u ≡ umax − umin as a function of the interaction width w, for a = b = 1, L = 100, D = 0.01, and different values of ν indicated on the figure. The dotted lines are a guide to the eyes, the vertical ones indicate the critical value predicted by Eq. (9). the range of parameters used. For instance in Ref. [8], wc/L would be of the order of 10−3. The critical value wc decreases with ν, indicating that a shorter influence range w is required when the dispersion passes from sub- diffusive to superdiffusive. Then superdiffusion favors pattern formation and the amplitude of the patterns is larger. For w = L/2 (or its multiples), the nonlocal term becomes J[u(x, t)] = J(t), that follows the equa- tion dJ/dt = (a − bJ)J. Then, in the long time limit J → u0, implying that the homogeneous state should also be attained in this extreme case. Furthermore, ν can have strong effects on patterns shape. As depicted in Fig. 1, subdiffusion (ν > 1) in- duces fragmentation (clusterization). Solutions that van- ish outside a finite support are typical of nonlinear subd- iffusion [12]. In the opposite case ν < 1 (super-diffusion), the effects are not so striking concerning pattern shape, for moderate values of the diffusion coefficient. The re- gion between crests assumes larger values the smaller ν. Fragmentation also emerges for any kind of diffusion when D is small enough (as discussed in connection to Fig. 5) or also if w becomes large enough (not shown). The shape of the clusters depends on ν. Their amplitude decays and their width increases as ν increases. It is noteworthy that the distance between crests (wavelength), ℓ = L/m ≃ 1.4w, is larger than the interac- tion range w, however if the cluster size 2x0(w) is large enough, there can be influence of one cluster over the two neighboring ones. When clusters are disconnected, 2x0 . ℓ − w = 0.4w means that one cluster does not influence the neighbors. Otherwise they do, even if dis- connected. 6 motivating the introduction of a state dependent diffu- sion coefficient, as in Eq. (3). We have shown how pattern formation is altered in the presence of anomalous diffu- sion. Moreover, in all cases, the initially fastest growing mode remains selected at longer times. This observation allows to obtain theoretical predictions that we verified through numerical integration of the evolution equation. Then, it is clear that diffusion is not a necessary in- gredient for the onset of patterns, nor has impact on the characteristic wavelength, that depends only on the in- teraction range w. Furthermore, diffusion imposes a crit- ical threshold of the model parameters for pattern forma- tion. The type of diffusion regime has impact on patterns shape, even if the characteristic mode is kept unchanged. An important qualitative change in the shape of patterns occurs mainly for ν > 1, in which case fragmentation of the population is induced. This effect is also observed for very small diffusion constant D and/or large interaction width w. The occurrence of fragmentation may have im- portant consequences in diseases dissemination and other spreading processes triggered by contacts between indi- viduals. Superdiffusion (ν < 1) facilitates pattern forma- tion, which can occur even for shorter interaction width w and manifesting larger amplitudes than in normal dif- fusion. Beyond the initial motivation of introducing nonlinear diffusion to the nonlocal FE, we uncovered aspects that apply also to the normal diffusion case previously studied. The identification of the main mode selected at long times allows to perform analytical predictions, which may be extended to tackle other variants of the model. III. FINAL REMARKS Acknowledgments Nonlinear diffusion is expected in the spreading of bi- ological populations rather than normal diffusion, hence We are grateful to Welles A.M. Morgado for useful dis- cussions. We acknowledge partial financial support from CNPq and Capes (Brazilian Government Agencies). [1] T.E. Woolley, R.E. Baker, E.A. Gaffney, P.K. Maini, [11] E. Brigatti, V. Schwammle, M.A. Neto, Phys. Rev. E 77, Phys. Rev. E 84, 046216 (2011). 021914 (2008). [2] J.A.R. da Cunha, A.L.A. Penna, F.A. Oliveira, Phys. [12] J.D. Murray, Mathematical Biology: An introduction, In- Rev. E 83, 015201 (2011). terdisciplinary Applied Mathematics (Springer, 2003). [3] R.F. Costantino, R.A. Desharnais, J.M. Cushing, B. Den- [13] M.E. Gurtin, R.C. MacCamy, Math. Biosciences 33, 35 nis, Science 275, 389 (1997). (1977). [4] N. Perry, J. Royal Soc. Inteface 4, 379 (2005). [5] L. Giuggioli and V.M. Kenkre, Physica D 183, 245 (2003). [6] R.A. Fisher, Ann. Eugen. 7, (1937). [7] V.M. Kenkre, M.N. Kuperman, Phys. Rev. E 67, 051921(2003). [8] M.A. Fuentes, M.N. Kuperman, V.M. Kenkre, Phys. Rev. Lett. 91, 158104 (2003). [9] M.A. Fuentes, M.N. Kuperman, V.M. Kenkre, J. Phys. Chem. B 108, 10505 (2004). [10] E. Brigatti, M. N´unez-L´opez, M. Oliva, Eur. Phys. J. B 81, 321-326 (2011). [14] S.E. Mangioni, Physica A 391, 113 (2012). [15] M. Muskat, The Flow of Homogeneous Fluids through (McGraw-Hill, New York, 1937); L.A. Porous Media in: H. Amann, N. Bazley, K. Kirchgassner Peletier, (Eds.), Applications of Nonlinear Analysis in the Physi- cal Sciences, (Pitman, London, 1981). [16] C. Tsallis, D.J. Bukman, Phys. Rev. E 54, R2197 (1996). [17] J. Buckmaster, J. Fluid Mech. 81, 735 (1977); E.W. Larsen and G.C. Pomraning, SIAM (Soc. Ind. Appl. Math.) J. Appl. Math. 39, 201 (1980); W.L. Kath, Phys- ica D 12 375 (1984); H. Spohn, J. Phys. I 3, 69 (1993). [18] P. Rosenau, Phys. Rev. Lett. 74, 1056 (1995); A. Compte, D. Jou, and Y. Katayama, J. Phys. A 30, 1023 (1997); L. Borland, Phys. Rev. Lett. 89, 098701 (2002). [19] E.K. Lenzi, C. Anteneodo, L. Borland, Phys. Rev. E 63 051109 (2001). [22] L.C. Malacarne, R.S. Mendes, I.T. Pedron, E.K. Lenzi, Phys. Rev. E 65 052101 (2002). [23] E.K. Lenzi, G.A. Mendes, R.S. Mendes, L.R. da Silva, and L. S. Lucena, Phys. Rev. E 67, 051109 (2003). [20] C. Anteneodo, Phys. Rev. E 76, 021102 (2007). [21] D.A. Stariolo, Phys. Rev. E 55, 4806 (1997); L. Borland, [24] E. Hern´andez-Garc´ıa, C. L´opez, S. Pigolotti, K.H. An- dersen, Phil. Trans. Royal Soc. A 367, 3183 (2009). Phys. Rev. E 57, 6634 (1998). 7
1107.0086
2
1107
2011-08-04T18:06:15
An Assessment of "What does photon energy tell us about cellphone safety" by Dr. William Bruno
[ "physics.bio-ph", "physics.med-ph" ]
Dr. William Bruno asserts the well-known fact that cell phones radiate microwaves in the classical regime. This, he says, means that the photon energy is not relevant to assessing safety. Citing optical tweezers as an example of biologically relevant non-thermal effects of electromagnetic radiation, Bruno concludes that all other reports of non-thermal effects from microwaves are likely valid. He seeks safety thresholds based upon requiring that cell phone energy density be less than kBT. This proposal and related ideas produce thresholds many orders of magnitude below present values. While Dr. Bruno is correct that cell phone microwave radiation is generally in the classical regime, he uses peculiar estimates (number of photons per cubic wavelength) that overstate the circumstance by more than 20 factors of ten. He misunderstands the operation of optical tweezers and ignores their significant thermal effects. He credulously accepts poorly supported claims of non-thermal effects. He mistakenly believes that kBT is the average thermal energy (per cubic wavelength or per cell) in materials. It is not. It is twice the average energy per molecule per degree of freedom in the material. The thermal energy density is (1/2)kBT X (average number of degrees of freedom of the molecules) X (Avogadro's number). (Avogadro's number is 6 X 1026 molecules/mole.) Thus, Bruno's proposed safety thresholds are more than 1025 too low. Using the correct value for the average thermal energy would place the thresholds close to today's standards. Throughout his analysis he neglects the index of refraction of living tissue, n \cong 9, or the absorption length, {\alpha} \cong 1 cm-1, in the microwave region.
physics.bio-ph
physics
An Assessment of “What does photon energy tell us about cellphone safety” by Dr. William Bruno By Bernard Leikind August 3, 2011 Published on EMF and Health: http://www.emfandhealth.com/Do%20Cell%20Phones%20Cause%20Cancer.html Also at: arXiv:1107.0086 [physics.bio-ph] Abstract: Dr. William Bruno asserts the well-known fact that cell phones radiate microwaves in the classical regime. This, he says, means that the photon energy is not relevant to assessing safety. Citing optical tweezers as an example of biologically relevant non-thermal effects of electromagnetic radiation, Bruno concludes that all other reports of non-thermal effects from microwaves are likely valid. He seeks safety thresholds based upon requiring that cell phone energy density be less than kBT. This proposal and related ideas produce thresholds many orders of magnitude below present values. While Dr. Bruno is correct that cell phone microwave radiation is generally in the classical regime, he uses peculiar estimates (number of photons per cubic wavelength) that overstate the circumstance by more than 20 factors of ten. He misunderstands the operation of optical tweezers and ignores their significant thermal effects. He credulously accepts poorly supported claims of non-thermal effects. He mistakenly believes that kBT is the average thermal energy (per cubic wavelength or per cell) in materials. It is not. It is twice the average energy per molecule per degree of freedom in the material. The thermal energy density is (1/2)kBT X (average number of degrees of freedom of the molecules) X (Avogadro’s number). (Avogadro’s number is 6 X 1026 molecules/mole.) Thus, Bruno’s proposed safety thresholds are more than 1025 too low. Using the correct value for the average thermal energy would place the thresholds close to today’s standards. Throughout his analysis he neglects the index of refraction of living tissue, n ≅ 9, or the absorption length, α ≅ 1 cm-1, in the microwave region. Dr. William Bruno asks “What does photon energy tell us about cellphone safety?” in a pre -print submitted to the physics arXiv pre-print database on April 24, 2001.1 Some in the scientific press noticed the paper and drew attention to it. Here is an example of friendly coverage from MIT’s Technology Review, “Cell Phones, Microwaves and The Human Health Threat”, on April 28, 2011.2 A pre-print is a research paper that scientists send to their friends and colleagues while simultaneously submitting the manuscript to a journal for publication. As the formal publication may take months, the pre-print is a way for a researcher to draw attention to his or her work rapidly. Unlike articles in the best scientific journals, professionals have not reviewed a pre-print for quality or importance. Therefore, it is rare for one to produce wide news coverage. Coverage of a pre-print may be the result of the author’s reputation as an important and distinguished scientist or may result from the importance of the results. Dr. Bruno describes himself as a physicist or as a theoretical biologist. He appears, however, to be nutty about the subject of electromagnetic fields and health effects. Santa Fe science writer George Johnson describes some of Bruno’s activities, including a literal metallic hat for protection against the ever present electromagnetic fields of modern civilization, in this witty account from Slate, “On Top of Microwave Mountain: I tried to sauté my brain at the base of a cell phone tower. It didn’t work.”3 The coverage is not the result of Bruno’s scientific reputation. The media interest arises from the significance of Bruno’s claims: that there are important, so-called non-thermal effects that play a role when organisms absorb or interact with microwave radiation, and that scientists and regulators do not recognize these effects when they establish safety limits for cell phones, cell phone towers, or WiFi and other equipment. Hence, Bruno argues safety limits are much too high. Physicists and other scientists have studied the interactions between electromagnetic fields and matter in all its forms from atoms and molecules to big chunks of stuff, pure and mixed up, for more than a hundred years. A major sub-field of this research deals with organisms and health effects. With the rise of cell phones, many people wonder if the phones’ microwave radiation might have bad effects on our brains. People are particularly concerned about cancers. After all, the microwaves definitely penetrate the brain and, generally, people haven’t been exposed to such radiation in their brains at any time in human existence. A vast industry has arisen in response to these concerns involving epidemiological research, bio-chemical and biological research, medical and clinical research, studies of rats and mice, studies of cells living in Petri dishes, and commercial enterprises to measure and protect us from the supposed dangers. What do physicists know about this? Some forms of electromagnetic radiation are harmful and cause cancer. These are ultraviolet radiation, X-rays, and gamma rays. Ultraviolet radiation causes skin cancers. No other form of electromagnetic radiation causes any cancer. These other forms include visible light, infrared radiation (heat radiation, as some would describe it), microwaves, radio waves, and on to lower frequencies and longer wavelengths. While the biological scientists don’t know everything about carcinogenesis, physicists know everything about how organisms absorb electromagnetic radiation and specifically microwaves. Physicists point out that the forms of electromagnetic energy that cause cancers all have sufficient energy in their photons (like particles of light) that they can break chemical bonds and ionize atoms and molecules. No other forms of the radiation can do this. At frequencies below and wavelengths longer than the visible spectrum, physicists know that when any material (in the conditions of living organisms) absorbs electromagnetic radiation, the energy goes directly into the incessant and random jostling, vibrating, and twisting of the molecules. This is heating, or a thermal effect. There is no missing energy. Since everyone knows that wearing a ski cap, swallowing a hot coffee, or jogging do not cause cancer, clearly cell phones, which produce less, sometimes much less, heating than these activities, cannot cause cancer. Some biological and epidemiological researchers, however, believe that they find deleterious effects of cell phone radiation. Therefore, they conclude that they must be seeing non- thermal effects.4 Sometimes these researchers are making an elementary blunder, assuming that when they do not detect a temperature increase in their experiments the effects must not be thermal in nature. It may be that they do not have sufficiently precise thermometers or do not measure the temperature in their experiments. Unfortunately, this error is common. Furthermore, when ice melts in a summer drink or water boils in a teakettle, the water’s temperature doesn’t change, yet these are both definitely thermal effects. A thermal effect may result in a detectable temperature increase, but it may not. Other researchers seek non-thermal effects in otherwise well-known phenomena associated with terms such as resonance, multi-photon effects, or nonlinear effects. Many researchers have proposed such phenomena or effects, and this brings us to Bruno’s paper. Bruno points to the statements of some physicists that only the photons of ionizing radiation have sufficient energy to break chemical bonds. He says this is not the correct way to think about microwave radiation. While its photons are individually very weak, they are present in cell phone radiation in very large numbers. Using a peculiar measure of the photon population density, photons per cubic wavelength, he shows that X-rays and ultraviolet radiation from common sources (medical X-ray machines and the sun’s radiation) have a relatively small number of photons per cubic wavelength, while microwave radiation from cell phones, cell phone towers have a relatively large number of them. This, he says, makes the individual photon energy less relevant than the tremendous numbers of photons flowing through tissues. He chooses as an example of a non-thermal biological effect based upon the flow of large numbers of non-ionizing photons, the remarkable devices known as optical tweezers. These use lasers and backward microscopes to focus visible or infrared radiation to tiny spots. With careful optical design, the experimenters can use variations in the intensity of the radiation to move, pull, or twist tiny objects, even individual molecules such as proteins and DNA. The United States Energy Secretary Steven Chu won his Nobel laurels for his work with these devices. Intensity of radiation is another way of saying that there are large numbers of photons flowing. Having established to his satisfaction that he has identified a non-thermal mechanism, because microwaves from a cell phone have many photons and optical tweezers involve many photons, Bruno then asserts that the existence of this mechanism shows that all of the researchers who claim to have found a non-thermal effect may be on to something. He lists putative breaches in the blood-brain barrier, microwave clicks, and several other effects. Accepting these effects as real and as occurring well below the official microwave safety thresholds, Bruno says that safety limits based upon heating are inadequate. He seeks possible new principles on which to base the limits. He considers the natural microwave radiation from the sun, which is orders of magnitude below the radiation levels of cell phones. He considers what he says is the average thermal energy present in organisms, which he calculates is also orders of magnitude below cell phone thermal levels. He sees no hope for safety in these thoughts, suggests future use of optical signals, and recommends hands-free use of cell phones as an otherwise inadequate move in the right direction. Bruno begins his attempt to show that the photon energy is not, in itself, sufficient to judge the effects or safety of cell phones by estimating the significance of the quantum, particulate nature of the radiation as compared to certain other sources. The information in his Table 1 is here in my Table 1, which contains some clarifications and additions. His data is in italics. The symbol ~ means that the value is representative of a range. Bruno demonstrates that there are many more photons per cubic wavelength in microwave radiation than in solar ultraviolet radiation at the earth’s surface or in a typical X ray. He concludes that cell phone microwaves are in the classical limit while medical X rays and solar UV radiation are in the quantum limit. His second column shows that there is a difference of 44 orders of magnitude between the X rays and the microwaves! In this way of thinking, the chunkiness of the X rays, the quantum nature of electromagnetic radiation, is important for them, but the waviness of the microwaves, the classical nature of the electromagnetic radiation, is important for them. No physicist would disagree that in certain contexts X-rays and UV tend to exhibit quantum effects, chunkiness, or that the microwaves tend to exhibit classical effects, waviness. But Bruno mistakes these tendencies with invariable behavior. Scientists use X ray diffraction, which is a classical wave effect as an important probe of crystal structure. Microwave engineers don’t worry about microwave photons for their purposes, but the photons are there nonetheless, and physicists studying the energy states of rotating molecules must consider microwave photons. In his desire to illustrate that microwaves are far from the quantum realm of X -rays and ultraviolet, Bruno uses a peculiar measure of the photon density, photons per cubic wavelength. (See the note at the end to convert between Bruno’s units and ones that are more usual.) Photons are an odd concept in quantum physics. It is sometimes allowable to imagine them as tiny little balls or perhaps even points. The famous Heisenberg Uncertainty Principle, the enforcer of quantum fuzziness, however, makes it hard to say just where a photon is in volumes smaller than a cubic wavelength. There is, however, no problem with estimating how many of a particular type of photon there might be in larger volumes. Bruno wishes to show that there are many, many more microwave photons around than X-ray photons. He counts how many microwave photons there are in a cubic microwave wavelength, which is, for his estimate, 0.3 meters, and cubed is 0.027 m3 = 2.7 X 10-2 m3. He also counts how many X-ray photons there are in a cubic X ray wavelength, which is, for his estimate, 10 -10 m, and cubed is 10-30 m3. He is mistaking the size of a photon for how accurately you can say where it is, and by this mistake, he gets a factor of 1028 increase in the number of microwave photons relative to the number of X ray photons. In Table 1 I have shown the Wavelength Cubed in parentheses in the second column so that you can see this factor. Any physicist seeking clarity in comparing the photon density or number of photons would have chosen the same volume for each case. I have added column 3 to show this quantity, the photon density in photons/m3. The physicist would also choose the same power or energy . I have added column 4, Photons/sec in 1 Watt = 1 Joule/sec. I can’t tell what X ray power level Bruno had in mind, but I believe that 1 Watt of X-ray power is in the ballpark for a typical medical exposure. Bruno says that solar ultraviolet radiation is about 10 W/m2, a flux or flow. (See the note at the end to convert between Bruno’s power flow, Watts/m2, and energy density, Joules/m3.) This is also in the ballpark for a 1 Watt nearby source. Cell towers broadcast more than 1 Watt, but not much more, and all exposures are from many meters away. Bruno takes 10 meters, as if a person were standing at the base of the tower. Cell phones broadcast about one Watt when they are transmitting. The fourth column shows the number of photons for each of the exemplary sources for the same emitted power. It is the case that microwaves have about ten orders of magnitude more photons than the same X ray power, but ten orders of magnitude are 34 orders of magnitude smaller than the exaggerated 44 orders Bruno gets from his volume factor. I have also added a row that describes optical tweezers, since Bruno chose them as an example of a non-thermal source of electromagnetic radiation. While objecting to Bruno’s calculation, accept, for the moment, his point that microwaves are usually in the classical, wave-like limit. He says that optical tweezers are an example of a optical system that operates in the classical limit and that is capable of exerting forces on molecules that might well damage them without heating; a non-thermal effect. Table 1* Photon Density Photons/m3 Photons/second in 1 Watt Notes 106 1.5 X 1015 X-ray: 30 cm from 1mA source, 1% efficient Wavelength = 1 nm = 10-10 m 4 X 1014 1.5 X 1017 ~10W/m2 Photons per cubic wavelength (Wavelength Cubed m3) ~ 1 X 10-24 (10-30 ) ~ 1 X 10-7 (2.7 X 10-24 ) ~ 1 X 10-1 (1 X 10-18 ) Wavelength = 300nm = 3 X 10-7 m 1017 1.5 X 1018 1 W visible or infrared laser beam focused to a 3 X 10-6 m radius beam. Wavelength = 3000 nm = 3 X 10-6 m Source Medical X- ray Sunlight UV Optical Tweezers (Visible or Infrared) Cell tower ~ 1 X 10+15 4 X 1016 1.5 X 1024 Frequency = 1 GHz (2.7 X 10-2) Wavelength = 3 X 10-1 m E = ~1V/m 10 meters from the antenna or about 33 feet Cell phone ~ 1 X 10+20 4 X 1021 1.5 X 1024 Frequency = 1 GHz (2.7 X 10-2) Wavelength = 3 X 10-1 m *Items, data and notes from Bruno’s Table 1 are in italics. I added other items, data and notes. E = ~300V/m (10 m / 300) = 0.03 m or about 1 inch from the antenna Figure 1 a) shows a simplified schematic of an optical tweezers system, a laser with its beam passing backwards through a microscope to focus to a tiny spot. Figure 1 b) shows the details of the radiation Figure 1. Optical Tweezers principles. How does it work? The most basic form o f an optical trap is diagramed in Fig 1a. A laser beam is focused by a high-quality microscope object ive to a spot in the specimen plane. This spot creates an "optical trap" which is able to hold a small part icle at its center. The forces felt by this particle consist of the light scattering and gradient forces due to the interaction of the particle with the light (Fig 1b, see Details). Most frequently, optical tweezers are built by modifying a standard optical microscope. These instruments have evo lved from simple tools to manipulate micron-sized objects to sophist icated devices under computer-control that can measure displacements and forces with high precision and accuracy. Figure 1. An illustration of the operation of an optical tweezers system and some explanatory text.5 beam in the spot or focused neck and a spherical object on which the beam exerts a force drawing the object to the center of the beam and to the narrowest part of the beam. Further details and a clear explanation of the operation of optical tweezers and of some of their applications are at the Stanford University site5. Considering round numbers, a typical optical tweezer system will use a 1 watt laser radiating in the visible or infrared region. It will pass the laser’s beam backwards through a microscope to shrink it to a small region, a few microns, millionths of a meter, in diameter. Let’s consider 6 microns in diameter, divide by two to get the radius, square it, and multiply by pi, to get the cross section of the beam. It is about 30 X 10-12 m2. Now compute the power flow (1 watt)/( 30 X 10-12 m2 ) = 3 X 1010 W/m2. This is 30 billion watts per square meter, a very intense beam. Bruno says that a medical X -ray has about 10 watts per square meter. The total power radiated from a cell phone is about one watt. Consider, just to estimate, that the cell phone’s antenna is a small source, and imagine a sphere 1 cm = 0.01 m in radius. The surface area of this sphere is (4/3) pi r2 = 4 X 10-4 m2. Now compute the power flow for this case (1 watt)/( 4 X 10-4 m2 ) = 0.25 X 104 w/m2 or about 250 W/m2. Two hundred and fifty is smaller than 30 billion by about a hundred million. The flow of photons per square meter radiated from a cell phone is smaller than the flow of photons per square meter in the tweezers’ focal spot by the same ratio. The effects of the optical tweezers, the ability of this intense beam of visible or infrared light, to exert forces on tiny objects depend upon spatial variations in the intensity of the tiny beam and on variations in the optical properties of the illuminated material. The intensity is highest along the axis of the optical system. In Figure 1 b) you can see the variation in the beam’s intensity as the curved mostly vertical line to the right of the big red arrow. Imagine the red lines rotated about the dotted line to picture the generally cylindrical form of the optical beam. The relevant optical properties of a material have two parts, one called the index of refraction, which is the part that matters for the laser to exert its force. This force would be the Bruno’s non-thermal effect if it were the only relevant effect. The other part of the transparency is the absorption coefficient. This coefficient measures how far a beam of radiation travels before about 2/3rds of it has been absorbed. Each time the beam travels a distance equal to the absorption coefficient, its intensity falls another 2/3rds. The energy that disappears from the beam appears in the material as thermal energy. The optical tweezers system depends upon the difference in the index of refraction in the object (the blue sphere shown as a circle in the figure) and in the background material. You can see that the variation in the intensity of the laser beam and the variation in the index or refraction , the size of the object, are roughly, the same scale. In the visible and infrared range chosen for optical tweezers, the transparent materials do not absorb much of the beam, but they absorb some. A beam might travel many meters before a noticeable amount is lost. The radiation flow is so intense, however, that researchers risk damaging the materials, molecules or organisms they study because of the heating. That is, Bruno’s example of a non-thermal effect often comes with important thermal side effects. The situation is different in the case of microwaves and the tissues of living organisms. Water molecules love microwaves, as do many other biologically important molecules. The absorption coefficient is much higher, a few hundred times higher, than it is for visible or infrared light. Therefore, the distance the radiation travels before the materials absorb it is much shorter. Nearly all of the radiation from a cell phone that enters an organism disappears in a few centimeters. All of this is a thermal effect. There is no non-thermal effect. All of the radiation moves from the microwave beam into the thermal motions of all of the molecules of the tissues. Having incorrectly convinced himself that optical tweezers provide an example of a relevant non- thermal effect, Bruno accepts , therefore, all other reports of non-thermal effects and concludes that experts setting radiation safety thresholds must consider these reported effects. He turns to investigating possible new safety thresholds, well below those set with regard to thermal effects. In this arena, he also chooses odd contexts and in one case makes an outright and major blunder. He considers natural microwave fluxes, such as those present from the sun, he says. But he notes that these levels would be even lower at night. He doesn’t think it would be practical to limit cell phone radiation to these levels, but he gives no detail. He considers the “average thermal energy, kbT, per cubic wavelength” in the organism. As long as the average energy from the cell phone is less than this quantity, he says, the organism should be safe. In this quantity kb is the Boltzmann constant, about 1.4×10−23 Joules/K, and T is the temperature measured in Kelvins. Body temperature is about 310 K. This is an amazing mistake. The average thermal energy density is not kbT. That value, kbT, is twice the energy per degree of freedom of a single molecule. Bruno should have multiplied this by the number of molecules in the volume he’s interested in and by the average number of degrees of freedom for the molecules in his volume. This latter is some small integer, let’s just say 5, which would be appropriate for a water molecule. Remember, too, that a cubic wavelength for Bruno’s cell phone (see Table 1) is about 27 liters (a cube 30 cm or a foot on a side). The immense error is neglecting the number of molecules in his volume. Just to get a value, let’s suppose a brain has the density of water, which would not be far off. A brain has a mass of about 1.2 kg and a volume of 1200 cm3. This is just about a cubic wavelength for a 12 cm 2.5 GHz microwave. Water has about 3 X 1022 molecules per gm or per cm3. In a cubic wavelength there would be about 4 X 1025 molecules. Bruno’s estimate for a threshold safe limit is too low by a factor about 40 million billion billion! (That’s US billions, 1,000,000,000.) That’s wrong by 40 septillions. Don’t forget that there is another factor of 10 or so for the degrees of freedom. If Bruno were to use the correct value here, he’d come up with something similar to modern day safety thresholds. Experts set those standards by requiring that the energy or temperature changes in organisms caused by microwaves should be less than normal variations that result from ordinary activities. Bruno cites various studies that claim to have seen effects, some deleterious, even at his wildly mistaken values. He cites reports of vague symptoms, headaches, depression, and sleeplessness, from people within a few miles of cell towers. These people, he says, experience microwave levels similar to his mistaken threshold. This tells us that those studies are foolishly in error, and the reports of symptoms are mistakenly attributing the symptoms to the cell tower radiation . We know that these studies and reports are foolish and mistaken because the reported effects are the result of radiation energy density many orders of magnitude less than what is naturally present within living organisms. In his entire consideration of microwaves propagating in living tissue, Bruno has neglected a major effect that I have also neglected so far because it would be too much to correct every error at once. I have discussed the value of the absorption coefficient but I have not mentioned the value of the index of refraction for microwaves in living tissue. Of course, in this case, we are considering a complex mate rial that includes skin, bone, brain, and so on, and the details would be complicated. The matter is further complicated because the absorbing material is within a few wavelengths of the antenna. But the value of the index of refraction that Bruno should be using in his analysis is about nine.6 This means that in the tissue, the wavelength will be, roughly speaking, about nine times shorter than it is in air. He has been over estimating the volume of a cubic wavelength, in an organism, by about 9 X 9 X 9 or about 700. Astonishingly, this error is so many orders of magnitude less than his other errors and miscues that neglecting it hasn’t changed my conclusions. Finally, Bruno points to the possibility of damaging effects arising from the modulation of the ce ll phone microwaves. This suggests that he, and the researchers he cites, have no idea about the modulation of cell phone signals. Modern digital cell phone signals use complex modulation schemes, but they are not turned on for a one and off for a zero. They shift between two frequencies that barely differ. The frequency might be a bit high for a one and a bit low for a zero. The main carrier frequency is on all the time that a phone call is in progress. In actual use, there are various nearby, in frequency , channels. A call may switch among channels during the call, and so on. But there is always a high frequency carrier signal. The modulation certainly makes no difference whatsoever in the absorption of a cell phone signal. Note about units and quantities To convert from Bruno’s photon number density, in photons per cubic wavelength, to a more usual photon number density, n = photons per cubic meter, multiply his quantity by the cube of the number of wavelengths in one meter. The number of wavelengths in one meter is one divided by the wavelength. Thus Photon density = # / wavelengths3 [photons / cubic wavelength] X (1/λ3) [wavelengths / meter]3 = n [photons/ meter3], where # represents the number of photons in the volume, λ is the wavelength, n is the photon density in #/m3, and I have put the units in brackets [] and multiplied out the units as if they were algebraic quantities. To convert from a density, n, to a flow, j, multiply the density by the speed of the flow, nv = j. In the case of electromagnetic radiation, the speed is c, the speed of light, 3 X 108 m/s. For example, to convert an energy density, in joules / m3, to an energy flow or power flow, in Joules/s/m2 or Watts/m2, where a Watt is a Joule/s, multiply n [Joules/m3] X 3X 108 [m/s] = (3 X 108) (n) [Joules/s/m2 = Watts/m2]. In this case, I’ve shown the conversion for an energy density to an energy (or power) flow. The conversion will work also for a photon density to a photon flow or flux, which would convert [photons/m 3] to [photons/s/m2]. References 2. 1. Bruno, William. “What does photon energy tell us about cellphone safety?”, April 24, 2001. arXiv:1104.5008v1 [q-bio.OT]. KFC. “Cell Phones, Microwaves And The Human Health Threat”, Physics arXiv blog at MIT’s Technology Review, April 28, 2011. http://www.technologyreview.biz/blog/arxiv/26708/ accessed June 30, 2011. 3. Johnson, George. “On Top of Microwave Mountain: I tried to sauté my brain at the base of a cell phone tower. It didn’t work.”3 Slate, April 21, 2010, http://www.slate.com/id/2251432/ accessed June 30, 2001. 4. Leikind, Bernard. “Do Cell Phones Cause Cancer?”, Skeptic Vol. 15, No. 4, 2010, p.30. Leikind, Bernard. “Do Cell Phones Cause Cancer?”, eSkeptic, June 9, 2010, arXiv:1007:4192 . http://www.skeptic.com/eskeptic/10-06-09/ , Shermer, Michael and Bernard Leikind, “Cell Phones and Cancer”, eSkeptic http://www.skeptic.com/eskeptic/10-12-08/ , both accessed June 30, 2011. 5. From http://www.stanford.edu/group/blocklab/Optical%20Tweezers%20Introduction.htm , accessed June27, 2011. Used by permission. 6. See Jackson, J. D. Classical Electrodynamics, 2nd ed., John Wiley & Sons, Inc., New York, 1975. Pages 290-2, which discuss the index of refraction and absorption length for water, including sea water, which will serve as a useful approximation to the value for living tissue.
1308.6830
1
1308
2013-08-30T19:45:13
The effects of psammophilous plants on sand dune dynamics
[ "physics.bio-ph", "nlin.CD", "q-bio.PE" ]
Psammophilous plants are special plants that flourish in sand moving environments. There are two main mechanisms by which the wind affects these plants: (i) sand drift exposes roots and covers branches--the exposed roots turn into new plants and the covered branches turn into new roots; both mechanisms result in an enhanced growth rate of the psammophilous plant cover of the dunes; (ii) strong winds, often associated with sand movement, tear branches and seed them in nearby locations, resulting in new plants and an enhanced growth rate of the psammophilous plant cover of the dunes. Despite their important role in dune dynamics, to our knowledge, psammophilous plants have never been incorporated into mathematical models of sand dunes. Here, we attempt to model the effects of these plants on sand dune dynamics. We construct a set of three ordinary differential equations for the fractions of surface cover of regular vegetation, biogenic soil crust and psammophilous plants. The latter reach their optimal growth under (i) specific sand drift or (ii) specific wind power. We show that psammophilous plants enrich the sand dune dynamics. Depending on the climatological conditions, it is possible to obtain one, two, or three steady dune states. The activity of the dunes can be associated with the surface cover--bare dunes are active, and dunes with significant cover of vegetation, biogenic soil crust, or psammophilous plants are fixed. Our model shows that under suitable precipitation rates and wind power, the dynamics of the different cover types is in accordance with the common view that dunes are initially stabilized by psammophilous plants that reduce sand activity, thus enhancing the growth of regular vegetation that eventually dominates the cover of the dunes and determines their activity.
physics.bio-ph
physics
The effects of psammophilous plants on sand dune dynamics Golan Bel∗ and Yosef Ashkenazy† Department of Environmental Physics, Blaustein Institutes for Desert Research, Ben-Gurion University of the Negev, Sede Boqer Campus 84990, Israel (Dated: July 30, 2018) Abstract Psammophilous plants are special plants that flourish in sand moving environments. There are two main mechanisms by which the wind affects these plants: (i) sand drift exposes roots and covers branches -- the exposed roots turn into new plants and the covered branches turn into new roots; both mechanisms result in an enhanced growth rate of the psammophilous plant cover of the dunes; (ii) strong winds, often associated with sand movement, tear branches and seed them in nearby locations, resulting in new plants and an enhanced growth rate of the psammophilous plant cover of the dunes. Despite their important role in dune dynamics, to our knowledge, psammophilous plants have never been incorporated into mathematical models of sand dunes. Here, we attempt to model the effects of these plants on sand dune dynamics. We construct a set of three ordinary differential equations for the fractions of surface cover of regular vegetation, biogenic soil crust and psammophilous plants. The latter reach their optimal growth under (i) specific sand drift or (ii) specific wind power. We show that psammophilous plants enrich the sand dune dynamics. Depending on the climatological conditions, it is possible to obtain one, two, or three steady dune states. The activity of the dunes can be associated with the surface cover -- bare dunes are active, and dunes with significant cover of vegetation, biogenic soil crust, or psammophilous plants are fixed. Our model shows that under suitable precipitation rates and wind power, the dynamics of the different cover types is in accordance with the common view that dunes are initially stabilized by psammophilous plants that reduce sand activity, thus enhancing the growth of regular vegetation that eventually dominates the cover of the dunes and determines 3 1 0 2 g u A 0 3 ] h p - o i b . s c i s y h p [ 1 v 0 3 8 6 . 8 0 3 1 : v i X r a their activity. PACS numbers: 87.23.Cc, 05.45.a, 45.70.n, 92.60.Gn ∗Electronic address: [email protected] †Electronic address: [email protected] 1 I. INTRODUCTION Sand dunes cover vast areas in arid and coastal regions [∼ 10%, 1 -- 3] and are considered to be an important component of geomorphological [4] and ecological [5, 6] systems. On one hand, active sand dunes are a threat to humans [7, 8], while, on the other hand, they are associated with unique ecosystems that increase biodiversity [9] and thus are important to humans. Human activities can affect sand dune ecosystems [5, 6]. Sand dunes may be sensitive to climate change [10, 11], and it has been claimed that they influence the climate system through changes in their albedo [12, 13]. The wind is the main driving force of sand dunes [14]. The migration rate of sand dunes is proportional to the wind power, which is a non-linear function of the wind speed [15]. Thus, dunes mainly migrate during a small number of extreme wind events. Dunes may be stabilized by vegetation and/or biogenic soil crust (BSC) [16]; since vegetation can only exist above a cer- tain precipitation threshold [typically ∼ 50mm/yr, 14], sand dune dynamics and activity in arid regions are strongly affected by the precipitation rate. Many experimental [2, 4, 15] and theoretical [4, 17 -- 22] works have been devoted to uncover- ing the mechanisms behind the geomorphology of sand dunes. Most of these models focused on the dune patterns and their corresponding scaling laws, on dune formation, and on the transition from one type of dune to another. These mathematical/ physical models usually require a long integration time, therefore only enabling the simulations of relatively small dune fields. An alter- native approach is to model the vegetation and BSC cover of the dunes, ignoring dune patterns and 3D dune dynamics, and to determine dune stability (active or fixed) according to the fraction of cover of vegetation and BSC; bare dunes are active, while vegetated and/or BSC covered dunes are fixed [23 -- 26]. Such models require a relatively short computation time and have been used to explain the bi-stability of active and fixed dunes under similar climatic conditions. In addition, it is possible to model the development of a 2D vegetation cover by considering the spatial effect of the wind and the diffusion of vegetation [26]. Both observations [6] and models [25] indicate that BSC plays an important role in dune stabilization in arid regions with relatively weak winds. The movement of windblown sand is a stress to "regular" vegetation (hereafter "vegetation"). Some species have evolved to tolerate, and even flourish in, moving-sand environments. These plants are are called "psammophilous plants" [27, 28]. Psammophilous plants have developed several physiological mechanisms to survive and benefit from sand drift. Here, we focus on the 2 following interactions of these plants with sand drift: i) exposure of roots due to sand movement; some of these plants can grow leaves on the exposed roots, thereby increasing their photosynthesis and their growth rates; ii) burial of branches by the windblown sand; some of these plants are able to use the buried branches as roots, thereby enhancing the growth rate of aboveground biomass without changing the root:shoot ratio; iii) tearing of branches/leaves by the wind and their burial by the sand; in some of these plants, this is a mechanism that enhances the clonal growth through the development of new plants from the buried branches. These interactions may be divided into two groups: interactions (i) and (ii) whose rate of occurrence and efficiency are determined by the actual sand drift (hereafter, we will refer to this group as mechanism I), and interaction (iii) whose rate and efficiency are determined by the wind drift potential (hereafter, mechanism II). We note that this is an oversimplified classification of the interactions of psammophilous plants with the wind and the sand drift. Psammophilous plants play an important role in dune stabilization. Due to their adaptation to sand moving environments, they are the first to develop (under suitable environmental condi- tions) in bare and active sand dunes [28]. Once sufficiently dense psammophilous plant cover is established, the sand movement is reduced accordingly, enabling the development of vegetation and BSC. This development further reduces the sand activity, suppressing the growth of psam- mophilous plants, and further enhancing the growth of vegetation and BSC. This process may continue until the dunes become fixed and reach a steady state associated with the environmental conditions. Despite their important role in dune stabilization, to our knowledge, psammophilous plants have never been incorporated into mathematical models of sand dunes. The major goal of this study is to investigate the dynamics of psammophilous plants on sand dunes when coupled to vegetation and BSC dynamics. The model suggested below is a natural extension of the model of [23] and others [24, 24, 25]. The model describes the development of vegetation, BSC, and psammophilous plants on sand dunes, taking into account the effects of the wind and the precipitation. We suggest two ways to model psammophilous plants. The first approach aims to describe "mechanism I," in which the growth of the psammophilous plants is optimal under a specified sand flux. The second approach describes "mechanism II," in which psammophilous plants reach their optimal growth under a specified optimal wind power (or drift potential, defined below). The setup up of the models of mechanisms I and II is different, since the drift potential, used to model the optimal growth due to mechanism II, is not affected by the actual dune cover, while the sand flux that is used to model mechanism I is strongly affected 3 by the dune cover. The modeling of mechanism I yielded a richer bifurcation diagram (steady states map) compared to the modeling of mechanism II. Both modeling approaches show that for some climatic conditions (a region in the drift potential and precipitation rate parameter space), the psammophilous plants act as pioneers in colonizing sand dunes, followed by vegetation and/or BSC that dominates the sand dune cover toward its stabilization. This dynamics is in agreement with the scenario suggested by [28]. II. THE MODEL Our model for psammophilous plants (coupled to vegetation and BSC) follows previously sug- gested mean field models [23 -- 25] for the dynamics of vegetation and BSC cover of sand dunes. The dynamical variables in our model are the fractions of regular vegetation cover, v, BSC cover, b, and psammophilous plant cover, vp, where vp is a new variable added to the model described in [25]. The effects considered in the previous models [23 -- 25], as well as in this model, may be divided into three categories: effects that are not related to the wind, effects that are directly related to the wind, and effects that are indirectly related to the wind (representing aeolian effects). The effects that are not related to the wind include the growth and mortality of the different cover types. We assume a logistic type growth [29]. The natural growth rate, αj (p) (j stands for the cover type, either b, v or vp), depends on the precipitation rate, p; for simplicity and consistency with previous works, we adopt the form of [23 -- 26], αj (p) ≡ α maxj(cid:18)1 − exp(cid:18)p − p minj cj (cid:19)(cid:19) j ∈ {v, vp, b} . (1) α maxj is the maximal growth rate of the j'th cover type. This maximal growth rate is achieved when the precipitation rate, p, is high enough not to be a growth limiting factor and when the other climatic conditions are optimal. In addition, we consider the spontaneous growth of the cover types (growth occurring even in bare dunes) due to effects not modeled here, such as the soil seed bank, underground roots and seed dispersal by the wind and animals. These effects are characterized by spontaneous growth rates, ηj. The wind-independent mortality is accounted for by an effective mortality rate for each cover type, µj. In modeling the direct and indirect effects of the wind, we use the wind drift potential, Dp, as a measure of the wind power [15]; Dp is linearly proportional to the sand drift.The wind drift 4 potential is defined as Dp ≡ hU 2 (U − Ut)i, (2) where U is the wind speed (at 10m height above the ground) measured in knots (1knot = 0.514m/s), Ut = 12knots is the threshold wind speed necessary for sand transport, and the h·i denotes a time average. When the wind speed, U, is measured in knots, Dp, is measured in vector units, V U. Dp provides only the potential value of sand drift; in the case of unidirectional wind, it coincides with the resultant wind drift potential (RDP), which also takes into account the wind direction. Here, we assume that the winds are unidirectional and use Dp instead of RDP. Two important, direct and indirect, wind effects are considered in our model. The direct dam- age/mortality by the wind is proportional to the square of the wind speed (which is proportional to the wind stress). For simplicity, and in order to minimize the number of the parameters in the model, we assume that the direct damage by the wind is proportional to D2/3 effect is the movement of windblown sandthat is, sand drift. The sand drift is equal to the drift . The indirect wind p potential multiplied by the amount of sand multiplied by a function, g(v, vp), which accounts for the sand-drift shading by the vegetation [30, 31]. The sand-drift shading function is assumed to be a step-like function that, above some critical value of the vegetation cover, vc, drops to zero, while for values of the vegetation cover much lower than vc, it obtains its maximal value, 1 [30]. For simplicity, it is assumed that g(v, vp) is a function of the difference between the actual fraction of vegetation cover, v + vp, and the critical value vc. In the previous models [23 -- 26], the sand drift was considered as a damaging effect, increasing the mortality of regular vegetation and BSC due to root exposure and aboveground biomass burial by the sand. Here, we focus on the role of psammophilous plants, and therefore, we consider their cover fraction, vp, as an additional dynamical variable with a unique interaction with the sand drift. The psammophilous plants reach their maximal growth rate under optimal sand drift [28], provid- ing them with the necessary rate of sand cover and/or exposure and branch/leaf seeding (the two mechanisms described in the introduction). These unusual optimal conditions yield a different dy- namics of psammophilous plants. This dynamics, when coupled with the dynamics of the regular vegetation and BSC, leads to interesting and complex bifurcation diagrams (steady states) of the sand dunes and their cover types. The complete set of equations describing the dynamics of the 5 sand dune cover types is: ∂tv = αv (p) (v + ηv) s − γvD2/3 p v − ǫvvD − µvv, ∂tvp = αvp (p)(cid:0)vp + ηvp(cid:1) s − γvpD2/3 ∂tb = αb (p) (b + ηb) s − ǫbbD − µbb. p vp − ǫvpvpfi (Dp, v, b, vp) − µvpvp, We introduce the following notations: s is the fraction of bare sand The sand drift, D, is defined as: s ≡ 1 − v − vp − b. D ≡ Dp × g (v + vp − vc) × s. The sand drift shading function is defined as: (3a) (3b) (3c) (4) (5) (6) 1 x < −1/d 0.5 (1 − xd) −1/d < x < 1/d 0 x > 1/d g (x) ≡  where the parameter d determines the sharpness of the transition from total sand-drift shading to its absence. The effect of sand drift on psammophilous plants is different than its effect on the other types of sand cover. Here, we consider two different options of modeling this unique interaction of psammophilous plants with sand drift, mechanisms I and II which were explained above. These two mechanisms are modeled using different forms of the function fi (Dp, v, b, vp). In the first approach (mechanism I), only the sand drift is assumed to affect the dynamics of vp. Therefore, fI (Dp, v, b, vp) = Q (D). Q (D) obtains its minimal value, Q (Dopt) = 0, for D = Dopt. Away from the optimal sand drift conditions, Q (D) is larger than zero and hence introduces a mortality term due to the non-optimal sand drift conditions. This form of the function Q (D) reflects the fact that the maximal growth rate, α maxvp, is assumed to be under optimal sand drift conditions. We thus used the following form of Q (D) 1 σ2 (D − Dopt)2 1 D − Dopt < σ D − Dopt > σ (7) Q (D) ≡  This choice of the function reflects the behavior described above. Other choices of the function Q (D) (such as 1−exp(cid:0)(D − Dopt)2 /σ2(cid:1)) yielded similar results. Therefore, we decided to focus 6 on this specific, rather simple, choice. It is important to note that this form of the function ensures that vp is never high enough (compared to vc) to create a total sand-drift shading (this would result in a vanishing sand drift and therefore, in far from optimal conditions). The second approach to model the enhanced growth of psammophilous plants under sand drift conditions (mechanism II) is simpler and assumes that the optimal growth conditions are achieved under an optimal wind drift potential rather than under optimal sand drift conditions. Therefore, we assume that fII (Dp, v, b, vp) = D × R (Dp), where R (Dp) ≡ 1 − ρ exp (Dp − Dp,opt)2 2σ2 ! . (8) The parameter ρ > 1 and is set to ensure that the maximal value of vp won't exceed vc. Note that when (Dp − Dp,opt)2 ≫ 2σ2, this vegetation growth term turns into an indirect mortality term, similar to the interaction of regular vegetation with the sand drift. Below, we refer to the different approaches as model I and II, respectively (corresponding to mechanisms I and II for the enhanced net growth of psammophilous plants). For consistency with previous studies [23 -- 25], we use the following values for the parameters:α maxv = 0.15/yr, p minv = 50mm/yr, cv = 100mm/yr, ηv = 0.2, µv = 0, γv = 0.0008V U 3/2/yr, ǫv = 0.001/V U/yr. α maxb = 0.015/yr, p minb = 20mm/yr, cb = 50mm/yr, ηb = 0.1, ǫb = 0.0001/V U/yr. α maxvp = 0.15/yr, p minvp = 50mm/yr, cvp = 100mm/yr, ηvp = 0.2, σ = 100V U, γvp = 0.0006V U 3/2/yr, vc = 0.3, d = 15/2.35 ≈ 6.383 (for a discussion of the choice of the value of d, see [24]). In model I, ǫvp = 0.2/yr, µvp = 0 and Dopt = 300V U, while in model II, ǫvp = ǫv = 0.001/V U/yr, µvp = 1.2/yr, Dp,opt = 300V U, and ρ = 5.0072. In what follows, the drift potential, Dp, will be measured in units of V U, the precipitation rate, p, in units of mm/yr and the mortality rate of the BSC, µb, in units of 1/yr. Time is measured in years. For convenience, we drop the units hereafter. Justification regarding the choice of the parameters and the model setup can be obtained from [23 -- 26]. Note that for simplicity, we do not include direct competition terms between the different models variables, unlike [25]. Below, we present results for different values of Dp, p and µb. III. RESULTS We started our model analysis by studying the number of physical solutions (0 < v, vp, b < 1) for given wind conditions characterized by Dp, and precipitation rate, p; these are the two 7 (a) (c) p D 4000 3000 2000 1000 p D 4000 3000 2000 1000 100 200 300 400 p 100 200 300 400 p 3 2 1 5 4 3 2 1 (b) (d) 4000 3000 p D 2000 1000 800 600 400 200 p D 3 2 1 3 2 1 100 200 300 400 p 100 200 300 400 p FIG. 1: Maps of the number of solutions as a function of the precipitation rate, p and the wind drift potential Dp. Panels (a)-(c) correspond to model I with BSC mortality rates µb = 0.001, 0.006, 0.01, respectively. Panel (d) corresponds to model II and BSC mortality rate µb = 0.006. main climatic factors that affect sand dune dynamics. We found that the number of physical solutions strongly depends on the maximal value of the BSC cover, which is determined by the BSC mortality rate, µb, and the other parameters. In Fig. 1, we show maps of the total number of physical solutions (both stable and unstable) for different values of p and Dp. Panels (a)-(c) correspond to model I and BSC mortality rates µb = 0.001, 0.006, 0.01, respectively. Panel (a) corresponds to a small value of the BSC mortality rate, µb = 0.001, and it shows the existence of a typical bi-stability region (the region with three solutions, two of which are stable and one is unstable). Panel (b) corresponds to a higher value of the BSC mortality rate, µb = 0.006, for which we have two regions of bi-stability. Panel (c) corresponds to an even higher value of the BSC mortality rate, µb = 0.01, for which we obtain two regions of bi-stability and, in addition, a region of tri-stability (the total number of steady states is 5). Panel (d) corresponds to model II with µb = 0.006. It shows the existence of a bi-stability range. For much smaller BSC mortality rates, model II doesn't show a bi-stability region; higher values of µb change the location (in the parameter space) of the bi-stability region but do not result in a qualitatively different bifurcation diagram. 8 Fig. 1 shows that the two approaches, adopted in models I and II respectively, result in different numbers of physical steady state solutions. In model I, for all three values of µb considered here, we have at least one range with three physical solutions (as shown in Fig. 1). Two of the three solutions are stable and one is unstable. As we increase the mortality rate of the BSC (see Fig. 1(b)), and therefore, reduce the maximal value of b, a second region of bi-stability appears. A further increase of µb results in an overlap of the two bi-stability regions, and therefore, in a range of tri-stability in which we have five physical solutions (three stable solutions and two unstable ones, see Fig. 1(c)). The two bi-stability regions are due to the different actions of the sand-drift shading on the regular and the psammophilous plants. Small values of µb allow for high values of b, and therefore, the only possible bi-stability is due to low or high values of vp which, by shading, reduces the sand drift even for high values of Dp. It is important to note that in this case, one of the states corresponds to active dunes, while the other one corresponds to marginally stable dunes. The psammophilous plants can never cover the dunes to the extent to which there is no sand drift because they cannot survive away from the optimal sand drift, Dopt. For higher values of µb, the previously observed bi-stability of active and stable dunes, due to sand-drift shading by regular plants [25], appears and creates the second range of bi-stability for lower values of Dp. Further increasing the BSC mortality rate results in an overlap of the two bi-stability ranges, and therefore, in a range of tri-stability. In model II, there is, at most, one region of bi-stability, as shown in Fig. 1(d). For the parameters explored here, we could not identify two distinct mechanisms of bi- stability. For much smaller values of µb, there is no bi-stability range, and for all values of p and Dp, there is only one physical solution. For larger values of mub, the bifurcation diagram is qualitatively the same as the one presented in Fig. 1(d). Similar behavior is obtained when considering only vegetation [24] and when considering BSC in addition to regular vegetation [25]. To better understand the complex steady state phase space, we present in Fig. 2 the bifurcation diagrams, as predicted by model I, against the drift potential. These bifurcation diagrams show cross-sections along the vertical dashed lines in Fig. 1(c). The two columns correspond to two values of the precipitation rate. The different rows correspond to the different cover type fractions. We also present: (i) the total vegetation cover (v + vp) which determines the stability of the dunes, and (ii) the fraction of exposed sand. Obviously, these two variables may be extracted from v, b, and vp and are only shown for clarity. For the higher value of the precipitation rate, p = 300, there are two Dp ranges of a single stable state (for low and high values of Dp). In addition, there are two ranges of bi-stability -- one in which the dunes may be exposed or densely covered and a 9 s b v p v t v 0.9 0.6 0.3 0 0.15 0.1 0.05 0 0.9 0.6 0.3 0 0.3 0.2 0.1 0 0.9 0.6 0.3 0 0 p=100 p=300 500 1000 Dp 1500 0 1000 2000 Dp 3000 vc FIG. 2: Bifurcation diagrams versus the drift potential, Dp, as predicted by model I along the vertical dashed lines indicated in Fig. 1(c). The left column corresponds to precipitation rate, p = 100, and the right column to p = 300. The rows (from top to bottom) correspond to the fractions of uncovered sand, s, BSC cover, b, psammophilous plant cover, vp, regular vegetation cover, v, and the total sand-drift shading vegetation, vt ≡ vp + v (the dotted line marks the critical value of the vegetation cover for sand-drift shading, vc). The solid (dashed) lines correspond to stable (unstable) states. The BSC mortality rate is µb = 0.01. second range in which the dunes may be exposed or partially covered (vt ∼ vc). In between the two bi-stability ranges, there is a range of Dp for which we have tri-stability. Namely, the dunes may be exposed, densely covered or partially covered. For the smaller precipitation rate, p = 100, the tri-stability range disappears. The value of the BSC mortality was set equal to the value used in Fig. 1(c) to capture the more complicated bifurcation diagrams. To complete the picture of the bifurcation diagrams, as predicted by model I, we show in Fig. 3 the bifurcation diagrams against the precipitation rate for a fixed value of the drift potential (Dp = 500, set to ensure that all of the five physical solutions are captured). These diagrams correspond to a cross-section along the dashed horizontal line in Fig. 1(c). In these diagrams, one can see the onset of bi-stability, followed by the onset of tri-stability which later on disappears as the precipitation rate increases. Fig. 4 depicts the bifurcation diagrams versus the drift potential, as predicted by model II, for 10 s b v p v t v 0.9 0.6 0.3 0 0.15 0.1 0.05 0 0.6 0.4 0.2 0 0.3 0.2 0.1 0 0.6 0.4 0.2 0 0 100 200 300 400 p vc FIG. 3: Bifurcation diagrams versus the precipitation rate, p, as predicted by model I along the horizontal dashed line of Fig. 1(c). The different rows correspond to the cover type fractions as in Fig. 2. The drift potential was set to Dp = 500, to capture all the solution branches. The BSC mortality rate was set to µb = 0.01. two values of the precipitation rate, p = 100 and p = 300. The bifurcation diagrams are taken along cross-sections corresponding to the dashed vertical lines in Fig. 1(d). For both values of the precipitation rate, there is only one bi-stability range. However, its width, shape and location (in the parameter space) are affected by the value of p. A significant difference between model II and model I is the lack in the former of a steady state corresponding to partially covered dunes (vt ∼ vc). Bifurcation diagrams versus the precipitation rate as predicted by model II are presented in Fig. 5. The drift potential was set to the optimal value for psammophilous plants according to this model, Dp = 300 (corresponding to the horizontal dashed line in Fig. 1(d)). In these diagrams, one can see the onset of bi-stability and its disappearance as the precipitation rate increases. The bifurcation diagrams alone do not elucidate all the information provided by the models. The dynamics is of relevance and importance to understanding the role of psammophilous plants in sand dune dynamics. We started exploring the dynamics predicted by the models by investigating the steady state reached from different initial conditions. In Fig. 6, we show the steady states reached by different initial conditions calculated using model I. Columns (a)-(d) correspond to the initial conditions of bare sand dunes (v(t = 0) = b(t = 0) = vp(t = 0) = 0), vegetation- 11 s b v p v t v 0.9 0.6 0.3 0 0.3 0.2 0.1 0 0.9 0.6 0.3 0 0.09 0.06 0.03 0 0.9 0.6 0.3 0 0 p=100 p=300 100 200 Dp 300 400 0 200 400 600 800 Dp vc FIG. 4: Bifurcation diagrams as predicted by model II along the vertical dashed lines of Fig. 1(d). The bifurcation parameter is the drift potential, Dp. The different rows correspond to the cover type fractions, as in Fig. 2. The left column corresponds to precipitation rate, p = 100, and the right column corresponds to precipitation rate, p = 300. s b v p v t v 0.9 0.6 0.3 0 0.2 0.1 0 0.9 0.6 0.3 0 0.09 0.06 0.03 0 0.9 0.6 0.3 0 0 100 200 300 400 p vc FIG. 5: Bifurcation diagrams as predicted by model II. The bifurcation parameter is the precipitation rate, p. The different rows correspond to the cover type fractions. The drift potential was set to, Dp = 300. 12 covered sand dunes (v(t = 0) = 1; b(t = 0) = vp(t = 0) = 0), BSC- covered sand dunes (v(t = 0) = 0; b(t = 0) = 1; vp(t = 0) = 0) and psammophilous plant-covered sand dunes (v(t = 0) = b(t = 0) = 0; vp(t = 0) = 1), respectively. The different rows correspond to the cover type fractions. Here again, for convenience, we show the exposed sand fraction. The different initial conditions resulted in different steady state maps. For all initial conditions, we found that for a low drift potential and a not too low precipitation rate (the lowest part of the panels of the first row in Fig. 6), the vegetation cover dominates and stabilizes the sand dunes. For the full vegetation cover initial condition, the vegetation remains dominant, even at higher values of the drift potential (see column (b) of Fig. 6). For all initial conditions and climatic conditions, except for a small regime of intermediate drift potential and low precipitation, the fraction of BSC cover is relatively small. The steady state with a maximal fraction of BSC cover is obtained for intermediate values of the drift potential and a not too small precipitation rate for an initial condition of full vegetation cover (see column (b) of Fig. 6). The psammophilous plant cover fraction is significant for intermediate and high values of the drift potential for all initial conditions except for the full vegetation cover initial condition for which vp only dominates at high values of the drift potential. The maximal value of vp is obtained for the vp = 1 initial condition. These results suggest that the basin of attraction of the steady state solution with vp ∼ vc is relatively small if there is a stable state with a high value of v. The bottom row in Fig. 6 shows that for a bare dune initial condition, stabilization of the dunes is only possible at a high enough precipitation rate and a low drift potential. For a small range of intermediate drift potential, the steady state is partially covered dunes, namely, vt ∼ vc. For the v = 1 initial condition, we found that the dunes remain stabilized for a high enough precipitation rate and an intermediate or weak drift potential. Note that for this initial condition, the partially covered dune steady state does not appear (for any climatic condition). For the b = 1 initial condition the steady state map is very similar to the map obtained for the bare dune initial condition. However, the region of partially covered dunes in steady state is larger and extends to higher values of the drift potential. For the vp = 1 initial condition, the steady state map shows a large region of partially covered dunes in steady state. Here, this region extends to very high values of the drift potential. Fig. 7 shows the steady states of model II reached by different initial conditions. Column (a) corresponds to the bare dune initial condition and column (b) corresponds to the full vegetation cover initial condition. Initial conditions of full psammophilous plant and BSC cover resulted 13 (a) 2 3 1 1 2 4000 3000 2000 1000 0 4000 3000 2000 1000 0 4000 3000 2000 1000 0 p D 4000 3000 2000 1000 0 0 100 200 300 400 (b) 100 200 300 400 0 (c) 100 200 300 400 (d) 100 200 300 400 0 v b vp s 0.75 0.5 0.25 0 0.12 0.08 0.04 0 0.3 0.2 0.1 0 0.75 0.5 0.25 0 0 p FIG. 6: The steady states corresponding to different initial conditions as predicted by model I as a function of p and Dp. Column (a) corresponds to the bare dune initial condition, v(t = 0) = b(t = 0) = vp(t = 0) = 0. Column (b) corresponds to the vegetation-covered dune initial condition, v(t = 0) = 1; b(t = 0) = vp(t = 0) = 0. Column (c) corresponds to the crust-covered dune initial condition, v(t = 0) = 0; b(t = 0) = 1; vp(t = 0) = 0. Column (d) corresponds to the psammophilous plant-covered dune initial condition, v(t = 0) = b(t = 0) = 0; vp(t = 0) = 1. The different rows correspond to the different cover type fractions as indicated. The dashed lines mark the edges of the multi-stability regimes, namely, the crossing lines between regions with different numbers of physical solutions as shown in Fig. 1(c). The numbers in the top left panel indicate the number of stable physical solutions in each region. in the same steady state as the bare dune initial condition. These results suggest that the basins of attraction of the states in the bi-stability regime are determined by the value of v and are less sensitive to the values of b and vp in this model. Similarly to model I, we found that in model II, for all initial conditions, a low drift potential and a not too low precipitation rate, the vegetation cover dominates and stabilizes the sand dunes. For the full vegetation cover initial condition, the vegetation remains dominant even at higher values of the drift potential (see column (b) of Fig. 7). Note that for the parameters used here, these values of the drift potential are significantly lower than the values predicted by model I; it is possible to extend these regions by choosing a larger 14 (a) (b) p D 800 600 400 200 800 600 400 200 800 600 400 200 800 600 400 200 1 2 1 100 200 300 400 p v b 0.75 0.5 0.25 0 0.24 0.16 0.08 0 0.075 0.05 0.025 vp 0 0.75 0.5 0.25 0 s 100 200 300 400 FIG. 7: The steady states corresponding to different initial conditions as a function of p and Dp, as predicted by model II. The left column corresponds to the bare dune initial condition, v(t = 0) = b(t = 0) = vp(t = 0) = 0, and the right column corresponds to the full vegetation cover initial condition, v(t = 0) = 1; b(t = 0) = vp(t = 0) = 0. The different rows correspond to the different cover type fractions as indicated. The dashed lines mark the edges of the bi-stability regime as shown in Fig. 1(d). The numbers in the top left panel indicate the number of stable solutions in each region. maximal growth rate, αv,max. For the bare dune initial condition, the fraction of BSC cover is small in all climatic conditions except for a small region of low precipitation and drift potential at which only the BSC can grow. For the v = 1 initial condition and not too low values of the drift potential, the values of b are significant; yet, these values are smaller than the values of v under the same climatic conditions. We see that for the parameters used in model II, the maximal values of vp are significantly smaller than those obtained for model I. The only regions with significant values of vp in steady state are around Dp,opt. For the bare dune initial condition, the region extends to high precipitation rates, while for the full vegetation cover initial condition, this region is truncated at low precipitation rates. The bottom row of Fig. 7 shows that the stability map of the sand dunes is similar to one obtained when the psammophilous plants are neglected and only the vegetation and BSC are considered as dynamical variables. A common paradigm for the stabilization of sand dunes under high sand drift is that the psam- mophilous plants act as pioneers [28]. Due to their ability to flourish under significant sand drift, they are the first to colonize bare dunes. Their growth reduces the sand drift by wind shading 15 b v vp p v , b , v 0.6 0.3 0 0 20 40 time 60 3 6 9 time 12 15 18 p v , b , v 0.4 0.3 0.2 0.1 0 0 FIG. 8: Time evolution of the cover type fractions as predicted by model I. The precipitation rate was set to p = 300. The dashed lines correspond to drift potential, Dp = 200, and the solid lines correspond to Dp = 300. The solid lines show dynamics in which the psammophilous plants initially dominate, and later on, the normal vegetation dominates. The dashed lines show an evolution in which the psammophilous plants and the normal vegetation grow equally at first, and later on, the normal vegetation dominates. and enables the growth of regular vegetation, eventually resulting in stabilized dunes in which the fraction of vegetation cover dominates. In order to test if our models are capable of reproducing this paradigm, we investigated the temporal dynamics. In Fig. 8, we show the values of v, b and vp versus the time for the bare dune initial condition. The precipitation rate was set to p = 300. The dashed lines correspond to Dp = 200, and the solid lines correspond to Dp = 300. Our results show that under some climatic conditions, the dynamics follows the paradigm (the solid lines), while under other climatic conditions, the initial growth of the vegetation is identical to the initial growth of the psammophilous plants (the dashed lines). The results presented in Fig. 8 were calculated using model I. For the same climatic conditions, model II yields qualitatively the same results. IV. SUMMARY AND DISCUSSION We have studied the effect of psammophilous plants on the dynamics of sand dunes using a simple mean field model for sand dune cover dynamics (vegetation, BSC and psammophilous plants). Two main mechanisms of interaction of the psammophilous plants with the wind and the sand drift were modeled separately. Root exposure and the covering of branches/leaves by the sand drift enhance the net growth of psammophilous plants and result in maximal growth under optimal sand drift conditions (model I). Branches/leaves torn off by the wind and buried by the 16 sand develop new plants and increase the growth rate of psammophilous plants under an optimal wind drift potential (model II). These two approaches resulted in qualitatively different steady states and dynamics; the sand drift optimal growth model (model I) shows a richer steady state map, with up to three stable dune states (extensive sand cover, moderate sand cover and small sand cover) while the wind drift potential optimal growth rate (model II) shows up to two stable dune states (extensive and small sand cover). While there are examples for the coexistence of active and fixed dunes under similar climate conditions [23, 24], we are not aware of observations that may be associated with the "new" dune state predicted by model I of moderate cover in which the vegetation cover is close to the critical vegetation cover, vc (it is important to note that according to this model, the basin of attraction of this state is very small, and therefore, it may not be easily realized in observations). Identifying such a state in observations will provide strong support for the model's setup. We hope to explore this and other features of the model in the future. The model proposed here does not aim to be operative. It aims to provide a qualitative under- standing of the dynamics of psammophilous plants in the presence of regular vegetation and BSC. Nevertheless, a comparison with observations of the bi-stability of sand dunes reported in [23, 24] indicates that model I fits the observations better than model II; model II exhibits only an active dune state for drift potential values that are larger than 700 (see Fig. 1), while stable dunes exist in nature for higher values of Dp. Model II can to be tuned to fit these observations by increasing α maxv or decreasing ǫv. We do not present these results here, in order to use the same parameters in models I and II wherever possible. Another observation that is worth noting concerns the frac- tion of BSC cover. Both model I and II resulted in BSC cover that does not exceed 25%. Studies have reported almost complete BSC cover on sand dunes for small plots (order of meters) [e.g., Fig. 5B and 7 of 32]. This discrepancy between the model and observations can be attributed to the mean field nature of the model, representing only scales of kms that usually contain several types of dune surface covers. The separation of the two growth mechanisms -- by assuming optimal growth conditions under either (i) an optimal sand drift or (ii) an optimal wind drift potential -- helps us to understand the effect of each of these mechanisms. Yet, a more realistic model should include a combination of these two mechanisms, resulting in, most probably, a richer dynamics and steady state map. Observations regarding psammophilous plants may help to determine the role of each mechanism, which one is more favorable, and if a combination of the two is plausible. 17 For the parameters used here, model I predicts that the psammophilous plant cover can reach the critical value for sand-drift shading, while model II predicts that their fraction of cover is very small. According to both models, there are climate conditions under which the steady state picture is not affected by the presence of the psammophilous plant cover fraction as an additional dynam- ical variable in the model. For example, under a low wind drift potential and a high precipitation rate, regular vegetation is the dominant cover type, and the BSC and the psammophilous plants may be ignored. Under extremely dry conditions, p < 50mm/yr, and a low wind drift poten- tial, the BSC will be the dominant cover type, and both regular vegetation and psammophilous plants may be ignored. Under an extremely high wind drift potential, the dunes will be fully ac- tive without any surface cover. Psammophilous plants may be the dominant cover type under a high enough precipitation rate (p > 50mm/yr) and a strong wind drift potential (the definition of strong depends on the model (I or II) and the parameters). While some of the cover types can be ignored in the steady state, they can still greatly influence the dynamics leading to the observed steady state. We have demonstrated that starting from a bare dune state, psammophilous plants may be the first to grow, reducing the sand drift and thus enabling the growth of regular vegetation which eventually dominates and stabilizes the dunes [28]. This, however, is not always the case as our model predicts that under different climate conditions (wind drift potential), the regular vegetation and psammophilous plants grow equally at first, cooperating in reducing the sand drift, followed by a faster growth of the regular vegetation, which eventually dominates and stabilizes the dunes. Our preliminary numerical results suggest the possibility of a Hopf bifurcation leading to os- cillatory behavior of the different cover types. This behavior and its relevance to observations, as well as the incorporation of spatial effects in the model, as was done in [26], are left for future research. In addition, we plan to use this model and its extensions to study the response of sand dunes to different scenarios of climate change. Acknowledgments The research leading to these results has received funding from the European Union Seventh Framework Programme (FP7/2007-2013) under grant number 293825. This research was also supported by the Israel Science Foundation. We thank Dotan Perlstein for his contribution during 18 the first stages of this research and Shai Kinast for helpful discussions. [1] K. Pye, Geografiska Annaler. Series A. Physical Geography 64, 213 (1982). [2] K. Pye and H. Tsoar, Aeolian Sand and Sand Dunes (Unwin Hyman, London, 1990). [3] D. S. G. Thomas and G. F. S. Wiggs, Earth Surface Processes and Landforms 33, 1396 (2008). [4] R. A. Bagnold, The physics of blown sands and desert dunes (Chapmann and Hall, London, 1941). [5] H. Tsoar, in Arid Dune Ecosystems, edited by S. W. Breckle, A. Yair, and M. Veste (Springer, Berlin, 2008), vol. 200 of Ecological Studies, pp. 79 -- 90. [6] M. Veste et al., in Sustainable Land Use in Deserts, edited by S. Breckle, M. Veste, and W. Wucherer (Springer, 2001), pp. 357 -- 367. [7] Z. Dong et al., J. Arid. Environ. 57, 329 (2005). [8] M. Khalaf and D. Al-Ajmi, Geomorphology 6, 111 (1993). [9] U. Shanas et al., Biological Conservation 132, 292 (2006). [10] D. S. G. Thomas, M. Knight, and G. F. S. Wiggs, Nature 435, 1218 (2005). [11] Y. Ashkenazy, H. Yizhaq, and H. Tsoar, Climate Change 112, 901 (2012). [12] J. Otterman, Science 186, 531 (1974). [13] J. G. Charney, P. H. Stone, and W. J. Quirk, Science 187, 434 (1975). [14] H. Tsoar, Physica A 357, 50 (2005). [15] S. G. Fryberger, in A study of global sand seas, edited by E. D. McKee (U.S. Geol. Surv., 1979), vol. 1052, pp. 137 -- 169. [16] A. Danin, Y. Bar-Or, I. Dor, and T. Yisraeli, Ecologica Mediterranea 15, 55 (1989). [17] B. Andreotti, P. Claudin, and S. Douady, Eur. Phys. J 28, 341 (2002). [18] O. Dur´an and H. J. Herrmann, Phys. Rev. Lett. 97, 188001 (2006). [19] M. C. M. de M. Luna et al., Physica A 388, 4205 (2009). [20] M. D. Reitz et al., Geophys. Res. Lett. 37, L19402 (2010). [21] J. M. Nield and A. C. W. Baas, Earth Planet. Sci. Lett. 33, 724 (2008). [22] J. F. Kok et al., Reports on Progress in Physics 75, 106901 (2012). [23] H. Yizhaq, Y. Ashkenazy, and H. Tsoar, Phys. Rev. Lett. 98, 188001 (2007). [24] H. Yizhaq, Y. Ashkenazy, and H. Tsoar, J. Geophys. Res. 114, F01023 (2009). [25] S. Kinast et al., Phys. Rev. E 87, 020701(R) (2013). 19 [26] H. Yizhaq et al., Physica A 392, 4502 (2013). [27] A. Danin, J. Arid Environments 21, 193 (1991). [28] A. Danin, Plants of desert dunes (Springer, Berlin, 1996). [29] M. Baudena et al., Advances in Water Resources 30, 1320 (2007). [30] B. E. Lee and B. F. Soliman, Transact. ASME, J. Fluid Eng. 99, 503 (1977). [31] S. A. Wolfe and W. C. Nickling, Prog. Phys. Geog. 17, 50 (1993). [32] M. Veste, K. Breckle, S. W. Eggert, and T. Littmann, Basic and Applied Dryland Research 5, 1 (2011). 20
1209.2588
2
1209
2013-04-18T22:13:38
Demographic noise and resilience in a semi-arid ecosystem model
[ "physics.bio-ph", "cond-mat.stat-mech", "q-bio.PE" ]
The scarcity of water characterising drylands forces vegetation to adopt appropriate survival strategies. Some of these generate water-vegetation feedback mechanisms that can lead to spatial self-organisation of vegetation, as it has been shown with models representing plants by a density of biomass, varying continuously in time and space. However, although plants are usually quite plastic they also display discrete qualities and stochastic behaviour. These features may give rise to demographic noise, which in certain cases can influence the qualitative dynamics of ecosystem models. In the present work we explore the effects of demographic noise on the resilience of a model semi-arid ecosystem. We introduce a spatial stochastic eco-hydrological hybrid model in which plants are modelled as discrete entities subject to stochastic dynamical rules, while the dynamics of surface and soil water are described by continuous variables. The model has a deterministic approximation very similar to previous continuous models of arid and semi-arid ecosystems. By means of numerical simulations we show that demographic noise can have important effects on the extinction and recovery dynamics of the system. In particular we find that the stochastic model escapes extinction under a wide range of conditions for which the corresponding deterministic approximation predicts absorption into desert states.
physics.bio-ph
physics
Demographic noise and resilience in a semi-arid ecosystem model John Realpe-Gomeza,b, Mara Baudenac,d, Tobias Gallaa, Alan J. McKanea, Max Rietkerkc aTheoretical Physics, School of Physics and Astronomy, The University of Manchester, Manchester M13 9PL, United Kingdom bGrupo de Ciencia Transdisciplinar, Informaci´on, y Complejidad, Instituto de Matem´aticas Aplicadas, Universidad de Cartagena, Bol´ıvar, cDepartment of Environmental Sciences, Copernicus Institute, Utrecht University, P.O. Box 80155 TC Utrecht, The Netherlands dGrupo Interdisciplinar de Sistemas Complejos (GISC), Departamento de Matem´aticas, Universidad Carlos III de Madrid, Avenida de la Universidad 30, 28911 Legan´es, Madrid, Spain Colombia 3 1 0 2 r p A 8 1 ] h p - o i b . s c i s y h p [ 2 v 8 8 5 2 . 9 0 2 1 : v i X r a Abstract The scarcity of water characterising drylands forces vegetation to adopt appropriate survival strategies. Some of these generate water-vegetation feedback mechanisms that can lead to spatial self-organisation of vegetation, as it has been shown with models representing plants by a density of biomass, varying continuously in time and space. However, although plants are usually quite plastic they also display discrete qualities and stochastic behaviour. These features may give rise to demographic noise, which in certain cases can influence the qualitative dynamics of ecosystem models. In the present work we explore the effects of demographic noise on the resilience of a model semi-arid ecosystem. We introduce a spatial stochastic eco-hydrological hybrid model in which plants are modelled as discrete entities subject to stochastic dynamical rules, while the dynamics of surface and soil water are described by continuous variables. The model has a deterministic approximation very similar to previous continuous models of arid and semi-arid ecosystems. By means of numerical simulations we show that demographic noise can have important effects on the extinction and recovery dynamics of the system. In particular we find that the stochastic model escapes extinction under a wide range of conditions for which the corresponding deterministic approximation predicts absorption into desert states. Keywords: Semi-arid ecosystems, Resilience, Vegetation patterns, Extinction, Recovery, Stochastic processes 1. Introduction In arid and semi-arid ecosystems water constitutes the main limiting resource for vegetation. The harsh envi- ronmental conditions, related to frequent droughts, limit plant recruitment and survival and often prevent veg- etation from fully covering the ground. In such ar- eas vegetation largely occurs in patches of high den- sity surrounded by bare soil, and forming spatial vegeta- tion patterns (e.g. Aguiar and Sala, 1999; Barbier et al., 2006; Deblauwe et al., 2008). These spatial structures emerge from the system’s dynamics as a consequence of scale-dependent water-vegetation feedback mechanisms even in absence of any underlying spatial heterogene- ity (e.g. Klausmeier, 1999; von Hardenberg et al., 2001; Rietkerk et al., 2002; Meron, 2012). in arid areas the vegetation increases the infiltration ca- pacity of the soil as compared to bare ground because of root penetration (Rietkerk and van de Koppel, 1997; Walker et al., 1981) and because it prevents the forma- tion of biogenic crusts that would form in bare soil For example, Email addresses: [email protected] (John Realpe-Gomez), [email protected] (Mara Baudena), [email protected] (Tobias Galla), [email protected] (Alan J. McKane), [email protected] (Max Rietkerk) (Casenave and Valentin, 1992). While vegetation patches compete for water, vegetation enhances its own growth locally within the patches. This so-called infiltration feed- back is known to be one of the most important scale- dependent water-vegetation feedback mechanisms, lead- ing to self-organization and pattern formation phenomena (e.g. Rietkerk et al., 2002). Scale-dependent resource concentration mechanisms of this type are also connected to the possible occurrence of catastrophic shifts in ecosystems. For example, a vegetated patchy state may turn into a degraded state with mostly bare soil if rainfall decreases below a cer- tain threshold. A subsequent increase in rainfall above the threshold may not be enough to recover the pre- vious vegetation state (e.g. von Hardenberg et al., 2001; Rietkerk et al., 2004; Rietkerk and van de Koppel, 2008; Baudena and Rietkerk, 2012). This is an example of how the resilience of such ecosystems is strongly interrelated with spatial structure. By ‘resilience’ we mean here the ability of an ecosystem to remain in a given domain of attraction and to return quickly to the same state after disturbances (Rietkerk and van de Koppel, 2008). In models used to represent dryland ecosystems both water and plants are often represented as density fields, varying continuously in time and space, and their dy- namics is typically described by deterministic differential Preprint submitted to Ecological Complexity May 2, 2014 equations (e.g. Gilad et al., 2004; Rietkerk et al., 2002). In other models the plant dynamics is modelled us- ing stochastic differential equations (Manor and Shnerb, 2008), or instead the vegetation is represented by dis- crete states in a drastically simplified way (e.g. K´efi et al., 2007a,b). Continuous models tend to be simple and pro- vide powerful insight through the use of analytical tech- niques. Moreover, in the case of semi-arid ecosystems, they very successfully point out the importance of scale- dependent water-vegetation feedbacks in the self-organised pattern formation processes (e.g. von Hardenberg et al., 2001; Rietkerk et al., 2002). The spatial self-organization of drylands ecosystems has not yet been explored with models representing single plants individually, even though such individual-based models are now increasingly common in other areas of ecology and biology. A reason for this may be that plants are not always conceived as discrete individuals since, unlike animals, they are extremely plastic and can respond quite continuously to environmental changes, for instance by partially dying and recovering later on (Crawley, 1990). However, plants have also discrete features: in their seed stage, they behave as discrete entities that can fall and perhaps germinate in some random position. The birth and establishment of a single plant is also subject to a collection of random events. Furthermore, they can live alone in patches composed essentially of a single plant. Finally, the death of a whole plant is also a possible (unpredictable) event. These features can be readily accommodated in a stochastic individual-based modelling approach. To be more precise, plants in drylands often have a modular structure and are composed of multiple hydraulically independent stems (Schenk et al., 2008). Except when the plant is in a seed stage such stems could be seen as the relevant individual entities rather than the whole plant. In plant ecology, forest models are a successful example of individual-based modelling (Botkin et al., 1972; Shugart, 1984; Pacala et al., 1993, 1996; Moorcroft et al., 2001; Perry and Enright, 2006). The approach has been extended to grasslands as well (Jørgensen, 2011; Coffin and Lauenroth, 1990; Peters, 2002; Rastetter et al., 2003), and it has also been used as a method to parametrise mean-field differential equations from fine-scale data (Moorcroft et al., 2001). In principle, it is possible to capture the most salient features of an IBM by a suitable continuous model (Black and McKane, 2012). How to do this in general is an active area of research in the modelling of complex systems (San Miguel et al., 2012). In the case of forest models, good progress has been made in achieving this (Strigul et al., 2008). Formally speaking, an IBM can coincide with a deterministic continuous model only in the limit of an infinite number of individuals. It might be imagined that a system with a relatively large number of individuals would simply lead to a small amount of noise around the deterministic solution. In some situations this is the case, but in others, however, results 2009; Biancalani et al., 2011; Biancalani et al., the stochastic effects in an IBM can produce qualita- (McKane and Newman, 2005; tively different 2010; Butler and Goldenfeld, Butler and Goldenfeld, 2011; Rogers et al., 2012; Ramaswamy et al., 2012). Still these effects become negligible with a ‘sufficiently’ large number of individuals, but how large is ‘sufficiently’ large? It can actually be unrealistically large and the stochastic effects may turn out to be necessary to account for the observed behaviour of real systems. It is not clear a priori how large the numbers need to be, and the system of interest needs to be carefully analysed before reaching a conclusion. Following the previous discussion it is understandable that a deterministic continuous model can be a good ap- proximation to an individual-based model of forests, where the vegetation is rather dense. In drylands, on the other hand, vegetation is composed of a relatively small number of plants coexisting with regions of empty land in which the above-ground biomass density is essentially zero. Un- der these conditions the intrinsic stochastic behaviour of individual plants may turn out to be relevant for under- standing the behaviour of the system (Black and McKane, 2012). Even more so if the drylands are close to extinction, where the number of individuals is rather scarce. In arid areas, vegetation is strictly dependent on water availability, whose dynamics is effectively captured by a deterministic continuous approach. A natural approach to represent the dynamics of drylands is in terms of so- called ‘hybrid models’, in which the dynamics of water is represented by differential equations, and in which the plant dynamics are described by an IBM (Vincenot et al., 2010). Thus, these models combine both continuous and discrete variables. In our work we use such a hybrid model to study the resilience of semi-arid ecosystems against desertification. The model includes the water-vegetation infiltration feed- back, and thus we expect vegetation patterns to emerge as previously observed in other types of models of semi- arid ecosystems (e.g. Rietkerk et al., 2002). The use of individual-based models can be computationally demand- ing if too many details are included, and this can ob- struct the identification of the underlying mechanisms rel- evant for the behaviour of the system. An example is the continuous range of values characterising each indi- vidual, e.g. mass and heterogeneity: every plant having a different mass. Our approach constitutes a compromise between realism, tractability, and insight, concentrating on what we think could be the most relevant vegetation features. We extend an IBM, recently investigated by Realpe-Gomez et al. (2012) to include spatial interactions. In particular, we neglect the growth phase and assume all plants have the same mass. We expect this to be a rea- sonable approximation if the time it takes for a plant to establish is much smaller than the time scale in which the spatial distribution of biomass changes appreciably and the growth phase of each individual plant is not very rel- 2 evant for the system dynamics. We study the resilience and stability of the model dryland ecosystem when the in- dividual nature of plants and its intrinsic stochasticity are considered. Our aim is to investigate the role of demo- graphic noise and to understand whether noise enhances recovery or whether it may drive the system to extinction. To do so, we will compare the outcomes of a stochastic IBM and a deterministic continuous model. 2. Model definitions 2.1. Stochastic model 2.1.1. General considerations In this section we introduce our stochastic hybrid model for semi-arid ecosystems. The model describes the dy- namics of vegetation, soil and surface water in a given area to which the model is applied. The source of sur- face water is rainfall. Surface water then infiltrates the soil where plants can take it up. Plants are consid- ered as discrete entities following a stochastic dynamics (Realpe-Gomez et al., 2012), while soil and surface wa- ter are modelled by continuous variables whose dynamics would be deterministic (Pueyo et al., 2008; Rietkerk et al., 2002; HilleRisLambers et al., 2001) were they not influ- enced by plants. Figure 1 summarises the dynamics of the model. For computational simplicity we will work mostly in one spatial dimension. Occasionally we will also con- sider the case of two dimensions; this is to illustrate that the results of the one-dimensional model are still valid in this more realistic case. We will define the model in one spatial dimension only, the modifications required to ex- tend it to two dimensions are straightforward. In the one- dimensional model we assume that the land is partitioned into L uniform square cells, each labelled by an index i. Thus the model describes a linear array of square cells. The length of the side of a cell is denoted by h. With ev- ery cell i we associate three variables corresponding to the discrete (integer) number of plants, ni, in the cell and the continuous quantities of soil and surface water, ωi and σi, respectively, on that piece of land. We will assume that the density of biomass per unit area in cell i is given by ρi = µ ni, where µ represents the mass of one plant indi- vidual divided by the area of the cell, i.e. µ = mP /h2. We have made here the simplifying assumption that all plants have identical mass mP . Thus, µ characterises the contri- bution of an individual plant to the biomass density. We consider the cells as homogeneous, i.e. we do not take into account any structure within the cell, such as the spatial distribution of plants within a cell. We also neglect prop- erties of individual plants, such as for example different plant sizes or stages of growth. The detailed dynamics of the various components of the model are explained below, where we also introduce the relevant model parameters. The units and numerical values of all the parameters in the model are summarised in Table 1. 2.1.2. Water dynamics For all cells, i, the dynamics of the depths of soil and surface water, ωi and σi (measured in mm) are described by differential equations that are deterministic when the biomass density ρi (measured in g m−2) is kept constant; namely dωi dt dσi dt = Fω(ρi, ωi, σi) + Dω ∆ ωi , = Fσ(ρi, ωi, σi) + Dσ ∆ σi . (1a) (1b) In each of these equations the first term on the right-hand side describes the water dynamics within a cell. These are captured by the functions Fω and Fσ, to be specified below (see Eqs. (3) and (4)). The second term in each of the two equations describes spatial transport processes, that we model here as standard diffusive processes by using the discrete diffusion operator ∆ωi = ∆σi = 1 h2 X h2 X j∈N (i) j∈N (i) 1 (ωj − ωi) , (σj − σi) . (2a) (2b) Here the sum is over all elements j ∈ N (i), i.e. the set of cells j which are nearest neighbours of cell i. The co- efficients Dω and Dσ of the diffusion operator in Eqs. (1) are the diffusion constants corresponding to the transport of soil and surface water, respectively. This is usually a good approximation (see e.g. Rietkerk et al., 2002) of the more accurate description derived from shallow water the- ory (Gilad et al., 2004; Meron, 2011). Indeed, recent work (van der Stelt et al., 2013) suggests that there is no qual- itative difference between these two descriptions. Following HilleRisLambers et al. (2001) we define the functions Fω and Fσ, describing the on-site water dynam- ics, as Fω(ρ, ω, σ) = α(ρ) σ − β(ω) ρ − r ω , Fσ(ρ, ω, σ) = R − α(ρ) σ . (3a) (3b) Here, α(ρ) = a ρ + k2W0 ρ + k2 , β(ω) = b ω ω + k1 , (4) describe the infiltration rate of surface water into the soil, and the soil water uptake due to plants, respectively. These rates saturate for large values of ρ and ω, respec- tively. The model parameters k1 and k2 determine the exact shape of the saturation curves and are known as ‘half-saturation’ constants (HilleRisLambers et al., 2001). The dependence of α on the biomass density, ρ, introduces the infiltration-water-vegetation feedback, known to gen- erate spatial vegetation patterns in existing deterministic 3 models (e.g. Rietkerk et al., 2002). The parameter r char- acterises the loss of soil water due to evaporation (Eq. (3a)), while the parameter R in Eq. (3b) is the rainfall rate. The dynamics of water are illustrated in Fig. 1, see the left-most pair of cells (A) and the pair of cells in the middle (B). 2.1.3. Plant dynamics We model individual plants with an IBM. Individual plants Ii in cell i are represented as discrete entities whose dynamics are given by the stochastic transition rules Γb(ωi) −−−−→ 2Ii , Γd−→ Ei, Γs(ωj ) −−−−→ Ii + Ij , Ii Ii Ii (5a) (5b) (5c) j ∈ N (i) . Here the symbols over the arrows refer to the probabilities per unit time, or rates, of the corresponding transitions. These rates may depend on the amount of available soil water ω (see Eqs. (6) below). The first transition rule, Eq. (5a), corresponds to an individual plant Ii in cell i giving birth to another plant in the same cell i at a rate Γb(ωi). The second rule, Eq. (5b), refers to the death of a plant in cell i. The quantity Ei on the right-hand side of this reaction stands for a vacancy in cell i, so that this transition rule indicates that an individual plant Ii in a cell i dies at a rate Γd and leaves behind an empty place Ei in cell i. The third rule, Eq. (5c), describes a process in which an individual plant in cell i gives birth to another plant in a neighbouring cell j at a rate Γs(ωj). This captures the dispersal of the seeds of an individual plant to neighbouring cells and includes the probability of germination of the seedling. Notice that the rate of such dispersion processes depends on the amount of soil water, ωj, in the cell in which the new plant is born (see Eq. (6) below). We define the transition rates in analogy with previous deterministic models, e.g. HilleRisLambers et al. (2001), Rietkerk et al. (2002), and Pueyo et al. (2008). Specifi- cally, we use Γd = d , Γb(ωi) = ec β(ωi), Γs(ωj) = K ec β(ωj) . (6a) (6b) (6c) The plant birth rate Γb is proportional to the plant water uptake β (Eq. (4)), and ec is a constant that characterises the conversion rate at which water uptake is turned into newly established plants (Eq. (6a)). The death rate Γd is assumed to be a constant d (Eq. (6b)). The rate Γs with which new plants are established in neighbouring cells is proportional to the water uptake β, to the parameter ec and to the constant K, representing the probability that a seed falls into a neighbouring cell and manages to survive (Eq. (6c)). As described above the water uptake rate β, 4 and thus Γs, depend on the soil water concentration in the colonised cell (ωj). This reflects the fact that the probabil- ity of germination of the seedling in the neighbouring cells is a function of the local availability of soil water. In our simple model we assume that plants can colonise only near- est neighbour cells (see e.g. K´efi et al., 2007b). This can be generalised to take into account longer-range dispersal kernels, which may be appropriate when the characteristic scale of seed dispersal is appreciably greater than h, the lateral extension of a cell (Pueyo et al., 2008). It is worth pointing out that the transition rates do not depend on the whole history of the system, but only on its current state. This type of stochastic hybrid model is usually re- ferred to as a piecewise deterministic Markov process in the literature (Davis, 1984; Faggionato et al., 2009, 2010; Realpe-Gomez et al., 2012). 2.1.4. Coarse-graining: cell dynamics Given that we neglect the spatial structure within a cell, the relevant transition rules involve the state of cells rather than that of particular individual plants, so our model is of an effective nature; it provides a coarse grained description of the underlying dynamics. According to the rules for individual plants, Eq. (5), the only possible transitions for a cell i with ni plants are ni → ni ± 1, i.e. the birth or death of a plant in cell i. Hence, the possible transitions that can occur in cell i are fully specified by the expressions (see e.g. Black and McKane (2012) for similar models) Tb(ni + 1ni; ωi) = ni Γb(ωi), Td(ni − 1ni) = ni Γd, (7a) (7b) T i→j s (ni, nj + 1ni, nj; ωj) = ni Γs(ωj), j ∈ N (i), (7c) for all i. The notation Tb(ni+1ni; ωi), for instance, stands for the probability per unit time (or rate) that the num- ber of plants ni in a cell i increases by one, given that the amount of soil water in i is ωi at the time the transition takes place. Since each plant within a cell i has the po- tential to give birth to a new plant in i with a rate Γb(ωi) independently of all other plants, the total rate for a tran- sition ni → ni + 1 in cell i is given by ni Γb(ωi). Similar considerations apply for the other two processes described in Eq. (7) above. Notice that the arguments before the vertical bar in the above rate functions, Eqs. (7), refer to the state of the system after the transition, while those to the right of the vertical bar refer to the state before the transition, following the standard notation for condi- tional probabilities. This stochastic dynamics of cells is illustrated in Fig. 1, see the pair of cells in the middle (B) and the three right-most pairs of cells (C, D, E). The cells at the top (C) correspond to an on-site birth event, those in the middle (D) correspond to a death event and those at the bottom (E) correspond to a birth event induced by a neighbouring cell. 2.2. Deterministic approximation 2.2.1. General considerations The dynamics of plants in our model are stochastic, so we cannot predict the exact state of the system at a future time. At best, it might be possible to obtain the probabil- ity of finding the system in a given state, specified by the number of plants, ni, and the soil and surface water depths ωi and σi, in all cells i. From such probability distributions it would then be possible to compute the expected aver- age behaviour. Another way to study stochastic systems is to perform a certain number S of independent stochas- tic simulations. In doing so we could, for each cell i and for a given time t, estimate the average density of biomass or the average depths of soil and surface water. We will denote these quantities by Pi, Wi and Oi, respectively. The transition rules governing the dynamics of plants (Eqs. (6)-(7)) along with the dynamical laws for the soil and surface water (Eqs. (1)-(4)) induce a corresponding deterministic equation for the dynamics of these averages in the limit of a very large number of simulations, formally S → ∞. At the same time this deterministic approxima- tion also describes the behaviour of the stochastic model for very small µ. The equivalence is formally exact in the limit µ → 0 (cf. Realpe-Gomez et al., 2012). In this sense µ controls the strength of the noise in the stochastic model: if µ → 0 the noise is irrelevant and the dynamics becomes deterministic. Since ρi = µni the number of plants tends to be very large (ni → ∞) in this limit in regions with a non-zero density of biomass, ρi > 0. We also refer to Realpe-Gomez et al. (2012) for more details of a non- spatial version of the model; see also Black and McKane (2012) for a general review of individual-based models and the connection with deterministic limiting descriptions. The deterministic model governing the average be- haviour of the stochastic model turns out to be sim- ilar to models based on partial differential equations, well known in the literature (HilleRisLambers et al., 2001; Rietkerk et al., 2002; Pueyo et al., 2008; Dakos et al., 2011). The average biomass density, Pi, and the average depths of soil and surface water, Wi and Oi, respectively, follow the equations dPi dt dWi dt dOi dt = Fρ(Pi, Wi, Oi) + Dρ (Wi) ∆Pi, = Fω(Pi, Wi, Oi) + Dω ∆ Wi , = Fσ(Pi, Wi, Oi) + Dσ ∆ Oi . (8a) (8b) (8c) Here Fω and Fσ are given by Eqs. (3)-(4) and Fρ(P, W, O) = cβ (W ) P − dP, (9) where we have defined a rescaled conversion rate c = (1 + z K)ec, with z the number of nearest neighbours of a cell, e.g. z = 2 in one dimension and z = 4 for two di- mensions. The diffusion operator is given by Eq. (2) as be- fore, and so Eq. (8a) above is effectively a diffusion equa- tion with a diffusion coefficient Dρ(Wi) = h2 K ec β (Wi). 5 The biomass diffusion coefficient thus depends on the av- erage amount of soil water in the cell under considera- tion. The reason for this water dependence is that the ‘diffusion’ of biomass occurs when a plant in a given cell j produces a newly established plant in a neighbouring cell (i ∈ ∂j). The assumption underlying our model is that the establishment of a new plant depends on the amount of water available in the cell in which the new plant is es- tablished. Indeed, a water-dependent diffusion coefficient is also implicit in the work by Pueyo et al. (2008). In the special case in which the range of the dispersal ker- nel is so short that only dispersal to neighbouring cells occurs, Eqs. (8) above coincide with the equations ob- tained in Pueyo et al. (2008). On the other hand, such a water dependence of the biomass ‘diffusion’ coefficient constitutes the only difference between the deterministic equations (8), and a discretisation of the set of equations introduced by HilleRisLambers et al. (2001). 2.2.2. Homogeneous fixed points The deterministic equations introduced above, Eq. (8), i.e. equilibrium states, have homogeneous fixed points, such that Pi = P , Wi = W and Oi = O, for all i, and dP dt = 0, dW dt = 0, dO dt = 0. (10) Such fixed points may or may not be stable under small spatially homogeneous perturbations. Depending on the choice of parameters there can be either one out of two possible stable fixed points (Pueyo et al., 2008) or none (Realpe-Gomez et al., 2012). The two types of stable fixed points correspond to a bare soil state (also referred to as a ‘desert state’ in the following) Pd = 0, Wd = R r , Od = R a W0 , (11) or to a state with non-zero vegetation Pv = c R d − r c k1 c b − d , Wv = d k1 c b − d , Ov = R α(Pv) . (12) In the region where there are no stable fixed points, a nu- merical integration of the corresponding non-spatial equa- tions shows the existence of homogeneous limit cycles, as it has been reported recently by Realpe-Gomez et al. (2012). These oscillations become unstable once spatial structure is taken into account giving way to vegetation patterns as discussed below. In this work we will not focus on this pa- rameter regime since we are interested in a regime where the deterministic dynamics can lead to extinction (see be- low). 2.2.3. Spatial patterns and resilience Besides the homogeneous states given by Eqs. (11)-(12) above, the deterministic system defined by Eqs. (8) can also display stationary spatial patterns (Rietkerk et al., 2002). This is not surprising, given the similarity of Eqs. (8) to those investigated by HilleRisLambers et al. (2001) and Pueyo et al. (2008), which are known to display spa- tially patterned solutions. For certain values of the pa- rameters, these patterns can ‘emerge’ out of the homoge- neous vegetated state given by Eq. (12) when a small heterogeneous perturbation is applied (Rietkerk et al., 2002). This results in Turing patterns (Turing, 1952; Cross and Greenside, 2009) and can be studied mathemat- ically within a linear approximation of Eqs. (8) around the fixed point with homogeneous vegetation. However, numerical integration of the full set of nonlinear equations has revealed that spatial patterns exist in a region of pa- rameters much wider than that explained by such a lin- ear analysis (Rietkerk et al., 2002). These non-trivial so- lutions are inherently due to nonlinearities and so they cannot be captured in a linear approximation. In a pa- rameter range relevant for dryland ecology (i.e. at low rainfall values) a relatively large region of bistability be- tween spatial patterns and the desert state exists. In the region we investigate here the deterministic system can converge either to the desert state given by Eq. (11) or to a stable vegetation pattern, depending on the initial con- ditions from which the dynamics are started. The basin of attraction of the state with vegetation patterns, i.e. the set of all initial conditions which lead to such a state, can be used to characterise the resilience of the deterministic model: the larger the size of the basin of attraction of a ‘desirable’ state the more resilient is the system (Holling, 1973; Grimm and Calabrese, 2011; Martin and Calabrese, 2011). Performing such a study quantitatively is not an easy task given that the space of all possible initial con- ditions is huge. We will therefore restrict our work to an analysis involving only one type of realistic initial condi- tions. This will be discussed in more detail below. 3. Analysis All results discussed in this paper were obtained from spatially explicit numerical simulations, mostly in one di- mension, with a few two-dimensional examples for illus- trative purposes. We used periodic boundary conditions throughout. In order to obtain sample simulations for the stochastic model we used an adaptation of an algorithm due to Gillespie (1976) (see also Gillespie, 1977), details are discussed in Realpe-Gomez et al. (2012). To obtain solutions of the deterministic approximation, Eq. (8), we integrated these equations numerically using the Euler- forward method with a time step of ∆t = 0.01 days. In all cases we used the same initial conditions for both the stochastic model and its deterministic approximation obtained as follows: (a) soil water ωi, and surface water σi in all cells i were assigned the values corresponding to the homogeneous desert state given by Wd and Od in Eq. (11); (b) biomass density was assigned a value ρ0 > 0 in a fraction f of the cells picked at random, and a value of zero in the remaining fraction (1 − f ) of cells. We emphasise that in each run the random choice of populated cells is the same in both models. Notice that ρ0, along with the parameter µ, fixes the initial number of individual plants per initial populated cell n0 = ρ0/µ, which has to be an integer. The possible choices of ρ0 and µ are therefore constrained. Before continuing we would like to comment on why we have chosen this particular type of initial conditions here. The main reason for this is that we are dealing with spatial models and to define a generic initial condition we would need a very large number of parameters, i.e. initial biomass, soil and surface water in each cell. It is to avoid this that we decided to focus here on a subset of initial conditions that can be easily specified by a few parameters, as we have explained above. First we analysed the differences between the determin- istic and the stochastic approach comparing the behaviour of the system in single runs. Then, in order to study how generic the observed outcomes were, we estimated the probability of extinction, Pext, in both the stochastic and the deterministic models as a function of the differ- ent model parameters. This probability was obtained from simulations of the corresponding dynamics for several dif- ferent initial conditions, picked at random as described above, and counting the fraction of samples in which the systems became extinct. In the case of the deterministic system the only difference between simulation runs is the randomly chosen initial spatial distribution of the vege- tated cells. Once the initial condition is fixed, the sub- sequent dynamics is fully deterministic, with no further random elements. Similarly, the simulation runs of the stochastic system also involved a random element due to the initial conditions, but further stochasticity then en- tered during the actual run due to the random processes determining the sequence and timing of the transitions of the individual-based model. We estimated the probability of extinction in terms of what we expect to be the relevant parameters of the set of initial conditions: (a) the initial cover f , i.e. the frac- tion of cells with ρ0 > 0 in the initial conditions; (b) the initial amount of biomass ρ0 introduced in those cells; (c) the mass of a plant per unit area µ; and (d) the rainfall rate R. We chose f and ρ0 as relevant parameters because they determine the amount of initial biomass, µ because it controls the strength of the stochasticity in the model and R in order to investigate the models under different conditions of water availability. In the limit µ → 0, the stochastic results are expected to show similar behaviour to the deterministic approximation, while for larger values of µ we would expect stochastic effects to be more com- mon. To estimate the probabilities of extinction, 50 simula- tions were run for each model and for each set of parame- ters. Each of these simulations was run for a period of time of up to T = 5000 days. Simulation runs of the stochas- tic model were identified as becoming extinct if all cells i reached a state with ni = 0 within this time. In the deter- ministic model we applied a threshold criterion Pi < ǫ, i.e. 6 a given cell is assumed to contain no plants when its plant density is below a set threshold ε. In order to be consistent with the original stochastic model we chose ǫ = µ, as we assumed that µ was the contribution of a single plant to the biomass density. The introduction of a threshold was necessary because the deterministic approximation works with a genuinely continuous density, and the state with exactly zero biomass is reached only asymptotically at in- finite times. We carried out some consistency checks, and observe that results depend little on the exact choice of the value of the threshold. Findings remained essentially the same even for ǫ much smaller than µ. 4. Results 4.1. Dynamics We compared the dynamics of the stochastic model with the dynamics of the corresponding deterministic approx- imation. Both models were initialised with homogeneous initial conditions (i.e. f = 1) with ρ0 = 10 g m−2. For all cells, soil and surface water took values corresponding to the desert state, given by Eq. (11). We compared the time dynamics of the total biomass density for both mod- els. As expected from the stability analysis of the deter- ministic non-spatial model, we observed that in this case the deterministic approximation converges asymptotically to the desert state (Fig. 2, red thick line). In contrast the stochastic model did not reach extinction for the same homogeneous initial conditions and parameter values. The stochastic model initially followed a path close to the de- terministic approximation and approached extinction, but then some regions displayed an explosive growth and the system finally was found in a quasi-stationary pattern. As expected the stochastic dynamics deviated substantially from its deterministic approximation especially when the number of plants in the system was small (see Fig. 2). We will refer to this effect of escaping a path to extinc- tion as ‘self-recovery’. From a mathematical perspective this phenomenon appears similar to the one investigated in Black and McKane (2011). For clarity the figure shows the results up to time t = 500 days even though we ran the simulations for much longer (up to time T = 5 × 104 days). We observed that the total biomass in the stochastic model remains essentially the same up to stochastic deviations, i.e. it reaches a stationary state. Still there was a nonzero probability that the stochastic model, after having escaped the initial path to extinction, e.g. after, say, t ≈ 300 days in Fig. 2, reached extinction. Such events occur because of large stochastic deviations, and the probability of such events can be expected to be rather small (Frey, 2010). The phenomenon of self-recovery in the stochastic model can also be seen in a comparison of the time-dependent spatial biomass density profile along the line of cells in both models, see Fig. 3. Initial conditions for the stochas- tic and the deterministic models were chosen as f = 1/2 and ρ0 = 10 g m−2. For all cells the soil and surface water 7 densities were initialised from the values of the desert state given by Eq. (11). In Fig. 3 we compare the dynamics of the vegetation profile in the stochastic model with that of the corresponding deterministic approximation. Again, we can see that the deterministic approximation leads to extinction while the stochastic model recovers. If we look at the stochastic simulation for a longer time (5×104 days) we clearly see that vegetation persists, see Fig. 4 below the horizontal dashed line. We have also used the last configu- ration reached by the stochastic model (horizontal dashed line in Fig. 4) as initial conditions for the deterministic model; the corresponding time dynamics is shown in Fig. 4, above the horizontal dashed line. The patterns also re- main stable under the deterministic dynamics, but it was demographic stochasticity which allowed the system to es- cape extinction in the first place; running the deterministic dynamics alone leads to extinction. The stochastic system thus generated suitable initial conditions for the determin- istic system to converge to a spatial pattern. In Fig. 4 one can also observe that the patches reached by the deter- ministic model are fully frozen at long times while those generated by the stochastic model remain dynamic, they can split, merge or become extinct. Similar observations can also be made in the two- dimensional system, as we show in Fig. 5 and in the en- closed supplementary video animation. As expected, both the stochastic and the deterministic models display spatial vegetation patterns, similar to those observed previously with other models, and in an analogous range of rainfall values (HilleRisLambers et al., 2001; Rietkerk et al., 2002; Pueyo et al., 2008). Finally we notice that in many simulations the stochas- tic system follows a regime of ‘explosive’ local growth soon after it has escaped the path to extinction. During this relatively short period of time some cells can get to carry a relatively large number of individuals with a maximum of about 120 plants observed in simulations. Afterwards plants spread out in space to a certain extent, and even- tually the system appears to reach a stochastic stationary state (see supplementary video). 4.2. Probability of extinction Next we will present the results of a more systematic in- vestigation of the effect of ‘self-recovery’. Figure 6 shows the probability of extinction Pext as a function of f , for two different values of ρ0 (10 g m−2 and 50 g m−2). We observe that the stochastic model almost never reached ex- tinction, while the deterministic approximation did so in almost all samples with f > 0.4. These observations may appear counter-intuitive at first sight: in the deterministic model we find that the more plants are in the system ini- tially, i.e. the larger the initial cover f , the more likely it is that the system reaches extinction. However, this effect is already seen in the deterministic non-spatial model which predicts the extinction of homogeneous initial configura- tions (i.e. f = 1). The spatial model instead predicts sta- ble patterns (HilleRisLambers et al., 2001; Rietkerk et al., 2002; Pueyo et al., 2008). Even more surprisingly perhaps, is that the larger the initial amount of biomass in each populated cell, ρ0, the larger the probability of extinction tends to be. To be more precise, this only happens if the initial cover is not too small (f & 0.1). From the simula- tions, it seems that if the amount of biomass in a cell is too large, a fast spatial spread of plants is triggered, which favours a more homogeneous distribution of biomass. This amounts to an effective increase of the initial cover, f , and thus the previous effect takes over. This behaviour appears to run contrary to the infiltration feedback mechanism for pattern formation: local biomass increase favours infiltra- tion in detriment of water availability in the surroundings. However, a similar effect already exists in the determin- istic model as well. Indeed, in the region of parameter values that we considered here (see Tab. 1), vegetation patterns are not the only stable stationary solutions to the deterministic system: homogeneous desert is also one such solutions; in other words, there is bistability between vegetation patterns and desert. This suggests that, in this regime, the infiltration feedback mechanism is not always effective in inducing the formation of patterns. To de- velop some intuition of the kind of processes that could lead to these phenomena, it is useful to imagine an extreme case with relatively large total reproduction rate and a not so large average transport of water. We can then expect that plants manage to spread quickly before water hetero- geneities consolidate. This might lead to a rather homo- geneous state with a behaviour similar to what one would see in the non-spatial model in this regime: all plants die together. Although this situation may not be realistic, it indicates that the infiltration feedback mechanism may break down at some point, leading the system to a desert state rather than to pattern formation. This appears to be consistent with our simulations, in which we find that a larger initial cover or a larger value of ρ0 imply a larger total reproduction rate. We study the variations of the probability of extinction Pext as a function of µ for three different values of f : 1/8, 1/2, and 7/8 in Fig. 7. We can observe that Pext ≈ 0 for almost all values of µ except for µ = 0.1 g m−2 and µ > 6 g m−2, which correspond to very small and large plant biomass, respectively. Next, we study the probability of extinction Pext as a function of the rainfall rate R, for a range corresponding to typical dryland values, using three different values for µ: 1 g m−2, 5 g m−2 and 10 g m−2 (Fig. 8). In the stochastic model, the effect of ‘self-recovery’, or escaping the path to extinction, is more appreciable the larger the amount of rainfall. The corresponding probability of extinction for the deterministic system is equal to one in almost the whole regime investigated (not shown). Under certain conditions, the opposite effect can hap- pen, with demographic noise promoting extinction in cases in which the deterministic system survives instead. How- ever, this needs a range of the parameter µ which we expect not to be relevant for real-world situations. In particular, we have observed such behaviour only for large values of the parameter µ and for low initial cover. More specifi- cally, we have observed that this can happen if f . 10% and µ & 10 g m−2, for R & 0.5 mm d−1, or µ & 5 g m−2, for 0.4 . R . 0.5 mm d−1. 5. Discussion Two general approaches to modelling ecosystems, and complex systems in general, are continuous models and in- dividual based models. The former tends to be relatively simple and amenable to mathematical analysis, while the latter tends to be more realistic and rely heavily on compu- tational simulations due to its usually higher complexity. In particular, individual based models of ecosystems are mainly stochastic reflecting the random behaviour com- monly observed in natural systems. On the other hand, continuous models are mostly deterministic or include a noise term added ad hoc. In some circumstances these can be seen as describing the most relevant features of indi- vidual based models. However, it is well known that de- mographic stochasticity can have important consequences that are usually neglected when modelling ecosystems in terms of continuous densities (see e.g. Black and McKane, 2012). In this work, we represented semi-arid ecosystems with two parallel models: a hybrid stochastic model, in- cluding individual-based vegetation dynamics and deter- ministic water dynamics, and a deterministic model, gov- erning the behaviour of the average stochastic variables. The models represented explicitly the water-vegetation in- filtration feedback. Thus, for the parameter values we studied, both models contained solutions corresponding to stable vegetation patterns, analogously to what was ob- served previously with other models of this type of ecosys- tem (e.g. Rietkerk et al., 2002). While continuous models assume that stochastic effects are irrelevant and so neglect them altogether, we have also made two assumptions to simplify the computational anal- ysis, i.e. we have assumed the ecosystem is not very sensi- tive to the growth phase of each individual plant so we can neglect it, and we have taken the same mass for all plants. We expect this to be a reasonable approximation if the time it takes for a plant to establish is much smaller than the time scale on which there is an appreciable change of the spatial distribution of biomass, e.g. the characteristic time for pattern formation. Conditioned on the validity of these assumptions, our analysis showed that the resilience of the vegetation patterns changed when taking into account the discrete nature of plants and the intrinsic stochasticity of their behaviour. Including such stochasticity, the modelled ecosystems did not turn into deserts in a wide range of cases in which the deterministic representation would predict the ecosystem to become extinct. This is an important observation, given that semi-arid ecosystems are characterised by a rather scarce number of plants, 8 scattered across regions of empty land. In a finite pop- ulation, the number of individuals varies because of the intrinsic stochastic behaviour of the individuals. Usually, the fewer the individuals, the stronger the effects of the de- mographic noise (Ramaswamy et al., 2012; Rogers et al., 2011; Butler and Goldenfeld, 2012; Biancalani et al., 2010; Butler and Goldenfeld, 2011; Biancalani et al., 2009; McKane and Newman, 2005). Intuitively, this is expected to promote extinction when the number of plants is small. It is therefore remarkable that we ob- served a relevant regime of parameters in which including the stochastic individual nature of the plants actually increased the likelihood of vegetation pattern emergence, and of escaping the desert state. The opposite might also occur, i.e. a larger probability of becoming extinct in the stochastic model, especially when considering large individual plants, covering initially only a small fraction of soil. Under these conditions, the probability of a large stochastic deviation that brings the system suddenly to extinction, could be non-negligible. However, we expect that our approach is better justified for intermediate values of µ, i.e. the biomass of a plant (per unit area), where the results were rather robust. Stochastic be- haviour was not observed when µ was too small, in which case our model coincided with a deterministic continuous model, whereas if µ was too large the growth phase and heterogeneity was more noticeable. When the individual plant biomass was relatively high (µ & 8 g m−2), or very low (µ → 0), the outcomes of the stochastic model tended to have higher probability of being desert than for intermediate µ values. Demographic noise seemed to be more important in an intermediate range of single-plant biomass, which corresponds to a realistic range for herbs or grass biomass (see e.g. Peters, 2002). In order to study the resilience of the vegetation pat- terns, we addressed the issue of how and how easily they were reached in time, i.e. for which type of initial conditions the system evolved towards the pattern (e.g. Eppinga, 2009). This gave an indication of how “attrac- tive” the patterns were. In the case of the deterministic approximation, this was equivalent to studying the basin of attraction of the patterned stable states. This notion, how- ever, is not applicable to stochastic systems, where there is no notion of equilibrium point, and no deterministic dy- namics uniquely leading from a certain initial condition to a final state. For this reason, we focused on the probability of extinction as a function of the initial conditions and for different parameter values. For a large set of initial con- ditions and parameters, we observed that the stochastic model almost never led the vegetation to extinction, while the vegetation in the deterministic model almost always became extinct. When varying rainfall in a realistic range for drylands, the effect of ‘self-recovery’ in the stochastic model was promoted for the highest rainfall values, where the contrast with the deterministic case was sharper. Given the success of deterministic continuous models to date, demographic noise is expected to play a relevant role only when the number of plants is rather low. Indeed we have observed (see e.g. Fig. 2) that initially the two mod- els follow essentially the same dynamics towards deserti- fication, and only when they are close to extinction the two dynamics diverge: the deterministic dynamics follows the path to extinction, while the stochastic one recovers. This particular difference between the two models was es- pecially important for vegetation initially covering more than 30-40%, where the deterministic case would practi- cally always become extinct. We should be careful, how- ever, not to give too much weight to the importance of the initial conditions, since they were arbitrarily chosen in simulations. The most robust statements of this work had to do with comparing what happened with both the de- terministic and stochastic models after they had followed their own dynamics for a while. Nevertheless, such a be- haviour of the deterministic model might appear counter- intuitive, in the sense that it might be concluded that a semi-arid ecosystem with many plants, e.g. due to plant- ing, risks extinction. The stochastic model, on the other hand, did not show this counter-intuitive behaviour since the probability of extinction was equally small (less than 0.1) for all values of initial cover. In the case of the deter- ministic model, the more homogeneously distributed the plants were, the more difficult it was for any particular vegetation patch to actively increase its own soil water. The water-vegetation feedback, in this case, needed con- trasts between vegetated and bare patches to take place effectively. For the same reasons, in this regime the homo- geneous vegetated state was not stable, and full homoge- neous vegetation cover would quickly evolve into a desert state. Not so for the stochastic model, though, which could escape extinction even in this case. On the other hand, for very low initial vegetation cover (less than 20-30%), the results of the deterministic and the stochastic approaches were very similar. We must underline that these low frac- tions of vegetation cover are quite realistic in the most arid ecosystems. One may wonder how it is that for large initial cover (e.g. larger than 40%) the deterministic sys- tem has a large probability of extinction, while for small cover (e.g. smaller than 20%) it almost always survives; after all the dynamics is continuous and to reach extinc- tion the system first needs to go through states of small cover. However, the spatial structure of the intermediate states thus reached does not necessarily coincide with that of a state with homogeneous distribution of soil and sur- face water and a random distribution of constant biomass ρ0, as were used here as initial conditions. This stresses the importance of being careful about interpreting our par- ticular choice of initial conditions, as mentioned above. The deterministic model we investigated in this work describes exactly the typical behaviour of the stochastic system when the individual plant biomass was negligible (µ → 0). In this case we could have a very large num- ber of plants per cell. This deterministic model was a good approximation to the kernel model investigated by Pueyo et al. (2008) in the case that the range of the ker- 9 nel of seed dispersal in the latter was of the order of one or two cells (∼ 2-4 m). Under these conditions, the kernel could be approximated with an effective diffusive term and with a diffusion coefficient depending on the state of the soil water. If, additionally, we could neglect the spatial variation of the soil water, the two models became similar to the model studied by HilleRisLambers et al. (2001). In this sense, we could say that the stochastic model intro- duced here was close to these well-studied models and, in particular, it was not unexpected to observe spatial pat- terns in a similar range of parameters. What could change drastically is the transitory dynamics while reaching a sta- tionary state, as we showed with the effect of self-recovery. In principle, a more complete approach would have to deal with each single plant individually, with all its at- tributes and ongoing processes, as has been done for in- stance in the study of forests, see e.g. Perry and Enright (2006). However, this would make the problem far more complex from a computational point of view, and it might become difficult to gain insight. Indeed, as has been dis- covered in the investigations on forests, a far simpler an- alytical approximation, called the perfect plastic approxi- mation, may capture most of the relevant details of the dy- namics observed in simulations (Strigul et al., 2008). This case, however, corresponds to a regime of high vegetation density, where fluctuations of the average behaviour are ex- pected to be irrelevant. This raises the question of which would be the ‘best’ way of modelling a plant in the study of semi-arid ecosystems in the sense of, paraphrasing Ein- stein, ‘keeping it simple, but not too simple’. In our stochastic modelling approach, we did not con- sider the heterogeneity in plant size. In particular, we did not include the different plant life stages. In a sense, plants were instantaneously created as adult individuals. This might appear as an unrealistic feature that might promote the effect of ‘self-recovery’ we discussed, because a plant could start increasing the availability of water locally as soon as it was created, and therefore instantly start pro- moting its own survival. However, this feature would also lead to an overestimation of water uptake. Since in our model plants died suddenly, in detriment of self-recovery, mortality was also over-estimated. Furthermore, in the stochastic model we investigated, a plant could produce another plant only in the neighbouring cells, thus limiting the impact the new birth had on the state of the ecosys- tem. A next step in the complexity of the modelling could seedlings and be to introduce two type of individuals: established plants. In this way we would be able to take into account, for instance, the high asymmetry in mortality between these two. The question is then how robust are the results we have discussed in this work under this more realistic scenario. At first sight, one could expect that the stochastic model becomes closer to the deterministic approximation, but experience in this field of research has shown that counter-intuitive effects are not uncommon (Ramaswamy et al., 2012; Rogers et al., 2012; Biancalani et al., 2011; Biancalani et al., 2009; McKane and Newman, 2005). 2011; Butler and Goldenfeld, 2010; Butler and Goldenfeld, The model we presented did not include environmen- tal heterogeneity and stochasticity, or topography (see e.g. Sheffer et al., 2013). We also discarded rainfall in- termittence. Vegetation in arid areas is well adapted to the occasional occurrence of rainfall (Baudena et al., 2007; Kletter et al., 2009; D’Odorico et al., 2007), al- though the effect of rainfall intermittency may not be too relevant when spatial feedbacks are represented (see Baudena and Provenzale, 2008). Besides water, we did not consider any other limiting resource, such as nutri- ent or light limitations. Moreover, we did not include another water-vegetation feedback mechanism, which is known to play a role in drylands, namely the effect of root length. Plant water availability increases with the root extent, which in turn increases with the biomass it- self, thus favouring plant growth and generating vegeta- tion patterns (Gilad et al., 2004; Lefever, R. and Lejeune, 1997; Barbier et al., 2008). Another relevant issue is how to validate our findings with observations. For example, we could analyse patch dynamics to see whether it displays the ‘wiggly’ behaviour observed in the model results. This nevertheless would re- quire time series of spatial patterns, with an appropriate spatial and temporal resolution, which may not be easily obtainable. In principle, it could even be possible to com- pute the statistics of this patch dynamics, in order to have quantitative predictions. Despite not being conclusive in any sense, this inves- tigation indicated that, in certain regimes, including de- mographic noise could lead to a larger estimate of the re- silience of semi-arid ecosystems. Our model results sug- gested that demographic noise may be more important in the less arid ecosystems, with higher rainfall and veg- etation cover. Since changes in rainfall regimes are ex- pected, for example as a consequence of climate change, it may be necessary to take into account individual-based dynamics to evaluate the resilience and resistance of these ecosystems to such forcing. In summary, we think the study of semi-arid ecosystems might benefit from the ap- proach taken for instance in the research on forests, where quite detailed IBM’s have been extensively used. Indeed, in contrast to forests which are characterised by a rather dense vegetation, the typical number of plants in semi-arid ecosystems is comparatively quite low and so the stochas- tic effects implicit in such a modelling approach are ex- pected to be more relevant. Acknowledgements The authors thank M. Eppinga for useful discussions. This research is supported by the ERA Complexity Net through the RESINEE project (‘Resilience and interaction of networks in ecology and economics’). UK support for this project to JRG, TG and AJM is administered by the 10 Engineering and Physical Sciences Research Council EP- SRC (grant reference EP/I019200/1). TG acknowledges funding by RCUK (EP/E500048/1). The research of MB and MR is also supported by the project CASCADE (Sev- enth Framework Programme FP7/2007-2013 grant agree- ment 283068). 11 References Aguiar, M., Sala, O. E., 1999. Patch structure, dynamics and impli- cations for the functioning of arid ecosystems. Trends in Ecology & Evolution 14, 273–277. Barbier, N., Couteron, P., Lefever, R., Deblauwe, V., Lejeune, O., 2008. Spatial decoupling of facilitation and competition at the origin of gapped vegetation patterns. Ecology 89, 1521–31. Barbier, N., Couteron, P., Lejoly, J., Deblauwe, V., Lejeune, O., 2006. Self-organized vegetation patterning as a fingerprint of cli- mate and human impact on semi-arid ecosystems. Journal of Ecol- ogy 94, 537–547. Baudena, M., Boni, G., Ferraris, L., von Hardenberg, J., Proven- zale, A., 2007. Vegetation response to rainfall intermittency in drylands: Results from a simple ecohydrological box model. Ad- vances in Water Resources 30, 1320–1328. Baudena, M., Provenzale, A., 2008. Rainfall intermittency and veg- etation feedbacks in drylands. Hydrology and Earth System Sci- ences 12, 679–689. Baudena, M., Rietkerk, M., 2012. Complexity and coexistence in a simple spatial model for arid savanna ecosystems. Theoretical Ecology. Biancalani, T., Fanelli, D., Patti, F. D., 2010. Stochastic Turing patterns in a Brusselator model. Physical Review E 81, 046215. Biancalani, T., Galla, T., McKane, A. J., 2011. Stochastic waves in a Brusselator model with nonlocal interactions. Physical Review E 84, 026201. Black, A. J., McKane, A. J., 2011. WKB calculation of an epidemic outbreak distribution. Journal of Statistical Mechanics, P12006. Black, A. J., McKane, A. J., 2012. Stochastic formulation of ecologi- cal models and their applications. Trends in Ecology & Evolution 27, 337–345. Botkin, D. B., Janak, J. F., Wallis, J. R., 1972. Rationale, limita- tions, and assumptions of a northeastern forest growth simulator. IBM Journal of Research & Development 16, 101–116. Butler, T., Goldenfeld, N., 2009. Robust ecological pattern formation induced by demographic noise. Physical Review E 80, 030902(R). Butler, T., Goldenfeld, N., 2011. Fluctuation-driven Turing patterns. Physical Review E 84, 011112. Casenave, A., Valentin, C., 1992. A runoff capability classification system based on surface features criteria in semi-arid areas of West Africa. Journal of Hydrology 130, 231–249. Coffin, D., Lauenroth, W., 1990. A gap dynamics simulation model of succession in a semiarid grassland. Ecological Modelling 49, 229–266. Crawley, M. J., 1990. The population dynamics of plants. Phil. Trans. R. Soc. Lond. B 330, 125–140. Cross, M. C., Greenside, H. S., 2009. Pattern formation and dy- namics in non-equilibrium systems. Cambridge University Press, Cambridge. Dakos, V., K´efi, S., Rietkerk, M., van Nes, E. H., Scheffer, M., 2011. Slowing down in spatially patterned ecosystems at the brink of collapse. The American Naturalist 177, E155–E166. Davis, M. H. A., 1984. Piecewise-deterministic Markov processes: a general class of non-diffusion stochastic models (with discussion). J. R. Stat. Soc. B 46, 353–388. Deblauwe, V., Barbier, N., Couteron, P., Lejeune, O., Bogaert, J., 2008. The global biogeography of semi-arid periodic vegetation patterns. Global Ecology and Biogeography 17, 715–723. D’Odorico, P., Laio, F., Porporato, A., Ridolfi, L., Barbier, N., 2007. Noise-induced vegetation patterns in fire-prone savannas. Journal of Geophysical Research 112, G02021–G02021. Eppinga, M., 2009. Amazing pattern. Ph.D. thesis, Universiteit Utrecht. Faggionato, A., Gabrielli, D., Crivellari, M. R., 2009. Non- equilibrium thermodynamics of piecewise deterministic Markov processes. Journal of Statistical Physics 137, 259–304. Faggionato, A., Gabrielli, D., Crivellari, M. R., 2010. Averaging and large deviation principles for fully-coupled piecewise deteministic Markov processes and applications to molecular motors. Markov Processes Related Fields 16, 497–548. Frey, E., 2010. Evolutionary game theory: Theoretical concepts and applications to microbial communities. Physica A: Statistical Me- chanics and its Applications 389, 4265–4298, section 2.1. Gilad, E., von Hardenberg, J., Provenzale, A., Shachak, M., Meron, E., 2004. Ecosystem engineers: from pattern formation to habitat creation. Physical Review Letters 93, 98105. Gillespie, D. T., 1976. A general method for numerically simulat- ing the stochastic time evolution of coupled chemical reactions. Journal of Computational Physics 22, 403–434. Gillespie, D. T., 1977. Exact stochastic simulation of coupled chem- ical reactions. Journal of Physical Chemistry 81, 2340–2361. Grimm, V., Calabrese, J. M., 2011. What is resilience? A short introduction. In: Deffuant, G., Gilbert, N. (Eds.), Viability and resilience of complex systems: concepts, methods and case studies from ecology and society. Springer, Heidelberg, Ch. 1, p. 15. HilleRisLambers, R., Rietkerk, M., van den Bosch, F., Prins, H. H. T., de Kroon, H., 2001. Vegetation pattern formation in semi- arid grazing systems. Ecology 82, 50–61. Holling, C., 1973. Resilience and stability of ecological systems. An- nual Review of Ecology and Systematics 4, 1–23. Jørgensen, S. E. (Ed.), 2011. Handbook of Ecological Models used in Ecosystem and Environmental Management. CRC Press, Boca Raton. K´efi, S., Rietkerk, M., Alados, C. L., Pueyo, Y., Papanastasis, V. P., Elaich, A., de Ruiter, P. C., 2007a. Spatial vegetation patterns and imminent desertification in Mediterranean arid ecosystems. Nature 449, 213–217. K´efi, S., Rietkerk, M., van Baalen, M., Loreau, M., 2007b. Local facilitation, bistability and transitions in arid ecosystems. Theo- retical Population Biology 71, 367–379. Klausmeier, C., 1999. Regular and irregular patterns in semi-arid vegetation. Science 284, 1826–1828. Kletter, A. Y., von Hardenberg, J., Meron, E., Provenzale, A., 2009. Patterned vegetation and rainfall intermittency. Journal of Theo- retical Biology 256, 574–583. Lefever, R. and Lejeune, O., 1997. On the origin of tiger bush. Bul- letin of Mathematical Biology 59, 263–294. Manor, A., Shnerb, N. M., 2008. Facilitation, competition, and vege- tation patchiness: from scale free distribution to patterns. Journal of Theoretical Biology 253, 838–842. Martin, S., Calabrese, G. D. J. M., 2011. Defining resilience mathe- matically: from attractors to viability. In: Deffuant, G., Gilbert, N. (Eds.), Viability and resilience of complex systems: concepts, methods and case studies from ecology and society. Springer, Hei- delberg, Ch. 2, p. 3. McKane, A. J., Newman, T. J., 2005. Predator-prey cycles from res- onant amplification of demographic stochasticity. Physical Review Letters 94, 218102. Meron, E., 2012. Pattern-formation approach to modelling spatially extended ecosystems. Ecological Modelling 234, 70–82. Meron, E., 2011. Modeling dryland landscapes. Mathematical Mod- elling of Natural Phenomena 6, 163–187. Moorcroft, P. R., Hurtt, G. C., Pacala, S. W., 2001. A method for scaling vegetation dynamics: the ecosystem demography model (ED). Ecological Monographs 71, 557–586. Pacala, S. W., Canham, C. D., Saponara, J., Jr., J. A. S., Kobe, R. K., Ribbens, E., 1996. Forest models defined by field mea- surements: Estimation, error analysis and dynamics. Ecological Monographs 66, 1–43. Pacala, S. W., Canham, C. D., Silander Jr., J. A., 1993. Forest mod- els defined by field measurements: I. the design of a northeastern forest simulator. Canadian Journal of Forest Research 23, 1980– 1988. Perry, G. L. W., Enright, N. J., 2006. Spatial modelling of vege- tation change in dynamic landscapes: a review of methods and applications. Progress in Physical Geography 30, 47–72. Peters, D. P. C., 2002. Plant species dominance at a grassland- shrubland ecotone: an individual-based gap dynamics model of herbaceous and woody species. Ecological Modelling 152, 5–32. Pueyo, Y., K´efi, S., Alados, C. L., Rietkerk, M., 2008. Dispersal strategies and spatial organization of vegetation in arid ecosys- 12 tems. Oikos 117, 1522–1532. Ramaswamy, R., Gonzalez-Segredo, N., Sbalzarini, I. F., Grima, R., 2012. Discreteness-induced concentration inversion in mesoscopic chemical systems. Nature Communications 3, 779. Rastetter, E. B., Aber, J. D., Peters, D. P. C., Ojima, D. S., Burke, I. C., 2003. Using mechanistic models to scale ecological processes across space and time. BioScience 53, 68–76. Realpe-Gomez, J., Galla, T., McKane, A. J., 2012. Demographic noise and piecewise deterministic Markov processes. Physical Re- view E 86, 011137. Rietkerk, M., Boerlijst, M. C., van Langevelde, F., HilleRisLambers, R., van de Koppel, J., Kumar, L., Prins, H. H. T., de Roos, A. M., 2002. Self-organization of vegetation in arid ecosystems. American Naturalist 160, 524–530. Rietkerk, M., Dekker, S. C., de Ruiter, P. C., van de Koppel, J., 2004. Self-organized patchiness and catastrophic shifts in ecosystems. Science 305, 1926–1929. Rietkerk, M., van de Koppel, J., 1997. Alternate stable states and threshold effects in semi-arid grazing systems. Oikos 79, 69–76. Rietkerk, M., van de Koppel, J., 2008. Regular pattern formation in real ecosystems. Trends in Ecology & Evolution 23, 169–175. Rogers, T., McKane, A. J., Rossberg, A. G., 2012. Demographic noise can lead to the spontaneous formation of species. Euro- physics Letters 97, 40008. San Miguel, M., Johnson, J. H., Kertesz, J., Kaski, K., Daz-Guilera, A., MacKay, R.S., Loreto, V., rdi, P., Helbing, D., 2012. Chal- lenges in complex systems science. The European Physical Journal Special Topics 214, 245–271. Schenk, H. J., Espino, S., Goedhart, C. M., Nordenstahl, M., Mar- tinez Cabrera, H. I., Jones, C. S., 2008. Hydraulic integration and shrub growth form linked across continental aridity gradients. Pro- ceedings of the National Academy of Sciences of The United States of America 105, 11248–11253. Sheffer, E., von Hardenberg, J., Yizhaq, H., Shachak, M., Meron, E., 2013. Emerged or imposed: a theory on the role of physical templates and self-organisation for vegetation patchiness. Ecology Letters 16, 127–139. Shugart, H., 1984. A theory of forest dynamics. Springer-Verlag, New York. van der Stelt, S., Doelman, A., Hek, G., Rademacher, J. D. M., 2013. Rise and fall of periodic patterns for a generalized Klausmeier- Gray-Scott model. Journal of Nonlinear Science 23, 39–95. Strigul, N., Pristinski, D., Purves, D., Dushoff, J., Pacala, S., 2008. Scaling from trees to forests: tractable macroscopic equations for forest dynamics. Ecological Monographs 78, 523–545. Turing, A. M., 1952. The chemical basis of morphogenesis. Phil. Trans. R. Soc. B (London) 237, 37–72. Vincenot, C. E., Giannino, F., Rietkerk, M., Moriya, K., Mazzoleni, S., 2010. Theoretical considerations on the combined use of sys- tem dynamics and individual-based modeling in ecology. Ecologi- cal Modelling 222, 210–218. von Hardenberg, J., Meron, E., Shachak, M., Zarmi, Y., 2001. Di- versity of vegetation patterns and desertification. Physical Review Letters 87, 198101. Walker, B., Ludwig, D., Holling, C., Peterman, R., 1981. Stability of semi-arid savanna grazing systems. Journal of Ecology 69, 473– 498. 13 STOCHASTIC BIRTH OR DEATH OF PLANTS AT t + τ DETERMINISTIC EVOLUTION OF WATER BETWEEN t AND t + τ niΓb(ωi) Time t Time t + τ Eqs. (1)-(4) ni + 1 ωi, σi C ni nj ω ′ ′ i, σ i ω ′ ′ j , σ j A ni nj ωi, σi ωj , σj niΓd ni − 1 ωi, σi B D ni nj + 1 ωi, σi ωj , σj niΓs(ωj) E Figure 1: (Colour online) Illustration of the stochastic hybrid model for semi-arid ecosystems. For clarity we show only two neighbouring cells i and j but the same applies to all the other cells. Suppose that, at time t, the system is in a state where the number of plants and the depths of soil and surface water in cell i are given by ni, ω′ j . Suppose, furthermore, that the next birth or death of a plant takes place at time t + τ ; these events happen at random, and so τ is itself a stochastic variable. Since there are no transition events between t and t + τ , soil and surface water in all cells evolve deterministically according to Eqs. (1)-(4) in this time interval. Suppose now that their new state at t + τ is given by ωi and σi, for all cells i (B). At t + τ a stochastic transition happens, there are three possible types of transitions: a plant at a cell i gives birth to a new plant in the same cell (C), it dies (D) or it gives birth to a new plant in a neighbouring cell j (E). These transitions happen with rates niΓb(ωi), niΓd, and niΓs(ωj ), respectively (see Eqs. (6) and (7)). Immediately after t + τ , and until the next birth or death of a plant takes place, soil and surface water in all cells again evolve deterministically as before. i, respectively (A). The analogous quantities in cell j are nj , ω′ i and σ′ j, σ′ 14 ] 2 - m g [ l l e c r e p s s a m o b e g a r e v A i 25 20 15 10 5 0 Stochastic Deterministic 0 100 200 300 400 500 Time [d] Figure 2: (Colour online) Comparison of the dynamics of the to- tal biomass density in the stochastic model (continuous green line) with that in the corresponding deterministic approximation (thick red line). While the deterministic approximation leads to extinc- tion, the full stochastic model recovers. Simulations were run on a line with 128 cells, periodic boundary conditions and the same ran- dom initial conditions in all cells: soil and surface water depths were given by the values in the desert state, Eq. (11), while the initial biomass was ρ0 = 10 g m−2 (f = 1). Other key parameter values were: µ = 1 g m−2, R = 0.6 mm d−1. See Table 1 for the remaining parameter values. I C I T S I N M R E T E D C I T S A H C O T S ] d 4 0 1 x [ e m T i 7 6 5 4 3 2 1 0 90 60 ] 2 - m g [ s s a m o B i 30 0 0 128 256 m STOCHASTIC DETERMINISTIC 90 1000 ] d [ e m T i 500 Figure 4: (Colour online) Dynamics of the vegetation profile pre- sented on the left side of Fig. 3 (stochastic model) for a longer period of time. Below the horizontal dashed line, i.e. from t = 0 days to t = 5 × 104 days, we show the dynamical behaviour of the stochastic model, and observe that the system indeed escapes the path to extinction and finally reaches a stationary pattern. Vege- tation patches, however, appear to follow dynamics on their own: they can split, merge, and become extinct. In order to compare this with the dynamics of the deterministic model, we use the con- figuration reached by the stochastic dynamics at t = 5 × 104 days (horizontal dashed line) as an initial condition for the deterministic model. The outcome of the corresponding deterministic dynamics from t = 5 × 104 are shown above the horizontal dashed line, i.e. days to t = 7 × 104 days. In other words, at time t = 5 × 104 days we switch the dynamics from the stochastic to the deterministic model. Clearly the patterns remain stable under the deterministic dynam- ics. Simulations were run on a line with 128 cells, periodic boundary conditions and the same initial conditions for both the stochastic model and its deterministic approximation: soil and surface water depths in all cells were given by the values in the desert state, Eq. (11), while biomass was ρ0 = 0 g m−2 in half of the cells, chosen randomly, and ρ0 = 10 g m−2 in the other half (f = 1/2). Key parameter values were: µ = 1 g m−2, R = 0.6 mm d−1. See Table 1 for the remaining parameter values. ] 2 - m g 60 [ s s a m o B i 30 0 256 0 0 256 0 128 m 128 m Figure 3: (Colour online) Comparison of the dynamics of the vegeta- tion profile in the stochastic model (left) with that in the correspond- ing deterministic approximation (right). While the deterministic ap- proximation leads to extinction, the full stochastic model recovers. Simulations were run on a line with 128 cells, periodic boundary con- ditions and the same initial conditions for both the stochastic model and its deterministic approximation: soil and surface water depths in all cells were given by the values in the desert state, Eq. (11), while biomass was ρ0 = 0 g m−2 in half of the cells, chosen randomly, and ρ0 = 10 g m−2 in the other half (f = 1/2). Key parameter values were: µ = 1 g m−2, R = 0.6 mm d−1. See Table 1 for the remaining parameter values. 15 STOCHASTIC DETERMINISTIC 0 days 90 60 30 ] 2 - m g [ s s a m o B i 64 m 64 m 64 m 64 m 128 0 50 days 128 0 100 days 128 0 500 days 64 m 64 m 64 m 0 128 90 60 30 ] 2 - m g [ s s a m o B i 0 128 90 ] 2 - m g 60 [ s s a m o B i 30 0 90 128 ] 2 - m g 60 [ s s a m o B i 30 0 128 0 64 m 128 128 m 64 0 0 128 m 64 0 0 128 m 64 0 0 128 m 64 0 0 Figure 5: (Colour online) Comparison of the dynamics of the vegeta- tion profile in the stochastic model in two dimensions (left) with that in the corresponding deterministic approximation (right); the axes correspond to the two spatial dimensions. See also the supplementary video which shows the full dynamics of these same vegetation pro- files. While the deterministic approximation leads to extinction, the full stochastic model recovers. The simulations were run on a square grid of 64 ×64 cells and with periodic boundary conditions. Both the stochastic model and its deterministic approximation were started with the same initial conditions: soil and surface water depths in all cells were given by the values in the desert state, Eq. (11), while biomass was ρ0 = 0 g m−2 in half of the cells, chosen randomly, and ρ0 = 10 g m−2 in the other half (f = 1/2). Key parameter values were: µ = 1 g m−2, R = 0.6 mm d−1. See Table 1 for the remaining parameter values. 1 0.8 0.6 0.4 0.2 n o i t c n i t x e f o y t i l i b a b o r P 0 0 0.4 0.8 0.2 Initial fraction of cells populated f 0.6 1 Figure 6: (Colour online) Probability of extinction in the stochastic model (blue triangles and magenta rhombi) and in the corresponding approximation to a deterministic model (red squares and green cir- cles). We can observe a very strong contrast between the two: while the deterministic model almost always becomes extinct for cover val- ues f & 0.4, the stochastic system almost never does, for any value of f . For the stochastic case, the probabilities were estimated from 50 numerical simulations for each point. The horizontal axis indi- cates the initial fraction f of populated cells, i.e. cells with initial biomass ρ0 > 0. Simulations were run on a line with 128 cells and with periodic boundary conditions. Both the stochastic model and its deterministic approximation were started with the same initial conditions: in all cells soil and surface water depths were given by the values in the desert state, Eq. (11), while biomass was ρ0 = 10 g m−2 (green circles and magenta rhombi), and ρ0 = 50 g m−2 (red squares and blue triangles) in a fraction f of randomly chosen cells and zero in the remaining cells. Key parameter values were: µ = 1 g m−2, R = 0.6 mm d−1. See Table 1 for the remaining parameter values. 1 0.8 0.6 0.4 0.2 0 n o i t c n i t x e f o y t i l i b a b o r P 0 2 4 6 8 10 Biomass per plant per cell area, µ [g m-2] Figure 7: (Colour online) Probability of extinction in the stochastic model as a function of the parameter µ, i.e. the average mass of a plant divided by the area of a cell. The probabilities were estimated from 50 numerical simulations for each point. Simulations were run on a line with 128 cells with periodic boundary conditions and initial conditions chosen as follows: in all cells initial soil and surface water depths were given by the values in the desert state, Eq. (11), while biomass was ρ0 ≈ 10 g m−2 in a fraction f of randomly chosen cells and zero in the remaining cells. The three curves correspond to three different values of f : 1/8 (red squares), 1/2 (green circles) and 7/8 (blue triangles). See Table 1 for the remaining parameter values. For the same parameter values and initial conditions the probability of extinction in the deterministic model is essentially one throughout the whole regime investigated (not shown). 16 n o i t c n i t x e f o y t i l i b a b o r P 1 0.8 0.6 0.4 0.2 0 0.35 0.4 0.45 0.5 0.55 0.6 Rainfall, R [mm d-1] Figure 8: (Colour online) Probability of extinction in the stochastic model as a function of the rainfall rate R. Extinction probabilities were estimated from 50 numerical simulations for each point. Sim- ulations were run on a line with 128 cells with periodic boundary conditions. In all cells the initial depths of soil and surface water were given by the values in the desert state, Eq. (11), while the initial biomass was ρ0 = 10 g m−2 in half of the cells (f = 1/2) and zero in the remaining half. The parameter µ took three different values for each of the three curves shown: 1 g m−2 (red squares), 5 g m−2 (green circles) and 10 g m−2 (blue triangles). Note that for the same parameters and initial conditions the probability of extinction of the deterministic model is essentially one over the whole regime investigated (not shown). See Table 1 for the remaining parameter values. 17 Parameter Description a b c d r h k1 k2 µ Dω Dσ K L R W0 maximum infiltration rate maximum specific water uptake conversion of water uptake to plants plant mortality rate water loss due to drainage and evaporation length of the side of a cell half-saturation constant of water uptake half-saturation constant of water infiltration mean contribution of a plant to biomass density diffusion coefficient for soil water diffusion coefficient for surface water probability of a seed moving to a neighbouring cell number of cells rainfall rate water infiltration rate in the absence of plants Units d−1 mm g−1 m2 d−1 g mm−1 m−2 d−1 d−1 m mm g m−1 g m−2 m2 d−1 m2 d−1 – – mm d−1 – Value 0.2 0.05 10 0.25 0.2 2 5 5 0.1-10 0.1 100 0.02 128, 64 × 64 0.35-0.60 0.1 Table 1: Parameter values for the models studied in this paper. 18
1602.02423
1
1602
2016-02-07T20:49:06
Catch bond mechanism in Dynein motor driven collective transport
[ "physics.bio-ph", "cond-mat.soft", "q-bio.SC" ]
Recent experiments have demonstrated that dynein motor exhibits catch bonding behaviour, in which the unbinding rate of a single dynein decreases with increasing force, for a certain range of force. Motivated by these experiments, we propose a model for catch bonding in dynein using a threshold force bond deformation (TFBD) model wherein catch bonding sets in beyond a critical applied load force. We study the effect of catch bonding on unidirectional transport properties of cellular cargo carried by multiple dynein motors within the framework of this model. We find catch bonding can result in dramatic changes in the transport properties, which are in sharp contrast to kinesin driven unidirectional transport, where catch bonding is absent. We predict that, under certain conditions, the average velocity of the cellular cargo can actually increase as applied load is increased. We characterize the transport properties in terms of a velocity profile phase plot in the parameter space of the catch bond strength and the stall force of the motor. This phase plot yields predictions that may be experimentally accessed by suitable modifications of motor transport and binding properties. Our work necessitates a reexamination of existing theories of collective bidirectional transport of cellular cargo where the catch bond effect of dynein described in this paper is expected to play a crucial role.
physics.bio-ph
physics
Catch bond mechanism in Dynein motor driven collective transport Anil Nair,1 Sameep Chandel,2 Mithun K Mitra,3 Sudipto Muhuri,1 and Abhishek Chaudhuri2 1Department of Physics, Savitribai Phule Pune University, Ganeshkhind, Pune 411007, India 2Indian Institute of Science Education and Research Mohali, Knowledge City, Punjab 140306, India 3Department of Physics, IIT Bombay, India (Dated: July 9, 2021) Recent experiments have demonstrated that dynein motor exhibits catch bonding behaviour, in which the unbinding rate of a single dynein decreases with increasing force, for a certain range of force. Motivated by these experiments, we propose a model for catch bonding in dynein using a threshold force bond deformation (TFBD) model wherein catch bonding sets in beyond a critical applied load force. We study the effect of catch bonding on unidirectional transport properties of cellular cargo carried by multiple dynein motors within the framework of this model. We find catch bonding can result in dramatic changes in the transport properties, which are in sharp contrast to kinesin driven unidirectional transport, where catch bonding is absent. We predict that, under certain conditions, the average velocity of the cellular cargo can actually increase as applied load is increased. We characterize the transport properties in terms of a velocity profile phase plot in the parameter space of the catch bond strength and the stall force of the motor. This phase plot yields predictions that may be experimentally accessed by suitable modifications of motor transport and binding properties. Our work necessitates a reexamination of existing theories of collective bidirectional transport of cellular cargo where the catch bond effect of dynein described in this Letter is expected to play a crucial role. PACS numbers: 87.16.Nn Motor protein driven transport of cellular cargoes along polar microtubule (MT) filaments is one of the principal mechanisms by which active long distance transport is achieved within an eukaryotic cell [1, 2]. This mechanism plays a vital role in keeping the cell spatially organized and maintaining the uneven distributions of the various cellular components[1, 2]. While single motor properties has been extensively studied, both in experi- ments and theory [3], a large class of cooperative trans- port processes depends critically on the interaction of various motors and their collective behaviour, which can give rise to a whole new class of emergent phenomena [4 -- 9]. The mechanism of this cooperative transport, however, remains an important open question. Experiments have revealed that unidirectional transport of cellular cargo involves teamwork of motor proteins of a single type, e.g; kinesin, dynein and myosin family of motors, while bidi- rectional transport requires team of oppositely directed kinesin and dynein motors [2, 4, 10]. While molecular architecture and transport properties of these different classes of motors are significantly diverse and different, existing theoretical studies, have used the kinesin motor as a paradigm for motor-driven transport [5, 6]. Cru- cially, the single motor unbinding rate is modeled as an exponentially increasing function of force (slip-bond) for both plus-end and minus-end directed motors [5, 6]. However, recent experiments have shown that dynein un- like kinesin, exhibits catch bond behaviour, where beyond a certain threshold force, the detachment rate of a single dynein from MT filament decreases with increasing load force [11, 12]. Catch bond behaviour [13 -- 15] has been observed in various biological protein receptor-ligand complexes, such as the complex of leukocyte adhesion molecule P- selectin with the ligand PSGL-1 [16] and actin/myosin complex [17, 18] as well as microtubule-kinetochore at- tachments [19]. Different mechanisms have been pro- posed for the catch bond such as the two-state two- pathway model [20, 21], the one-state two-pathway model [22 -- 24] and the bond deformation model [25]. The one- state two-pathway model provides two dissociation path- ways, a catch barrier which increases with force, and a slip barrier which decreases with force. If the catch barrier is initially lower than the slip barrier, the sys- tem demonstrates a catch-slip transition with increasing force. The deformation model, on the other hand, pro- poses that force alters the conformation space in a fash- ion that strengthens receptor-ligand binding, and hence decreases the detachment rate. If the minimum of the po- tential decreases faster with force than the height of the barrier, one again obtains a catch-slip behaviour [13]. In this work, motivated by recent experimental and simulation studies of dynein structure, we focus on the deformation model of catch bonds. Cytoplasmic dynein has two heads that walk processively along the micro- tubule stalk in discrete steps. Each head has a globular region consisting of six AAA domains [12, 26, 27]. This globular region has two elongated structures emerging from it, the stalk, which binds to the microtubule, and the stem, which binds to the cargo [12]. It has been pro- posed that the globular head region contracts under ap- plied load, which in turn causes tension to develop along the MT-binding stalk [12, 26]. Beyond a certain critical 6 1 0 2 b e F 7 ] h p - o i b . s c i s y h p [ 1 v 3 2 4 2 0 . 2 0 6 1 : v i X r a 2 The bond deformation model proposes that catch bond behavior occurs by lowering the bound state due to force induced deformation of the bond [25]. The deformation energy is given by, Ed(f ) = α[1 − exp(− f )], where α characterizes the strength of the deformation energy and f0 sets a force scale. f0 Since in-vitro dynein exhibits catch bond behaviour above a threshold force f > fm, and the load force F is shared between k motors, we introduce the Threshold Force Bond Deformation (TFBD) model with the defor- mation energy now given by, (cid:20) (cid:18) − f − fm f0 (cid:19)(cid:21) , Ed(f = F/k) = Θ(f − fm)α 1 − exp (2) and the unbinding rate of the cargo carried with k at- tached motors attached to filaments is εk(f = F/k) = kε0 exp [−Ed(f ) + f /fd] (3) where the second term represents the usual slip contribu- tion which exponentially grows with applied load. This TFBD Model exhibits a slip-catch-slip behavior for a sin- gle motor unbinding rate as a function of applied force on the motor. In a more general context, we also study the Bond Deformation (BD) model, by setting fm to zero in Eq.( 3). A cargo which is bound by k motors moves with a velocity vk. With increasing load force, the velocity of the cargo is expected to decrease until it comes to a rest at some critical stall force fs. The decrease is approx- imately linear, as has been measured for kinesin [3, 28] and cytoplasmic dynein [29, 30]. This is modelled by the following force velocity relation (cid:20) (cid:18) F (cid:19)(cid:21) kfs vk = v0 1 − (4) where v0 is the zero-force velocity for k bound motors which is assumed to be independent of the number of bound motors. The steady state solutions to the master equation yield the probabilities for the unbound and the various bound motor states. The different transport properties are ob- tained after normalizing the probabilities with respect to the bound motor states alone [5]. Results: Having described the model, we first fit the experimental data [11] of unbinding rate of single dynein from the MT filament with our proposed TFBD model. This model is able to capture the essential functional be- haviour of unbinding rate of a single dynein as function of applied load force that has been observed in experiments, where the unbinding rate is seen to initially increase with increasing load starting with zero load, and subsequently decrease with increasing load beyond stall force fs as shown in Fig. 2(a) [11]. In fact, we predict that the single motor unbinding rates should eventually increase FIG. 1: (Color online) Schematic representation of dynein walking on an MT filament, and catch bonding under applied load. The magnified region shows the MT binding domain of the dynein stalk, which can undergo a conformational change under applied load. load, this can lead to allosteric deformations in the recep- tor region of the dynein stalk and the ligand domain on the MT surface which lock them together (Fig. 1), result- ing in a catch bond [12]. At low/intermediate loads, the catch bond cannot be activated and differential stepping is required to advance against load [12]. Direct experi- mental evidence comes from recent in-vitro experiments on single-molecule dynein detachment kinetics [11]. The detachment rate of a dynein motor is found to initially increase, and then decrease with force beyond a critical threshold. At extremely large forces, we should eventu- ally regain a slip bond, thus exhibiting a slip-catch-clip behaviour over the entire force range. This we model through a Threshold Force Bond Deformation (TFBD) Model, where the deformation pathway activates beyond a certain threshold force. Model: We consider a cargo which is transported on a filament by N dynein motors against a constant exter- nal load force F . The state of the cargo is characterized by the number of bound motors k (0 ≤ k ≤ N ). Mo- tors are irreversibly attached to the cargo but undergo attachment (detachment) to (from) the filament. With k bound motors, the rates of attachment and detachment are given by πk = (N − k)πad and εk = kε respectively. The load force is assumed to be shared equally by the k bound motors [5], so that each motor experiences a force f = F/k. The dynamics of the attachment/detachment process is given by the temporal evolution of the prob- ability pk of having k bound motors, expressed as the one-step master equation, = εk+1pn+1 + πk−1pk−1 − (εk + πk)pk. (1) dpk dt The corresponding master equation for the unbound vesi- cle (k = 0) is, dp0 N bound motors, it is dpN dt = ε1p1 − π0p0, while for vesicle with dt = πN−1pN−1 − εN pN . CargoCargoAAA DomainStemStemStalkStalk stiffensAAA Domain contractsF++−−LRLRLoad Force 3 FIG. 2: (Color online) (a) Variation of unbinding rate of a sin- gle dynein with constant load force (F): The circles correspond to data points obtained from Ref.[11]. The solid curve is ob- tained from TFBD model with α/kBT = 68, fo = 40.7pN , fm = 1.4pN , fd = 0.67pN with fs = 1.25pN [11] and ε0 = 1s−1 [11]. Transport properties of cargo carried by 5 dyneins as function of load force (F):(b) Effective Unbinding rate (K) vs F . (c) The average number of attached motors vs F, (d) Average velocity vs F. For (b), (c) and (d), pa- rameter values chosen are same as (a) with πad = 1.6s−1[6], vo = 0.65µm/s[6]. with increased load force for forces higher than those accessed in this set of experimental studies. Next we study how the catch bonding behaviour exhibited by sin- gle dynein for unidirectional transport affects the steady state transport properties of cellular cargo that is being transported by multiple dynein motors. We consider a cellular cargo being carried by 5 dynein motors [31]. We use the set of fitted model parameter values of the afore- mentioned experiment, to find the functional behaviour of various transport properties as a function of applied load force F . Fig. 2(b) shows the effective unbinding rate of the cargo from the MT filament increases initially and then it decreases for forces upto about 10 pN , fol- lowing which it again starts increasing with increasing load force. Fig. 2(c) shows that the average number of attached motors also increases after an initial dip and finally for forces large than about 20 pN, the average number of attached motors decreases, and exponentially approaches 1. Fig. 2(d) shows that the mean velocity of the cargo decreases monotonically with increasing load in this parameter regime. Experimentally, it is known that the various motor properties can vary significantly for different classes of dynein motors. For instance, weak dynein is known to have a stall force of fs = 1.1pN while for strong dynein fs = 7pN [30]. We next explore the different plausible scenarios resulting from the ramifications of generic catch FIG. 3: (Color online) Probability distribution for the num- ber of bound motors in a system with N = 5 for the (a) TFBD and (b) BD model. Compared to the zero force case (green crosses), where the probability distribution is peaked at zero bound motors and then decreases monotonically, for finite forces, both the TFBD and the BD model show a non-monotonic probability distribution, with the distribution peaking at larger n in the BD case. Panels (c) and (d) shows the effective unbinding rate and the average number of bound motors as a function of force for a slip bond (solid blue line), as opposed to the TFBD (green dashed curve) and the BD model (red dotted line). Data is for α/kBT = 20 and fs = 2pN with v0 = 0.65µm/s [6] , ε0 = 1.0s−1 [11], f0 = 7pN, fm = 1.4pN, fd = 0.67pN and πad = 0.1s−1. bond behaviour observed not only for dynein motors but also for myosin motors [17, 18] by studying the transport properties both for the TFBD model as well as the Bond Deformation (BD) model, by setting fm to zero in Eq. 3. We shall focus on the variation of the transport proper- ties resulting from changes in stall force fs , binding rates πad and catch bond strength α. The effect of the catch bonding on the probability dis- tribution of n bound motor state P (n) is to shift the peak value of the distribution towards higher number of bound motors for certain range of forces both for the TFBD model (Fig. 3a) and the BD model (Fig. 3b). This shift results from the fact that when a larger number of mo- tors is bound to the filament, the force on each motor is low enough that they are in the catch regime, resulting in a decrease of propensity of individual motors to de- tach from filament in this state and hence a consequent increase in the probability of the states with higher num- ber of attached motors. Conversely, when fewer number of motors bind to the filament, the load on each motor is higher, so that motors in this state are more likely to detach, and hence have a lower probability. This shifting of peak of the probability distribution towards higher n states manifests as an increase in the average number of 05101548120510150.20.40.6010203040123402460.20.40.6(a)(b)(c)(d)Unbinding rateK V(µm /s)N( s )-1-1( s )Force ( pN )Force ( pN )01234500.20.40.60.81Probability P ( n )01234500.20.40.60.81Probability P ( n )051015Force ( pN )12Effective unbinding rate ( s )0102030Force ( pN )123Avg. no. of bound motorsNumber of bound motors ( n )Number of bound motors ( n )-1F = 9 pNF = 0 pNF = 0 pNF = 9 pN( a )( b )( d )( c ) 4 (Color online) Force-velocity curves: FIG. 4: (a) TFBD Model for two different binding rates, πad = 0.1s−1 (solid blue curve) and πad = 1.0s−1 (dashed red curve). The solid curve shows four velocity humps which reduces to a single- hump velocity profile on increasing πad. (b) BD Model for πad = 0.01s−1 (solid blue curve) and πad = 0.1s−1 (dashed In this case, a four-hump profile reduces to a red curve). monotonically decreasing velocity profile on increasing πad. Data is for N = 5 motors with v0 = 0.65µm/s [6], 0 = 1.0s−1[11], f0 = 7pN, fm = 1.4pN, fd = 0.67pN, α/kBT = 35 and fs = 2pN . bound motors and a decrease in effective unbinding rates for certain range of forces for both TFBD and BD model. This is shown in Figs. 3(c) and 3(d) and the behaviour contrasts with the case of slip bonds where the effective unbinding rate monotonically increases and the average number of bound motors monotonically decreases with increasing load. Next we consider the mean velocity profile of the cargo as a function of load force. Catch bond behaviour mani- fests in rather remarkable behaviour for the velocity pro- file of the cargo - the mean velocity of the cargo can actually increase with increase in opposing load force for certain range of fs, the strength of catch bond α and binding rate πad for both the TFBD and BD model (Fig. 4). This unique behaviour can be understood as a direct consequence of the catch bond effect, which tends to stabilize the bound motor states with higher attached motors which move with a higher velocity so that the average velocity of the cargo actually increases. We systematically study the effect of variation of the stall force fs and α by constructing a velocity profile phase diagram with different regions characterised by the number of maxima of mean velocity in the force-velocity profiles. Fig. 5 shows the resulting phase diagrams for two different πad values each for the TFBD (5(a) 5(b)) and the BD (5(c) 5(d)) models. The different regions in the phase diagram corresponds to different velocity-force profiles having maximas ranging from zero to 4 ( for N = 5). For sufficiently weak catch bond strength α, the force-velocity profiles is such that the mean velocity al- FIG. 5: Velocity phase diagrams in the α − Fs plane for the TFBD model with πad = (a) 0.1s−1, (b) 1.0s−1, and for the BD model with πad = (a) 0.01s−1, (b) 0.1s−1. The violet re- gion corresponds to a monotonically decreasing force velocity profile. The red, green blue and yellow regions corresponds to velocity profiles with one, two, three and four humps respec- tively (see Fig. 4 for examples of individual velocity profiles). Parameter values are the same as in Fig. 4. Stall forces is in pNs. α is in in units of kbT . ways decreases for increasing load force, similar to the behaviour in the absence of catch bond. However for both TFBD and BD model, increasing α and lowering fs have the effect of modifying the velocity-force profiles in a manner that they have one or more maxima of the mean velocity as illustrated in Fig. 5. The parameter space explored is a plausible biological regime and in princi- pal should be observable by suitable biochemical means which can alter and/or the stall forces, the catch bond strength and binding rates of the motors to the filament. The TFBD Model proposed in this Letter captures the experimental results for dynein driven cargo transport. In contrast to canonical slip-bond models, the TFBD model for the catch bond has non-trivial consequences for the transport properties, in particular for the velocity profiles in response to applied loads, which should in prin- ciple be observable in experiments, and hence provides a testable prediction for our model. The more generic BD model also has similar dramatic differences in the transport properties that might be relevant for other mo- tor driven systems. This work necessitates a reexamina- tion of existing models of cellular cargo transport to take into account the catch bond mechanism described here. In particular, cooperative bi-directional cargo transport through the simultaneous action of oppositely directed motors (with one or both types of motors having a catch bond) is expected to have significantly different charac- teristics as compared to those described by existing the- 02468100.10.20.30.40.50.6036912150.10.20.30.40.50.6Force ( pN )Velocity ( m/s )µ( a )( b ) ories, and will be discussed in a forthcoming publication. [14] R. P. McEver and C. Zhu, Annu Rev Cell Dev Biol 26, 5 Acknowledgements MKM acknowledges financial support from the Ra- manujan Fellowship, Department of Science and Tech- nology, India and the IRCC Seed Grant, IIT Bom- bay. SM acknowledges DBT RGYI Project No: BT/PR6715/GBD/27/463/2012 for financial support. SC and AC acknowledges DST, India for financial sup- port. The authors would also like to thank the organ- isers of the SMYIM conference, Pondicherry, where part of this work was done. 363 (2010). [15] S. Rakshit and S. Sivasankar, Phys. Chem. Chem. Phys. 16, 2211 (2014). [16] B. T. Marshall, M. Long, J. W. Piper, T. Yago, R. McEver and C. Zhu, Nature 423, 193 (2003). [17] B. Guo and W. H. Guilford, Proc. Natl. Acad. Sci. 103, 9844 (2006). [18] A. Yamada, A. Mamane, J. Lee-Tin-Wah, A. D. Cicco, C. Prevost, D. Levy, J. F. Joanny, E. Coudier and P. Bassereau, Nat. Commun. 5, 3624 ( 2014) [19] B. Akiyoshi, K. K. Sarangapani, A. F. Powers, C. R. Nelson, S. L. Reichow, H. Arellano-Santoyo, T. Gonen, J. A. Ranish, C. L. Asbury, and S. Biggins, Nature 468, 576 (2010). [20] E. Evans, A. Leung, V. Heinrich, and C. Zhu, Proc. Natl. Acad. Sci. 101, 11281 (2004). [21] V. Barsegov and D. Thirumalai, Proc. Natl. Acad. Sci. 102, 284 (2005). [22] D. Bartolo, I. Derenyi and A. Ajdari, Phys. Rev. E 102, [1] B. Alberts at. al , Molecular Biology of the cell ( Garland 1835 (2005). Science, New York, 2002, 4th ed) [2] M. A. Welte, Curr.Biol. 14, R525 (2004). [3] J. Howard, Mechanics of motor proteins and the cy- toskeleton (Sinauer Associates, Sunderland, 2001) [23] Y. V. Pereverzev, O. V. Prezhdo, M. Forero, E. V. Sokurenko, and W. E. Thomas, Biophys. J. 89, 1446 (2005). [24] E. A. Novikova and C. Storm, Biophys. J. 105, 1336 [4] A. K. Rai, A. Rai, A. J. Ramaiya, R. Jha, and R. Mallik, (2013). Cell 152, 172 (2013) [25] O. V. Prezhdo and Y. V. Pereverzev, Phys. Rev. E 73, [5] S. Klumpp and R. Lipowsky, Proc. Natl. Acad. Sci. 102, 050902 (2006). 284 (2005) [26] R. Mallik, B. C. Carter, S. A. Lex, S. J. King and S. P. [6] M. J. I. Muller, S. Klumpp and R. Lipowsky, Proc. Natl. Gross, Nature 427, 649 (2004) Acad. Sci. 105, 4609 (2008) [27] M. P. Singh, R. Mallik, S. P. Gross, and C. C. Yu, Proc. [7] V. Soppina, A. K. Rai, A. J. Ramaiya, P. Barak and R. Natl. Acad. Sci. 102, 19381 (2005) Mallik, Proc. Natl. Acad. Sci. 106, 19381 (2009) [8] D. Bhat and M. Gopalakrishnan, Phys. Biol. 9, 046003 (2012) [9] W. O. Hancock, Nature 15, 615 (2014) [10] V. Levi, V. Gelfand et.al : Biophys. J. 90, 318 (2006) [11] A. Kunwar et. al, Proc. Natl. Acad. Sci. 108, 18960 (2011) [28] N. J. Carter and R. A. Cross, Nature 435, 308 ( 2005) [29] S. Toba, T. M. Watanabe, L. Okimoto-Yamaguchi, Y. Y. Toyoshima and H. Higuchi, Proc. Natl. Acad. Sci. 103, 5741 (2006) [30] M. J. I. Muller, S. Klumpp and R. Lipowsky, J. Stat. Phys. 133, 1059 (2008) [31] It has been reported in Ref.[7] that for endosome trans- [12] R. Mallik, A. K. Rai, P. Barak, A. Rai, and A. Kunwar, port four to eight dyneins are involved. Trends in Cell Biol. 23, 575 (2013) [13] O. V. Prezhdo and Y. V. Pereverzev, Accounts of Chem. Research 42, 693 (2009).
1010.1207
2
1010
2010-11-23T14:12:57
Near-critical fluctuations and cytoskeleton-assisted phase separation lead to subdiffusion in cell membranes
[ "physics.bio-ph", "cond-mat.soft" ]
We address the relationship between membrane microheterogeneity and anomalous subdiffusion in cell membranes by carrying out Monte Carlo simulations of two-component lipid membranes. We find that near-critical fluctuations in the membrane lead to transient subdiffusion, while membrane-cytoskeleton interaction strongly affects phase separation, enhances subdiffusion, and eventually leads to hop diffusion of lipids. Thus, we present a minimum realistic model for membrane rafts showing the features of both microscopic phase separation and subdiffusion.
physics.bio-ph
physics
Near-Critical Fluctuations and Cytoskeleton-Assisted Phase Separation Lead to Subdiffusion in Cell Membranes 0 1 0 2 v o N 3 2 ] h p - o i b . s c i s y h p [ 2 v 7 0 2 1 . 0 1 0 1 : v i X r a Jens Ehrig, Eugene P. Petrov,∗ and Petra Schwille Biophysics, BIOTEC, Technische Universitat Dresden, Tatzberg 47/49, 01307 Dresden, Germany To appear in Biophysical Journal Vol. 99, Issue 12, 2010. ∗Corresponding author: [email protected] Abstract We address the relationship between membrane microheterogeneity and anomalous subdif- fusion in cell membranes by carrying out Monte Carlo simulations of two-component lipid membranes. We find that near-critical fluctuations in the membrane lead to transient subdif- fusion, while membrane -- cytoskeleton interaction strongly affects phase separation, enhances subdiffusion, and eventually leads to hop diffusion of lipids. Thus, we present a minimum realistic model for membrane rafts showing the features of both microscopic phase separation and subdiffusion. PACS: 87.14.Cc, 87.15.ak, 87.16.dj, 87.16.dt. Keywords: Lipid membranes, DMPC, DSPC, phase separation, membrane rafts, subdiffu- sion. 2 Introduction Anomalous subdiffusion in cell membranes is an intriguing phenomenon whose molecular ori- gins are still a subject of debate [1 -- 3]. From the general viewpoint, the phenomenon is related to (dynamic) microheterogeneities in the properties of the cell membrane creating a rugged energy landscape for the diffusing protein or lipid molecule. The current understanding of membrane microdomains embraces the concepts of lipid rafts [4] and cytoskeleton-based picket fence [5]. The general concept of membrane rafts [6] implies a dynamic interplay be- tween the membrane local composition and phase, and local membrane-protein interactions, which can result in local lipid demixing and phase separation. This view is supported by the recent experimental evidence demonstrating that local lipid demixing and phase separation can be induced by crosslinking [7], change in the local membrane curvature [8], and inter- action with a cytoskeleton [9]. Additionally, the presence of the membrane-associated actin network leads to a shift in the membrane phase transition temperature and, potentially, to broadening of the phase transition [9]. The efficiency of external perturbations in changing the local properties of the mem- brane should be strongly enhanced in the vicinity of the membrane critical point. This is indeed supported by recent experimental observations [10]. Moreover, it was suggested that dynamic microdomains in the cell membrane are nothing but near-critical fluctuations in the local composition and phase of the membrane [11]. All this is in agreement with the observation that the composition of the cell membrane is adjusted to keep it above the phase transition temperature, and is regulated in response to environmental changes to maintain this condition [12]. In cases where this mechanism fails, cold shock damage takes place [13]. In this paper, we address the relationship between membrane microheterogeneity and anomalous subdiffusion in cell membranes by carrying out Monte Carlo simulations of two- component lipid membranes on experimentally relevant spatial scales (∼1 µm) and time intervals (∼1 s), with a special emphasis on interactions with the model membrane skeleton. We demonstrate that (i) near-critical fluctuations in a free lipid membrane can lead to transient anomalous subdiffusion, (ii) phase separation in two-component (and, hence, mul- ticomponent) lipid membranes can be strongly affected by interaction with the membrane skeleton, which, depending on the temperature and membrane composition, can either lead to precipitation of highly dynamic membrane domains (rafts), or prevent large-scale phase separation, and (iii) interaction with the membrane skeleton enhances anomalous subdiffu- sion and eventually leads to hop-diffusion of lipids. Thus, we construct a minimum realistic model for membrane rafts showing the features of both microscopic phase separation and anomalous subdiffusion. The binary lipid system consisting of 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and 1,2-distearoyl-sn-glycero-3-phosphocholine (DSPC) is chosen as a generic model of a two-component membrane. The particular choice of this system is motivated by the fact that it is well studied both in vitro [14 -- 21] and in silico (see [20 -- 22] and refs. therein). Even though the exact temperature range of phase separation in this membrane is too high for most living organisms, the main qualitative conclusions of the study related to behavior of cell membranes are not affected by the particular thermodynamic parameters of 3 the lipid mixture. From the physical perspective, it is assumed that living organisms try to maintain the composition of cell membranes to keep them above the phase transition temperature [12]. Therefore, to be biologically relevant, in the present paper we mostly focus on the upper part of the phase diagram. Materials and Methods Experimental The saturated phospholipids DMPC and DSPC with the melting temperatures 297 and 328 K were purchased from Avanti Polar Lipids (Alabaster, AL). Multilamellar vesicle (MLV) suspensions were obtained by rehydration of a dry lipid film. Excess heat capacity curves of DMPC/DSPC MLV suspensions in a 10 mM Hepes buffer, pH 7.4, were obtained using a VP-DSC calorimeter (MicroCal, Northampton, MA) at a scan rate of 2 -- 3 K/h. The empirical phase diagram based on the temperatures of the onset and completion of the phase transition determined from the scanning calorimetry data as described [20, 21] are in agreement with previously reported results (see Supporting Material). Monte Carlo simulations Our approach to lattice-based Monte Carlo (MC) simulations of a two-component membrane is generally similar to the one described previously [20, 21]. To facilitate efficient simula- tions on experimentally relevant spatial scales (∼1 µm) and time intervals (∼1 s), where a particular type of lipid packing and fine molecular details should be of little importance, we further simplified the model and represented the membrane as a square lattice, each node of which represents a molecule of one of the two lipid types, which can be in either gel or fluid conformational state. An elementary MC step consists of two substeps: (i) an attempt to change the state of a randomly chosen lipid and (ii) an attempt of position exchange of a randomly chosen next-neighbor pair of lipids. As in [21], a rate function was introduced for the next-neighbor exchange step to ensure ∼40 times slower lipid diffusion in the gel phase. For an L × L lattice, an MC cycle consists of a chain of L2 elementary MC steps. For every lipid composition and temperature studied, the membrane was first equilibrated, if required, and simulations were run for (6 − 20) × 106 MC cycles to collect the necessary data. More details on the simulation procedure are given in the Supporting Material. Simulations were carried out on an L × L = 600 × 600 (400 × 400 when modeling the effects of the membrane skeleton) square lattice with periodic boundary conditions. By assuming the average lipid headgroup size of 0.8 nm, this corresponds to a membrane with an experimentally relevant size of 0.48 × 0.48 µm2 (0.32 × 0.32 µm2). By comparing the DMPC diffusion coefficient obtained in our simulations of pure DMPC in the fluid phase at 304 K with its experimental values at the same temperature (3 − 6 µm2/s [23 -- 25]), we found that one MC cycle corresponds to ∼50 ns, and thus, our simulations cover processes on timescales up to ∼0.3 -- 1 s. 4 Model for the membrane skeleton The membrane skeleton was modeled by a random Voronoi tessellation satisfying the pe- riodic boundary conditions. For simulations with L = 400, random tessellations with N = 36 compartments were used, which gives the average linear size of the compartment (cid:96) = LN−1/2 = 66.6 lattice units ≈ 53 nm. The generated filament meshwork was projected onto the square lattice thereby creating a set of pixels representing the locations of the fila- ments. Each of these locations can be assigned to be a cytoskeleton pinning site. To mimic the membrane -- cortical skeleton interaction, a simple rule inspired by experimental data on lipid interactions with transmembrane proteins [26] was followed: a lipid located at a fila- ment pinning site is forced to assume the gel conformation with no explicit restrictions on its mobility. Thus, the pinning site does not present an obstacle for lipid diffusion: its effect on diffusion is indirect and takes place solely due to a lower lipid mobility in the gel-state local environment. The effect of the varying strength of the membrane-skeleton interaction was modeled by randomly assigning a fraction of filament position pixels to be filament pinning sites. Simulations were carried out with the filament pinning density set to 25%, 50%, and 100%; in these cases the total number of pinning sites amounted to approximately 1%, 2%, and 4% of the total membrane area. Our choice of immobile pinning sites is justified by experimental observations demon- strating that band 3 and ankyrin strongly bound to the membrane skeleton show a very low diffusion coefficient of ∼ 10−4− 10−3 µm2/s over time intervals up to tens of seconds [27, 28]. It should be pointed out that the approach used in the present work to account for interactions of lipid molecules with membrane proteins is conceptually similar, though not identical, to the one used in several previous MC simulation-based studies [29 -- 32]. In these works, it was assumed that proteins, preferentially wetted by one of the membrane phases, could freely diffuse in the membrane, which results in their accumulation in this phase and, in case of two-component membranes, in fact enhances large-scale phase separation. (The situation is more complicated in case of active proteins -- for details, see [32] and refs. therein.) It was found out that, under certain conditions, interaction of a single-component membrane with small transmembrane proteins or peptides can even lead to emergence of a closed loop of gel -- fluid coexistence with a lower critical point [33]. It should be pointed out, however, that these works focused only on structural properties of protein-loaded membranes, and did not address the issues of diffusion in the membrane. What sets the present work apart from the above-mentioned studies, is that here we consider interaction of the membrane with the immobile cytoskeleton and study its effects on phase separation and diffusion of lipids in the membrane. Analysis of lipid diffusion data Positions of a small fraction of lipid molecules (150 for simulations with L = 600 and 50 for simulations with L = 400, which amounted to 0.04% and 0.03%, respectively) were recorded, and the time- and ensemble-averaged mean-square displacement (MSD) was determined as 5 follows: MSD(τ ) = 1 tmax − τ (cid:8)[x(t) − x(t + τ )]2 + [y(t) − y(t + τ )]2(cid:9) , tmax−τ(cid:88) t=1 (1) where τ , t, and tmax are times measured in units of MC cycles; here, τ denotes the time lag, and tmax is the total length of the lipid molecule trajectory. No difference between time- and ensemble-averaged MSD (Eq. 1) and time-only- and ensemble-only-averaged MSDs beyond the statistical error level was observed. In case of normal diffusion, the MSD grows in a linear fashion with time: MSD(τ ) = 4Dτ in 2D, where D is the translational diffusion coefficient. In case of anomalous subdiffusion, the MSD shows a slower sublinear power-law growth MSD(τ ) ∼ τ β with 0 < β < 1 (see, e.g., [34]). An alternative description of diffusion showing deviations from the normal behavior (also in cases where it cannot be described in terms of subdiffusion) can be provided using an effective time-dependent diffusion coefficient D(τ ). Therefore, to characterize the behavior of the MSD curves, the local exponent of the mean-square displacement βMSD(τ ) = d log MSD(τ )/d log τ, and the effective time-dependent diffusion coefficient D(τ ) = 1 4 dMSD(τ )/dτ were calculated. Simulation of FCS experiments (2) (3) To simulate fluorescence correlation spectroscopy (FCS) [35] measurements, the tracked par- ticles (as above, in the amount of 150 for simulations with L = 600 and 50 for simulations with L = 400) were assumed to be fluorescent. Fluorescence intensity fluctuations δF (t) = F (t)−(cid:104)F(cid:105) about the mean intensity (cid:104)F(cid:105) in a 2D Gaussian detection spot exp(−2r2/r2 0) were recorded, and their autocorrelation function G(τ ) = (cid:104)δF (t)δF (t + τ )(cid:105)/(cid:104)F(cid:105)2 was calculated. The detection spot size was set to r0 = 31 lattice units ≈ 25 nm, the size experimentally achievable using the stimulated emission depletion (STED) FCS technique [36]. We addi- tionally note that this detection spot is much smaller than the lattice size (r0/L ≈ 0.05 for L = 600 and r0/L ≈ 0.08 for L = 400), which allowed us to avoid artifacts in the fluorescence autocorrelation function and additionally ensured that our simulations are experimentally relevant. FCS curves were averaged over nine different positions on the lattice; when studying the effects of the membrane skeleton, the results were additionally averaged over five random realizations of the filament network. Simulated FCS curves were analyzed using the model G(τ ) = G(0)/(cid:2)1 + (τ /τD)βFCS(cid:3) . 6 (4) For βFCS = 1 this expression corresponds to normal diffusion, while for 0 < βFCS < 1 it provides a simple way to describe anomalous subdiffusion in FCS [35]. In case of normal diffusion, τD is related to the diffusion coefficient of fluorescent particles D and detection spot size r0: τD = r2 0/(4D). Notice that since FCS is sensitive not only to the MSD(τ ), but also to the higher moments of the distribution of displacements of a diffusing particle, the connection between G(τ ) and MSD(τ ) is straightforward only in case of a Gaussian distribution of displacements [35]. As a result, in case of anomalous diffusion of particles, generally βFCS (cid:54)= βMSD. Results and Discussion Phase and component separation in the membrane With appropriate tuning of lipid interaction parameters, the empirical heat capacity-based phase diagram obtained from our MC simulation is in agreement with our experimental data (Fig. 1), as well as with previously published experimental and simulation data on the same lipid system (see Supporting Material). At the same time, the empirical phase diagram constructed on the basis of heat capacity data may differ from the real phase diagram of the system and thus not provide an insight into the microscopic structure and the dynamics of the membrane. More details on the phase diagram can be obtained based on binodal and spinodal curves of the system. The binodal curves, also known as coexistence curves, are the boundaries of the region in the phase diagram in which the equilibrated system shows a complete separation of the two phases. The spinodal encloses the region where the mixture is unstable with respect to small local fluctuations of the composition and always lies inside the area enclosed by the binodal, with the exception of a critical point, where the binodal and spinodal touch. To reconstruct the binodals and spinodals of the DMPC/DSPC lipid mixture, we ana- lyzed the static structure factors [37] for lipids SL(k) and lipid states SS(k) (see Supporting Material), reflecting the character of spatial fluctuations of the membrane composition and state. We found that outside the phase coexistence region, the Ornstein -- Zernike (OZ) approx- imation [37] SOZ(k) = S(0)/[1 + (kξ)2], where k is a wavenumber, and ξ is a correlation length, provided an excellent description of the lipid-state structure factors SS(k). The tem- perature dependences of SS(0) and ξS were used to estimate spinodal and binodal curves. In particular, the spinodals were determined by extrapolating 1/SS(0) dependence to zero crossing [37, 38], and binodals were determined from the condition d(1/ξS)/dT = 0 [38]. Interestingly, in some regions of the phase diagram, the binodal deviates quite strongly from the phase coexistence boundary estimated from the excess heat capacity data (Fig. 1). This discrepancy indeed shows that the analysis of heat capacity data does not necessarily recover the real phase diagram of the system. Here, this behavior reflects the continuous charac- ter of the phase transition in the membrane, whose proper description requires a combined component and state phase diagram [39]. What process is reflected by the excess heat capacity curves where they fail to describe 7 Figure 1: Left-hand panel: Component and lipid state separation phase diagram of DMPC/DSPC lipid mixtures. Phase transition temperatures as determined from experi- mental ((cid:52)) and simulated (◦) excess heat capacity curves. Lipid state spinodal (- - -), lipid state binodal ( -- -- -- ), and lipid demixing curves ( -- ). Right-hand panel: representative snap- shots of equilibrium membrane configurations at temperatures and membrane compositions corresponding to filled squares in the left-hand panel. Lattice size: 600× 600; scale bar: 200 lattice units ≈ 160 nm. L1 (k) + SOZ the fluid -- gel phase separation? An analysis of the structure factors SL(k) for the lipid species helps to answer this question. It appears that, generally, two OZ components are required to describe these data: SL(k) = SOZ L2 (k). Parameters of component 1 only weakly depend on the temperature and reflect demixing of lipids in the same state. Parameters of the second component SL2(0) and ξL2 show strong temperature dependences similar to those of SS(0) and ξS, and thus describe appearance of dynamic microscopic domains, which cannot be treated as distinct thermodynamically stable phases [22]. We therefore define the lipid demixing curves as a set of points on the phase diagram satisfying the condition SL1(0) = SL2(0). The fact that the lipid demixing curves are in excellent agreement with the calorimetry-based empirical phase diagrams, supports the above reasoning, as do the snapshots presented in Fig. 1. Notice that in this region the membrane undergoes near- critical fluctuations. On the other hand, in the coexistence region, equilibrated membranes show complete phase separation and form large-scale lipid domains (Fig. 1). The above analysis of the structure factors SL(k) and SS(k) and equilibrated membrane snapshots shows that in the part of the phase diagram where the lipid demixing curve closely approximates the binodal, the phase transition from the fluid state to the fluid -- gel coexis- tence has a quasi-abrupt character. On the other hand, in the region where the lipid demixing curve strongly deviates from the binodal, the membrane shows characteristic near-critical 8 fluctuations (Fig. 1), and the transition becomes continuous as the system approaches the critical point (DMPC/DSPC ≈ 20:80, Tc = 320.5± 0.2 K -- see [40] for details). Remarkably, this behavior is qualitatively similar to recent experimental observations on three-component lipid membranes [11]: depending on the membrane composition, the transition to the two- phase coexistence takes place either in an abrupt manner -- when the membrane does not pass through a critical point, or via critical fluctuations -- when the membrane does pass through a critical point. A more detailed consideration of the phase diagram of the DMPC/DSPC system based on MC simulations will be published elsewhere [40]. In the fluid -- gel phase coexistence region, large-scale phase separation takes place. Since no explicit or implicit penalties are imposed [41 -- 43] on the domain size in our model, and domain growth is driven by minimization of the line tension energy, phase separation results in formation of a single circular-shaped domain of the minority phase (Fig. 1). Effect of the cytoskeleton on phase separation in the membrane In the region of near-critical fluctuations the membrane is expected to be very sensitive to external perturbations, including the interaction with the membrane skeleton. It appears that in the vicinity of the critical point the interaction with the membrane skeleton leads to immediate condensation of the gel phase on the skeleton filaments, and formation of mem- brane domains, in a good agreement with results of experiments on lipid bilayers with a reconstituted actin skeleton [9]. These domains dynamically change their shape, but nev- ertheless stay pinned to the filaments (Fig. 2). Notice that the effect of the filaments is remarkably robust with respect to the filament pinning density. We strongly believe that these domains represent the minimal model of membrane rafts [6]. The minimum character of this model stems from the fact that in the present scenario domain formation is thermo- dynamically driven, and does not require any active processes like chemical cross-linking of membrane components or lipid recycling. If the membrane is abruptly cooled down from the all-fluid state to a temperature in the fluid -- gel coexistence region of the phase diagram, phase separation takes place, and domains of the fluid and gel phase start to nucleate and coarsen with time. In a free membrane, domains grow according to the power law R(t) ∼ tn with the growth exponent n depending on the particular domain growth mechanism [44]. In our simulations for a free membrane, depending on the lipid composition and temperature, n takes values from 1/4 to 1/3 (Fig. 3), consistent with general expectations for the domain growth in 2D [44] and in agreement with experimental results (see, e.g., [45]) (for a more detailed discussion, see [40]). In a finite-size system with a linear dimension L, domain coarsening eventually stops, and a single circular-shaped domain is produced (Fig. 1) with the radius R∞ (cid:39) (X/π)1/2L, where 0 < X < 1/2 is the fraction of the minority phase. Remarkably, we found that the presence of the membrane skeleton strongly inhibits or even eventually prevents large-scale phase separation in the phase coexistence region (Fig. 2). As a result, the radius of the membrane domains in this case is largely determined by the characteristic compartment radius Rcomp (cid:39) (cid:96)/2 of the filament network. The way we account for interaction of lipid molecules with filaments at their pinning sites is similar to a theoretical 9 Figure 2: Effect of the membrane skeleton on the phase separation in a DMPC/DSPC membrane. Representative snapshots of membrane configurations are shown for the free membrane (first row ) and membrane interacting with a network of filaments at 100% (sec- ond row ), 50% (third row ) and 25% pinning density (fourth row ). Snapshots for the free membrane, as well as for the membrane interacting with filaments at T = 321 and 322 K, represent fully equilibrated configurations; snapshots for the membrane interacting with fil- aments at T = 310 and 317.5 K correspond to equilibration time of 6 × 106 MC cycles (see text for discussion). For presentation purposes, the filaments are drawn thicker than they are in reality. Lattice size: 400 × 400; scale bar: 125 lattice units ≈ 100 nm. model (the random-field Ising model [46]) implying the presence of static random position- dependent perturbations in a 2D system of spins. This model predicts that the initial power- law domain growth R(t) ∼ tn is strongly slowed down at intermediate stages, and crosses over to an extremely slow logarithmic growth R(t) ∼ log t; eventually the domains are expected to reach the perturbation strength-dependent maximum size R(cid:48) < R∞ in an exponential In our system, the perturbation time to create an equilibrium disordered state [47, 48]. strength is determined by the filament pinning density and the average compartment radius Rcomp. Therefore, if the interaction of the membrane with the cytoskeleton is strong enough, and R∞2 (cid:29) R2 comp, one can expect that the domain growth stops when domains reach the characteristic size Rcomp (cid:46) R(cid:48) < R∞. We indeed observed an extreme slowing down of the domain growth in the presence of the membrane skeleton and cross-over to the slow logarithmic growth (Fig. 3). As is evident from the Figure, at the end of a simulation run of 6 × 106 MC cycles, the domain sizes are indeed R(t) ∼ Rcomp. Although some growth is still observed at this stage, it is so slow, that at present it is unclear, whether the system will evolve toward a disordered equilibrium state featuring a number of small domains, as suggested in [47, 48], or the slow logarithmic growth will continue further to produce eventually (after an extremely long time) a single domain with the size R∞. What is clear, though, that the time required for complete 10 :::: Figure 3: Effect of the membrane skeleton on the the domain growth in DMPC/DSPC 50:50 membrane abruptly cooled from the all-fluid state down to T = 310 K in the fluid -- gel phase coexistence region. Kinetics of domain growth when the membrane is free (upper curve) and in the presence of membrane skeleton with the filaments pinning density of 50% (lower curve). Black solid line shows the power law dependence R(t) ∼ tn with n = 0.32; dashed line shows the stage of the slow logarithmic growth R(t) ∼ ln t in the presence of membrane skeleton. Representative membrane configurations obtained in our MC simulations at time instants 104, 105, and 106 MC cycles are shown at the corresponding curves. Lattice size: 400 × 400. equilibration is so long, that from the practical viewpoint one can state that indeed the presence of the membrane skeleton prevents large-scale phase separation in the membrane. This becomes especially clear if one takes into account that a membrane in a live cell is not in the equilibrium state, and a number of other processes affecting the membrane state and composition, e.g., lipid recycling, take place in parallel on different spatial and time scales. Thus, the cytoskeleton-induced inhibition of large-scale phase separation can serve as one of the possible explanations why micrometer-scale membrane domains are observed in giant plasma membrane vesicles [49], but not in cell membranes. In addition, our observations suggest that the established cryoprotective role played by the cytoskeleton [50] may consist in delaying phase separation-induced cold shock damage of living cells. Diffusion of lipids in the free membrane At higher temperatures, away from the phase transition and the near-critical fluctuations region, the membrane is in the homogeneous all-fluid state, and, not unexpectedly, lipid diffusion is normal (Fig. 4). This picture changes upon approaching the critical point in the region of near-critical 11 : Figure 4: Effect of the proximity to the phase transition on diffusion of DMPC lipids in a DMPC/DSPC 20:80 membrane. Mean-square displacement MSD(τ ) (a), local expo- nent βMSD(τ ) (b), time-dependent diffusion coefficient D(τ ) (c), and FCS autocorrelation G(τ )/G(0) (d). Panels (a) -- (c) top to bottom: T = 328, 322, and 321 K. Panel (d): T = 328 (left) and 321 K (right). For clarity, data for 322 K are omitted in panel (d). Panel (d) additionally shows fits to the FCS diffusion model Eq. (4) giving βFCS = 1.01 at 328 K and βFCS = 0.86 at 321 K ( -- -- -- ). For comparison, a fit of 321 K data with fixed βFCS = 1.0 is shown (- - -). fluctuations where interpenetrating fluctuating domains form in the membrane. The anal- ysis of the immediate environment of the DMPC and DSPC lipids shows that, while the DSPC lipid does not have a pronounced preference for the state of its local environment and therefore is evenly distributed between fluid and gel domains, the DMPC lipid shows a strong preference for the fluid local environment and is thus predominantly partitioned into fluid domains. Under these conditions, the DSPC lipid shows no significant deviation from the normal diffusion behavior (data not shown). By contrast, quite unexpectedly, the DMPC lipid was found to demonstrate very pronounced anomalous subdiffusion (Fig. 4). At a first glance, this subdiffusive behavior is rather surprising, especially in the light of the seminal work by Kawasaki [51] demonstrating that the diffusion coefficient of a two- component mixture does not vanish upon approaching the critical point, which implies nor- mal diffusion, at least in the long-time asymptotic regime. A closer look at the data shows, 12 20:8020:8020:8020:80 however, that the subdiffusive behavior is in fact transient, though it covers several orders of magnitude in time, and at longer times the cross-over to the normal diffusion takes place. One can also notice that at very short time lags, diffusion is normal as well. This behavior is also evident from the plots of the local exponent βMSD(τ ) and effective time-dependent diffusion coefficient D(τ ) extracted from the MSD(τ ) dependences (Fig. 4b, c). How this cross-over from normal diffusion to a subdiffusive behavior and back to normal diffusion at long times can be explained? By recalling that the DMPC lipid is preferentially partitioned into fluid domains and having a look at the corresponding membrane snapshots (Fig. 1), we realize that DMPC molecules diffuse on a dynamically rearranging fractal- like fluid phase pattern. Therefore, at early times, when the DMPC molecule explores its immediate environment, normal diffusion should take place. Later, though at times shorter than a characteristic time of rearrangement of this dynamic fractal structure, the motion of DMPC lipid molecules is subdiffusive, similar to what is expected in case of diffusion on fractal-like percolation clusters [52]. At much longer times, evolution of these clusters leads to dynamic percolation behavior (see, e.g., [53]), which results in a crossover from the subdiffusive motion to normal diffusion. In contrast to diffusion on static percolation clusters, where cross-over to normal diffusion occurs only above the percolation transition [52], in the case of lipid diffusion in a near-critical membrane, cross-over to normal diffusion will always take place irrespectively of whether the fluid phase is above or below the percolation threshold. This subdiffusive behavior is also clearly observed in our simulations of FCS experiments (Fig. 4d). Upon a closer approach to the critical point of the system, the autocorrelation functions of fluorescence intensity fluctuations G(τ ) progressively stronger deviate from the dependence expected for normal diffusion (i.e., Eq. (4) with βFCS = 1). A good description of G(τ ) data in this case can only be obtained with βFCS < 1. For example, we find that for DMPC/DSPC 20:80 mixture at T = 321.0 K the best fit is obtained with βFCS = 0.86 (Fig. 4d), and a smaller βFCS = 0.79 is required to fit G(τ ) at T = 320.7 K, which is closer to the critical point (data not shown). A new important result of the present work is that transient anomalous subdiffusion spanning several orders of magnitude in time can be observed in a system close to its critical point. We emphasize that the appearance of the transient subdiffusive behavior in the region of near-critical fluctuations is not related to any specific properties of the system and should be observed close to criticality in various systems independent of their origin, including, of course, multicomponent lipid membranes. It is well known that phase separation in lipid bilayers produces domains with typical sizes ranging from a few to several tens of micrometers (provided that there are no restrictions on the domain growth). In agreement with that, we observe in our simulations that within the coexistence region an equilibrated membrane always shows complete phase separation resulting in a single circular-shaped domain of the minority phase. Therefore, in this case, the only reasonable way to carry out FCS measurements requires parking the detection spot into the bulk of the majority phase away from the inter-phase boundary, or into the center of the minority phase domain. Exactly this approach is used in experimental FCS studies on membranes showing large-scale phase separation (see, e.g., [54 -- 56]). Not unexpectedly, 13 this approach results in normal diffusion in both phases with phase-dependent diffusion coefficients. This behavior is also observed in our simulations (data not shown). In contrast to that and quite surprisingly, an MC simulation study of Hac et al. [21] reported FCS curves strongly deviating from the normal diffusion model exactly in the phase coexistence region, i.e., where large-scale phase separation takes place. These results were attributed by the authors of [21] as resulting from subdiffusive motion of lipids, which is in contradiction with the macroscopic phase separation in the system. This seeming contradiction is resolved upon careful examination of the approach to simulations in [21]. There, the diameter of the FCS detection spot was approximately equal to the simulation box size. In this case, the FCS detection spot covers effectively the whole simulated membrane, and, even in the presence of large-scale phase separation, motion of lipids in both fluid and gel phases contributes to the FCS results, which explains deviations of FCS curves from the normal diffusion model observed in [21] in the fluid -- gel phase coexistence region and rule out the interpretation of these results in terms of subdiffusion. Moreover, the situation simulated in [21] is very unlikely for experiments on single lipid bilayers within the phase coexistence region, where membrane domains typically have radii of a few to several tens of micrometers, whereas the typical FCS detection spot size is ∼ 200 nm for the standard FCS [35], and down to 25 nm for STED FCS [36]. Here we should also mention a previous lattice-based simulation study [57] addressing the character of lateral diffusion of lipid molecules in DMPC/DSPC 50:50 membranes. There, in agreement with our results, diffusion is normal at T > 320 K, i.e. in the all-fluid membrane state (cf. Fig. 1). Deviations from the normal diffusion for the DMPC/DSPC 50:50 mixture found in [57] within the temperature range T = 300 − 320 K (i.e. exactly where the large- scale phase separation for this lipid composition takes place -- see Fig. 1) were interpreted in terms of subdiffusive motion of lipids. As in [21], the spatial scales on which deviations from normal diffusion were observed in [57] approach the system size. Hopping of lipid molecules between the macroscopic gel and fluid domains definitely leads to deviations from normal diffusion at these large observation scales, although these deviations can hardly be interpreted in terms of anomalous subdiffusion. We emphasize that our results presented in this paper unambiguously show that local component and phase fluctuations in a near-critical membrane lead to the transient anoma- lous subdiffusion spanning several orders of magnitude in time. Under suitable experimental conditions, we believe, this subdiffusive behavior can be observed experimentally using, e.g., the (STED) FCS technique. Effects of membrane -- cytoskeleton interaction on lipid diffusion Interaction with the membrane skeleton strongly affects the character of diffusion in the re- gion of near-critical fluctuations (Fig. 5). With increasing the filament pinning density, the interaction of the membrane with filaments becomes more pronounced: it first enhances the anomalous diffusion and eventually leads to hop-diffusion of lipid molecules. In the presence of filaments, the faster diffusion process clearly corresponds to Brownian motion within com- partments defined by the membrane skeleton, whereas the slower diffusion is due to hopping 14 Figure 5: Effect of the membrane skeleton on diffusion of DMPC lipids in a DMPC/DSPC 20:80 membrane at 322 K. Mean-square displacement MSD(τ ) (a), local exponent βMSD(τ ) (b), time-dependent diffusion coefficient D(τ ) (c), and FCS autocorrelation G(τ )/G(0) (d). Panels (a) -- (c) top to bottom, and panel (d) left to right: filament pinning density equals 0% (free membrane), 25%, 50%, and 100%. between the compartments, and therefore it strongly depends on the filament pinning den- sity. Notice that gel phase condensation at the membrane skeleton substantially increases the effective thickness of filaments and thus enhances their influence on lipid diffusion. Remarkably, in this case our results for the mean-square displacement MSD(τ ) (Fig. 5a) and effective time-dependent diffusion coefficient D(τ ) (Fig. 5c) are in good qualitative agreement with results of single-molecule tracking experiments in cell membranes [58], where results were interpreted in terms of a picket-fence model. Our MSD(τ ) data also qualita- tively agree with a theoretical model for diffusion on an infinite periodic square meshwork of penetrable barriers [59], except for in our case the short-time motion of a molecule within a compartment is subdiffusive due to near-critical fluctuations. We again emphasize that even a relatively low pinning density of the cytoskeleton filaments leads to appearance of rather strong barriers for diffusing lipids due to filament-induced local condensation of the gel phase. Our simulations of STED FCS experiments with the experimentally achievable detection spot size [36, 60] show (Fig. 5d) that, while at early times all FCS curves exhibit behavior 15 20:8020:8020:8020:80 characteristic of subdiffusion, interaction with the membrane skeleton leads to appearance of a second slow diffusion component. The contribution of this slow component increases with the filament pinning density. Clearly, this component is related to the large-scale diffusion that involves crossing the filament-induced barriers on the membrane. This two-component character of G(τ ) is in qualitative agreement with the recent STED FCS measurements of lipid diffusion in cell membranes [60]. It would be very interesting to follow the dependence of the FCS autocorrelation functions and the respective apparent diffusion coefficients on the spatial scale using the variable detection spot FCS technique [61, 62], as has been done in [60]. This work is currently in progress. Here, we point out once again that in our model the pinning sites do not represent obstacles for diffusing lipids, and affect lipid diffusion only indirectly, via the lower lipid mobility in the local gel-state environment (see Materials and Methods). We anticipate that in case where pinning sites would be presented by immobile particles preferentially wetted by gel-phase lipids, the effects of the membrane -- filament interaction on diffusion of lipids should become even more pronounced. This issue will be addressed in our future work. Conclusion In conclusion, in this paper, we carried out Monte Carlo simulations of model two-component lipid membranes on experimentally relevant spatial scales and time intervals. This allowed us to better understand the details of the dynamics of phase and component separation, and by this means to address the relationship between membrane microheterogeneity and anomalous subdiffusion in cell membranes. We observed that interaction of the near-critical membrane with the cytoskeleton strongly affects phase separation: for the membrane in the state of near-critical fluctuations it leads to local phase separation and formation of dynamic domains with a size of a few tens of nanometers, which conform to the current understanding of membrane rafts [6]. We find that in a membrane showing near-critical fluctuations lipids show subdiffusive behavior covering several orders of magnitude in time. The interaction of the membrane with the cytoskeleton enhances subdiffusion and eventually leads to hop-diffusion of lipids. In the fluid -- gel phase coexistence region of the phase diagram, interaction of the mem- brane with the cytoskeleton is found to tremendously slow down large scale phase separation in the membrane, which may explain the established connection between the cytoskeleton and cold shock resistance of organisms [50]. The concepts of lipid rafts [4] and cytoskeleton-related picket fence [5] are frequently dis- cussed as the alternative viewpoints on the origin of anomalous diffusion in cell membranes. In our paper, we bring these two concepts together to show that not only they do not contra- dict one another, but rather work in synergy, resulting in formation of cytoskeleton-induced dynamic lipid domains in the near-critical membrane. By this means we construct what, we believe, is a minimum raft model, since the domain formation is driven solely by thermody- namics and does not require either chemical cross-linking of membrane components, or lipid recycling. 16 Acknowledgements We acknowledge inspiring discussions with H. Rigneault and C. Favard at the early stage of the project. The work was supported by the Deutsche Forschungsgemeinschaft via Research Group FOR 877. 17 References [1] Saxton, M. J. 2007. A biological interpretation of transient anomalous subdiffusion. I. Quantitative model. Biophys. J. 92:1178 -- 1191. [2] Destainville, N., A. Sauli`ere, and L. Salom`e. 2008. Comment to the Article by Michael J. Saxton: A biological interpretation of transient anomalous subdiffusion. I. Quantitative model. Biophys. J. 95:3117 -- 3119. [3] Saxton, M. J. 2008. Response to Comment by Destainville et al. Biophys. J. 95:3120 -- 3122. [4] Simons, K., and E. Ikonen. 1997. Functional rafts in cell membranes. Nature. 387:569 -- 572. [5] Fujiwara, T., K. Ritchie, H. Murakoshi, K. Jacobson, and A. Kusumi. 2002. Phos- pholipids undergo hop diffusion in compartmentalized cell membrane. J. Cell Biol. 157:1071 -- 1082. [6] Pike, L. J. 2006. Rafts defined: A report on the Keystone symposium on lipid rafts. J. Lipid Res. 47:1597 -- 1598. [7] Hammond, A. T., F. A. Heberle, T. Baumgart, D. Holowka, B. Baird, and G. W. Feigenson. 2005. Crosslinking a lipid raft component triggers liquid ordered-liquid dis- ordered phase separation in model plasma membranes. Proc. Natl. Acad. Sci. USA. 102:6320 -- 6325. [8] Roux, A., D. Cuvelier, P. Nassoy, J. Prost, P. Bassereau, and B. Goud. 2005. Role of curvature and phase transition in lipid sorting and fission of membrane tubules. EMBO J. 24:1537 -- 1545. [9] Liu, A. P., and D. A. Fletcher. 2006. Actin polymerization serves as a membrane domain switch in model lipid bilayers. Biophys. J. 91:4064 -- 4070. [10] Sorre, B., A. Callan-Jones, J.-B. Manneville, P. Nassoy, J.-F. Joanny, J. Prost, B. Goud, and P. Bassereau. 2009. Curvature-driven lipid sorting needs proximity to a demixing point and is aided by proteins. Proc. Natl. Acad. Sci. USA. 106:5622 -- 5626. [11] Veatch, S. L., P. Cicuta, P. Sengupta, A. Honerkamp-Smith, D. Holowka, and B. Baird. 2008. Critical fluctuations in plasma membrane vesicles. ACS Chem. Biol. 3:287 -- 293. [12] Mouritsen, O. G. 1987. Physics of biological membranes. In Physics of Living Matter. Lecture Notes in Physics, Vol. 284. D. Baeriswyl, M. Droz, A. Malaspinas, and P. Martinoli, editors. Springer, Berlin Heidelberg New York. 76 -- 109. [13] Drobnis, E. Z., L. M. Crowe, T. Berger, T. J. Anchordoguy, J. W. Overstreet, and J. H. Crowe. 1993. Cold shock damage is due to lipid phase transitions in cell membranes: A demonstration using sperm as a model. J. Exp. Zool. 265:432 -- 437. 18 [14] Shimshick, E. J., and H. M. McConnell. 1973. Lateral phase separation in phospholipid membranes. Biochemistry. 12:2351 -- 2360. [15] Mabrey, S., and J. M. Sturtevant. 1976. Investigation of phase transitions of lipids and lipid mixtures by high sensitivity differential scanning calorimetry. Proc. Natl. Acad. Sci. USA. 73:3862 -- 3866. [16] Lentz, B. R., Y. Barenholz, and T. E. Thompson. 1976. Fluorescence depolarization studies of phase transitions and fluidity in phospholipid bilayers. 2. Two-component phosphatidylcholine liposomes. Biochemistry. 15:4529 -- 4537. [17] van Dijck, P. W. M., A. J. Kaper, H. A. J. Oonk, and J. de Gier. 1977. Miscibility prop- erties of binary phosphatidylcholine mixtures. A calorimetric study. Biochim. Biophys. Acta. 470:58 -- 69. [18] Wilkinson, D. A., and J. F. Nagle. 1979. Dilatometric study of binary mixtures of phosphatidylcholines. Biochemistry. 18:4244 -- 4249. [19] Nibu, Y., T. Inoue, and I. Motoda. 1995. Effect of headgroup type on the miscibility of homologous phospholipids with different acyl chain lengths in hydrated bilayer. Biophys. Chem. 56:273 -- 280. [20] Sug´ar, I. P., T. E. Thompson, and R. L. Biltonen. 1999. Monte Carlo simulation of two-component bilayers: DMPC/DSPC mixtures. Biophys. J. 76:2099 -- 2110. [21] Hac, A. E., H. M. Seeger, M. Fidorra, and T. Heimburg. 2005. Diffusion in two- component lipid membranes -- A fluorescence correlation spectroscopy and Monte Carlo simulation study. Biophys. J. 88:317 -- 333. [22] Heimburg, T. 2007. Thermal Biophysics of Membranes. Wiley-VCH, Weinheim. [23] Kuo, A. L., and C. G. Wade. 1979. Lipid lateral diffusion by pulsed nuclear magnetic resonance. Biochemistry. 18:2300 -- 2308. [24] Vaz, W. L. C., R. M. Clegg, and D. Hallmann. 1985. Translational diffusion of lipids in liquid clystalline phase phosphatidylcholine multibilayers. A comparison of experiment with theory. Biochemistry. 24:781 -- 786. [25] Dolainsky, C., P. Karakatsanis, and T. M. Bayerl. 1997. Lipid domains as obstacles for lateral diffusion in supported bilayers probed at different time and length scales by two- dimensional exchange and field gradient solid state NMR. Phys. Rev. E. 55:4512 -- 4521. [26] Kinnunen, P. K. J. 1991. On the principles of functional ordering in biological mem- branes. Chem. Phys. Lipids. 57:375 -- 399. [27] Tomishige, M., Y. Sako, and A. Kusumi. 1998. Regulation mechanism of the lateral diffusion of band 3 in erythrocyte membranes by the membrane skeleton. J. Cell Biol. 142:989 -- 1000. 19 [28] Cairo, C. W., R. Das, A. Albohy, Q. J. Baca, D. Pradhan, J. S. Morrow, D. Coombs, and D. E. Golan. 2010. Dynamic regulation of CD45 lateral mobility by the spectrin-ankyrin cytoskeleton of T cells. J. Biol. Chem. 285:11392 -- 11401. [29] Sperotto, M. M., and O. G. Mouritsen. 1991. Monte Carlo simulation studies of lipid order parameter profiles near integral membrane proteins. Biophys. J. 59:261 -- 270. [30] Heimburg, T., and R. L. Biltonen. 1996. A Monte Carlo simulation study of protein- induced heat capacity changes and lipid-induced protein clustering. Biophys. J. 70:84 -- 96. [31] Sabra, M. C., and O. G. Mouritsen. 1998. Steady-state compartmentalization of lipid membranes by active proteins. Biophys. J. 74:745 -- 752. [32] Gil, T., J. H. Ipsen, O. G. Mouritsen, M. C. Sabra, M. M. Sperotto, and M. J. Zucker- mann. 1998. Theoretical analysis of protein organization in lipid membranes. Biochim. Biophys. Acta. 1376:245 -- 266. [33] Zhang, Z., M. M. Sperotto, M. J. Zuckermann, and O. G. Mouritsen. 1993. A micro- scopic model for lipid/protein bilayers with critical mixing. Biochim. Biophys. Acta. 1147:154 -- 160. [34] Metzler, R., and J. Klafter. 2000. The random walk's guide to anomalous diffusion: a fractional dynamics approach. Phys. Rep. 339:1 -- 77. [35] Petrov, E. P., and P. Schwille. 2008. State of the art and novel trends in fluorescence correlation spectroscopy. In Standardization and Quality Assurance in Fluorescence Measurements II. Springer Series on Fluorescence, Vol. 6. U. Resch-Genger, editor. Springer, Berlin Heidelberg New York. 145 -- 197. [36] Kastrup, L., H. Blom, C. Eggeling, and S. Hell. 2005. Fluorescence correlation spec- troscopy in subdiffraction focal volumes. Phys. Rev. Lett. 94:178104. [37] Fisher, M. E. 1964. Correlation functions and the critical region of simple fluids. J. Math. Phys. 5:944 -- 962. [38] Strobl, G. 1997. The Physics of Polymers. 2nd ed. Springer, Berlin Heidelberg New York. [39] Michonova-Alexova, E. I., and I. P. Sug´ar. 2002. Component and state separation in DMPC/DSPC lipid bilayers: A Monte Carlo simulation study. Biophys. J. 83:1820 -- 1833. [40] Ehrig, J., E. P. Petrov, and P. Schwille. 2010. Phase separation and near-critical fluctuations in two-component lipid membranes: Monte Carlo simulations on exper- imentally relevant spatial scales and time intervals. New J. Phys. submitted; e-print arXiv:1009.4860. 20 [41] Baumgart, T., S. T. Hess, and W. W. Webb. 2003. Imaging coexisting fluid domains in biomembrane models coupling curvature and line tension. Nature. 425:821 -- 824. [42] Semrau, S., T. Idema, L. Holtzer, T. Schmidt, and C. Storm. 2009. Membrane-mediated interactions measured using membrane domains. Biophys. J. 96:4906 -- 4915. [43] Ursell, T. S., W. S. Klug, and R. Phillips. 2009. Morphology and interaction between lipid domains. Proc. Natl. Acad. Sci. USA. 106:13301 -- 13306. [44] Furukawa, H. 1985. A dynamic scaling assumption for phase separation. Adv. Phys. 34:703 -- 750. [45] Jensen, M. H., E. J. Morris, and A. C. Simonsen. 2007. Domain shapes, coarsening, and random patterns in ternary membranes. Langmuir. 23:8135 -- 8141. [46] Imry, Y., and S.-k. Ma, 1975. Random-field instability of the ordered state of continuous symmetry. Phys. Rev. Lett. 35:1399 -- 1401. [47] Grant, M., and J. D. Gunton. 1987. Metastable states in the random-field Ising model. Phys. Rev. B. 35:4922 -- 4928. [48] Anderson, S. R. 1987. Growth and equilibration in the two-dimensional random-field Ising model. Phys. Rev. B. 36:8435 -- 8446. [49] Baumgart, T., A. T. Hammond, P. Sengupta, S. T. Hess, D. A. Holowka, B. A. Baird, and W. W. Webb. 2007. Large-scale fluid/fluid phase separation of proteins and lipids in giant plasma membrane vesicles. Proc. Natl. Acad. Sci. USA. 104:3165 -- 3170. [50] Clark, M. S., and M. R. Worland. 2008. How insects survive the cold: molecular mech- anisms -- a review. J. Comp. Physiol. B. 178:917 -- 933. [51] Kawasaki, K. 1966. Diffusion constants near the critical point for time-dependent Ising models. III. Self-diffusion constant. Phys. Rev. 150:285 -- 290. [52] ben-Avraham, D., and S. Havlin. 2000. Diffusion and Reactions in Fractals and Disor- dered Systems. Cambridge University Press, Cambridge. [53] Kutner, B., and K. W. Kehr. 1983. Diffusion in concentrated lattice gases IV. Diffusion coefficient of tracer particle with different jump rate. Philos. Mag. A. 48:199 -- 213. [54] Bacia, K., P. Schwille, and T. Kurzchalia. 2005. Sterol structure determines the separa- tion of phases and the curvature of the liquid-ordered phase in model membranes. Proc. Natl. Acad. Sci. USA. 102:3272 -- 3277. [55] Kahya, N., D. Scherfeld, K. Bacia, B. Poolman, and P. Schwille. 2003. Probing lipid mobility of raft-exhibiting model membranes by fluorescence correlation spectroscopy. J. Biol. Chem. 278:28109 -- 28115. 21 [56] Chiantia, S., J. Ries, N. Kahya, P. and Schwille. 2006. Combined AFM and two-focus SFCS study of raft-exhibiting model membranes. ChemPhysChem. 7:2409 -- 2418. [57] Sug´ar, I. P., and R. L. Biltonen. 2005. Lateral diffusion of molecules in two-component lipid bilayer: A Monte Carlo simulation study. J. Phys. Chem. B. 109:7373 -- 7386. [58] Murase, K., T. Fujiwara, Y. Umemura, K. Suzuki, R. Iino, H. Yamashita, M. Saito, H. Murakoshi, K. Ritchie, and A. Kusumi. 2004. Ultrafine membrane compartments for molecular diffusion as revealed by single molecule techniques. Biophys. J. 86:4075 -- 4093. [59] Wieser, S., M. Moertelmaier, E. Fuertbauer, H. Stockinger, and G. J. Schutz. 2007. (Un)Confined diffusion of CD59 in the plasma membrane determined by high-resolution single molecule microscopy. Biophys. J. 92:3719 -- 3728. [60] Eggeling, C., C. Ringemann, R. Medda, G. Schwarzmann, K. Sandhoff, S. Polyakova, V. N. Belov, B. Hein, C. von Middendorff, A. Schonle, and S. W. Hell. 2009. Di- rect observation of the nanoscale dynamics of membrane lipids in a living cell. Nature. 457:1159 -- 1163. [61] Masuda, A., U. Kiminori, and T. Okamoto. 2005. New fluorescence correlation spec- troscopy enabling direct observation of spatiotemporal dependence of diffusion constants as an evidence of anomalous transport in extracellular matrices. Biophys. J. 88:3584 -- 3591. [62] Wawrezinieck, L., H. Rigneault, D. Marguet, and P.-F. Lenne. 2005. Fluorescence corre- lation spectroscopy diffusion laws to probe the submicron cell membrane organization. Biophys. J. 89:4029 -- 4042. 22 Near-Critical Fluctuations and Cytoskeleton-Assisted Phase Separation Lead to Subdiffusion in Cell Membranes Jens Ehrig, Eugene P. Petrov,∗ and Petra Schwille Biophysics, BIOTEC, Technische Universitat Dresden, Tatzberg 47/49, 01307 Dresden, Germany -- Supporting Material -- ∗Corresponding author: [email protected] Empirical phase diagram based on differential scanning calorimetry data A single-component membrane undergoes a phase transition from the gel state to the fluid state at its melting temperature. On the other hand, for a two-component lipid membrane the transition from the all-gel state to the all-fluid state is not immediate. (In fact, the two-component membrane undergoes two phase transitions: one from the gel state to the fluid -- gel coexistence and, at a higher temperature, another one from the fluid -- gel coexistence to the fluid state.) Therefore, from the practical point of view, one can speak about a broad- ened gel -- fluid transition in a two-lipid system (compared to a single-lipid system) [S1 -- S3]. This broadened transition can be characterized by the onset and completion temperatures, which can be determined experimentally. By plotting the experimentally determined onset and completion temperatures as a function of the membrane composition, one obtains an empirical experimental phase diagram of the system. To determine the onset and completion temperatures from the experimental DSC data, the empirical tangent method was used. The outer slopes of the C(T ) profiles of a range of compositions (DMPC/DSPC = 0/100, 10/90, 20/80, ..., 90/10, 100/0) were fit with straight lines passing through the corresponding inflection points. The onset and completion temperatures in this case are defined as the respective intersections of the tangential lines with the zero-line. The experimental phase diagram of the DMPC/DSPC lipid mixture obtained in the present study is in a good agreement with the ones published earlier by other groups [S1 -- S9] (Fig. S1). We remark here that evaluation of the onset and completion temperatures is not always unambiguous. In particular, our reanalysis of the experimental C(T ) data from [S8] for DMPC/DSPC 20/80 and 10/90 mixtures gives onset temperatures lower than those reported in [S8] and very close to the ones we obtained from our data (Fig. S1). Monte Carlo simulations The fundamental step of the MC simulation consists of two sub-steps. In the first sub-step, an attempt is made to change the state of a randomly chosen lipid (from gel to fluid or vice versa). The second sub-step is an attempt to exchange the positions of a randomly chosen pair of next neighbors on the lattice. For each sub-step the change in the Gibbs free energy ∆G =∆N F + ∆N GG 1 (∆E1 − T ∆S1) + ∆N F 12 wFF 12 + ∆N FF 12 wGG 2 (∆E2 − T ∆S2) + ∆N GF 11 wGF 21 wGF 21 , 12 + ∆N GF 12 wGF 12 + ∆N GF 11 + ∆N GF 22 wGF 22 (S1) is calculated. Here, ∆Ei and ∆Si are the changes in the internal energy and entropy of a molecule of lipid i when it switches its state from gel to fluid, and wmn are the next- neighbor interaction parameters of lipids i and j being in states n and m, respectively. In the simulation, the attempts of the state change and next-neighbor exchange are accepted with probability p = 1 for ∆G < 0 and p = exp (−∆G/RT ) for ∆G ≥ 0. ij S2 Figure S1: Empirical phase diagram of DMPC/DSPC constructed from excess heat capacity MC simulation data. Experimental data: present work (•), Ref. curves measured experimentally on multilamellar vesicle suspensions and calculated from [S2] (◦), Ref. [S3, S8] ((cid:46)), our reanalysis of experimental C(T ) data for DMPC/DSPC 20/80 and 10/90 from [S8] ((cid:46)· ), Ref. [S9] (×). Monte Carlo simulation data: present work ((cid:4)), Refs. [S3, S8] (+), Ref. [S9] (∗). [S4] ((cid:5)), Ref. [S5] ((cid:79)), Ref. [S6] ((cid:77)), Ref. [S1] ((cid:3)), Ref. [S7] ((cid:47)), Refs. For an L × L lattice, one MC cycle consists of a chain of L2 elementary MC steps. After each cycle, the enthalpy of the lattice is calculated as follows: H =N F 1 ∆E1 + N F + N GG 12 wGG 12 + N FF 12 wFF 2 ∆E2 + N GF 11 wGF 12 + N GF 11 + N GF 12 wGF 22 wGF 22 21 wGF 12 + N GF 21 . (S2) The excess heat capacity C(T ) is calculated from the variance of equilibrium fluctuations of the total lattice enthalpy H as follows: C(T ) =(cid:10)(H − (cid:104)H(cid:105))2(cid:11)/RT 2. (S3) The lipid interaction parameters were adjusted using the approach previously described by Sug´ar et al. [S3, S10]: temperature-dependent excess heat capacity curves were obtained from MC simulations for a range of membrane compositions and compared with experimen- tally measured heat capacity curves, and the parameters wmn ij were varied until a reasonable agreement with experimental DSC data was achieved. Lipid interaction parameters for our model are listed in Table S1. Since a simpler lattice representation of the lipid system (lipid molecules arranged on a square lattice) was used in our simulations compared to the previ- ous studies [S3, S8 -- S10] (lipids represented as dimers of acyl chains arranged on a triangular lattice), not surprisingly, the lipid interaction parameters wmn in our study differ from those used in previous publications. ij S3 Table S1: Thermodynamic parameters of lattice-based MC simulations of the DMPC/DSPC lipid membranes used in the present work. Thermodynamic Parameters ∆E1 (Jmol-1) ∆E2 (Jmol-1) ∆S1 ∆S2 (Jmol-1K-1) (Jmol-1K-1) wGF 11 (Jmol-1) wGF 22 (Jmol-1) wGF 12 (Jmol-1) wGF 21 (Jmol-1) wGG 12 (Jmol-1) wFF 12 (Jmol-1) 26330 50740 88.653 154.695 1827 1622 4025 4460 1412 502 As is evident from Fig. S1, with this set of lipid interaction parameters, the heat capacity- based empirical phase diagram obtained in our MC simulations agrees well with experimental data (both ours and previously published), as well as with previously published results of lattice-based MC simulations of the same lipid system. For every lipid composition and temperature studied, the membrane was first equili- brated, if required. To do that, a system being initially at T = ∞ was brought to the equilibrium at a target temperature T by exercising the Monte Carlo procedure which ad- ditionally included a third MC substep consisting in position exchange of two randomly chosen lipid molecules. This approach, which was suggested in [S11] and successfully ap- plied in [S3, S8, S12] is known to be very efficient in driving the system toward the equilibrium configuration. This procedure was propagated typically for 1.5 × 105 MC cycles, which is substantially longer than the typical time required by the total lattice enthalpy H to reach its equilibrium value at a given temperature in the presence of the random lipid exchange sub- step. After the completion of this procedure, the random lipid exchange substep is switched off, and the equilibrated system is ready for studies of its equilibrium properties. The simulation code was written in Fortran and compiled with Compaq Visual Fortran Compiler Version 6.6A (Compaq Computer Corporation, Houston, TX). Results were ana- lyzed using original dedicated routines written in MATLAB (The MathWorks, Nattick, MA). A reliable random number generator is essential for the success of an MC simulation. There- fore, we used the Mersenne Twister routine [S13] providing sequences of pseudo-random numbers equidistributed in 623 dimensions and characterized by an extremely long period of 219937 − 1 ≈ 4.3 × 106001. Monte Carlo simulations were carried out on a workstation (Intel Core2 Quad Extreme CPU X9770 3.2 GHz, 4 GB RAM) running under Windows XP. Under these conditions, a simulation run on a 600 × 600 square lattice for 2 × 107 MC cycles takes about 700 h of CPU time. Calculation of static structure factors for lipids and lipid states To gain information on the character of spatial fluctuation of the local membrane composition and lipid state in the equilibrium, the static structure factors for lipids (DMPC and DSPC) SL(k) and lipid states (fluid and gel) SS(k) were calculated independently. The structure factor S(k) = 1 + ¯ρg(k) (S4) S4 is expressed via the Fourier transform g(k) of the pair distribution function of particles g(r) = ¯ρ−2 δ(ri)δ(rj − r) , (S5) (cid:42)(cid:88) (cid:88) j(cid:54)=i i (cid:43) where ¯ρ is the spatially averaged number density of particles [S14]. To avoid artifacts in spatial and angular averaging, the fact that the particles cannot take arbitrary positions, but rather occupy sites on a square lattice, was taken into account. References [S1] Shimshick, E. J., and H. M. McConnell. 1973. Lateral phase separation in phospholipid membranes. Biochemistry. 12:2351 -- 2360. [S2] Wilkinson, D. A., and J. F. Nagle. 1979. Dilatometric study of binary mixtures of phosphatidylcholines Biochemistry. 18:4244 -- 4249. [S3] Sug´ar, I. P., T. E. Thompson, and R. L. Biltonen. 1999. Monte Carlo simulation of two-component bilayers: DMPC/DSPC mixtures. Biophys. J. 76:2099 -- 2110. [S4] Mabrey, S., and J. M. Sturtevant. 1976. Investigation of phase transitions of lipids and lipid mixtures by high sensitivity differential scanning calorimetry. Proc. Natl. Acad. Sci. USA. 73:3862 -- 3866. [S5] Lentz, B. R., Y. Barenholz, and T. E. Thompson. 1976. Fluorescence depolarization studies of phase transitions and fluidity in phospholipid bilayers. 2. Two-component phosphatidylcholine liposomes. Biochemistry. 15:4529 -- 4537. [S6] van Dijck, P. W. M., A. J. Kaper, H. A. J. Oonk, and J. de Gier. 1977. Miscibil- ity properties of binary phosphatidylcholine mixtures. A calorimetric study. Biochim. Biophys. Acta. 470:58 -- 69. [S7] Nibu, Y., T. Inoue, and I. Motoda. 1995. Effect of headgroup type on the miscibility of homologous phospholipids with different acyl chain lengths in hydrated bilayer. Biophys. Chem. 56:273 -- 280. [S8] Sug´ar, I. P., and R. L. Biltonen. 2000. Structure-function relationships in two- component phospholipid bilayers: Monte Carlo simulation approach using a two-state model. Methods Enzymol. 323:340 -- 372. [S9] Hac, A. E., H. M. Seeger, M. Fidorra, and T. Heimburg. 2005. Diffusion in two- component lipid membranes -- A fluorescence correlation spectroscopy and Monte Carlo simulation study. Biophys. J. 88:317 -- 333. S5 [S10] Sug´ar, I.P., R. L. Biltonen, and N. Mitchard. 1994. Monte Carlo simulations of mem- branes: Phase transition of small unilamellar dipalmitoylphosphatidylcholine vesicles. Methods Enzymol. 240:569 -- 593. [S11] Sun, H., and I. P. Sug´ar. 1997. Acceleration of convergence to the thermodynamic equilibrium by introducing shuffling operations to the Metropolis algorithm of Monte Carlo simulations. J. Phys. Chem. B. 101:3221 -- 3227. [S12] Sug´ar, I. P., and R. L. Biltonen. 2005. Lateral diffusion of molecules in two-component lipid bilayer: A Monte Carlo simulation study. J. Phys. Chem. B. 109:7373 -- 7386. [S13] Matsumoto, M., and T. Nishimura. 1998. Mersenne Twister: a 623-dimensionally equidistributed uniform pseudo-random number generator. ACM Trans. Model. Com- put. Simul. 8:3 -- 30. [S14] Allen, M. P., and D. J. Tildesley. 1991. Computer Simulations of Liquids. Oxford University Press, New York. S6
1010.5928
1
1010
2010-10-28T11:56:23
Localised IR spectroscopy of hemoglobin
[ "physics.bio-ph" ]
IR absorption spectroscopy of hemoglobin was performed using an IR optical parametric oscillator laser and a commercial atomic force microscope in a novel experimental arrangement based on the use of a bottom-up excitation alignment. This experimental approach enables detection of protein samples with a resolution that is much higher than that of standard IR spectroscopy. Presented here are AFM based IR absorption spectra of micron sized hemoglobin features
physics.bio-ph
physics
Localised IR spectroscopy of hemoglobin Fiona Yarrow and James H. Ricea NanoPhotonics Research Group, School of Physics, University Co llege Dublin, Belfield, Dublin, Ireland a) Electronic mail: james.rice@ucd. ie Abstract IR absorption spectroscopy o f hemoglobin was performed using an IR optical parametric oscillator laser and a commercial atomic force microscope in a novel experimental arrangement based on the use o f a bottom-up excitat ion alignment. This experimental approach enables detection o f protein samples with a reso lut ion that is much higher than that of standard IR spectroscopy. Presented here are AFM based IR absorption spectra of micron sized hemoglobin features. Introduction Absorption spectroscopy is a widely applied technique for chemical characterisat ion. This method is able to detect both luminescent and non-luminescent materials and provide chemical specific informat ion. An extensively used form o f absorption spectroscopy is infrared absorption (IR) spectroscopy. This measures specific frequencies in the infrared region o f the electromagnet ic spectrum at which constituent parts of mo lecules corresponding to specific types o f mo lecular bonds vibrate. This makes possible structural elucidat ion and compound identificat ion o f materials. As a consequence, IR absorption is extensively used as an analyt ical tool. IR spectroscopy has been widely applied for the mo lecular structure characterizat ion of lipids and proteins. As outlined IR spectroscopy detects mo lecular vibrat ions accompanied by changing mo lecular dipo le moments. As a consequence the vibration frequencies that are detected are sensit ive to mo lecular conformat ion. Thus, spectroscopy is a useful method to invest igate lipid structure and conformat ion. While IR absorption studies of biosystems have been well established in the study o f proteins there remains a number o f limitations in the informat ion that IR absorption spectroscopy can provide. For example, when studying mircon or nanosized features IR absorption spectroscopy is limited by the diffraction limit (Abbe 1873; Rice 2007). The maximum image reso lution in optical microscopy is found to be ca. /2. As a result of diffract ion, the image reso lut ion will be 2.5 m when imaging in the infrared using electromagnet ic radiat ion at 5m (corresponding to an IR absorption frequency of 2000 cm-1). This means that IR absorption spectroscopy techno logy cannot be applied to study features that are approximately smaller than a few microns. A newly emerging method for IR absorption spectroscopy using a photothermal based methodology enables IR spectroscopy o f smaller amounts of materials than is currently possible using established IR absorption methods. This new method uses an Atomic Force Microscope (AFM) cant ilever tip as the detection mechanism (Hammiche et al 2004; Hammiche et al 1999; Hammiche et al 2004). One particular method based on this approach, referred to as AFMIR, measures IR absorption direct ly via measuring local transient deformation in the sample via the AFM cant ilever which is induced by an IR pulsed laser tuned at a vibrat ion absorbing wavelength (Dazzi et al 2005; Houel et al 2007; Mayet et al 2008; Hill et al 2009; Rice 2010). This enables IR absorption spectra to be recorded of features as small as the AFM cant ilever t ip. Studies using this approach have reported the study of quantum dot nanomaterials with a spat ial reso lution of 60 nm (Houel et al 2007). To date the experimental methodology o f AFMIR has utilised attenuated total interna l reflect ion arrangement in combinat ion with IR cyclotron radiat ion or a top down configuration using a customised IR laser source. Here, we outline work performed on an experimental method for IR surface spectroscopy that samples direct ly via a novel bottom-up optical arrangement using a commercial laser system. The advantage of this set-up is that is uses a commercia l laser rather than synchrotron radiat ion and allows the use of commonly used substrates such as glass or mica. In this letter, this novel AFMIR set-up is applied for the first time to study hemoglobin. The work presented here demonstrates that AFMIR can be applied to study micron sized aggregations of protein. Methods The AFMIR experimental set-up is shown in Fig 1a. The experimental configurat ion consist ing o f an optical parametric oscillator (OPO) laser and an AFM. The excitat ion IR radiation is directed upward in a novel configuration using gold coated mirrors to direct the laser light. The sample was mounted onto a glass side to facilitate this optical arrangement. IR radiat ion was generated using an OPO laser (Cohesion) based on a periodically poled LiNbO3 crystal emitting IR laser radiation that is tuneable over > 3.0 to 3.6 m. The output power was c.a. 2 mW. The laser was focused to a relat ively large spot of c.a. 500 m on the sample in order to cover the ent ire area probed. The energy was low enough to avoid damaging the sample. An AFM (Veeco Explorer system) was used with a scanner with lateral and vertical dimensions of 100 x 100 m and 8 m respectively. The AFM is operated in contact mode enabling simultaneous IR and topography measurements. Silicon nitride t ips mounted on a V-shaped cant ilever with a nominal spring constant of 0.05 N/m (Veeco) were used. A force setpoint of 1 - 3 nN was used. Samples were prepared on standard microscope glass slides. A Fourier Transform IR (FTIR) spectrometer (Varian model 3100) was used to record a reference IR spectrum. Results The AFM tip was posit ioned over the sample with the tip in contact with the sample surface. Fo llowing absorption o f the incident radiat ion, the energy absorbed is dissipated through thermal and acoustic mechanisms. Propagating acoustic waves create a deformation in the surface topography which can be detected by displacement of the AFM tip (Dazzi et al 2005; Rice 2010). As an IR laser source is tuned into resonance with a vibrat ion mode, absorption of IR radiat ion increases. The response of the cant ilever tip was monitored fo llowing the applicat ion o f the IR radiat ion. AFMIR studies of a deposited layer o f hemoglobin on a glass slide were undertaken. An AFM topography image of the surface was recorded (as shown in Fig 1b). A small area of the sample surveyed in the AFM topography image (marked  in Fig 1b) was selected for study. The hemoglobin feature in this area was c.a. 1 m higher than the surrounding layer and has a diameter of c.a. 200 nm. The lateral size of the sample area probed is proportional to the size of the AFM tip (i.e. around 20 nm). Fig 2a shows the oscillat ion of the AFM cant ilever fo llowing absorption of the IR laser pulse by the sample. The intensity of the cant ilever oscillat ion changes on resonance (2960 cm-1) and o ff resonance (2810 cm-1) with the C-H stretching mode o f hemoglobin. The intensity o f the oscillat ion o f the cant ilever oscillat ion as a funct ion of wavelength was recorded. The result ing AFMIR spectrum of hemoglobin is shown in Fig 2b. The AFMIR spectrum is shown alongside the FTIR spectrum of hemoglobin. The two spectra show very similar features. The AFMIR spectra shown in Fig 2.b possess a wavelength resolut ion o f 15 cm-1, while the FTIR based spectrum has a wavelength resolut ion of 2 cm-1. Comparing the AFMIR and FTIR spectra shows that they possess similar spectral features. Both spectra show the presence o f peaks (marked on a broad background. The position o f these peaks corresponds to crystalline hemoglobin. Kuenstner et al (2000) reported the position o f peaks for crystalline hemoglobin to be 2871.5, 2960.2 and 3060.5 cm-1 which match the posit ion o f the peaks seen in both the AFMIR and FTIR spectra. The bands seen in the spectra arise from N-H and C-H vibrat ions. The amide B band at 3061 cm-1 is assigned to an intramo lecular hydrogen-bonded N-H stretching or to an overtone band (i.e. 2 x 1541 cm-1) (Kuenstner et al 2000). The band at 2960 cm-1 is due to aliphat ic C-H stretching. Changes in the relat ive intensit ies in some peaks are present when comparing the AFMIR and FTIR spectra. Normalising to the peak at  the peak at  is reduced in intensity, while peak at  is reduced compared to . These changes in relat ive intensity may be associated with the microsized particle measured inducing small changes in the conformat ion o f the protein. Studies o f smaller nanosized protein layer features were undertaken. Fig 3a shows an AFM image o f hemoglobin sample on a glass surface. Two regions were probed which were c.a. 150 nm different in height (denoted by  and  . Studies o f these two regions showed that difference in the AFM tip oscillat ions on (region ) and off (region ) the heme ledge could be seen at a wavelength of 2960 cm-1, which corresponds to on resonance (see Fig 3b). The thinner part of the sample showed a weaker AFM tip oscillat ion than the thicker part of the sample (i.e. regions denoted by  and   This shows that small nanosized features can be discerned using AFMIR. Changing the wavelength to a wavelength that corresponds to off resonance (i.e. 2810 cm-1) led to the same effects in the oscillat ions as shown in Fig 2a. In conclusion, IR absorption spectroscopy o f hemoglobin was performed using an IR optical parametric oscillator laser and a commercial atomic force microscope in a novel configurat ion. This experimental approach enables detection of micron sized protein features. This reso lut ion cannot be achieved by standard IR spectroscopy. This  methodology for IR spectroscopy can potentially be applied to study other protein structures on the micro- and nanometre length scales. Acknowledgements The authors would like to acknowledge Science Foundation Ireland for supporting this research. References Abbe, E. 1873. Beitrage zur theorie des Mikroskops und der mikroskopischer Wahrnehmung. Arch. Mikroskop. Anat. 9:413-487. Charbonneau, D., M. Beauregard, and H. A. Tajmir-Riahi. 2009. Structural analysis of human serum albumin complexes with cationic lipids. J. Phys. Chem. B. 113:1777- 1784. Dazzi, A., R. Prazeres, F. Glotin, and J. M. Ortega. 2005. Analysis o f nano-chemical mapping performed by an AFM-based (“AFMIR”) acousto-optic technique. Ultramicroscopy. 107:1194-1200. Hammiche, A., L. Bozec, H. M. Pollock, M. German, and M. Reading. 2004. Progress in near-field photothermal infra-red microspectroscopy. J. Microsc. 213:129-134. Hammiche, A., M. H. Pollock, M. Reading, M. Claybourn, P. M. Turner, and K. Jewkes. 1999. Photothermal FT-IR spectroscopy: A step towards FT-IR microscopy at a resolution better than the diffract ion limit. Appl. Spectrosc. 53:810-815. Hammiche A., L. Bozec, M. J. German, J. M. Chalmers, N. J. Everall, G. Poulter, M. Reading, D. B. Grandy, F. L. Mart in, and H. M. Pollock. 2004. Mid-infrared microspectroscopy photothermal near-field using samples difficult of microspectroscopy. Spectrosc. 19:20-40. Hill, G., J. H. Rice, S. R. Meech, P. Kuo, K. Vodopyanov, and M. Reading. 2009. Nano-infrared surface imaging using an OPO and an AFM. Optics Lett. 34:431-433. Houel, J., S. Sauvage, P. Boucaud, A. Dazzi, R. Prazeres, F. Glotin, J.M. Ortega, A. Miard, and A. Lemaıtre. 2007. Ultraweak-absorption microscopy of a single semiconductor quantum dot in the midinfrared range. Phys. Rev. Lett. 99:217404. Kuenstner, J. T., K. Norris, and V. F. Kalasinsky. 2000. Spectrophotometry of Human Hemoglobin in the Midinfrared Region. Biospectroscopy. 3:225-232. Mayet, C., A. Dazzi, R. Prazeres, F. Allot, F. Glotin, and J. M. Ortega. 2008. Sub-100 nm IR spectromicroscopy o f living cells. Optics Lett. 33:1611-1613. Rice, J. H. 2010. Nanoscale optical imaging by atomic force infrared microscopy. Nanoscale. 2:660-667. Rice, J. H. 2007. Beyond the diffraction limit : far-field fluorescence imaging with ultrahigh reso lut ion. Mol. BioSyst. 3:781-793. Vincent, J. S., C. J. Steer, and I. W. Levin. 1984. Infrared spectroscopic study of the pH-dependent secondary structure of brain clathrin. Biochemistry. 23:625-631. Fig 1. a) Schematic drawing of the experimental set-up, b) AFM image of hemoglobin sample on a glass surface (100 x 100 m area). Left, topography and right, height profile recorded at the posit ion o f the white line. The arrow shows the region probed by AFMIR, which corresponds to . Fig 2. a) Plot of AFM tip oscillat ions on (2960 cm-1) and off (2810 cm-1) resonance, b) IR absorption spectra for hemoglobin. Fig 3.a) AFM image of hemoglobin sample on a glass surface (500 x 500 nm area). Top, topography and bottom, height profile recorded at the posit ion o f the black line. The areas probed are denoted by  and . b) Plot of AFM t ip oscillat ions on resonance (2960 cm-1) on and off the heme ledge (and respectively).
1802.00986
1
1802
2018-02-03T15:12:02
Lipid-protein interaction induced domains: kinetics and conformational changes in multicomponent vesicles
[ "physics.bio-ph", "cond-mat.soft", "cond-mat.stat-mech", "physics.chem-ph" ]
The spatio-temporal organization of proteins and the associated morphological changes in membranes are of importance in cell signaling. Several mechanisms that promote the aggregation of proteins at low cell surface concentrations have been investigated in the past. We show, using Monte Carlo simulations, that the affinity of proteins for specific lipids can hasten its aggregation kinetics. The lipid membrane is modeled as a dynamically triangulated surface with the proteins defined as in-plane fields at the vertices. We show that, even at low protein concentrations, strong lipid-protein interactions can result in large protein clusters indicating a route to lipid mediated signal amplification. At high protein concentrations the domains form buds similar to that seen in lipid-lipid interaction induced phase separation. Protein interaction induced domain budding is suppressed when proteins act as anisotropic inclusions and exhibit nematic orientational order. The kinetics of protein clustering and resulting conformational changes are shown to be significantly different for the isotropic and anisotropic curvature inducing proteins.
physics.bio-ph
physics
AIP/123-QED Lipid-protein interaction induced domains: kinetics and conformational changes in multicomponent vesicles K. K. Sreeja1, a) and P. B. Sunil Kumar2, b) 1)Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India, Department of Chemical and Bio-molecular Engineering, University of Pennsylvania, Philadelphia, PA 19104, USA. 2)Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India, Department of Physics, Indian Institute of Technology Palakkad, Palakkad, 678557, India. (Dated: February 6, 2018) The spatio-temporal organization of proteins and the associated morphological changes in membranes are of importance in cell signaling. Several mechanisms that promote the aggregation of proteins at low cell surface concentrations have been investigated in the past. We show, using Monte Carlo simulations, that the affinity of proteins for specific lipids can hasten its aggregation kinetics. The lipid membrane is modeled as a dynamically triangulated surface with the proteins defined as in-plane fields at the vertices. We show that, even at low protein concentrations, strong lipid-protein interactions can result in large protein clusters indicating a route to lipid mediated signal amplification. At high protein concentrations the domains form buds similar to that seen in lipid-lipid interaction induced phase separation. Protein interaction induced domain budding is suppressed when proteins act as anisotropic inclusions and exhibit nematic orientational order. The kinetics of protein clustering and resulting conformational changes are shown to be significantly different for the isotropic and anisotropic curvature inducing proteins. Keywords: membrane remodeling, domain formation, lipid-protein interaction, monte carlo simulations 8 1 0 2 b e F 3 ] h p - o i b . s c i s y h p [ 1 v 6 8 9 0 0 . 2 0 8 1 : v i X r a a)Electronic mail: [email protected] b)Electronic mail: [email protected] 1 I. INTRODUCTION Protein redistribution and clustering on the cell surface are important for signal trans- duction pathways1. At its low physiologically relevant cell surface concentrations, direct interaction between proteins cannot be the primary cause for clustering, and many ac- tive and passive mechanisms, that indirectly aid protein clustering have been proposed2–5. The specialized membrane domains known as rafts, which is the result of a sterol and sphingolipid enriched aggregation, are believed to be one of the precursors for the pro- tein clustering process1,6,7. Such membrane domains are often associated with peripheral and integral proteins8–13. Rafts are known to make a suitable platform for aggregation of GPI(glycosylphosphatidylinositol)- anchored proteins, which correspond to a set of exoplas- mic, eukaryotic proteins exhibiting specific intracellular sorting and signaling properties5,14. Caveolin and clathrin are some of the non raft proteins associated with lipid domains2,15,16. Caveolae are glycolipid enriched domains, that are flask like invaginations formed by the as- sembly of Caveolin proteins. It is not clear if these lipid-protein domains arise from the direct interactions between the proteins or due to the interaction between non-protein membrane constituents and the affinity of proteins to certain membrane composition17,18. Another factor that has lead to considerable interest in understanding the mechanisms behind lipid-protein sorting in biological membranes is the asymmetric distribution of lipids and proteins in the intercellular organelles such as the golgi and endoplasmic reticulum19,20. Lipids can dynamically vary the constituents of a membrane by selectively recruiting various proteins which in turn can change the functionality of the membrane, and similarly proteins can sort lipids to specific membrane locations through steric or electrostatic interactions21–23. The membrane protein aggregation due to lipid-protein interactions has been studied using coarse grained molecular dynamics approaches when the length scale of interest are of few tens of nanometers23. Since our aim here is to explore the role of lipid-protein interac- tion in the formation of domains and since the length scale of the resulting conformational changes are much larger than the membrane thickness, we consider a mesoscale computa- tional approach. Existing computational studies on the equilibrium or dynamic properties of membranes membranes mostly deals with lipid phase separation following a quench into the two phase coexistence regime24–27. Experimental validation of the results from mesoscale simulations28 has motivated further studies on the dynamics of these domains29. However, 2 there are very few attempts to understand how the interaction of proteins with other mem- brane constituents can lead to compositional inhomogeneities and clustering of proteins. Here we use a Monte Carlo model for a three component fluid vesicle, with one component representing protein inclusions and the others two different compositions of lipids, to inves- tigate how lipid-protein interactions affect the kinetics of domain formation and associated conformational changes in the vesicle. To study protein clustering in detail, in our model, we also account for the direct protein-protein interactions and the membrane curvature inducing properties of the proteins. The paper is organized as follows. In Section II, we describe the Monte Carlo model for a multi-component membrane. In the results and discussions given in Section III, we first focus on the lipid domain formation. A comparison of the kinetics of lipid-protein interaction induced domain formation with that due to lipid-lipid interaction is presented in Section III A. Conformational changes of the membrane for different protein concentrations are also discussed here. In Section III A 1, we present results on the effect of combined lipid- lipid and lipid-protein interaction on domain growth. In the second part of Section III, we present our results on protein cluster growth with varying lipid and protein concentrations. Section III B 2 and III B 3 is dedicated, respectively, to discussion on how explicit protein- protein interactions and the curvature remodeling activity of proteins affect cluster growth. Our main results are the following. We show that the affinity of proteins for certain type of lipids can lead to formation of large protein clusters even at low concentration of proteins. The increase in protein cluster growth rate, due to strong lipid-protein interaction, indicate a route to lipid mediated signal amplification. It is pointed out that the absence of line tension at the domain boundary, at low protein concentration, is the primary reason for enhanced kinetics of domain growth. The domain growth is slower at high protein concentrations and when proteins act as anisotropic inclusions to exhibit nematic orientational order. II. MODEL In the Monte Carlo simulations carried out here, the conformation of a lipid membrane is approximated to be that of an elastic sheet represented by a randomly triangulated surface. In this scheme, a vesicle of spherical topology is represented by Nv vertices, NL = 3(Nv − 2) links and NT = 2(Nv − 2) triangles and the triangulation is changed randomly to simulate 3 the in-plane fluidity of the membrane. In the case of an isotropic and homogeneous mem- brane, the energetics of the elastic sheet is described using a discretized form of the Helfrich Hamiltonian, Helastic = κ 2 Nv(cid:88) v=1 (2Hv − C0)2Av, (1) where summation is over all the vertices of the triangulated membrane. Here κ is the bending rigidity of the membrane, C0 is the spontaneous curvature resulting from lipid asymmetry of protein induced deformations in the bilayer, Hv is the mean curvature defined at the vertex and Av is the area associated with the vertex v, computed as in Ramakrishnan et. al.30. To model a multicomponent vesicle with two coexisting lipid compositions and one type of protein, we introduce a lipid composition field φv and a protein field pv = 1, both positioned on the membrane vertices. The lipid field φv = 1 when the composition of lipid at vertex v is labelled A and φv = −1 if the lipid composition is labelled B. Similarly pv = 1 in the presence of a protein at the vertex and pv = 0 otherwise. A vertex can be simultaneously occupied both by the protein and the lipid fields. Lipid-lipid interactions are assumed to be Ising like. The two component lipid vesicle model has been previously used to study the lipid induced phase separations and budding of domains25. The Hamiltonian describing explicit lipid-lipid interactions is given as Hφ = −Jφ (cid:88) (cid:104)vv(cid:48)(cid:105) φvφv(cid:48), where the summation runs over all the neighboring pairs. The lipid-protein interaction is modeled using the Hamiltonian, (cid:88) Hpφ = −Jpφ pvφv(cid:48). (2) (3) (cid:104)vv(cid:48)(cid:105)(cid:48) The prime on the summation indicates that the protein field at any vertex is allowed to interact with the lipids within the one ring neighborhood including its own vertex. When we choose Jpφ > 0, the lipid-protein interaction is attractive between the proteins and type A lipids. Since A lipids are miscible in B lipids and is the minority component, in this article we will often refer to lipid composition of type A as co-lipid. The presence of curvature active proteins modulate local membrane shapes by inducing spontaneous curvature. We study two classes of curvature generating proteins: the first kind of proteins induce a uniform mean curvature on the membrane and the second kind 4 Figure 1. An illustration of a multicomponent vesicle: type A lipids are marked as spheres while the unmarked vertices correspond to type B lipids. The protein inclusions are shown by black lines. are structurally anisotropic proteins, inducing directional curvatures. To model an isotropic curvature generating protein we consider the spontaneous curvature C0 (see Eqn.1) at any vertex to be nonzero in presence of a protein. The second class of proteins or protein complexes have an extended structure and cannot be considered as point like objects31. To incorporate the structural anisotropy of such a protein into the model we introduce a unit orientation vector nv such that the in-plane protein field is now a vector pv = pv nv. The anisotropic shape of the protein is reflected by the rotational asymmetry of nv. In this paper we consider only the case wherein nv has a π rotational symmetry representing elongated protein inclusions30. The protein field thus acts like a nematic orientational field on the membrane. The explicit orientational interaction between the proteins on the membrane is modeled using the Lebwohl-Lasher model32, given by HLL = −LL pvpv(cid:48) cos2 Φ(nv, n(cid:48)v) − 1 2 (cid:88) (cid:104)vv(cid:48)(cid:105) (cid:110)3 2 (cid:111) , (4) where, Φ(nv, n(cid:48)v), the angle between the two in-plane field vectors on the tangent planes at neighboring vertices v and v(cid:48), is computed using a parallel transport operation30. The protein orientation field is coupled to the membrane curvature using a discretized version of the Hamiltonian proposed by Frank and Kardar33, − C(cid:107)0 )2 + Hanis = (Hn,(cid:107) κ⊥ 2 (Hn,⊥ (cid:104) κ(cid:107) 2 (cid:88) v − C⊥0 )2(cid:105) Av. (5) Hn,(cid:107) and Hn,⊥ are the curvatures along the directions nv and n⊥v respectively. C(cid:107)0 and C⊥0 5 are the local directional spontaneous curvatures and κ(cid:107) and κ⊥ are the bending stiffness along nv and n⊥v , respectively. The anisotropic protein inclusions considered in our study generate additional stiffness and curvatures only along the direction n, i.e., we consider only the cases with κ⊥ = 0. The multicomponent vesicle is equilibrated through a set of Monte Carlo (MC) moves with the total effective Hamiltonian: Htotal = Helastic + Hpφ + Hφ + Hanis + HLL. (6) We only consider the case of conserved in-plane fields and the MC moves include the vertex moves, link flips, p field rotations and the exchange of p and φ fields 30,34. Unless otherwise stated, all moves are accepted through the Metropolis algorithm. (a) Vertex move: Here the position of a vertex is updated to a new position chosen randomly within a cutoff distance. This cutoff is fixed such that 50% of the moves are accepted. This move allows for shape changes of the vesicle. (b) Link flip: A randomly selected edge, connecting two triangles, is disconnected and a new connection between the unconnected vertices of the same triangles is constructed. This move changes the triangulation/connectivity and physically models the fluidity of the bilayer membrane by allowing the vertices to diffuse through the surface. (c) Exchange of φ fields: The diffusion of lipid composition field on the membrane surface is captured using a Kawasaki move that allows for an exchange between type A and type B vertices. (d) Exchange of p fields: The diffusion of protein field on the membrane surface is captured using a Kawasaki exchange move. (e) p field rotation: In this step, a vertex is chosen randomly and the orientation field at the chosen vertex, if nonzero, is rotated to a new, randomly chosen direction in the tangent plane. This rotation of the field allows for the relaxation of the orientational order of the field. The multicomponent membrane system described here is studied using vesicles with 2030 vertices and 4056 triangles. We consider a vesicle with bending rigidity κ = 10 kBT . Ini- tial configurations of the vesicle are generated by randomly assigning φ% of the membrane vertices to have lipids of type A and the rest of the vertices are assigned to have type B lipids. The proteins, whose number fraction is represented by p%, are also placed at ran- domly chosen vertices with random in-plane orientations. A patch of the membrane with co-existing lipid and protein fields is shown in Fig. 1 and we follow the same representation for further discussions. It should be noted that even when the proteins are not anisotropic in 6 nature they are represented by solid black lines in order to distinguish them from lipid-type specification on the vertices. III. RESULTS AND DISCUSSIONS A. Lipid clustering due to protein-lipid interaction To investigate membrane inhomogeneities induced by the lipid-protein interactions, we first consider the case of isotropic inclusions, such that Hanis and HLL = 0, no direct lipid- lipid interactions (i.e. Hφ = 0) and focus on the aggregation kinetics that is solely driven by lipid-protein interaction Jpφ > 0. For these parameters, the proteins only serve to promote the aggregation of lipids and do not have any direct impact on local membrane shapes. In our model the lipids and proteins can occupy the same vertex and lipid-protein interaction is limited to nearest neighbor sites. The relative values of φ% and p% is thus another important parameter. Below we first analyze the cluster growth when p% < φ%. Figure 2. A comparison of domain formation with p − φ and φ − φ interactions as a function of MC steps. (a) Conformations of a vesicle when lipids aggregate through p − φ interactions for Jφ = 0, Jpφ = 2, φ% = 30 and p% = 20. (b) Lipid clusters induced by φ - φ interactions for Jφ = 2, φ% = 30 and p% = 0. 7 MC steps(a)(b) Clustering kinetics: The evolution of membrane inhomogeneities and the associated vesi- cle conformations, as a function of MC time, are shown in Fig. 2. Panel (a) shows p − φ interaction driven clustering when Hpφ > 0, Hφ = 0, φ% = 30 and p% = 20. For comparison panel (b) shows membrane conformations for lipid-lipid interaction induced aggregation, when Hpφ = 0, Hφ > 0, φ% = 30 and p% = 0. As can be inferred from the figure, the kinetics of clustering and conformational changes in the membrane are significantly different in these two cases. The evolution of vesicle shape, resulting from direct lipid-lipid inter- actions, shown in panel (b) of Fig. 2, is similar to that observed in previous studies25. At early times (MC steps ≤ 104), in both cases, domains of co-lipids nucleate and grow into stable clusters as shown in Fig. 2. As can be seen in panel (a), p− φ interaction is sufficient to induce an effective attraction between the lipids and trigger the formation of co-lipid patches. The late time growth of domains, however, is strongly dependent on the nature of interactions. When φ% > p%, in the case of protein induced segregation, shown in panel (a), the domains remain flat and their aggregation is fast. While in panel (b), the presence of explicit lipid-lipid interaction leads to budding and slowing down of coarsening. Figure 3. Comparison of lipid cluster growth with p − φ and φ − φ interactions when φ% = 30. Average cluster size of type A domains for different values of Jφ and Jpφ is given. When Jφ = 0 (filled symbols) protein concentration is fixed at p% = 20 while for Jφ > 0 (open symbols), it is taken to be p% = 0. A quantitative comparison of domain growth can be obtained by analyzing the average 8 103104105106MCsteps(t)102103CsJφ=1;Jpφ=0Jφ=2;Jpφ=0Jφ=3;Jpφ=0Jφ=0;Jpφ=1Jφ=0;Jpφ=2Jφ=0;Jpφ=3t1/2 lipid cluster size as a function of time. The average cluster size for different values of the interaction strengths Jφ and Jpφ are given in Fig. 3. The cluster sizes (Cs) correspond to the number of vertices of type A that form a continuous map on the triangulated surface. The early time domain growth is similar in all cases. When the coarsening is only through lipid-lipid interaction strength (i.e., for Jφ > 0 and Jpφ = 0), for all values of Jφ, and for the system sizes considered here, the growth rate remains low and the system does not enter into a scaling regime. The coarsening is considerably faster with lipid-protein interaction Jpφ > 0 and we see a power law regime with Cs ∝ t1/2, which could result from domain diffusion snd coalescence (see Appendix A), as the diffusion coefficient of domains, in the Rouse dynamics, varies as the inverse of the domain size (data not shown). The corresponding configurations are shown in Fig. 4. It is important to note that, for Jpφ ≤ 3, Jφ = 0 and with φ% > p%, the domains remain flat. This is evident from the Ls ∝ t1/4 dependence of interfacial length Ls on time as shown in Fig. 5. On the other hand, in the case of clustering induced by direct lipid-lipid interactions, when the value of Jφ is higher, the line tension is usually significant enough to induce budding. This regime can be easily identified in Fig. 5 as one with sudden fast decrease of interface length. The movement of domains, which now involve membrane shape changes, significantly slows down domain coarsening. Figure 4. Coalescence of protein rich domains on the vesicle surface due to p− φ interactions. The cluster sizes shown are for φ% = 30, p% = 20, Jpφ = 2 and Jφ = 0. It is clear from the above discussions that the ability of domains to remain flat is important for fast clustering of proteins. In the model, positive values of Jpφ favor type A lipids to 9 103104105106MCsteps(t)102103CsJ=0;Jp=2 Figure 5. Comparison of lipid domain interfacial length for φ− φ and p− φ interactions for various values of Jφ and Jpφ when φ% = 30. When Jφ = 0 protein concentration p% = 20 and Linterface is computed as the number of vertices occupied by the A lipids with atleast one type B lipid vertex as a neighbor. occupy a vertex with a protein on it or in the one ring neighborhood of it. When the fraction of vertices occupied by the protein is much smaller than that of A lipids (φ% > p%), there are enough A lipids to occupy the one ring neighborhood even when there are many small protein clusters. In this regime, when Jφ = 0, co-lipids at the boundary of the domains will act as surfactants and we expect the interfacial tension at the lipid-protein domain boundaries to be negligible resulting in flat domains. Such a surfactant lined domain could stabilize small clusters and prevent coarsening. But our simulations show a complete coalescence of domains, indicating that the entropy gain from release of excess co-lipids is significant enough to drive domain coarsening. Thus the protein induced domains show faster coalescence compared to the domains formed by direct lipid-lipid interaction. Conformational changes: When Jφ = 0, there are two main factors that affect the conformational changes of the membrane; (i) the fraction of vertices occupied by the proteins (p%) in comparison to that occupied by co-lipids (φ%) and (ii) the interaction strength Jpφ. Fig. 6 shows the representative equilibrium conformations for different values of p% and Jpφ for a fixed value of φ% = 30. When p% = 10, large clusters are formed, but these clusters do not initiate budding even for large value of Jpφ. As described in the previous section, 10 103104105106MCsteps(t)102LinterfaceJφ=1;Jpφ=0Jφ=2;Jpφ=0Jφ=3;Jpφ=0Jφ=0;Jpφ=1Jφ=0;Jpφ=2Jφ=0;Jpφ=3t−1/4 this is due to the presence of additional type A vertices at the interface that shields the protein from type B vertices and reduces the line tension. The domains start to deform when p% = 20 and Jpφ = 2. At p% = 30 the number of proteins become equal to the number of type A vertices, and there are no additional co-lipids to line the domain interface. In this case when the lipid-protein domain size reaches a certain value it starts to bud. Figure 6. Phase separation and shape changes as a function of p% and Jpφ when φ% = 30 and Jφ = 0. Lipid domains starts to bud when p% = 20 and Jpφ = 2. The snapshots are taken after 2 × 107 MC steps. 1. Effect of lipid-protein interaction at low co-lipid concentration In biological membranes both lipid-lipid and lipid-protein interactions are expected to affect the protein phase-segregation and hence a study of combined Hφ and Hpφ interactions are of relevance to cellular membranes. Here we study the effect of lipid-protein interactions by holding the lipid-lipid interaction fixed at Jφ = 1. The average cluster size in this case when φ% = 30 and p% = 20 is shown in Fig. 7. Here as Jpφ increases the growth rate of lipid cluster size decreases in a monotonic fashion. When there is not enough co-lipids to cover all protein neighborhood and the protein domain boundary is occupied by both A and B lipids, non zero protein B lipid interaction, parameterized through Jpφ, will increase the line tension at the boundary, resulting in domain shape changes. Such morphological changes 11 Jp=1Jp=2p%=10p%=20p%=30 reduces domain diffusion and leads to a slow down in coalescence of domains and increases the life time of smaller domains. Figure 7. Average cluster size with lipid-protein and lipid-lipid interactions when φ% = 30 and p% = 20 for Jφ = 1 and Jpφ = 0, 0.5, 1 and 2. B. Protein clustering due to lipid-protein interaction In this section we investigate how the lipid-protein interaction can change the effective protein clustering and the structural properties of protein inclusions affect protein cluster- ing. In the following discussion we consider domain formation without explicit lipid-lipid interactions and constant lipid-protein interactions, i.e., we take Jφ = 0 and Jpφ = 2. 1. Concentration of proteins and lipids Here we study the aggregation of proteins at a smaller protein concentration compared to the previous cases. We keep p% = 10 and study the domain formation by varying φ%. The resulting conformations and the largest cluster sizes are shown in Fig. 8. As shown in panel(b), the domains bud for equal concentration of co-lipids and proteins. When φ% > p% the line tension decreases and hence the domains remain flat. The increase in cluster sizes due to the flat domains can be seen in Fig. 8(a). When φ% ≥ 20, at early times we observe a domain growth that depends on the fraction of co-lipids present and at late times a complete 12 103104105106MCsteps(t)101102CsJφ=1;Jpφ=0Jφ=1;Jpφ=0.5Jφ=1;Jpφ=1Jφ=1;Jpφ=2 Figure 8. Protein aggregation at low protein concentration as a function of φ% for p% = 10, Jφ = 0 and Jpφ = 2. Panel (a) shows the size of largest protein cluster as a function of time. Panel (b) shows representative vesicle conformations at MC steps= 5 × 106. clustering of proteins. Even when φ% >> p% we observe a large cluster instead of small clusters. This could be due to the increase in entropy due to the configurational freedom of proteins in a larger cluster compared to small domains. Complete protein clustering, even at low values of protein concentration and in the ab- sence of any direct protein-protein interaction, can be achieved through lipid-protein inter- action is one of the main results presented in this paper. To understand this more quanti- tatively we also looked at lipid-protein aggregation as a function of p% when φ% = 30. In Fig. 9(a) we show Clp/p% to represent the fraction of proteins clustered and in Fig. 9(b), the corresponding conformations at 5 × 106 MC steps. For low protein concentrations (p% = 5, 10) a complete clustering of proteins is observed. The fraction of protein clustered is mini- mum when p% = 30, which is equal to φ%. This is expected as the line tension and membrane deformations are maximum when p% = φ%. 13 103104105106MCsteps(t)102Clp%=5%=10%=20%=30%=40(a)(b) Figure 9. Protein clustering as a function of p% when φ% = 30, Jφ = 0 and Jpφ = 2. Panel (a) shows Clp/p% as a function of time and (b) the vesicle conformations at MC steps= 5 × 106. 2. Explicit protein-protein orientational interactions In this section we focus on the effect of a direct interaction between the proteins in ad- dition to its interaction with lipids. Orientational interaction between proteins are relevant for elongated protein inclusions. The proteins are now treated as in-plane vector fields and their nematic like orientational interaction is modeled through the Lebwohl-Lasher energy functional35, which is given in Eqn. 4. This interaction is short ranged and is only between in-plane fields (proteins) within the connected neighborhood and captures the symmetry property of the nematic vectors. In an earlier study it has been shown that this orienta- tional interaction (HLL) alone can aggregate proteins on the surface and induce significant membrane conformational changes30. In Fig. 10 we show the time evolution of protein domains on the vesicle with orientational order and fixed Hpφ interaction; panel (a) shows the largest protein cluster size and panel (b) the vesicle conformations as a function of time when LL = 3. We consider the case with p% = 20 and φ% = 30, to compare with the case discussed in Sec III A. From Fig. 10(a) it can be seen that as LL increases the cluster growth slows down. As LL increases the additional 14 103104105106MCsteps(t)101100Clp/p%p%=5p%=10p%=20p%=30p%=40(a)(b) attractive interaction between the components reduces the diffusivity of the protein cluster which in turn reduces the rate of growth of clusters. When LL = 0 (see panel (a) of Fig. 2), Hpφ interaction induces domains that are nearly circular in shape which, at a later stage, coalesce and bud off. When LL > 0 (see panel (b) of Fig. 10), domains with non-circular boundaries are formed first. Within a domain the protein fields are aligned in one direction. As time progress the domain boundaries take a circular shape. However, due to the energy cost in maintaining parallel orientation of the protein field, the domains do not bud as in the case with LL = 0. Figure 10. Formation of protein clusters with explicit protein-protein interaction when Jpφ = 2, Jφ = 0, p% = 20 and φ% = 30. Panel (a) is the largest protein cluster size and (b) the vesicle conformations when LL = 3. 3. Curvature active proteins Collective interactions of curvature generating membrane associated proteins, like BAR domain proteins, caveolin and clathrin can lead to interesting shape transformations in or- ganelle membranes. Protein induced lipid sorting has been shown with compelling evidences 15 103104105106MCsteps(t)102Clp✏LL=0✏LL=1✏LL=3✏LL=5MC steps(a)(b) in the trans-golgi and endosomal membranes and it is believed that sorting happens in re- sponse to the curvature induced by the proteins2,15,16,36. The presence of curvature inducing and curvature sensing proteins in lipid domains can significantly alter the composition of the domain since these proteins are known to have a strong affinity for certain class of lipids37–39. Here we discuss how the protein field induced curvature along with lipid-protein interactions, can alter the protein domain growth and the shape of the vesicle. The domain formation due to two different classes of proteins: isotropic and anisotropic curvature generating proteins, are considered. Aggregation due to isotropic curvature generating proteins: The spontaneous curvature generated by proteins is accounted via Eqn. 1 by assigning C0 > 0 to the vertices occupied by the proteins. The clustering kinetics and the corresponding conformations when φ% = 30, p% = 20 are shown in Fig. 11. As can be seen in Fig. 11(a), at early times (MC steps ≤ 105) the growth curve is nearly independent of the value of C0. For MC steps ≥ 105 the protein cluster size saturates at a value that decreases with increasing C0. The membrane deformations due to curvature active proteins are observed only when these proteins form a cluster. When the cluster size reaches a threshold, at a value which depend on C0, membrane starts to deform and this is reflected in the diffusion and growth of the clusters. In Sec III B 1 we have seen bud formation, induced by line tension, takes place only when φ% ≤ p%. Here we show that such buds can be formed even when φ% > p% if the proteins are curvature active. Earlier such effects, of spontaneous curvature on vesicle conformations, have been studied only with direct lipid-lipid interaction induced clustering25. Aggregation due to anisotropic curvature generating proteins: In order to study the role of anisotropic curvature induction in protein clustering we introduce curvature effects through the anisotropic curvature Hamiltonian given in Eqn. 5. We concentrate on the conformations of the membrane when φ% = 30, p% = 20 and κ⊥ = 0. It was shown earlier that, at these concentrations of anisotropic proteins with LL = 0 protein fields cannot induce large scale aggregation through membrane curvature mediated interaction alone34 and Hpφ interactions are necessary for them to aggregate. The cluster formation for different values of C(cid:107)0 are shown in Fig. 12(a). The largest protein cluster size Clp at short time scales are found to be strongly dependent on the induced curvature. For example, Clp for C(cid:107)0 = 0.75, LL = 0, exhibits the fastest growth, showing that the proteins can enhance the aggregation through a membrane mediated interaction 16 Figure 11. Cluster formation in the presence of isotropic curvature generating proteins as a function of time when Jφ = 0 and Jpφ = 2. Panel (a) shows the largest protein cluster size for different values of C0 and panel (b) the conformations of the vesicle for C0 = 0.3. between the protein fields. The membrane conformations in the presence of curvature active proteins when Jpφ = 2 are shown in Fig. 12(b). Though no explicit orientational interaction between the proteins are included, the proteins in a domain tend to align due to similar induced curvature. At early time the clusters grow faster than those seen for C(cid:107)0 = 0 (refer Fig. 2(a)). At later times, these domains coalesce resulting in larger clusters until they are big enough to deform the membrane to form tubular structures. Such a large scale aggregation is not observed in the case of proteins that induce isotropic curvature as the domain induced budding occurs at an earlier stage. IV. CONCLUSIONS We propose a Monte Carlo based multicomponent vesicle model that includes the effect of lipid-lipid, lipid-protein and protein-protein interactions to study (1) the kinetics of protein induced membrane domain formation and (2) the domain induced conformational changes 17 103104105106MCsteps(t)102ClpC0=0C0=0.3C0=0.5C0=0.75MC steps(a)(b) Figure 12. Cluster formation in presence of anisotropic curvature generating proteins as a function of time when Jφ = 0, Jpφ = 2, LL = 0 and κ(cid:107) = 5. (a) Clp for different values of C(cid:107)0 . (b) The vesicle conformations for induced curvature C(cid:107)0 = 0.75. of the membrane. Our study finds that the lipid-protein interaction induced aggregations to be significantly different from that resulting from lipid-lipid interactions, with the former showing faster cluster formation. The cluster sizes resulting from lipid-protein interactions depend on the protein concentration and on their interaction strength. For small protein fractions (p% < 20) and low lipid-protein interaction strengths (Jpφ < 2), the absence of line tension at the domain boundaries lead to fast clustering kinetics. At low co-lipid protein composition ratio and high lipid-protein interactions the effect of line tension is prominent and budding of membrane domains and slowing down of domain growth is observed. Our simulations explored the effect of lipid-protein interactions on protein clustering and showed that even at small protein concentrations it is possible to get complete protein clus- tering. We also examined the role of the structural properties of the proteins on the protein aggregation. An explicit protein-protein interaction which prefers parallel alignment of pro- teins is shown to reduces diffusion of clusters and limits budding of domains. The clustering of curvature active proteins is shown to have significant dependence on the anisotropy of the 18 MC steps103104105106MCsteps(t)102ClpCk0=0Ck0=0.3Ck0=0.5Ck0=0.75(a)(b) induced curvature, with stronger anisotropy suppressing budding and favoring larger cluster formation. Appendix A: Scaling of domains Scaling assumption implies that the distance between domains d should scale the same way as the size of the domain size itself. For circular domains of radius R, this would imply that d ∝ R. If the domain coalescence takes place through diffusion of domains then the coalescence time tc scales as d2 ∝ Dtc, where D is the diffusion coefficient. Rouse dynamics implies that D ∝ 1/R2, i.e., D is inversely proportional to the number of vertices in the domain. Thus d2 ∝ tc/R2 which leads to R4 ∝ tc. Therefore R ∝ t1/4 and the size of the domains scales as t1/2. REFERENCES 1K. Simons and D. Toomre, "Lipid rafts and signal transduction," Nat Rev Mol Cell Biol 1, 31 (2000). 2R. Sarasij, S. Mayor, and M. Rao, "Chirality-induced budding: A raft-mediated mecha- nism for endocytosis and morphology of caveolae?" Biophys. J. 92, 3140–3158 (2007). 3D. Goswami, K. Gowrishankar, S. Bilgrami, S. Ghosh, R. Raghupathy, R. Chadda, V. Rand, M. Rao, and S. Mayor, "Nanoclusters of gpi-anchored proteins are formed by cortical actin-driven activity," Cell 135, 1085–1097 (2008). 4K. Gowrishankar, S. Ghosh, S. Saha, C. Rumamol, S. Mayor, and M. Rao, "Active re- modeling of cortical actin regulates spatiotemporal organization of cell surface molecules," Cell 149, 1353–1367 (2012). 5R. Raghupathy, A. Anilkumar, A. Polley, P. Singh, M. Yadav, C. Johnson, S. Suryawanshi, V. Saikam, S. Sawant, A. Panda, Z. Guo, R. Vishwakarma, M. Rao, and S. Mayor, "Transbilayer lipid interactions mediate nanoclustering of lipid-anchored proteins," Cell 161, 581–594 (2015). 6K. Simons and E. Ilkon, "Functional rafts in cell membranes," Nature 387, 569–572 (1997). 7K. Simons and W. L. Vaz, "Model systems, lipid rafts, and cell membranes," Annual Review of Biophysics and Biomolecular Structure 33, 269–295 (2004). 19 8M. Mercker and A. Marciniak-Czochra, "Bud-neck scaffolding as a possible driving force in escrt-induced membrane budding," Biophys. J. 108, 833–843 (2015). 9C. Barlowe, L. Orci, T. Yeung, M. Hosobuchi, S. Hamamoto, N. Salama, M. F. Rexach, M. Ravazzola, M. Amherdt, and R. Schekman, "Copii: A membrane coat formed by sec proteins that drive vesicle budding from the endoplasmic reticulum," Cell 77, 895–907 (1994). 10B. Antonny, "Membrane deformation by protein coats," Curr. Opin. Cell Biol. 18, 386–394 (2006). 11J. Poveda, A. Fernandez, J. Encinar, and J. Gonzalez-Ros, "Protein-promoted membrane domains," Biochim. Biophys. Acta 1778, 1583–1590 (2008). 12R. M. Epand, "Proteins and cholesterol-rich domains," Biochim. Biophys. Acta 1778, 1576–1582 (2008). 13Q.-Y. Wu and Q. Liang, "Interplay between curvature and lateral organization of lipids and peptides/proteins in model membranes," Langmuir 30, 1116–1122 (2014). 14S. Mayor and H. Riezman, "Sorting gpi-anchored proteins," Nat. Rev. Mol. Cell Biol. 5, 110–120 (2004). 15R. G. W. Anderson and K. Jacobson, "A role for lipid shells in targeting proteins to caveolae, rafts, and other lipid domains," Science 296, 1821–1825 (2002). 16M. G. J. Ford, I. G. Mills, B. J. Peter, Y. Vallis, G. J. K. Praefcke, P. R. Evans, and H. T. McMahon, "Curvature of clathrin-coated pits driven by epsin." Nature 419, 361–366 (2002). 17P. Drucker, M. Pejic, H.-J. Galla, and V. Gerke, "Lipid segregation and membrane budding induced by the peripheral membrane binding protein annexin a2," Journal of Biological Chemistry 288, 24764–24776 (2013). 18R. Oliva, P. D. Vecchio, M. I. Stellato, A. M. D'Ursi, G. D'Errico, L. Paduano, and L. Pe- traccone, "A thermodynamic signature of lipid segregation in biomembranes induced by a short peptide derived from glycoprotein gp36 of feline immunodeficiency virus," Biochimica et Biophysica Acta (BBA) - Biomembranes 1848, 510 – 517 (2015). 19G. van Meer, D. R. Voelker, and G. W. Feigenson, "Membrane lipids: where they are and how they behave," Nature 9, 112–124 (2008). 20G. van Meer and A. I. P. M. de Kroon, "Lipid map of the mammalian cell," Journal of Cell Science 124, 5–8 (2011). 20 21S. McLaughlin and D. Murray, "Plasma membrane phosphoinositide organization by pro- tein electrostatics," Nature 438, 605–611 (2005). 22A. Frost, V. M. Unger, and P. D. Camilli, "The bar domain superfamily: Membrane- molding macromolecules," Cell 137, 191–196 (2009). 23G. van den Bogaart, K. Meyenberg, H. J. Risselada, H. Amin, K. I. Willig, B. E. Hubrich, M. Dier, S. W. Hell, H. Grubmuller, U. Diederichsen, and R. Jahn, "Membrane protein sequestering by ionic protein-lipid interactions," Nature 479, 552–555 (2011). 24P. B. Sunil Kumar and M. Rao, "Shape Instabilities in the Dynamics of a Two-Component Fluid Membrane," Phys. Rev. Lett. 80, 2489–2492 (1998). 25P. B. Sunil Kumar, G. Gompper, and R. Lipowsky, "Budding dynamics of multicomponent membranes," Phys. Rev. Lett. 86, 3911–3914 (2001). 26M. Laradji and P. Sunil Kumar, "Dynamics of Domain Growth in Self-Assembled Fluid Vesicles," Phys. Rev. Lett. 93, 198105 (2004). 27M. Laradji and P. B. Sunil Kumar, "Domain growth, budding, and fission in phase-separating self-assembled fluid bilayers," J. Chem. Phys. 123, 224902 (2005), 10.1063/1.2102894. 28L. Li, X. Liang, M. Lin, F. Qiu, , and Y. Yang, "Budding dynamics of multicomponent tubular vesicles," Journal of the American Chemical Society 127, 17996–17997 (2005). 29S. Ramachandran, M. Laradji, and P. B. S. Kumar, "Lateral organization of lipids in multi-component liposomes," J. Phys. Soc. Jpn. 78, 041006 (2009). 30N. Ramakrishnan, P. B. S. Kumar, and J. H. Ipsen, "Monte carlo simulations of fluid vesicles with in-plane orientational ordering," Phys. Rev. E 81, 041922 (2010). 31B. J. Peter, H. M. Kent, I. G. Mills, Y. Vallis, P. J. G. Butler, P. R. Evans, and H. T. McMahon, "Bar domains as sensors of membrane curvature: The amphiphysin bar struc- ture," Science 303, 495–499 (2004). 32P. A. Lebwohl and G. Lasher, "Nematic-liquid-crystal order¯a monte carlo calculation," Phys. Rev. A 6, 426–429 (1972). 33J. R. Frank and M. Kardar, "Defects in nematic membranes can buckle into pseudo- spheres," Phys. Rev. E 77, 041705 (2008). 34N. Ramakrishnan, P. B. S. Kumar, and J. H. Ipsen, "Membrane-mediated aggregation of curvature-inducing nematogens and membrane tubulation," Biophys. J. 104, 1018–1028 (2013). 21 35Y. Han, Y. Shokef, A. M. Alsayed, P. Yunker, T. C. Lubensky, and A. G. Yodh, "Geometric frustration in buckled colloidal monolayers," Nature 456, 898–903 (2008). 36E. J. Ungewickell and L. Hinrichsen, "Endocytosis: clathrin-mediated membrane bud- ding," Curr. Opin. Cell Biol. 19, 417–425 (2007). 37R. M. Epand, B. G. Sayer, and R. F. Epand, "Caveolin scaffolding region and cholesterol- rich domains in membranes," Journal of Molecular Biology 345, 339 – 350 (2005). 38P. K. Mattila, A. Pykalainen, J. Saarikangas, V. O. Paavilainen, H. Vihinen, E. Jokitalo, and P. Lappalainen, "Missing-in-metastasis and irsp53 deform pi(4,5)p2-rich membranes by an inverse bar domain like mechanism," The Journal of Cell Biology 176, 953–964 (2007). 39Y. Zhao, J. Liu, C. Yang, B. Capraro, T. Baumgart, R. Bradley, N. Ramakrishnan, X. Xu, R. Radhakrishnan, T. Svitkina, and W. Guo, "Exo70 generates membrane curvature for morphogenesis and cell migration," Dev. Cell 26, 266–278 (2013). 22
1305.2332
2
1305
2013-12-20T17:51:12
On 1/f^alpha power laws originating from linear neuronal cable theory: power spectral densities of the soma potential, transmembrane current and single-neuron contribution to the EEG
[ "physics.bio-ph", "q-bio.NC" ]
Power laws, that is, power spectral densities (PSDs) exhibiting 1/f^alpha behavior for large frequencies f, have commonly been observed in neural recordings. Power laws in noise spectra have not only been observed in microscopic recordings of neural membrane potentials and membrane currents, but also in macroscopic EEG (electroencephalographic) recordings. While complex network behavior has been suggested to be at the root of this phenomenon, we here demonstrate a possible origin of such power laws in the biophysical properties of single neurons described by the standard cable equation. Taking advantage of the analytical tractability of the so called ball and stick neuron model, we derive general expressions for the PSD transfer functions for a set of measures of neuronal activity: the soma membrane current, the current-dipole moment (corresponding to the single-neuron EEG contribution), and the soma membrane potential. These PSD transfer functions relate the PSDs of the respective measurements to the PSDs of the noisy input currents. With homogeneously distributed input currents across the neuronal membrane we find that all PSD transfer functions express asymptotic high-frequency 1/f^alpha power laws. The corresponding power-law exponents are analytically identified as alpha_inf^I = 1/2 for the soma membrane current, alpha_inf^p = 3/2 for the current-dipole moment, and alpha_inf^V = 2 for the soma membrane potential. These power-law exponents are found for arbitrary combinations of uncorrelated and correlated noisy input current (as long as both the dendrites and the soma receive some uncorrelated input currents). Comparison with available data suggests that the apparent power laws observed in experiments may stem from uncorrelated current sources, presumably intrinsic ion channels, which are homogeneously distributed across the neural membranes and themselves exhibit ...
physics.bio-ph
physics
CONTENTS I. Introduction II. Models A. Cable equation for dendritic sticks B. Ball and stick neuron with single current input Distal part of dendritic stick Soma and proximal part of dendritic stick Full solution C. Ball and stick neuron with spatially distributed input Correlated current inputs Uncorrelated current inputs D. Summary of PSD transfer functions for ball and stick neuron E. From single-neuron current-dipole moments to EEG F. Numerical simulations III. Results A. Biophysically detailed neuron model vs. ball and stick model B. Power laws for ball and stick neuron C. Apparent power laws for experimentally relevant frequencies Soma current Current-dipole moment / EEG contribution Soma potential D. PSDs for varying biophysical parameters for ball and stick neuron IV. Discussion A. Summary of main findings B. Comparison with power laws observed in neural recordings C. Origin of noise D. Power laws for local field potentials (LFP) and ECoG signals E. Passive approximation F. Concluding remarks 1 4 8 9 10 11 12 14 15 17 18 21 22 24 26 26 27 30 31 33 33 35 37 37 37 40 41 43 43 Acknowledgments References Figure Legends Tables 44 44 47 58 On 1/f α power laws originating from linear neuronal cable theory: power spectral densities of the soma potential, soma membrane current and single-neuron contribution to the EEG Klas H. Pettersen a Center for Integrative Genetics, Dept. of Mathematical Sciences and Technology, Norwegian University of Life Sciences, As, Norway. Dept. of Computational Biology, School of Computer Science and Communication, Henrik Lind´en Royal Institute of Technology (KTH), Stockholm, Sweden and Dept. of Mathematical Sciences and Technology, Norwegian University of Life Sciences, As, Norway. Tom Tetzlaff Inst. of Neuroscience and Medicine (INM-6), Computational and Systems Neuroscience Research Center, Julich, Germany and Dept. of Mathematical Sciences and Technology, Norwegian University of Life Sciences, As, Norway. Gaute T. Einevoll Dept. of Mathematical Sciences and Technology, Norwegian University of Life Sciences, As, Norway (Dated: September 25, 2018) a Corresponding author: [email protected] 2 Abstract Power laws, that is, power spectral densities (PSDs) exhibiting 1/f α behavior for large frequen- cies f , have commonly been observed in neural recordings. Power laws in noise spectra have not only been observed in microscopic recordings of neural membrane potentials and membrane cur- rents, but also in macroscopic EEG (electroencephalographic) recordings. While complex network behavior has been suggested to be at the root of this phenomenon, we here demonstrate a possible origin of such power laws in the biophysical properties of single neurons described by the standard cable equation. Taking advantage of the analytical tractability of the so called ball and stick neu- ron model, we derive general expressions for the PSD transfer functions for a set of measures of neuronal activity: the soma membrane current, the current-dipole moment (corresponding to the single-neuron EEG contribution), and the soma membrane potential. These PSD transfer functions relate the PSDs of the respective measurements to the PSDs of the noisy input currents. With homogeneously distributed input currents across the neuronal membrane we find that all PSD transfer functions express asymptotic high-frequency 1/f α power laws. The corresponding power- law exponents are analytically identified as αI∞ = 1/2 for the soma membrane current, αp∞ = 3/2 for the current-dipole moment, and αV∞ = 2 for the soma membrane potential. These power-law exponents are found for arbitrary combinations of uncorrelated and correlated noisy input current (as long as both the dendrites and the soma receive some uncorrelated input currents). Compari- son with available data suggests that the apparent power laws observed in experiments may stem from uncorrelated current sources, presumably intrinsic ion channels, which are homogeneously distributed across the neural membranes and themselves exhibit pink (1/f ) noise distributions. The significance of this finding goes beyond neuroscience as it demonstrates how 1/f α power laws with a wide range of values for the power-law exponent α may arise from a simple, linear partial differential equation. We find here that the well-known cable equation describing the electrical properties of membranes transfers white-noise current input into 'colored' 1/f α-noise where α may have any half-numbered value within the interval from 1/2 to 3 for the different measurement modalities. Intuitively, the physical origin of these novel power laws can be understood in terms of the superposition of numerous low-pass filtered contributions with different cut-off frequencies (i.e., different time constants) due to the different spatial positions of the various current inputs along the neuron. As our model system is linear, the results directly generalize to any colored input noise, i.e., transferring 1/f β spectra of input currents to 1/f β+α output spectra. 3 Popular summary The common observation of power laws in nature and society, that is, that quantities or probabilities follow 1/xα distributions, has for long intrigued scientists. Such power laws have been seen in a wide range of situations including frequencies of differently sized earth quakes, distribution of links on the World Wide Web, and size scaling in animals. In the brain, power laws in the power spectral density (PSD) have been observed in electrophysiological recordings, both at the microscopic (single-neuron recordings) and macroscopic (EEG) levels. While these neural power laws have been suggested to stem from complex network behavior, we here demonstrate a possible origin of power laws in the basic biophysical properties of neurons, that is, in the standard cable-equation description of neuronal membranes. Taking advantage of the mathematical tractability of the so called ball and stick neuron model, we demonstrate analytically that high-frequency power laws in key experimental measures of neural activity will arise naturally when the noise sources are evenly distributed across the neuronal membrane. Comparison with available data further suggests that the apparent power laws observed in experiments may stem from uncorrelated current sources, presumably intrinsic ion channels, which are homogeneously distributed across the neural membranes and themselves exhibit pink (1/f ) noise distributions. The significance of this finding goes beyond neuroscience as it demonstrates how 1/f α power laws with a wide range of values for the power-law exponent α, i.e., any half-numbered value between 1/2 and 3, may arise from a simple, linear physics equation. I. INTRODUCTION The apparent ubiquity of power laws in nature and society, i.e., that quantities or prob- ability distributions y(x) satisfy the relationship y(x) ∝ x−α , (1) where α is the power-law exponent, has for a long time intrigued scientists [1]. Power laws in the tails of distributions have been reported in a wide range of situations including such 4 different phenomena as frequency of differently sized earth quakes, distribution of links on the World Wide Web, paper publication rates in physics, and allometric scaling in animals (see [1] and references therein). A key feature of power laws is that they are scale invariant over several orders of magnitude, i.e., that they do not give preference to a particular scale in space or time. There are several theories with such scale invariance as its fingerprint, among the most popular are fractal geometry [2] and the theory of self-organized critical states [3]. Conspicuous power laws have been observed also in the field of neuroscience. Ever since Hans Berger recorded the first human electroencephalogram (EEG) in 1924 [4], its features have been under extensive study, especially since many of them are directly related to disease and to states of consciousness. Moreover, in the last decades the underlying power spectral density (PSD) of the EEG has also attracted significant attention as the PSD is often well fitted by a 1/f α power law with α typically in the range from 1 to 2.5 [5, 6]. Power-law spectra are not only seen in macroscopic neural recordings such as EEG, they also appear at the microscopic level, i.e., in single-neuron recordings. PSDs of the subthreshold membrane potentials recorded in the somas of neurons often resemble a 1/f α power law, typically with a larger exponent α ranging from 2 to 3 [7 -- 11]. As for the EEG, this power law seems to be very robust: it has been observed across species, brain regions and different experimental set-ups, such as cultured hippocampal layer V neurons [7], pyramidal layer IV -- V neurons from rat neocortex in vitro [9, 10], and neocortical neurons from cat visual cortex in vivo [8, 11]. At present, the origin, or origins, of these macroscopic and microscopic power laws observed in neural recordings are poorly understood. Lack of sufficient statistical support have questioned the validity of identified power-law behaviors, and as a rule of thumb, a candidate power law should exhibit an approximately linear relationship in a log-log plot over at least two orders of magnitude [1]. Further, a mechanistic explanation of how the power laws arise from the underlying dynamics should ideally be provided [1]. In the present paper we show through a combination of analytical and numerical investigations how power laws naturally can arise in neural systems from noise sources homogeneously distributed throughout neuronal membranes. We further show that the mechanism behind microscopic (soma potential, soma current) power laws will also lead to power laws in the single-neuron contribution (current-dipole moment) to the EEG, More- over, if all single-neuron contributions to the recorded EEG signal exhibit the same power 5 law, the EEG signal will also exhibit this power law. We find that for different measurement modalities different power-law exponents naturally follow from the well-established, biophys- ical cable properties of the neuronal membranes: the soma potential will be more low-pass filtered than the corresponding current-dipole moment determining the single-neuron con- tribution to the EEG [12, 13], and as a consequence, the power-law exponent α will be larger for the soma potential than for the single-neuron contribution to the EEG [14] (see illustration in Fig. 1). When comparing with experimental data, we further observed that for the special case when uncorrelated and homogeneously distributed membrane-current sources themselves exhibit 1/f power laws in their PSD, the theory predicts power-law exponents α in accor- dance with experimental observations for the microscopic measures, i.e., the soma current and soma potential. The experimental situation is much less clear for the EEG signal. However, we note that under the assumption that such single-neuron sources dominate the high-frequency part of the EEG signal, the theoretical predictions are also compatible with the power-law-like behavior so far observed experimentally. Both synaptic noise and intrinsic channel noise will in general contribute to the observed noise spectra, cf. Fig. 1. While our theory per se is indifferent to the detailed membrane mechanism providing the noisy current, our findings suggests that the dominant noise source underlying the observed power spectra may be channel noise: prevalent theories for synaptic currents are difficult to reconcile with a 1/f power law, while potassium ion channels with such 1/f noise spectra indeed have been observed [15]. Through the pioneering work by Wilfred Rall half a century ago [16, 17] the ball and stick neuron model was established as a key model for the study of the signal processing properties of neurons. An important advantage is the model's analytical tractability, and this is exploited in the present study. We first demonstrate the relevance of this simplified model in the present context by numerical comparisons with results from a morphologi- cally reconstructed multicompartmental pyramidal neuron model. Then we derive analyti- cal power-law expressions for the various types of electrophysiological measurements. While a single current input onto a dendrite does not give rise to power laws, we here show that power laws naturally arise for the case with homogeneously distributed inputs across the dendrite and the soma [18], see Fig. 1. For this situation we show that the ball and stick neuron model acts as a power-law filter for high frequencies, i.e., the transfer function from 6 the PSD of the input membrane currents, s(f ), to the PSD of the output (soma potential, soma current, or current-dipole moment setting up the EEG), S(f ), is described by a power law: S(f )/s(f ) = 1/f α. Notably the analytically derived power-law exponents α for these transfer functions are seen to be different for the different measurement modalities. The analytical expressions further reveal the dependence of the PSDs on single-neuron features such as the correlation of input currents, dendritic length and diameter, soma diameter and membrane impedance. The theory presented here also contributes to 1/f -theory in general: it illustrates that a basic physics equation, the cable equation, can act as a 1/f α power-law filter for high frequencies when the underlying model has spatially distributed input. Furthermore, α may have any half-numbered value between 1/2 and 3, depending on the physical measure (some potential, soma current, single-neuron contribution to the EEG) under consideration, and the coherence of the input currents. Intuitively, the emergence of the power-law spectra can be understood as a result of a superposition of low-pass filters with a wide range of cutoff frequencies due to position-dependent intrinsic dendritic filtering [12, 13, 19] of the spatially extended neuron. The paper is organized as follows: In the next section we derive analytical expressions for the soma potential, soma current and current-dipole moment for the ball and stick neuron for the case with noisy current inputs impinging on the soma 'ball' and homogeneously on the dendritic stick. While these derivations are cumbersome, the final results are transparent: power laws are observed for all measurement modalities in the high-frequency limit. In Results we first demonstrate by means of numerical simulations the qualitative similarity of the power-law behaviors between the ball and stick model and a biophysically detailed pyramidal neuron. We then go on to analytically identify the set of power-law exponents for the various measurement modalities both in the case of uncorrelated and correlated current inputs. While the derived power laws strictly speaking refer to the functional form of PSDs in the high-frequency limit (Eq. 1), the purported power laws in neural data have typically been observed for frequencies less than a few hundred hertz. Our model study implies that the true high-frequency limit is not achieved at these frequencies. However, in our ball and stick model, quasi-linear relationships can still be observed in the characteristic PSD log-log plots for the experimentally relevant frequency range. These apparent power laws typically have smaller power-law exponents than their respective asymptotic value. The numerical 7 values of these exponents will depend on details in the neuron model, but the ball and stick model has a very limited parameter space: it is fully specified by four parameters, a dimensionless frequency, the dimensionless stick length, the ratio between the soma and infinite-stick conductances, and the ratio between the somatic and dendritic current density. This allows for a comprehensive investigations of the apparent power-law exponents in terms of the neuron parameters, which we pursue next. To facilitate comparison with experiments we round off the Results section exploring how PSDs, and in particular apparent power laws, depend on relevant biophysical parameters. In the Discussion we then compare our model findings with experiments and speculate on the biophysical origin of the membrane currents underlying the observed PSD power laws. FIG. 1 AROUND HERE II. MODELS In the present study the idealized ball and stick neuron model will be treated analytically, while simulation results will be presented for a reconstructed layer V pyramidal neuron from cat visual cortex [20] (Fig. 2). Both the ball and stick model and the reconstructed layer V neuron model are purely passive, ensuring that linear theory can be used. The input currents are distributed throughout the neuron models with area density ρd in the dendrite and ρs in the soma. The input currents share statistics, i.e., they all have the same PSD, denoted s = s(ω), and a pairwise coherence c = c(ω). The coherence is zero for uncorrelated input and unity for perfectly correlated input. For the ball and stick neuron, the cable equation is treated analytically in frequency space. We first provide a solution for a single current input at an arbitrary position, and then use this solution as basis for the case of input currents evenly distributed throughout the neuronal membrane. The resulting PSDs can be expressed as Riemann sums where the terms correspond to single-input contributions. In the continuum limit where the neuron is assumed to be densely bombarded by input currents, the Riemann sums become analytically solvable integrals. From these analytical solutions we can then extract the various transfer functions relating the output PSDs to the PSDs of the input current. Here the output 8 modalities of interest are the net somatic current, the soma potential and the single-neuron contribution to the EEG, see Figs. 1 and 2. Below we treat the ball and stick neuron analytically. For the pyramidal neuron (Fig. 2), the NEURON Simulation Environment [21] with the supplied Python interface [22] was used. FIG. 2 AROUND HERE A. Cable equation for dendritic sticks For a cylinder with a constant diameter d the cable equation is given by λ2 ∂2V (x, t) = τm ∂V (x, t) ∂x2 √ with the length constant λ = 1/ gmri =(cid:112)dRm/4Ri and the time constant τm = cm/gm = + V (x, t) , (2) ∂t RmCm. Rm, Cm and Ri denote the specific membrane resistance, the specific membrane capacitance and the inner resistivity, respectively, and have dimensions [Rm] = Ωm2, [Cm] = F/m2 and [Ri] = Ωm. Lower-case letters are used to describe the electrical properties per unit length of the cable: gm = 1/rm = πd/Rm, cm = πdCm and ri = 4Ri/πd2, with units [gm] = 1/Ωm, [cm] = F/m and [ri] = Ω/m. For convenience, the specific membrane conductance, Gm = 1/Rm, will also be used, see Table I for a list of symbols. With dimensionless variables, X = x/λ and T = t/τm, the cable equation, Eq. 2, can be expressed ∂2V (X, T ) ∂X 2 − ∂V (X, T ) ∂T − V (X, T ) = 0 . (3) Due to linearity, each frequency component of the input signal can be treated individually. For this, it is convenient to express the membrane potential in a complex (boldface notation) form, V = V(X, W )ejW T , (4) where V is a complex number containing the amplitude abs( V) and phase arg( V) of the signal, and the dimensionless frequency is defined as W = ωτm. The complex potentials are related to the measurable potential V (X, T ) through the Fourier components of the 9 potential, V (X, T ) = V0(X) + ∞(cid:88) k=1 Re{ V(X, Wk)ejWkT} , (5) where V0(X) is the direct current (DC) potential. The cable equation can then be simplified to d2 V dX 2 − q2 V = 0 , where q2 ≡ 1 + jW , see [12, 23]. The general solution to Eq. 6 can be expressed as V(X, W ) = C1 cosh(qL − qX) + C2 sinh(qL − qX) . The expression for the axial current is given by Ii(x, t) = − 1 ri ∂V (x, t) ∂x , (6) (7) (8) and is applied at the boundaries to find the specific solutions for the ball and stick neuron. In complex notation and with dimensionless variables this can be expressed as Ii(X, W ) = − 1 riλ ∂ V(X, W ) ∂X = −G∞ ∂ V(X, W ) ∂X , (9) where G∞ is the infinite-stick conductance. Similarly, the transmembrane current density (including both leak currents and capacitive currents) is given by im = −∂Ii(x, t) ∂x = 1 ri ∂2V (x, t) ∂x2 , (10) with its complex counterpart, im(X, W ) = − 1 λ ∂I(X, W ) ∂X = 1 riλ2 ∂2 V(X, W ) ∂X 2 = gm ∂2 V(X, W ) ∂X 2 . (11) FIG. 3 AROUND HERE B. Ball and stick neuron with single current input The ball and stick neuron [16] consists of a dendritic stick attached to a single-compartment soma, see Fig. 3A. Here we envision the stick to be a long and thin cylinder with diameter d and length l. The membrane area of the soma is set to be πd2 s , corresponding to the surface 10 area of a sphere with diameter ds, or equivalently, the side area of a cylindrical box with diameter and height ds. The solution of the cable equation for a ball and stick neuron with a single input current at an arbitrary dendritic position is found by solving the cable equation separately for the neural compartment proximal to the input current and the neural compartment distal to the input current, These solutions are then connected through a common voltage boundary condition V0 at the connection point. For the proximal part of the stick, Ohm's law in combination with the lumped soma admittance gives the boundary condition at the somatic site, and for the distal part of the stick, a sealed-end boundary is applied at the far end. In this configuration the boundary condition V0 acts as the driving force of the system. The potential V0 can, however, also be related to a corresponding input current Iin through the input impedance, i.e., Iin = V0 Yin. Distal part of dendritic stick First, we focus on the part of the stick distally to the input in Fig. 3A. Assume that the stick has V0 as a boundary condition at the proximal end and a sealed-end boundary at the distal end. We use the subscript 'd' for distal stick at the spatial coordinates, and shift the coordinate system so that the input is in Xd = 0. The boundary condition at the proximal end, i.e., at the position of the input current, then becomes V(Xd = 0) = V0, while a sealed end is assumed at the distal end of the stick, i.e., at Xd = Ld. Here Ld denotes the electrotonic length a the stick with physical length l, i.e., Ld = ld/λ. A sealed-end boundary corresponds to zero axial current, Eq. 9. With these boundary conditions the specific solution to the cable equation becomes [12, 23], Vd(Xd, W ) = V0 cosh(qLd − qXd) cosh(qLd) . The axial current Ii(Xd, W ) is given by Eq. 9, Ii,d(Xd, W ) = V0qG∞ sinh(qLd − qXd) cosh(qLd) . (12) (13) The dendritic input admittance, Yin,d(W ) = Ii,d(Xd = 0, W )/ Vd(Xd = 0, W ), will then be Yin,d(W ) = qG∞ tanh(qLd) . (14) 11 im,d(Xd, W ) = gmq2 cosh(qLd − qXd) (cid:90) Ld cosh(qLd) V0 . (15) (16) Since lim L→∞ tanh(qL) → 1, the infinite-stick admittance can be expressed as Y∞(W ) = G∞q = q/riλ, and the finite-stick admittance can be expressed as Yin,d(W ) = Y∞(W ) tanh(qLd). From Eqs. 11 and 12 it follows that the transfer function linking an imposed voltage V0 in the proximal end to a transmembrane current density in position Xd can be expressed as [12] The complex dipole-moment for a stick with a sealed end is then given by the integral pd(W ) = λ2 im,d(X, W )X dX = λG∞ V0[1 − 1/ cosh(qLd)] . 0 Soma and proximal part of dendritic stick Let us now consider a ball and stick neuron with an input current at the far end of the stick, effectively corresponding to the proximal part of the ball and stick neuron in Fig. 3A. We denote the coordinates with the subscript 'p' for proximal. Similar to the situation for the distal stick, we apply a boundary condition V0 to the site of the current input and put this in Xp = 0, i.e., Vp(Xp = 0) = V0. The stick is assumed to lie along the Xp-axis, to have electrotonic length Lp, and the soma site located at Xp = Lp. The lumped-soma boundary condition implies that the leak current out of the dendritic end is, through Ohm's law, proportional to the soma admittance, Ii,p(Lp, W ) = Is = Ys Vp(Lp, W ) = Ys Vs, where Is, Vs and Ys denote the somatic transmembrane current, soma potential and somatic membrane admittance, respectively. Thus, for Xp = 0 the boundary condition becomes: Vp(0, W ) = V0 , and, through Eq. 9, we have at Xp = Lp: Ii,p(Lp, W ) = −G∞ ∂ Vp(Xp, W ) ∂Xp (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xp=Lp = Ys Vs . The complex constant C2 in Eq. 7 is found from the boundary condition in Eq. 18, C2 = Ys Vs G∞q = Vs Ys Y∞ , which, combined with Eq. 17, gives C1: V0 C1 = cosh(qLp) − Vs Ys Y∞ tanh(qLp) . 12 (17) (18) (19) (20) By substituting the constants C1 and C2 and by using Vs = V(Lp, W ), Eq. 7 gives V0/ Vs = cosh(qLp)(1 + Y tanh(qLp)) , (21) where Y = Ys/Y∞. Next, Eq. 21 is used to substitute for Vs in the constants C1 and C2, and after some algebraic manipulations the solution for the cable equation with the given boundary conditions becomes, Vp(Xp, W ) = V0 cosh(qLp − qXp) + Y sinh(qLp − qXp) cosh(qLp) + Y sinh(qLp) . (22) The axial current is through Eq. 9 given by Ii,p(Xp, W ) = V0Y∞ sinh(qLp − qXp) + Y cosh(qLp − qXp) cosh(qLp) + Y sinh(qLp) , (23) and the input admittance is, through Ohm's law, given by Yin,p = Ii,p(0, W )/ V0, Yin,p = Y∞ sinh(qLp) + Y cosh(qLp) cosh(qLp) + Y sinh(qLp) . The axial current at Xp = Lp, i.e., the somatic transmembrane current, will then be Is = Ii,p(Lp, W ) = V0Ys cosh(qLp) + Y sinh(qLp) , and the transmembrane current density will be given by Eq. 11, (24) (25) im,p = V0gmq2 cosh(qLp − qXp) + Y sinh(qLp − qXp) (cid:21) cosh(qLp) + Y sinh(qLp) (cid:20) pstick(W ) = V0 λG∞ − lpYs + λG∞ . (26) By an integral similar to Eq. 16, the current-dipole moment for the stick is found to be cosh(qLp) + Y sinh(qLp) . (27) The contribution to the current-dipole moment from the somatic return current is the prod- uct of the somatic current, Eq. 25, and the fixed dipole length (i.e., distance between the position of the current input and the soma), here corresponding to the stick length lp, ps = lpIs = lp V0Ys cosh(qLp) + Y sinh(qLp) . (28) The total dipole moment for a ball and stick neuron with current input at the far end of the stick is therefore pp = ps + pstick = V0λG∞ − V0λG∞ cosh(qLp) + Y sinh(qLp) . (29) 13 Full solution The full solution for current inputs at arbitrary positions is achieved by superposition of the distal-stick solution and the solution for the proximal stick with a lumped soma, see Fig. 3A. We will now use the same notation and coordinate system as in Fig. 3A, i.e., Xp = −X + Lp and Xd = X − Lp, and introduce the sum of the stick lengths L = Lp + Ld. Thus, the stick is along the X-axis from X = 0 (soma end) to X = L (distal end), and the input current is assumed to be injected at position X(cid:48). By summation of Eqs. 16 and 29 the ball and stick dipole moment now becomes p = − V0λG∞ 1 cosh(qL − qX(cid:48)) − 1 cosh(qX(cid:48)) + Y sinh(qX(cid:48)) . (30) (cid:20) (cid:21) (cid:21) (cid:20)sinh(qLp) + Y cosh(qLp) cosh(qLp) + Y sinh(qLp) The total input admittance of the ball and stick neuron is given by the sum of the proximal admittance and the distal admittance, Yin = Yin,p + Yin,d = Y∞ + tanh(qLd) , (31) which, with the coordinates used in Fig. 3A, becomes (cid:20)sinh(qX(cid:48)) + Y cosh(qX(cid:48)) cosh(qX(cid:48)) + Y sinh(qX(cid:48)) (cid:21) + tanh(q(L − X(cid:48))) . (32) Yin = Y∞ From Eq. 30 we now find, by means of Ohm's law and this expression for the input ad- mittance, the following transfer function between input current Iin and dipole moment, p = TpIin, cosh(qL − qX(cid:48)) − Y sinh(qX(cid:48)) − cosh(qX(cid:48)) . (33) Y cosh(qL) + sinh(qL) Tp = λG∞ Y∞ in, Ts I = Is/Is Transfer functions for the other quantities of interest, TV = Vs/Iin, TI = Is/Iin, Ts V = Vs/Is in, can be found similarly. The superscript 's' denotes that this applies for an input current at the soma. By substituting for V0 in Eq. 25, the transfer function for the soma current becomes p = ps/Is in , Ts TI = Y cosh(qL − qX(cid:48)) Y cosh(qL) + sinh(qL) . (34) From Eq. 34 and by assuming Ohm's law for the soma membrane, the soma potential transfer function becomes TV = 1 Y∞ cosh(qL − qX(cid:48)) Y cosh(qL) + sinh(qL) . 14 (35) For a somatic input current, Iin = Is by its total neuron input impedance seen from soma, in, the soma potential is, through Ohm's law, described Ts V = 1 Yin(X(cid:48) = 0) = 1 Y∞ cosh(qL) Y cosh(qL) + sinh(qL) . (36) By comparison between Eq. 35 and Eq. 36, we see that Eq. 35 also applies for the special V = TV (X(cid:48) = 0). The net somatic transmembrane current in and the somatic return current) has to enter the stick axially in X = 0. s Yin,dLd=L, and the transfer case with somatic input, i.e., Ts (including both Is Thus, the net somatic current can be described by Is s = − Vs function becomes I = − Ts sinh(qL) , (37) I (cid:54)= TI(X(cid:48) = 0). The intracellular resistance between the soma and the start position X = 0 of the stick is assumed to be zero, and which differs from the result in Eq. 34, i.e., Ts Y cosh(qL) + sinh(qL) the soma potential will therefore be the same regardless of whether the input current is positioned at the proximal end of the stick (i.e., at X = 0) or in the soma. However, when estimating the net somatic membrane current this distinction is important: the current input will itself count as a part of the calculated soma current if it is positioned in the soma, but not if it is positioned at the proximal end of the dendritic stick. For somatic input, the finite-stick expression in Eq. 16 will apply to the dipole moment. However, the input admittance is now different, and the transfer function becomes i.e., the expression in Eq. 33 holds, Ts Ts p = λG∞ Y∞ cosh(qL) − 1 Y cosh(qL) + sinh(qL) p = Tp(X(cid:48) = 0). , (38) C. Ball and stick neuron with spatially distributed input Above we derived transfer functions T for the ball and stick neuron, connecting current input at an arbitrary position on the neuron to the various measurement modalities, i.e., the current-dipole moment (Tp), the soma potential (TV ) and the soma current (TI). We will now derive expressions for the PSDs when the ball and stick neuron is bombarded with multiple inputs assuming that all input currents have the same PSD and a pairwise coherence c(ω) [24]. The PSDs can then be divided into separate terms for uncorrelated (c(ω) = 0) and fully correlated (c(ω) = 1) input. 15 The PSD, S = S(ω), of the output can for the case of multiple current inputs be expressed as S = l=1 k=1 N(cid:88) N(cid:88) (cid:34) (1 − c) (1 − c) = s = s N(cid:88) N(cid:88) k=1 k=1 Ik inTk(Il inTl)∗ N(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) N(cid:88) k=1 k=1 Tk(Tl)∗ N(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 = sH , l=1 + c Tk Tk(Tk)∗ + c (cid:12)(cid:12)Tk(cid:12)(cid:12)2 (cid:35) (39) where s = s(ω) is the PSD of the input currents, c = c(ω) is their coherence and H = H(ω) is the transfer function between the PSD of the input and the PSD of the output. The complex conjugate is denoted by the asterisk. We now assume the first J of the N input currents to be positioned at the soma com- partment, and the rest of the input to be spread homogeneously across the dendritic stick. The transfer function for the soma compartment, Ts, is the same for all somatic inputs, Tk = Ts for k = 1, 2, . . . , J, while the input transfer function for the dendritic stick is position dependent, Tk = T(Xk, W ) for k = J + 1, J + 2, . . . , N . The PSD transfer function can then be expressed H = (1 − c) (cid:32) N(cid:88) (cid:12)(cid:12)Tk(cid:12)(cid:12)2 k=J+1 J Ts2 + (cid:33) + c (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)JTs + N(cid:88) k=J+1 Tk (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 . (40) To allow for analytical extraction of power laws, we next convert the sums into integrals. By assuming uniform current-input density (per membrane area) in the dendritic stick (given by ρd = (N − J)/lπd), it follows that the axial density of current inputs is 1/(ρdπd). In the continuum limit (N → ∞) we thus have (cid:90) L N(cid:88) F (Tk) → k=J+1 0 F (T(X)) ρd πd λdX (41) where the last factor λ comes from the conversion to dimensionless lengths. The PSD transfer function, H ≡ S/s, in Eq. 40 can then be split into three parts, H =(cid:0)1 − c(cid:1)(cid:0)Huc,s + Huc,d (cid:1) + c Hc , where Huc,s = ρsπd2 s Ts(W )2 16 (42) (43) (cid:90) L 0 (cid:90) L 0 (cid:12)(cid:12)(cid:12)(cid:12)2 is the PSD transfer function for uncorrelated input at the soma compartment, Huc,d = ρdπdλ T(X, W )2dX (44) is the PSD transfer function for uncorrelated input distributed throughout the dendritic stick, and Hc = (cid:12)(cid:12)(cid:12)(cid:12)ρsπd2 s Ts + ρdπdλ T(X, W )dX (45) is the PSD transfer function for correlated input distributed both across the dendritic stick and onto the soma. We have now derived (i) a general expressions for the PSD transfer function H expressed by the general, single-input transfer functions T and Ts, and (ii) specific analytical ex- pressions for the single-input transfer functions for the dipole moment, the soma potential and the soma current. We will next combine these results and analytically derive specific PSD transfer functions for the dipole moment, the soma potential and the soma current for distributed input. Correlated current inputs For correlated activity the somatic transfer function and the corresponding integral of the dendritic transfer function are summed, see Eq. 45. For the soma current the integral within Eq. 45 is given by(cid:90) L TI(X, W )dX = Y sinh(qL)/q Y cosh(qL) + sinh(qL) . 0 By defining the denominator D(ω) = Y cosh(qL) + sinh(qL) , the PSD transfer function for the soma current is after some algebra found to be with the squared norm of D given by c = (ρdπdλY/q − ρsπd2 H I = s (ρd − ρs)2 π2d4 2 s ) sinh(qL)2/D2 [cosh(2aL) − cos(2bL)]/D2 , D2 = (cid:2)(B2(a2 + b2) + 1) cosh(2aL) + 2aB sinh(2aL) + (B2(a2 + b2) − 1) cos(2bL) + 2Bb sin(2bL)(cid:3) , 1 2 17 (46) (47) (48) (49) with a and b denoting the real and imaginary parts of q, respectively, i.e., and a = ([(1 + W 2)1/2 + 1]/2)1/2 , b = ([(1 + W 2)1/2 − 1]/2)1/2 . (50) (51) In Eq. 49 the specific membrane conductance and capacitance are assumed to be the same in the soma and the dendrite. Thus, Ys = πd2 s q2Gm and Y∞ = q/(λri). The admittance ratio can then be expressed as where B = d2 s /(dλ). Y = qB , (52) The contribution to the soma potential from dendritic input is given by the same integral as in Eq. 46 divided by the somatic impedance. By adding the corresponding transfer function for the somatic input the PSD transfer function is found to be: s cosh(qL)]/Y∞2/D2 c = [ρdπdλ sinh(qL)/q + ρsπd2 H V = + cosh(2aL)(cid:0)d4 2 (a2 + b2)2 D2 s ρ2 s π2λ2r2 i (cid:2)cos(2bL)(cid:0)d4 (cid:0)a2 + b2(cid:1) − d2λ2ρ2 (cid:1) (cid:0)a2 + b2(cid:1) + d2λ2ρ2 s λρdρs(a sinh(2aL) + b sin(2bL))(cid:3) . + 2dd2 s ρ2 s d d (cid:1) (53) For the current-dipole moment, the integral within Eq. 45, combined with the transfer function from Eq. 33, has the following simple solution, Tp(X, W )dX = λG∞ Y∞qD Y [1 − cosh(qL)] , (54) (cid:90) L 0 and the PSD transfer function for the dipole moment for correlated input currents is found to be H p c = = (cid:12)(cid:12)(cid:12)(cid:12) πλG∞[1 − cosh(qL)](ρddλY/q − ρsd2 s ) s λ2(ρd − ρs)2(cos(bL) − cosh(aL))2 Y∞D π2d4 (cid:12)(cid:12)(cid:12)(cid:12)2 (a2 + b2)D2 . (55) Uncorrelated current inputs In the case of uncorrelated input currents, the squared norm of hyperbolic functions, as well as cross-terms of different hyperbolic functions, must be integrated from X = 0 to X = L 18 to get the contributions from the dendritic stick. These integrals can be solved by converting the hyperbolic functions to their corresponding exponential expressions and expanding the products before applying straight-forward integration of the different exponential terms. For example, the following integral has to be solved for all PSDs, both the soma current PSD, the soma potential PSD and the PSD of the single-neuron contribution to the EEG: I1 = cosh(qL − qX)2dX , (56) where I now denotes an integral, not a current. The integrand is translated to its exponential 0 counterpart, (cid:90) L 0 I1 = (cid:2)e(q+q∗)(L−X) + e−(q+q∗)(L−X) + e(q−q∗)(L−X) + e−(q−q∗)(L−X)(cid:3) dX , 1 4 (cid:90) L (cid:20) + I1 = 1 4 and the integral is straightforwardly evaluated and found to be: 1 1 q − q∗ q + q∗ − 1 − 1 q + q∗ + q + q∗ − e−(q+q∗)L e(q+q∗)L q + q∗ + (cid:18)sinh [(q + q∗)L] q − q∗ + q − q∗ − e−(q−q∗)L e(q−q∗)L q − q∗ (cid:19) q + q∗ + q − q∗ , I1 = 1 2 The expression can be transformed back to hyperbolic functions sinh [(q − q∗)L] (cid:21) . and simplified as where we have used I1 = sinh(2aL)/4a + sin(2bL)/4b , sinh(2jbL) = j sin(2bL) . (57) (58) (59) (60) (61) From the expressions for the single-input transfer functions for the soma potential, Eq. 35, and soma current, Eq. 34, it follows that H V uc,d and H I uc,d (cf. Eq. 44) are both proportional to I1, i.e., and H V uc,d = R2∞ sinh(2aL)/a + sin(2bL)/b 4 (a2 + b2) , H I uc,d = B2(a2 + b2)(a sin(2bL) + b sinh(2aL)) 4ab 19 (62) (63) . For H p uc,d the following integrals also appear: 0 (cid:90) L (cid:90) L (cid:90) L (cid:90) L (cid:90) L 0 0 0 I2 = I3 = I4 = I5 = I6 = cosh(qX)2dX , sinh(qX)2dX , cosh(qL − qX) cosh(q∗X)dX , cosh(qL − qX) sinh(q∗X)dX , cosh(qX) sinh(q∗X)dX , (64) (65) (66) (67) (68) (69) (70) (71) (72) (73) All integrals can be solved by a similar scheme as above, and the solutions are 0 I2 = sinh (2aL)/4a + sin (2bL)/4b , I3 = sinh(2aL)/4a − sin (2bL)/4b , I4 = sinh(aL) cos(bL)/2a + cosh(aL) sin(bL)/2b , I5 = sinh(aL) sin(bL)/2b − j sinh(aL) sin(bL)/2a , I6 = cosh(2aL)/4a − 1/4a + j cos(2bL)/4b − 1/4b . Note that the solutions to the integrals I5 and I6 are complex. In the expression for the 5 and I∗ dipole moment the complex conjugated versions of the integrals I5 and I6, i.e., I∗ 6, also appear. For these the results are found directly from Eqs. 72-73 with j replaced by −j. The PSD transfer function for the dipole moment with uncorrelated input at the dendrite only, H p uc,d, can then be expressed as H p uc,d = ρdπdλ3 q2D2 [I1 + I2 + Y2I3 − 2Re{I4} − 2Re{Y∗I5} + 2Re{Y∗I6}] . (74) The full expression of H p uc,d is then H p uc,d = ρdπdλ3 (a2 + b2)D2 [sinh (2aL)/2a + sin (2bL)/2b 1 + y2 2)(sinh (2aL)/4a − sin (2bL)/4b) + (y2 − sinh(aL) cos(bL)/a + cosh(aL) sin(bL)/b − y1 sinh(aL) sin(bL)/b + y2 sinh(aL) sin(bL)/a + y1(cosh(2aL) − 1)/2a + y2(cos(2bL) − 1)/2b] , 20 (75) where y1 = Re{Y} and y2 = Im{Y}. For the special case where the specific admittance of the soma is equal to the specific admittance of the dendrite, i.e., Y = qd2 s /λd, this simplifies to the expression given in Eq. 85. The somatic contributions to the uncorrelated PSD transfer functions are given by H I s [cosh(2aL) − cos(2bL)]/D2 , uc,s = ρsπd2 md2 ρsR2 s πd2λ2 2(a2 + b2)D2 cosh(2aL) + cos(2bL) uc,s = H V , and H p uc,s = see Eqs. 36-38. s λ2 ρsπd2 2(a2 + b2)D2 [cosh(2aL) − 2 cosh(aL) cos(bL) + cos(2bL) + 2] , (76) (77) (78) D. Summary of PSD transfer functions for ball and stick neuron For convenience we here summarize the results, now solely in terms of dimensionless s /(dλ), L ≡ l/λ, and variables (except for the amplitudes A), i.e., ρ ≡ ρs/(ρs + ρd), B ≡ d2 W ≡ ωτ (see Table II). The general expression for the PSD transfer functions reads: H = (1 − c)Huc + c Hc , (79) where Huc = Huc(W ) represents the contributions from uncorrelated current inputs, Hc = Hc(W ) represents the contributions from correlated inputs, and c = c(W ) is the pairwise coherence function. The contributions from uncorrelated input currents are in turn given as sums over contributions from somatic Huc,s = Huc,s(W ) and dendritic inputs Huc,d = Huc,d(W ), i.e., The contribution to the PSD transfer functions for correlated input currents are given by Huc = Huc,s + Huc,d . (80) H I H p c = cB2[cosh(2aL) − cos(2bL)]/D2 , c = AI cB2 Ap a2 + b2 [cosh(2aL)/2 −2 cosh(aL) cos(bL) + cos(2bL)/2 + 1] /D2 , (cid:2)cos(2bL)(cid:0)B2ρ2(cid:0)a2 + b2(cid:1) − (1 − ρ)2(cid:1) + cosh(2aL)(cid:0)B2ρ2(cid:0)a2 + b2(cid:1) + (1 − ρ)2(cid:1) 2(a2 + b2)2 AV c H V c = + 2B(1 − ρ)ρ(a sinh(2aL) + b sin(2bL))] /D2 , 21 (81) (82) (83) with the squared norm of D given by Eq. 49, and a and b defined by Eqs. 50 and 51, respectively. The contributions from uncorrelated dendritic inputs are: (cid:19) /D2 , + sin (2bL) 2b (cid:18)sinh (2aL) (cid:18)sinh (2aL) (cid:20)sinh (2aL) 2a 2a + uc,dB2(a2 + b2) H I uc,d = H p uc,d = AI √ √ Ap 2 (a2 + b2) uc,d 2 B2(a2 + b2) + 2 sin (2bL) 2b − sin (2bL) (cid:21) 2b 2a − cosh(aL) sin(bL) −sinh(aL) cos(bL) −Ba sinh(aL) sin(bL) a cosh(2aL) − 1 + B b 2 + B (cid:18)sinh (2aL) Bb sinh(aL) sin(bL) + a cos(2bL) − 1 /D2 , b 2 (cid:19) (cid:19) H V uc,d = √ AV uc,dB2 2(a2 + b2) sin (2bL) + /D2 . 2a 2b (84) (85) (86) In the special case with input to soma only, the PSD transfer functions are the same for uncorrelated (Eq. 43) and correlated input (Eq. 45), the only difference being the amplitudes, Huc,s = Hcρ=1 ρsπd2 s . (ρ = ρs/(ρs + ρd) = 1 implies that the input is onto soma only.) The corresponding PSD transfer functions from uncorrelated somatic input thus become uc,sB2[cosh(2aL) − cos(2bL)]/D2 , uc,s = AI uc,sB2 Ap a2 + b2 [cosh(2aL)/2 uc,s = H p H I − 2 cosh(aL) cos(bL) + cos(2bL)/2 + 1] /D2 , H V uc,s = AV c B2 2(a2 + b2) [cosh(2aL) + cos(2bL)]/D2 . E. From single-neuron current-dipole moments to EEG (87) (88) (89) In an infinite, homogenous, isotropic Ohmic medium with conductivity σ, the extracel- lular potential recorded at a given position (cid:126)r far away from a single-neuron current dipole is given by [14, 25]. Φ1((cid:126)r, t) = p1(t) cos θ1 4πσ((cid:126)r − (cid:126)r1)2 , 22 (90) where (cid:126)r1 designates the spatial position of the current dipole, p1 is the magnitude of the current-dipole moment, and θ1 is the angle between the dipole moment vector (cid:126)p1 and the position vector (cid:126)r−(cid:126)r1. An important feature is that all time dependence of the single-neuron contribution to the potential Φ lies in p1(t) so that Φ1((cid:126)r, t) factorizes as Φ1((cid:126)r, t) = p1(t)g1((cid:126)r) . (91) For the electrical potential recorded at an EEG electrode, the forward model in Eq. 90 is no longer applicable due to different electrical conductivities of neural tissue, dura matter, scull and scalp. Analytical expressions analogous to Eq. 90 can still be derived under certain cir- cumstances such as with three-shell or four-shell concentric spherical head models (see Nunez and Srinivasan [25], Appendix G), but the key observation for the present argument is that the single-neuron contribution to the EEG will still factorize, i.e., Φ1((cid:126)r, t) = p1(t)g1((cid:126)r) where g1((cid:126)r) here is an unspecified function. The compound EEG signal from a set of Nn single-neuron current dipoles is now given by Φ((cid:126)r, t) = pn(t)gn((cid:126)r) , (92) where the index n runs over all single-neuron current dipoles. For each Fourier component (frequency) we now have Φ((cid:126)r, f ) = For the special case where the different single-neuron current dipoles moments are uncorre- n=1 lated we find that the power spectral density SEEG U C (f ) of the EEG is of the form [26] pn(f )gn((cid:126)r) . (93) U C ((cid:126)r, f ) = Φ((cid:126)r, f )2 = SEEG pn(f )2 gn((cid:126)r)2 . (94) (We have here introduced the notation 'UC', i.e. capitalized, to highlight the difference be- tween the present assumption of uncorrelated single-neuron current dipoles and the separate assumption of uncorrelated membrane currents onto individual neurons in the above sec- n=1 tions.) If the single-neuron current dipoles have the same power-law behavior in a particular frequency range, i.e., pn(f )2 ≈ cn/f αp, it follows directly that the EEG signal will inherit this power-law behavior: Nn(cid:88) n=1 Nn(cid:88) Nn(cid:88) SEEG U C ((cid:126)r, f ) = /f αp = GU C((cid:126)r)/f αp , (95) Nn(cid:88) pn(f )2 gn((cid:126)r)2 ≈(cid:16) Nn(cid:88) cn gn((cid:126)r)2(cid:17) n=1 n=1 23 where GU C((cid:126)r) determines the PSD amplitude, but not the slope. The inheritance of the single-neuron power-law behavior also applies to the case of corre- lated sources, provided that the pairwise coherences are frequency independent. By similar reasoning as above we then find SEEG C ((cid:126)r, f ) = (cid:12)(cid:12)(cid:12) Nn(cid:88) n=1 (cid:12)(cid:12)(cid:12)2 ≈ GC((cid:126)r)/f αp . pn(f ) gn((cid:126)r) (96) Analogous expressions for the PSD for the EEG can also be derived when both correlated and uncorrelated single-neuron current dipoles contribute, but we do not pursue this here; see Lind´en et al. [24] and Leski et al. [26] for more details. F. Numerical simulations The NEURON simulation environment [21] with the supplied Python interface [22] was used to simulate a layer-V pyramidal neuron from cat visual cortex [20]. The main motiva- tion for pursuing this was to allow for a direct numerical comparison with results from the ball and stick neuron to probe similarities and differences, see Fig. 2. In addition, NEURON was also used on the ball and stick neuron model to verify consistency with the analytical results above. Both the layer-V pyramidal neuron and the ball and stick neuron had a purely passive membrane, with specific membrane resistance Rm = 3 Ωm2, specific axial resistivity Ri = 1.5 Ωm, and specific membrane capacitance Cm = 0.01 F/m2. Simulations were performed with a time resolution of 0.0625 ms, and resulting data used for analysis had a time resolution of 0.25 ms. All simulations were run for a time period of 1200 ms and the first 200 ms were removed from the subsequent analysis to avoid transient upstart effects in the simulations. The digital cell reconstruction of the layer-V pyramidal neuron was downloaded from ModelDB (http://senselab.med.yale.edu/), and the axon compartments were removed. To ensure sufficient numerical precision compartmentalization was done so that no dendritic compartment was larger than 1/30th of the electrotonic length at 100 Hz (using the function lambda f(100) in NEURON), which resulted in 3214 compartments. The soma was modeled as a single compartment. The ball and stick neuron was modeled with a total of 201 segments, one segment was the iso-potential soma segment with length 20 µm and diameter 20 µm, and 200 segments 24 belonged to the attached dendritic stick of length 1 mm and diameter 2 µm. Simulations were performed with the same white-noise current trace injected into each compartment separately. The white-noise input current was constructed as a sum of sinu- soidal currents [13] I(t) = I0 1000(cid:88) sin(2πf t + ϕf ) (97) f =1 where ϕf represents a random phase for each frequency contribution. Due to linearity of the cable equation, the contributions of individual current inputs could be combined to compute the PSD of the soma potential, the soma current and the dipole moment resulting from current injection into all N compartments. In correspondence with Eq. 39, the summation of the contributions from the input currents of different segments i with membrane areas Ai was done differently for uncorrelated and correlated input currents. The uncorrelated PSDs, Suc, were computed according to while the correlated PSDs, Sc, were computed according to i=1 N(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) N(cid:88) i=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 Suc(ω) = ρiAiyi(ω)2 , Sc(ω) = ρiAiyi(ω) . (98) (99) Here, yi(ω) denotes the Fourier components of the signal y(t) (either soma potential, soma current or dipole moment due to input in one segment), the product ρiAi gives the total number of input currents into one segment i, and the density ρi represents ρd for dendritic input and ρs for somatic input. The total dipole moment (cid:126)p was in the numerical computations assumed to equal the dipole moment in one direction only: the direction along the stick for the ball and stick model, and the direction along the apical dendrite for the pyramidal neuron model, both denoted as the x-component, px. For the pyramidal neuron this is an approximation as the dipole moment also will have components in the lateral directions. However, the prominent 'open-field' asymmetry of the pyramidal neuron in the vertical direction suggests that this is a reasonable approximation when predicting contributions to the EEG signal. The current- dipole moment is then given by px = xiIi(t) , (100) N(cid:88) i=1 25 where Ii is the transmembrane current of compartment i, and xi is the corresponding x- position. III. RESULTS A. Biophysically detailed neuron model vs. ball and stick model To establish the relevance of using the simple ball and stick neuron to investigate the biophysical origin of power laws, we compare in Fig. 2 the normalized power spectral densities (PSDs) of the transmembrane soma current (row 1), the current-dipole moment (row 2), and the soma potential (row 3) of this model (column 1) with the corresponding results for a biophysically detailed layer-V pyramidal neuron (column 2); the rightmost column gives a direct comparison of PSDs. Both neuron models have a purely passive membrane and receive spatially distributed current input. As described in the Models section (II.E), the PSD of the single-neuron contribution to the EEG will be proportional to the PSD of the neuronal current-dipole moment given the observation that the extracellular medium, dura matter, scull and scalp appear to be purely ohmic [13, 25]. We here stick to the term 'current-dipole moment' even if the term 'single-neuron contribution to the EEG' could equally be used. A first striking observation is that unlike single-input PSDs (thin gray lines in Fig. 2), the PSDs resulting from numerous, homogeneously distributed input currents (thick lines) have a linear or quasi-linear appearance for high frequencies in these log-log plots, resembling 1/f α power laws. This is seen both when the numerous current inputs are correlated (green thick lines) and uncorrelated (blue thick lines). We also observe that the decay in the PSD with increasing frequency is strongest for the soma potential, somewhat smaller for the current-dipole moment, and smallest for the soma current. This is reflected in the power-law exponents α estimated at 1000 Hz from these PSDs, see legend in Fig. 2. Here we observe that α is largest for the soma potential (bottom row) and smallest for the soma current (top row). In the example in Fig. 2 we have assumed constant input current densities across the neurons, i.e., ρs = ρd. For this special case, correlated current input will, at all times, change the membrane charge density equally across the neuron, and as a consequence the neuron will be iso-potential. In this case the axial current within the neuron will be zero, and likewise 26 the net membrane current (with the capacitive current included) for any compartment, including the soma. As a consequence the current-dipole moment vanishes, and the model can effectively be collapsed to an equivalent single-compartment neuron. For the soma current and dipole moment we thus only show results for uncorrelated inputs in Fig. 2. However, correlated current input will still drive the soma potential (green curves in columns 1 and 2). Here we observe that the exponent α is smaller for uncorrelated input than for correlated input both for the ball and stick neuron and for the pyramidal neuron. The results above pertains to the situation with white-noise current inputs, i.e., flat-band PSDs. However, the results are easily generalized to the case with current inputs with other PSDs. Since our neuron models are passive and thus linear, the PSDs simply multiply. This is illustrated in column 3 of Fig. 2 which shows how our PSDs for uncorrelated input change with varying PSDs of the current input, s(ω). The blue curves correspond to white-noise input and are identical to the blue curves in column 2. The pink and brown curves illustrate the case of pink (1/f ) and Brownian (1/f 2) input, respectively. Since the PSDs multiply, the power-law exponent of the input noise simply adds to the exponent α. Thus, the pink and Brownian input increase the slope α with 1 and 2, respectively, compared to white-noise input. Even though the dendritic structure of the reconstructed pyramidal neuron is very dif- ferent from the ball and stick neuron in that it has both a highly branched structure and a varying diameter along its neural sections (tapering), both models seem to produce lin- ear or quasi-linear high-frequency PSDs in the log-log representation. Also the power-law exponents are found to be fairly similar. This implies that the ball and stick neuron model captures salient power-law properties of the more biophysically detailed neuron model, and motivates our detailed analytical investigation of the power-law properties of the ball and stick neuron following next. B. Power laws for ball and stick neuron In the Models section above we derived analytical expressions for the PSD transfer func- tions of the soma current (H I), current-dipole moment (H p) and soma potential (H V ) for the ball and stick neuron for spatially distributed input currents. The resulting transfer 27 functions H(f ), summarized in Eqs. 79-89, were of the form H(f ) =(cid:0)1 − c(f )(cid:1)(cid:0)Huc,s(f ) + Huc,d(f )(cid:1) + c(f ) Hc(f ) , (101) where Huc,s(f ) and Huc,d(f ) represent the contributions from uncorrelated somatic and den- dritic inputs, respectively, and Hc(f ) represents the contribution from correlated inputs. c = c(f ) is the pairwise coherence of the current inputs, all assumed to have the same PSDs (s = s(f )). These mathematical expressions are quite cumbersome, but they are dramatically sim- plified in the high-frequency limit, f → ∞, in which the dominant power can be found analytically by a series expansions of the mathematical expressions for the transfer func- tions in Eqs. 81-89. The expressions for the PSD transfer functions contain terms which are both polynomial and superpolynomial (i.e., including exponentials/exponentially decaying functions) with respect to frequency. As these superpolynomial terms will dominate the polynomial terms in the high-frequency limit, it follows from Eq. 49 that for high frequencies the absolute square of the denominator D can be approximated by D2 ≈ sinh(2aL)(cid:2)coth(2aL)(B2(a2 + b2) + 1)/2 + aB(cid:3) , (102) where terms decaying exponentially to zero with increasing frequency have been set to zero. The frequency dependence is through a and b, see Eqs. 50 and 51. Note that limf→∞ coth(2aL) = 1 since limf→∞ a = ∞. In the high-frequency limit the PSD trans- fer functions Eqs. 81-89 become c/(a2 + b2 + 2a/B + 1/B2) , c/[(a2 + b2)(a2 + b2 + 2a/B + 1/B2)] , ρ2[B2(a2 + b2) + 1 − 2aB] + 2ρ(aB − 1) + 1 (a2 + b2)2[B2(a2 + b2) + 2aB + 1] √ 2a(a2 + b2 + 2a/B + 1/B2)] , uc,d(a2 + b2)/[ uc,d a2 + b2 − 2a/B + 2/B2 √ √ 2a(a2 + b2)(a2 + b2 + 2a/B + 1/B2) uc,d/[ 2a(a2 + b2)(a2 + b2 + 2a/B + 1/B2)] , , c c ≈ AI H I c ≈ Ap H p c ≈ AV H V uc,d ≈ AI H I uc,d ≈ Ap H p uc,d ≈ AV H V , (103) (104) (105) (106) (107) (108) where the amplitudes A are found in Table II. When the PSDs expressed in Eqs. 103-107 28 are expanded reciprocally for high frequencies, i.e., W = ωτm = 2πf τm (cid:29) 1, we get √ (109) (110) (111) (112) H I uc,d/AI H I H p uc,d/Ap H p H V uc,d/AV H V c /AV 2/B + (1/B2 + 1/2)W −1/2 + O(W −1)] , √ 2W/B + (B2 + 6)W 1/2/2B2 + O(W 0)] , uc,d ≈ 1/[W 1/2 + √ 2W 1/2/B + 1/B2 + O(W −1/2)] , c ≈ 1/[W + c /AI uc,d ≈ 1/[(W 3/2 + 2 √ c ≈ 1/[W 2 + 2W 3/2/B + W/B2 + O(W 1/2)] , c /Ap uc,d ≈ 1/[W 5/2 + W 2 c ≈ 1/[W 2/ρ2 + W 3/2 √ 2/B + W 3/2(1/B2 + 1/2) + O(W 1)] , √ (113) 2(2ρ − 1)/Bρ3 + W (1 − 2ρ)2/B2ρ4 + O(W 1/2)] ,(114) where ρ is the dimensionless relative density, ρ = ρs/(ρs + ρd), and B = d2 s /λd, with ds and d denoting the somatic and dendritic diameter, respectively, and λ denoting the dendritic length constant. The expansions were done in Mathematica (version 7.0), and a list of parameters used throughout the present paper is given in Table I (along with the default numerical values used in the numerical investigations in later Results sections). In Eqs. 109-114 terms which are exponentially decaying to zero for large W have been approximated to zero. Note that Eq. 114 does not apply in the special case of no somatic input, ρ = 0, for which the series expansion gives √ H V c /AV c ≈ 1/[W 3B2 + W 5/2 2B + W 2 + O(W 3/2)] . (115) The corresponding high frequency expansions of the PSD transfer functions for uncorre- lated somatic input, Huc,s/Auc,s, are not shown, as these expressions are identical to the corresponding transfer functions for correlated input into the soma only, Hc/Ac (i.e., equal to Eqs. 110, 112 and 114 with ρ = 1). Eqs. 109-115 show that, due to position-dependent frequency filtering of the numer- ous inputs spread across the membrane (cf. Fig. 3B), all PSD transfer functions express asymptotic high-frequency power laws. Moreover, these genuine 'infinite-frequency' power- law exponents, denoted α∞, span every half power from α∞ = 1/2 (for H I α∞ = 3 (for H V uc,d, Eq. 109) to c , Eq. 115) for the different transfer functions. The results are summarized in Table II. To obtain the power-law exponents in the general case with contributions from both correlated and uncorrelated current inputs, we need to compare the different terms in the the general expression for H(f ) in Eq. 101. With different leading power-law exponents α∞ 29 in their asymptotic expressions, the term with the lowest exponent will always dominate for sufficiently high frequencies. From Table II we see that for all three quantities of interest, i.e., H I(f ), H p(f ) and H V (f ), the lowest exponent always comes from contributions from uncorrelated inputs. Note that the correlated term in Eq. 101 also involves a frequency- dependent coherence term c(f ), but to the extent it modifies the PSD, it will likely add an additional low-pass filtering effect [26] and, if anything, increase the power-law exponent. If we assume that the coherence is constant with respect to frequency we identify the following asymptotic exponents αall∞ (i.e., with 'all' types of possible input) for H I, H p and H V : αall,I∞ = 1/2, αall,p∞ = 3/2, αall,V∞ = 2 . Note that these power-law exponents are unchanged as long as uncorrelated activity is distributed both onto the soma and the dendrite, but will increase to αI∞ = 1 and αp∞ = 2 if no uncorrelated input are present on the dendrite. Similarly, without input onto soma, the asymptotic value will change for the soma potential PSD: it becomes αV∞ = 2.5 if uncorrelated input is uniformly distributed on the dendrite, and αV∞ = 3 if the dendritic input is correlated. C. Apparent power laws for experimentally relevant frequencies Detailed inspection of the power-law slopes for the ball and stick model in Fig. 2 and comparison with the power-law exponents α∞ listed in Table II reveal that although the curves might look linearly decaying in the log-log plot for high frequencies, the expressed exponents α are still deviating from their high-frequency values α∞, even at 1000 Hz. As experimental power laws have been measured for much lower frequencies than this, we now go on to investigate apparent PSD power laws for lower frequencies. For this it is convenient to define a low-frequency (lf ) regime, an intermediate-frequency (if ) regime and a high- frequency (hf ) regime, as illustrated in Fig. 3C. The transition frequencies between the regimes are given by the frequencies at which α is 50% and 90% of αall∞, respectively. The log-log decay rates of the PSD transfer functions can be defined for any frequency by defining the slope α(W ) as the negative log-log derivative of the PSD transfer functions, α(W ) = −d(log H)/d(log W ) . (116) 30 In Figs. 4, 5, and 6 we show color plots of α(W ) for the soma current (αI(W )), current- dipole moment (αp(W )), and soma potential (αV (W )), respectively, both for cases with uncorrelated and correlated inputs. The depicted results are found by numerically evaluating Eq. 116 based on the expressions for H listed in Eqs. 81-89. Note that since our model is linear, the log-log derivative is independent of the amplitude A. Thus, with either completely correlated or completely uncorrelated input, the dimensionless parameters B, L, ρ and W span the whole parameter space of the model. The 2D color plots in Figs. 4-6 depict α as function of W and B for three different values of the electronic length L = l/λ (L=0.25, 1, and 4), i.e., spanning the situations from a very short dendritic stick (L = 0.25) to a very long stick (L = 4). Electrotonic lengths greater than L = 4 produced plots that were indistinguishable by eye from the plots for L = 4. The thin black contour line denotes the transition between the low- and intermediate-frequency regimes (α = 0.5α∞), whereas the thick black contour line denotes the transition between the intermediate- and high-frequency regimes (α = 0.9α∞). FIG. 4 AROUND HERE Soma current first row applies to correlated inputs (H I Fig. 4 shows the slopes α of the PSD transfer functions for the soma current, H I. The c ) for all values of ρs and ρd as long as ρs (cid:54)= ρd. This independence of ρ = ρs/(ρs + ρd) is seen directly in the transfer functions in Eqs. 81 and 82. (For the special case ρs = ρd there will be no net somatic current). The plot in row 1 also applies to the case of uncorrelated current inputs onto the soma only (H I uc,s). That these particular PSD transfer functions have identical slopes are to be expected: correlated result pertains also to the special case ρd = 0 for which all input is onto the soma, and changing from correlated to uncorrelated current inputs onto the soma will only change the overall amplitude of the resulting soma current, not the PSD slope. The first row of Fig. 4 illustrates how the slope α approaches the asymptotic value αI∞ = 1 for correlated input (ρs (cid:54)= ρd) (and uncorrelated input onto the soma) for high frequencies, see Table II. It also shows that this asymptotic value is reached for lower frequencies when 31 B = d2 s /(dλ) is large, i.e., when the soma area is large compared to the effective area λd of the dendrite. Row 2 correspondingly shows how α for large frequencies approaches the asymptotic value of αI∞ = 1/2 (row 2) for uncorrelated input uniformly spread over the dendrite. For the case depicted in row 3, i.e., uncorrelated input onto both the soma and dendrite with ρs = ρd, the asymptotic high-frequency expression is seen to eventually be dominated by the lowest power, i.e., α ≈ αall,I∞ = 1/2. The lf regime, that is, the area to the left of the thin contour line, is seen to be quite substantial in Fig. 4, and is also highly dependent on B. For the default parameters, depicted by the white horizontal line, the left column in Fig. 4 shows that the lf regime extends up to much more than 100 Hz for compact neurons (L=0.25), and even for L = 1 and L = 4 (two rightmost columns) the lf regimes are substantial. (For our default membrane time constant of 30 ms, 100 Hz corresponds to the middle vertical white line in the panels.) Such a prominent lf regime was also seen for the pyramidal neuron in Fig. 2 where the normalized PSD for the somatic membrane current with uncorrelated input was almost constant up to 1000 Hz. It is also interesting that in some situations the soma current is band-pass filtered with respect to the input currents. This is especially seen in Fig. 4 for intermediate (L = 1) and long (L = 4) sticks with uncorrelated dendritic input currents (row 2), where the substantial dark blue area represents a band of negative α-values which is turning positive for higher frequencies, and the PSD thus is band-pass filtered around the frequencies corresponding to α = 0. For the higher frequencies within the frequency interval typically recorded in experiments (up to a few hundred hertz), Fig. 4 shows that one could expect some low-pass filtering for the intermediate and long sticks (l > λ), in particular if the current input is (i) predominantly onto the soma or (ii) correlated, and the neuron has a large value of B. However, as indicated by Fig. 2, this effect may be very small for pyramidal neurons. FIG. 5 AROUND HERE 32 Current-dipole moment / EEG contribution Fig. 5 shows corresponding slope plots of the PSD for the current-dipole moment, H p, i.e., the single-neuron contribution to the EEG. The panels are organized as for the soma current in Fig. 4, and as for the soma current we observe that for high frequencies α approaches the asymptotic value αp∞=2 for the cases with either correlated input (ρs (cid:54)= ρd) or uncorrelated input onto the soma only (row 1), see Table II. Further, for the case with uncorrelated input on the dendrites, α is seen to approach the predicted αall,p∞ = 1.5 (rows 2 and 3). Moreover, as for the soma current the lf regime is seen to be large for compact neurons (L=0.25). For such neurons one would thus expect very little filtering within the frequency interval typically recorded for the EEG, typically up to 100 or 200 Hz (middle vertical white line in panels). For less compact neurons (L=1 and 4), the filtering is, however, seen to be substantial also within the frequency interval from 10 to 100 Hz, even for low values of B. This filtering is seen to be even more prominent for the pyramidal neuron in Fig. 2, suggesting that the filtering could be of considerable importance for the large pyramidal neurons in human cortex thought to dominate human EEG. The if regime is seen to be quite narrow in all panels in Fig. 5, implying that the PSD has a quite abrupt transition to the hf regime where the slope is quite constant and close to its asymptotic values αp∞. The pyramidal neuron receiving uncorrelated input in Fig. 2, however, is seen to obey an approximate power-law with αp of only about 1.25 at 1000 Hz. This is not within the range defined here as the hf regime, i.e., α ≥ 0.9αp∞ = 1.35, but rather within the upper range of the if regime. FIG. 6 AROUND HERE Soma potential In Fig. 6 the slopes α of the PSD of the soma potential are shown. Unlike H I c and for the soma potential with correlated input currents H p c , the PSD transfer function H V c varies with ρ = ρs/(ρs + ρd), and is also non-zero for ρs = ρd, cf. Eq. 83. More panels are thus needed to describe the model predictions properly: Row 1 corresponds to correlated 33 correlated (H V dendritic input (H V input onto the dendrite only (H V c (ρs = 0)), row 2 corresponds to somatic input only, either uc,s), while row 3 corresponds to uncorrelated uc,d). The two bottom rows correspond to homogeneous input onto the whole neuron, i.e., ρd = ρs, with uncorrelated input in row 4 and correlated input in row 5. c (ρd = 0)) or uncorrelated (H V The different panels of Fig. 6 display quite varied PSD slopes for the various scenarios of input current. Row 1 shows that for correlated input solely onto the dendrite, α is quite close to the asymptotic value αV∞=3 (cf. Table II) for modest frequencies, even for the compact neuron with L = 0.25. The narrow if region and large power-law exponent α in row 1 makes this case quite different from the results depicted in the other panels. With input instead onto the soma only (row 2), for example, a completely different slope pattern is observed: for compact neurons (L = 0.25) the log-log slope of the PSD is seen to have regions with a positive double derivative (concave slope), with the consequence that the if regime is divided into two distinct frequency regions with an intermediate hf interval. Row 3 depicts the case with uncorrelated input onto the dendrites. Qualitatively the results resemble the case with correlated dendritic inputs in row 1, except that here α approaches the asymptotic values αV∞ = 2.5 (cf. Table II), rather than 3. For the non- compact neurons (L = 1 and L = 4) the default parameters give an if region for uncorrelated dendritic input which goes up to almost 100 Hz. However, the thick contour line illustrates that the transition to the hf regime is highly dependent on the values B, and a slightly larger B is seen to substantially lower the transition frequency to the hf regime. With uncorrelated input homogeneously distributed over the whole neuron, i.e., ρs = ρd (row 4), we observe a similar pattern of power-law exponents as for somatic input only (row 2). Thus the contribution from the soma for which αV∞ = 2, dominates the contribution from the dendritic inputs where αV∞=2.5. Another observation is that for the non-compact neurons (L=1 and 4) the if regime is wide for a large range of B values. For the default parameters corresponding to B=0.2 we observe that the if interval stretches from less than 10 Hz to almost 1000 Hz. For the last example case in row 5 with correlated input spread homogeneously onto the whole neuron (ρs = ρd) we observe that α is independent of the parameter B. For homogenous correlated input the whole neuron is iso-potential and corresponds to a single- compartment neuron with zero dipole moment and zero net membrane current, as reflected in the vanishing amplitudes of AI c and Ap c in Table II. In this special case the spatial extension 34 of the dendritic stick will not affect the filtering properties of the neuron, and the PSD transfer function can be expressed as a simple Lorentzian, i.e., H V c (cid:12)(cid:12)ρ=0.5 ∝ 1/(1 + W 2). The slope α is thus solely determined by the membrane time constant τm hidden within the dimensionless frequency W = 2πf τm. D. PSDs for varying biophysical parameters for ball and stick neuron The 2D color plots in Figs. 4 -- 6 depicting the slopes α of the PSDs of the transfer functions H(f ), give a comprehensive overview of the power-law properties of the ball and stick model as they are given in terms of the three key dimensionless parameters W = ωτm = 2πf τm, B = d2 s/dλ, and L = l/λ. To get an additional view of how the model predictions depend on biophysical model parameters, we plot in Figs. 7 and 8 PSDs, denoted S(f ), for a range of model parameters for the soma current, current-dipole moment and soma potential when the neuron receives homogeneous white-noise current input across the dendrite and/or the soma. We focus on biophysical parameters that may vary significantly from neuron to neuron: the dendritic stick length l, the specific membrane resistance Rm, the dendritic stick diameter d, and the soma diameter ds. The specific membrane resistance may not only vary between neurons, but also between different network states for the same neuron [27, 28]. To predict PSDs S(f ) of the various measurements, and not just PSDs of the transfer functions H(f ), we also need to specify numerical values for the current-input densities ρd and ρs (and not only the ratio ρ = ρs/(ρs + ρd)), as well as the magnitude of the PSDs of the current inputs. These choices will only affect the magnitudes of the predicted PSDs, not the power-law slopes. As the numerical values of the slopes predicted by the present work suggest that channel noise from intrinsic membrane conductances rather than synaptic noise dominates the observed noise in experiments (see Discussion), we gear our choice of param- eters towards intrinsic channel noise. We first assume the input densities ρd and ρs (when they are non-zero) to be 2 µm−2, in agreement with measurements of the density of the large conductance calcium-dependent potassium (BK) channel [29]. Next we assume the magni- tude of PSD of the white-noise current input to be s(f )=const.=1 fA2/Hz. This choice for s gives magnitudes of predicted PDSs of the soma potential, assuming uncorrelated current inputs, in rough agreement with what was observed in [7], i.e., about 10−3 -- 10−2 mV2/Hz for low frequencies. 35 Figs. 7 and 8 show PSDs for uncorrelated and correlated input currents, respectively. A first observation is that the predicted PSD magnitudes are typically orders of magnitude larger for correlated inputs, than for uncorrelated inputs. With the present choice of pa- rameters, the cases with correlated inputs predict PSDs for the soma potential and soma current much larger than what is seen in experiments [7, 9, 10]. A second observation is that variations in the dendritic stick length (first column in Figs. 7-8) and membrane resistance (second column) typically have little effect on the PSDs at high frequencies, but may sig- nificantly affect the cut-off frequencies, i.e., the frequency where the PSD kinks downwards. This may be somewhat counterintuitive, especially that the PSDs for the current-dipole moment are independent of stick length l as one could think that a longer stick gives a larger dipole moment. For the ball and stick neuron, however, this is not so: input currents injected far away from both boundaries (ends) of a long stick will not contribute to any net dipole moment, as the input current will return symmetrically on both sides of the injection point and thus form a quadrupole moment. This symmetry is broken near the ends of the stick: for uncorrelated input a local dipole is created at each endpoint; for correlated input the dendrite will be iso-potential near the distal end of the stick, while a local dipole will arise at the somatic end if ρd (cid:54)= ρs. Note though that this is expected to be different for neurons with realistic dendritic morphology, since the dendritic cables typically are quite asymmetric due to branching and tapering. The effects of varying the dendritic stick diameter and soma diameter are quite different (cf., two rightmost columns in Figs. 7 -- 8). Here both the magnitudes and the slopes of the high-frequency parts are seen to be significantly affected. On the other hand, the cut-off frequency is seen to be little affected when varying the soma diameter ds, in particular for the current-dipole (Sp) and soma potential (SV ) PSDs. (Note that for the case with homogeneous correlated input, ρs = ρd (row 4 in Fig. 8), the ball and stick model is effectively reduced to a single-compartment neuron for which the PSD is independent of d and ds.) In Figs. 4 -- 6 regions in the log-log slope plots were observed to have positive double derivatives, i.e., concave curvature. The effect was particularly prevalent for the soma po- tential transfer function H V in the case of short dendritic sticks (L = 0.25) with dominant current input to the soma. This feature is also seen in the corresponding 'soma-input' curves (bottom rows of Figs. 7 -- 8), also for non-compact sticks, i.e., for the default value l=1 mm (L=1). 36 IV. DISCUSSION A. Summary of main findings In the present work we have taken advantage of the analytical tractability of the ball and stick neuron model to obtain general expressions for the power spectral density (PSD) transfer functions for a set of measures of neural activity: the somatic membrane current, the current-dipole moment (corresponding to the single-neuron EEG contribution), and the soma potential. With homogeneously distributed input currents both onto the dendritic stick and with the same, or another current density, onto the soma we find that all three PSD transfer functions, relating the PSDs of the measurements to the PSDs the noisy inputs currents, express asymptotic high-frequency 1/f α power laws. The corresponding power- law exponents are analytically identified as αI∞ = 1/2 for the somatic membrane current, αp∞ = 3/2 for the current-dipole moment, and αV∞ = 2 for the soma potential. These power- law exponents are found for arbitrary combinations of uncorrelated and correlated noisy input current (as long as both the dendrites and the soma receive some uncorrelated input currents). The significance of this finding goes beyond neuroscience as it demonstrates how 1/f α power laws with a wide range of values for the power-law exponent α may arise from a sim- ple, linear physics equation. We find here that the cable equation describing the electrical properties of membranes, transfers white-noise current input into 'colored' 1/f α-noise where α may have any half-numbered value within the interval from 1/2 to 3 for the different mea- surement modalities. Intuitively, the physical underpinning of these novel power laws is the superposition of numerous low-pass filtered contributions with different cut-off frequencies (i.e., different time constants) due to the different spatial positions of the various current inputs along the neuron. As our model system is linear, the results directly generalize to any colored input noise, i.e., transferring 1/f β spectra of input currents to 1/f β+α output spectra. B. Comparison with power laws observed in neural recordings Our ball and stick model expressions for the PSDs cover all frequencies, not just the high frequencies where the power-law behavior is seen. When comparing with results from neural 37 recordings, one could thus envision to compare model results with experimental results across the entire frequency spectrum. However, the experimental spectra will generally be superpositions of contributions from numerous sources, both from synapses [28] and from ion channels [7]. These various types of input currents will in general have different PSDs, i.e., different s(f ). A full-spectra comparison with our theory is thus not possible without specific assumptions about the types and weights of the various noise contributions, information which is presently not available from experiments. However, the presence of power-law behavior at high frequencies implies that a single noise process (or several noise processes with identical power-law exponent) dominates the others in this frequency range. In in vivo experiments from neocortical neurons where PSDs for frequencies up to 1000 Hz or more have been used to estimate power-law exponents, the soma potential has typically been seen to express power laws with αV exp close to 2.6 [8, 11, 30]. In an analogous experiment in hippocampal cell culture where the PSD for frequencies up to 500 Hz was measured, a value of αV but for voltage-clamped neurons in hippocampal cell cultures a power-law with αI exp of about 2.4 was estimated [7]. For the soma current the results are fewer, exp = 1.1 was seen in the high-frequency end of the PSD recorded up to 500 Hz [7]. For the pyramidal neuron depicted in Fig. 2 we correspondingly found αV =1.61 and αI=0.15 for the PSD of the transfer functions for uncorrelated current inputs. Thus if these uncorrelated input current sources themselves have a pink (1/f , i.e., β=1) power-law dependence of the PSD in the relevant frequency range, the power-law exponents of the model PSDs become αV +β=1.61+1=2.61 and αI + β=1.15, intriguingly close to the experimental observations. Note that while these model results pertain for a particular choice of model parameters for the pyramidal neuron, the results shown in Fig. 7 for the ball and stick neuron implies that moderate changes in the model parameters will yield modest changes in predicted power-law exponents. For the EEG, only experimental findings up to frequencies of 100 Hz are available, and here estimated power laws have exhibited a large variation in power-law exponents with αexp's varying between 1 and 2 [6]. If uncorrelated pink-noise (1/f ) input currents are assumed also here, the pyramidal neuron results in Fig. 2 imply αp + β=2.25, i.e., the single-neuron contribution to the EEG exhibits a power law with an exponent somewhat above the typical value for macroscopic recordings. Note, however, that even for pink-noise input, shorter dendritic sticks may imply power-law exponents as small as αp + β=1 for the 38 single-neuron EEG contribution within the lower frequency range typically probed in EEG recordings (Fig. 5, [6]). Note also that the link between our model predictions and putative EEG power laws is more tenuous as it involves the additional assumption that network dynamics do not affect the observed high-frequency power laws, i.e., the high-frequency tail of the input current PSD s(f ) of the neurons giving the dominant contributions to the EEG signal. On balance we think the comparison with presently available experiments supports an hypothesis that the observed microscopic (soma potential, soma current) power laws, or more precisely 'apparent' power laws, stem from (i) uncorrelated membrane current sources which are (ii) homogeneously distributed across the neural membranes and (iii) each have a pink (1/f ) noise distribution. Further, while experimental data presently are scarce, this hypothesis may also explain the presence of power laws in EEG recordings. Note, however, that power-law exponents alone are not sufficient to uniquely determine whether the dominant inputs are correlated or uncorrelated. As seen for the 'infinite- frequency' power-law exponents α∞ in Table II and Figs. 4-8, α's are equal to or larger for correlated inputs than for uncorrelated inputs for our ball and stick neuron; the typical difference for α∞ being 1/2. Thus correlated current inputs with power-law PSDs with an exponent β of about 1/2 (rather than the pink-noise value of β=1) would give about the same power-law exponent (α+β) in the various measurements. Note also that since the power-law exponents α with uncorrelated inputs are generally smaller than for correlated inputs, the uncorrelated contributions will in principle always dominate for sufficiently high frequencies. However, the contribution from correlated current inputs scales differently with the number density of input currents than for uncorrelated inputs: the PSD grows as the square of the input densities (ρs, ρd) for correlated inputs, while it grows only linearly with these input densities for uncorrelated inputs. Thus in experimental settings the relative contributions from correlated and uncorrelated current inputs will depend on the size of these densities as well as the value of the coherence c, parameters which cannot be expected to be universal, but rather depend on the biophysical nature of the underlying current noise source. It is thus difficult to a priori assess whether the noise spectra are dominated by correlated or uncorrelated input. As argued on biophysical grounds in the next subsection, however, we think that the explanation assuming uncorrelated inputs with pink noise is more likely. 39 C. Origin of noise Our supposition that homogeneously distributed, uncorrelated pink-noise current input, i.e. s = s(f ) ∼ 1/f input, may underlie the observed power-law behavior in the soma cur- rent, soma potential, and possibly also EEG spectra, applies regardless of the biophysical nature of the underlying noisy currents. Nevertheless, we will in the following discuss the origin of the noise which in previous modeling studies have been assumed to stem from synapses [10, 11], intrinsic ion channels [7], or a combination of the two [9]. Our argument presented below that the high-frequency power-law behavior predominantly stems from chan- nel noise follows from (i) the observed lack of change in power-law exponents when synaptic inputs are blocked [9], (ii) the previous experimental observation of pink noise in certain ion channels [15, 31 -- 33], and (iii) the difficulty of reconciling pink noise from synapses with prevailing synapse models [10, 11]. (Note, however, that this does not imply that channel noise rather than synaptic noise dominates the PSD at lower frequencies.) Most detailed studies of neuronal noise spectra have focused on the soma potential [7 -- 10, 30], where interestingly the same power-law exponent of about αV = 2.6 have been observed both in in vivo [8, 11, 30] and in vitro conditions [7, 9, 10]. The most parsimonious explanation is that the same noise process dominates under both conditions, i.e., that the higher spiking activity in in vivo conditions than in in vitro conditions is not the key process underlying the observed power law. This suggests that the dominant mechanism rather is noise stemming from intrinsic ion channels. This conjecture is supported by the observation that the slope of the soma potential power law in rat neocortical slices was not affected by application of synaptic blockers (DNQX, gabazine) [9]. (However, the synaptic blockers reduced the overall amplitude of the PSD, which would imply that a secondary effect is involved, e.g., that blocking of synaptic inputs indirectly affects the amplitude of the ion- channels noise by, for example, changing the intracellular calcium concentration.) Further, it has been difficult to account for 1/f input spectra in model studies based on assuming a synaptic origin. In [10] and [11] synapses were spread evenly across dendrites of morphologically reconstructed neurons and were activated by presynaptic spike trains assumed to have Poissonian distributions (cf. Fig. 1). With current-based exponential synapses, the PSD of the current noise source will then have the form of a Lorentzian, i.e., s(f ) ∝ 1/(1 + (2πf τs)2), where τs is the synaptic time constant. For high frequencies this 40 implies s ∼ 1/f 2 (β=2), cf. results for Brownian (1/f 2) input in right column of Fig. 2. As previous studies also found that this implies a too large value for the soma potential power-law exponent several approaches has been suggested to compensate for this: [34] showed that network correlations due to delay distributions can give non-Poissonian pre- synaptic spike-train statistics and thus change the power-law exponent. Alternatively, small synaptic time constants (τs = 2 ms in [10]) will give a higher cut-off frequency (f = 1/2πτs) for the transition to the high-frequency power-law regime. If this cut-off frequency is in the upper range of the recorded frequency interval, s(f ) will essentially be independent of frequency (i.e., white) and apparent soma-potentials power laws with smaller exponents can be obtained. In [11] it was instead suggested that a non-ideal membrane capacitance could have a compensatory effect. In contrast, several recordings of PSDs of the intrinsic channel noise in potassium channels have shown 1/f scaling [15, 31 -- 33], i.e., exactly the type of pink input-current noise spectrum required for our model prediction to be in accordance with the experimentally observed power-law exponents in the PSDs of the soma current, soma potential, and EEG. If the power law indeed stems from intrinsic channel noise, uncorrelated rather than correlated current sources are expected. Again this agrees with predictions from comparing our model with experiments. Further, it is tempting to speculate on what particular type of ion channel could give rise to the observed power-law spectra. Several experiments have hinted that potassium channels may be important sources of membrane noise [7, 15, 31, 32], and of those a natural candidate is the BK ('big') potassium channel which has a large single- channel conductivity and thus the potential for large current fluctuations. D. Power laws for local field potentials (LFP) and ECoG signals Power laws have also been reported in recordings of extracellular potentials inside (local field potential; LFP) and at the surface of cortex (electrocorticography; ECoG). However, the reported power-law exponents vary a lot, with αexp's between 1 and 3 for LFPs [35 -- 38] and between 2 and 4 for ECoG [39 -- 43]. From a modeling perspective the single-neuron contribution to putative power-law exponents for these signals is more difficult as, unlike the EEG signal, the single-neuron contributions are not determined only by the current- dipole moment: dominant contributions to these signals will come from neurons close to the 41 electrode (typically on the order of hundred or a few hundred micrometers [24]), so close that the far-field dipole approximation relating the current-dipole moment directly to the contributed extracellular potential [14] is not applicable [24]. A point to note, however, is that it may very well be that power laws observed in the LFP or ECoG are dominated by other current sources than the power laws observed in the EEG spectra: As observed in [24, 26] (see also [44]) the LFP recorded in a cortical column receiving correlated synaptic inputs can be very strong, and it is thus at least in principle conceivable that power laws in the LFP may stem from synaptic inputs from neurons surrounding the electrode, whereas the EEG signal, which picks up contributions from a much larger cortical area, may be dominated by uncorrelated noise from ion channels. Further, the soma potential and soma current of each single neuron may also still be dominated by uncorrelated channel noise, even if the the LFP is dominated by correlated synaptic activity. This is because correlated synaptic inputs onto a population of neurons add up constructively in the LFP, whereas the uncorrelated inputs do not [24, 26]. For single-neuron measures such as the soma potential and soma current there will be no such population effects, and the uncorrelated inputs will more easily dominate the power spectra. As a final comment it is interesting to note that in the only reported study we are aware of for the frequency range 300-3000 Hz, the PSD of the LFP exhibited a power law with a fitted exponent of α=1.1 [37]. This is very close to what would be predicted if the LFP was dominated by the soma current from uncorrelated (pink) noise sources: In Table II we see that the 'infinite-frequency' power-law exponent for the transfer function from dendritic current inputs to soma current is αI∞ = 0.5. With a pink (1/f ) PSD of the input noise current, the 'infinite-frequency' prediction for the soma current exponent will thus be 1.5. This is already fairly close to the experimental observation of 1.1. Further, from Fig. 4 it follows that the apparent power-law coefficient for the transfer-function power law may be somewhat smaller than 0.5 in the frequency range of interest, suggesting that the agreement between experiments and model predictions assuming uncorrelated noise may be even better. If so, it may be that the LFP power spectra are dominated by synaptic inputs for frequencies below a few hundred hertz (with rapidly decaying LFP contributions with increasing frequency, i.e., higher power-law exponents in accordance with [35, 36, 38]), while uncorrelated inputs, and thus power laws with smaller exponents, dominate at higher frequencies. 42 E. Passive approximation In the present analysis we have modeled the membranes of somas and dendrites as simple passive linear (RC) circuit elements. This implies a strictly linear response to the current inputs, allowing for the present frequency-resolved (Fourier) analysis. The present results also serve as a starting point for the exploration of non-linear effects, for example due to active membrane conductances. Close to the resting potential of the neuron, the active conductances can be linearized, and the neuron dynamics can be described by linear theory with quasi-active membrane modeled by a combination of resistors, capacitors and inductors [45, 46]. These extra circuit elements will change the PSD. For example, the inductor typically introduces a resonance in the system. In Koch [46] the impedance for this 'quasi- active' membrane was however found to coincide with the impedance for a purely passive membrane for frequencies above 200 Hz, implying that the predicted high frequency power laws will be about the same. This is in accordance with experimental results from neocortical slices, where blocking of sodium channels were shown mainly to affect the soma potential PSD for frequencies below 2 Hz [9]. Nevertheless, the investigation of the role of active conductance on PSDs is a topic deserving further investigations. F. Concluding remarks A key conclusion from the present work is that the power-law predictions from our models are in close agreement with experimental findings for the soma potential and the soma current provided the transmembrane current sources are assumed to be (i) homogeneously distributed throughout the whole neuron, (ii) uncorrelated, and (iii) have a pink (1/f ) noise distribution. It should be stressed that we do not argue against synaptic noise being a major component underlying neural noise spectra; the importance of synaptic inputs in setting the noise level has been clearly demonstrated, for example by the large difference in membrane potential fluctuation between in vivo and in vitro preparations [28, 30]. We rather suggest that the power-law behavior seen at the high-frequency end of these noise spectra are dominated by intrinsic channel noise, not synaptic noise. We also speculate that potassium channels with inherent noisy current with PSDs follow- ing a 1/f distribution in the relevant frequency range, underlie the observed power laws, and 43 the BK channel is suggested as a main contributor. If future experiments indeed confirm that the BK channel is a dominant source of membrane noise, this may have direct implica- tion of the understanding several pathologies. Not only has the BK channel been implicated as a source of increased neural excitability [47] and epilepsy [48], but also disorders such as schizophrenia [49], autism and mental retardation [50] have been linked to the BK channel through a decrease in its expression [51]. ACKNOWLEDGMENTS We thank Eivind Norheim for useful feedback on the manuscript and for testing the reproducibility of Figs. 4-6. This work was partially funded by the Research Council of Norway (eVita [eNEURO], NOTUR,ISP) and EU Grant 269921 (BrainScaleS). [1] M. P. H. Stumpf and M. A. Porter, Science 335, 665 (2012). [2] B. B. Mandelbrot, The Fractal Geometry of Nature (New York: W. H. Freeman and Co., 1977). [3] P. Bak, C. Tang, and K. Wiesenfeld, Phys Rev Lett 59, 381 (1987). [4] H. Berger, European Archives of Psychiatry and Clinical Neuroscience 87, 527 (1929). [5] G. Buzs´aki, Rhythms of the brain (Oxford University Press, 2006). [6] W. J. Freeman, M. D. Holmes, B. C. Burke, and S. Vanhatalo, Clin Neurophysiol 114, 1053 (2003). [7] K. Diba, H. A. Lester, and C. Koch, J Neurosci 24, 9723 (2004). [8] M. Rudolph, J. G. Pelletier, D. Par´e, and A. Destexhe, J Neurophysiol 94, 2805 (2005). [9] G. A. Jacobson, K. Diba, A. Yaron-Jakoubovitch, Y. Oz, C. Koch, I. Segev, and Y. Yarom, J Physiol 564, 145 (2005). [10] A. Yaron-Jakoubovitch, G. A. Jacobson, C. Koch, I. Segev, and Y. Yarom, Front Cell Neurosci 2, 3 (2008). [11] C. B´edard and A. Destexhe, Biophys J 94, 1133 (2008). [12] K. H. Pettersen and G. T. Einevoll, Biophys J 94, 784 (2008). [13] H. Lind´en, K. H. Pettersen, and G. T. Einevoll, J Comput Neurosci 29, 423 (2010). 44 [14] M. Hamalainen, R. Haari, R. J. Ilmoniemi, J. Knuutila, and O. V. Lounasmaa, Reviews of Modern Physics 65, 413 (1993). [15] H. E. Derksen and A. A. Verveen, Science 151, 1388 (1966). [16] W. Rall, Experimental Neurology 1, 491 (1959). [17] I. Segev, J. Rinzel, and G. M. Shepherd, eds., Theoretical Foundations of Dendritic Function: The Collected Papers of Wilfrid Rall with Commentaries (MIT Press, 1994). [18] H. C. Tuckwell and J. B. Walsh, Biological Cybernetics 49, 99 (1983). [19] K. H. Pettersen, H. Lind´en, A. M. Dale, and G. T. Einevoll, in Handbook of Neural Activity Measurement, edited by R. Brette and A. Destexhe (Cambridge University Press, 2012) pp. 92 -- 135. [20] Z. F. Mainen and T. J. Sejnowski, Nature 382, 363 (1996). [21] N. T. Carnevale and M. L. Hines, The NEURON Book (Cambridge University Press, 2006). [22] M. L. Hines, A. P. Davison, and E. Muller, Front Neuroinformatics 3, 1 (2009). [23] W. Rall and H. Agmon-Snir, in Methods in Neuronal Modeling: From Ions to Networks, edited by C. Koch and I. Segev (MIT Press, Cambridge, MA, 1998) 2nd ed., pp. 27 -- 92. [24] H. Lind´en, T. Tetzlaff, T. C. Potjans, K. H. Pettersen, S. Grun, M. Diesmann, and G. T. Einevoll, Neuron 72, 859 (2011). [25] P. L. Nunez and R. Srinivasan, Electric fields of the brain, 2nd ed. (Oxford University Press, Inc., 2006). [26] S. Leski, H. Lind´en, T. Tetzlaff, K. H. Pettersen, and G. T. Einevoll, PLoS Computational Biology 9, e1003137 (2013). [27] J. Waters and F. Helmchen, Journal of Neuroscience 26, 8267 (2006). [28] A. Destexhe and M. Rudolph-Lilith, Neuronal noise (Springer, 2012). [29] N. Benhassine and T. Berger, Eur J Neurosci 21, 914 (2005). [30] A. Destexhe, M. Rudolph, and D. Par´e, Nat Rev Neurosci 4, 739 (2003). [31] D. J. Poussart, Biophys J 11, 211 (1971). [32] H. M. Fishman, Proc Natl Acad Sci U S A 70, 876 (1973). [33] Z. Siwy and A. Fuli´nski, Physical Review Letters 89, 158101 (2002). [34] S. E. Boustani, O. Marre, S. B´ehuret, P. Baudot, P. Yger, T. Bal, A. Destexhe, and Y. Fr´egnac, PLoS Comput Biol 5, e1000519 (2009). [35] C. B´edard, H. Kroger, and A. Destexhe, Phys Rev Lett 97, 118102 (2006). 45 [36] J. Milstein, F. Mormann, I. Fried, and C. Koch, PLoS One 4, e4338 (2009). [37] J. Martinez, C. Pedreira, M. J. Ison, and R. Q. Quiroga, J Neurosci Methods 184, 285 (2009). [38] G. Baranauskas, E. Maggiolini, A. Vato, G. Angotzi, A. Bonfanti, G. Zambra, A. Spinelli, and L. Fadiga, Journal of Neurophysiology 107, 984 (2012). [39] W. J. Freeman, L. J. Rogers, M. D. Holmes, and D. L. Silbergeld, Journal of Neuroscience Methods 95, 111 (2000). [40] W. J. Freeman, Cogn Neurodyn 3, 105 (2009). [41] K. J. Miller, L. B. Sorensen, J. G. Ojemann, and M. den Nijs, PLoS Comput Biol 5, e1000609 (2009). [42] B. J. He, J. M. Zempel, A. Z. Snyder, and M. E. Raichle, Neuron 66, 353 (2010). [43] J. M. Zempel, D. G. Politte, M. Kelsey, R. Verner, T. S. Nolan, A. Babajani-Feremi, F. Prior, and L. J. Larson-Prior, Frontiers in Sleep and Chronobiology 3, 76 (2012). [44] G. Einevoll, H. Lind´en, T. Tetzlaff, S. Leski, and K. Pettersen, in Principles of Neural Coding, edited by R. Q. Quiroga and S. Panzeri (CRC Press, 2013) pp. 37 -- 59. [45] W. K. Chandler, R. Fitzhugh, and K. S. Cole, Biophys J 2, 105 (1962). [46] C. Koch, Biol Cybern 50, 15 (1984). [47] R. C. Gerkin, R. L. Clem, S. Shruti, R. E. Kass, and A. L. Barth, J Clin Neurophysiol 27, 425 (2010). [48] W. Du, J. F. Bautista, H. Yang, A. Diez-Sampedro, S.-A. You, L. Wang, P. Kotagal, H. O. Luders, J. Shi, J. Cui, G. B. Richerson, and Q. K. Wang, Nat Genet 37, 733 (2005). [49] L. Zhang, X. Li, R. Zhou, and G. Xing, Med Hypotheses 67, 41 (2006). [50] F. Laumonnier, S. Roger, P. Gu´erin, F. Molinari, R. M'rad, D. Cahard, A. Belhadj, M. Ha- layem, A. M. Persico, M. Elia, V. Romano, S. Holbert, C. Andres, H. Chaabouni, L. Colleaux, J. Constant, J.-Y. L. Guennec, and S. Briault, Am J Psychiatry 163, 1622 (2006). [51] U. S. Lee and J. Cui, Trends Neurosci 33, 415 (2010). 46 FIGURE LEGENDS Figure 1: Schematic illustration of the input-output relationship between transmembrane currents (input) and the different measurement modalities (output). The transmembrane currents are illustrated by synaptic currents and channel currents. A synaptic current is commonly modeled by means of exponentially decaying functions (synaptic kernel) trig- gered by incoming spike trains, whereas a channel current typically is modeled by a channel switching between an open state (o), letting a current with constant amplitude through the channel, or a closed state (c). The input currents are filtered by the neuronal cable, result- ing in a low-pass filtered output current in the soma with a power spectral density (PSD) designated SI. The PSDs of the other measurement modalities studied here, i.e., the soma potential (SV ) and the current-dipole moment giving the single-neuron contribution to the EEG (Sp), are typically even more low-pass filtered, as illustrated by the PSDs plotted in the lower right panel. Figure 2: Normalized power spectral densities (PSDs) for the soma current, the current- dipole moment (i.e., EEG contribution) and the soma potential for a ball and stick neuron and a pyramidal neuron. A homogeneous density of noisy input currents is applied through- out the neural membrane. Columns 1 (ball and stick neuron) and 2 (pyramidal neuron) show PSDs for white-noise input, the blue and green lines correspond to uncorrelated and correlated input currents, respectively. Note that there is no green line in the two upper rows, since a homogeneous density of correlated inputs throughout the neuron gives no net soma current or dipole moment. An ensemble of PSDs from 20 single input currents for the ball and stick neuron and 107 single input currents for the pyramidal neuron is shown in grey. The results for the most distal synapses are shown in dark grey and the results for the proximal synapses in light grey, corresponding to the color shown in the filled circles at the respective neuron morphology (between columns 1 and 2). Column 3 illustrates how different power-law spectra of the input currents change the output PSDs: the blue, pink and brown lines express the PSD for uncorrelated white (constant), pink (1/f ) and Brown- ian noise input (1/f 2), respectively. The values of α in legends denote estimated power-law exponents at 1000 Hz, i.e., the negative discrete log-log derivative, −∆(log S)/∆(log f ). In 47 the rightmost column the values of α correspond to pink noise input, for Brownian noise input and white-noise input the values are '+1' and '-1' with respect to the pink input, respectively, as indicated by the brown '+' and the blue '-'. The ball and stick neuron was simulated with 200 dendritic segments, while the pyramidal neuron was simulated with 3214 dendritic segments. Broken lines correspond to the ball and stick neuron, whole lines to the pyramidal neuron. Figure 3: Schematic illustration of the ball and stick neuron model and its filtering prop- erties. (A) Schematic illustration of the ball and stick neuron model with a single input at a given position X = X(cid:48). The lumped soma is assumed iso-potential and located at X = 0. (B) Frequency-dependent current-density envelopes of return currents for a ball and stick neuron with input at X = 0.8L. The somatic return currents are illustrated as current densities from a soma section with length 20 µm placed below the stick. For 1 Hz, 10 Hz, 100 Hz and 1000 Hz the amplitudes of the somatic return currents are about 1/7.3, 1/7.5, 1/22 and 1/3100 of the input current, respectively. Parameters used for the ball and stick neuron model: stick diameter d = 2 µm, somatic diameter ds = 20 µm, stick length l = 1 mm, specific membrane resistance Rm = 3 Ωm2, inner resistivity Ri = 1.5 Ωm and spe- cific membrane capacitance of Cm = 0.01 F/m2. This parameter set is the default parameter set used in the present study, see Table I. (C) Representative log-log plot for a PSD when input is homogeneously distributed across the entire neuron model. The low frequency (lf ), intermediate frequency (if ) and high frequency (hf ) regimes are stipulated. The regimes are defined relatively to αall∞ describing the asymptotic value of the exponent of the respective power-law transfer functions (H I, H p or H V ), with both uncorrelated and correlated input ('all' types of input) onto both the soma and the stick. Figure 4: Slopes αI for the PSD transfer function for the soma current for a ball and input currents (H I stick neuron in terms of its dimensionless parameters. Row 1 corresponds both to correlated c ) with any input densities ρs (cid:54)= ρd, and to uncorrelated input to soma only uc,s). Row 2 corresponds to the case of uncorrelated input currents solely onto the dendrite. Row 3 corresponds to uncorrelated input currents with equal density, ρs = ρd, throughout (H I the neuron. The dimensionless parameter B = d2 s /dλ is plotted along the vertical axes, while the dimensionless frequency W is plotted logarithmically along the horizontal axes. In 48 the left column the dimensionless length is L = 0.25, in the middle column L = 1 and the right column L = 4. The horizontal white line express the default value of the parameter B, B = 0.2 (soma diameter ds = 20 µm, stick diameter d = 2 µm, length constant λ = 1 mm), while the vertical white lines correspond to frequencies of 10 Hz, 100 Hz and 1000 Hz, respectively, for the default membrane time constant τm = 30 ms. The thin black line denotes α = 0.5αall∞ = 0.25 and the thicker black line denotes α = 0.9αall∞ = 0.45, with αall∞ = 0.5 denoting the asymptotic value for the case of both uncorrelated and correlated input onto the whole neuron. All plots use the same color scale for α, given by the color bar to the right. Figure 5: Slopes αp for the PSD transfer function for the current-dipole moment (single- neuron EEG contribution) for a ball and stick neuron in terms of its dimensionless param- eters. Row 1 corresponds both to correlated input currents (H p c ) with any input densities ρs (cid:54)= ρd, and to uncorrelated input to soma only (H p uc,s). Row 2 corresponds to the case of input currents solely onto the dendrite. Row 3 corresponds to uncorrelated white-noise input currents with equal density, ρs = ρd, throughout the neuron. The dimensionless pa- rameter B is plotted along the vertical axes, while the dimensionless frequency W is plotted logarithmically along the horizontal axes. In the left column the dimensionless length is L = 0.25, in the middle column L = 1 and the right column L = 4. The horizontal white line express the default value of the parameter B, B = 0.2 (soma diameter ds = 20 µm, stick diameter d = 2 µm, length constant λ = 1 mm), while the vertical white lines corre- spond to frequencies of 10 Hz, 100 Hz and 1000 Hz for the default membrane time constant τm = 30 ms. The thin black line denotes α = 0.5αall∞ = 0.75 and the thicker black line denotes α = 0.9αall∞ = 1.35, with αall∞ = 1.5 denoting the asymptotic value for the case of both uncorrelated and correlated input onto the whole neuron. All plots use the same color scale for α, given by the color bar to the right. Figure 6: Slopes αV for the PSD transfer function for the soma potential for a ball and stick neuron in terms of its dimensionless parameters. Row 1 corresponds to correlated input currents solely onto the dendrite. Row 2 corresponds to input currents solely onto soma, either correlated (H V uc,s). In row 3 uncorrelated input currents are applied homogeneously across the dendrite. Row 5 corresponds to uncorrelated c (ρd = 0)) or uncorrelated (H V 49 input currents with equal density, ρs = ρd, throughout the neuron. Row 6 shows results for correlated input currents with equal density, ρs = ρd, throughout the neuron. The dimen- sionless parameter B is plotted along the vertical axes, while the dimensionless frequency W is plotted logarithmically along the horizontal axes. In the left column the dimensionless length is L = 0.25, in the middle column L = 1 and the right column L = 4. The hor- izontal white line express the default value of the parameter B, B = 0.2 (soma diameter ds = 20 µm, stick diameter d = 2 µm, length constant λ = 1 mm), while the vertical white lines correspond to frequencies of 10 Hz, 100 Hz and 1000 Hz for the default membrane time constant τm = 30 ms. The thin black line denotes α = 0.5αall∞ = 1 and the thicker black line denotes α = 0.9αall∞ = 1.8, with αall∞ = 2 denoting the asymptotic value for the case of both uncorrelated and correlated input onto the whole neuron. All plots use the same color scale for α, given by the color bar to the right. Figure 7: Dependence of PSDs on biophysical parameters for uncorrelated input. PSDs of the soma current (row 1), current-dipole moment (row 2) and soma potential (row 3) for the ball and stick model with uncorrelated white-noise input currents homogeneously dis- tributed throughout the membrane. The input density is two inputs per square micrometer, and the input current is assumed to have a constant (white noise) PSD, s = 1 fA2/Hz. The columns show variation with stick length (first column), specific membrane resistance (sec- ond column), stick diameter (third column) and soma diameter (fourth column) with values shown in the legends below the panels. All other parameters of the ball and stick neuron have default values: stick diameter d = 2 µm, somatic diameter ds = 20 µm, stick length l = 1 mm, specific membrane resistance Rm = 3 Ωm2, inner resistivity Ri = 1.5 Ωm and a specific membrane capacitance Cm = 0.01 F/m2. The values of α printed in the legends describe the powers of the slopes at 1000 Hz. The upper α corresponds to the low value of the parameter varied (green), the middle α corresponds to the default parameter (red), while the lower α corresponds to the high value of the parameter varied (blue). Figure 8: Dependence of PSDs on biophysical parameters for correlated input. PSDs of the soma current (row 1), current-dipole moment (row 2) and soma potential (rows 3 to 5) for the ball and stick model with correlated white-noise input currents homogeneously distributed throughout the stick only (row 1 to 3), the soma only (row 5) or with equal den- 50 sity throughout the soma and the stick (row 4). The input density is two inputs per square micrometer, unless a zero density is indicated on the axis. The input current is assumed to have a constant (white noise) PSD, s = 1 fA2/Hz. The columns show variation with stick length (first column), specific membrane resistance (second column), stick diameter (third column) and soma diameter (fourth column) with values shown in the legends below the panels. All other parameters of the ball and stick neuron have default values: stick diam- eter d = 2 µm, somatic diameter ds = 20 µm, stick length l = 1 mm, specific membrane resistance Rm = 3 Ωm2, inner resistivity Ri = 1.5 Ωm and a specific membrane capacitance Cm = 0.01 F/m2. The values of α printed in the legends describe the powers of the slopes at 1000 Hz. The upper α corresponds to the low value of the parameter varied (green), the middle α corresponds to the default parameter (red), while the lower α corresponds to the high value of the parameter varied (blue). 51 FIG. 1. FIG. 2. 52 FIG. 3. FIG. 4. 53 FIG. 5. 54 FIG. 6. 55 FIG. 7. 56 FIG. 8. 57 TABLES 58 TABLE I. List of symbols in alphabetical order. In the column labeled Default (Unit) the default value of the parameter is given. If a default value is not listed, the unit is given in parenthesis. The specific electrical properties of the soma membrane and stick membrane are here assumed to be equal. Symbol B = d2 s /dλ Cm cm = πdCm d ds f Default (Unit) Description 0.2 relative soma to infinite-stick conductance 0.01 pF/µm2 specific membrane capacitance 0.0628 pF/µm membrane capacitance per unit length of cable 2 µm stick diameter 20 µm soma diameter (Hz) frequency Gm = 1/Rm 0.333 pS/µm2 specific membrane conductance gm = 1/rm = πd/Rm 2.09 pS/µm membrane conductivity per unit length of cable G∞ = 1/riλ L = l/λ √ l q = Ri 2.09 nS infinite-stick conductance 1 electrotonic length 1 mm stick length 1 + jW = Y∞/G∞ (1) frequency dependence of the infinite-stick admittance 1.5 MΩµm inner resistivity ri = 4Ri/πd2 0.477 MΩ/µm inner resistance per unit length of cable s T = t/τm W = ωτ X = x/λ Yin Y = Ys/Y∞ = qB Ys = πd2 s Gmq2 Y∞ = qG∞ √ gmri λ = 1/ 1 fA2/Hz power spectral density of input current (1) dimensionless time (1) dimensionless frequency (1) dimensionless position (S) input admittance (1) relative soma to infinitestick admittance (S) soma admittance (S) infinite -- stick admittance 1 mm neuron length constant ρ = ρs/(ρs + ρd) 0.5 relative input density ρd ρs τm = RmCm ω = 2πf 2/µm2 dendritic current-input number density 2/µm2 somatic current-input number density 30 ms membrane time constant 59 rad/s angular frequency TABLE II. PSD amplitudes and high-frequency power laws. The amplitudes A and the asymptotic powers α∞ for the different PSDs. The right column shows the amplitude A(cid:48) for the asymptotic PSDs expressed in terms of biophysical parameters. When W approaches infinity, the asymptotic value of all PSD transfer functions except for H V c two asymptotic values of non-standard form: H V c → AV H V not correspond to the given formula for A(cid:48), but rather to A(cid:48)ρ2 (left) and A(cid:48)B−2 (right). there are mW −2 for ρs (cid:54)= 0 (left) and mB−2W −3 for ρs = 0 (right). (∗): The values of the right column does c B−2W −3 = ρ2 c ρ2W −2 = ρ2 c → AV dR2 s R2 is given by H → AW −α∞. For H V c A(cid:48) = A × (f /W )α (ρd − ρs)2πd3/8RiCm ρdπ1/2d3/2/4R1/2 i C1/2 s Ri (ρd − ρs)2d4/64R2 ρdd5/2/32π1/2R3/2 d3ρs/8Cmd2 m m md2 i C2 m i C3/2 s R2 i md4 dd3/32π3C3 sR1/2 i C5/2 md2 s m ρs/4π3C2 d4ρs/64πC2 ρdd3/2/16π7/2d4 ρ2 s /4π2C2 m; ρ2 s Ri (∗) Case Amplitude (A) α∞ (W −α∞) H I c (ρd − ρs)2(πdλ)2 1 √ uc,d ρdπdλ/ H I 2 1/2 H I uc,s ρsπdλ/B 1 (ρd − ρs)2π2d2λ4 2 H p c √ uc,d ρdπdλ3/ uc,s ρsπdλ3/B 2 H p H p (ρd + ρs)2R2 m √ H V H V c uc,d ρdR2 uc,s ρsR2 H V m/πdλB m/ 2πdλB2 5/2 3/2 2 2; 3 2 60
1003.4253
2
1003
2010-04-13T13:20:10
Mobility induces global synchronization of oscillators in periodic extended systems
[ "physics.bio-ph", "nlin.CD" ]
We study synchronization of locally coupled noisy phase oscillators which move diffusively in a one-dimensional ring. Together with the disordered and the globally synchronized states, the system also exhibits several wave-like states which display local order. We use a statistical description valid for a large number of oscillators to show that for any finite system there is a critical spatial diffusion above which all wave-like solutions become unstable. Through Langevin simulations, we show that the transition to global synchronization is mediated by the relative size of attractor basins associated to wave-like states. Spatial diffusion disrupts these states and paves the way for the system to attain global synchronization.
physics.bio-ph
physics
Mobility induces global synchronization of oscillators in periodic extended systems Fernando Peruani,1, 2 Ernesto M. Nicola,2, 3 and Luis G. Morelli2, 4, 5 1CEA-Service de Physique de l'Etat Condens´e, Centre d'Etudes de Saclay, 91191 Gif-sur-Yvette, France 2Max Planck Institute for the Physics of Complex Systems, Nothnitzer Str. 38, 01187 Dresden, Germany 3IFISC, Institute for Cross-Disciplinary Physics and Complex Systems (CSIC-UIB), Campus Universitat Illes Balears, E-07122 Palma de Mallorca, Spain 4Departamento de F´ısica, FCEyN, UBA, Ciudad Universitaria, 1428 Buenos Aires, Argentina 5Max Planck Institute of Molecular Cell Biology and Genetics, Pfotenhauerstr. 108, 01307 Dresden, Germany (Dated: June 26, 2018) Abstract We study synchronization of locally coupled noisy phase oscillators that move diffusively in a one-dimensional ring. Together with the disordered and the globally synchronized states, the system also exhibits wave-like states displaying local order. We use a statistical description valid for a large number of oscillators to show that for any finite system there is a critical mobility above which all wave-like solutions become unstable. Through Langevin simulations, we show that the transition to global synchronization is mediated by a shift in the relative size of attractor basins associated to wave-like states. Mobility disrupts these states and paves the way for the system to attain global synchronization. PACS numbers: 05.45.Xt,87.18.Gh 0 1 0 2 r p A 3 1 ] h p - o i b . s c i s y h p [ 2 v 3 5 2 4 . 3 0 0 1 : v i X r a 1 Synchronization of oscillators is a widespread phenomenon in nature [1 -- 3]. In biology, synchronization can occur at scales that range from groups of single cells to ensembles of complex organisms [4]. When oscillators hold fixed positions in space and the interaction that drives synchronization is short ranged, spatial and temporal patterns can self-organize. Such is the case in cardiac tissue, where cells generate spiral patterns that shape the heartbeats [5]. Also in central pattern generators, the oscillating neural network self-organizes to produce coordinated movements of the body [6]. A different situation arises when the oscillators are not fixed in space but are able to move around. The problem of synchronization of moving oscillators has many applications in the domain of chemistry [7], biology [8], and technology [9]. Small porous particles loaded with the catalyst of the Belousov-Zhabotinsky reaction behave as individual chemical oscillators, undergoing a density-dependent synchronization transition as the stirring rate is increased [7]. The same particles support wave propagation in the form of dynamic target and spiral patterns when the particles are not moving [10]. This phenomenon illustrates a wider scenario: mobility and mixing remove local defects and patterns, enabling global order. This effect has far reaching consequences in finite systems. For example, in ecosystems of competing populations with cyclic interactions, biodiversity can be sustained if dispersal is local, but it is lost when dispersal occurs over large length scales [11]. The dynamics of such cyclic competition was described by a complex Ginzburg-Landau equation near a Hopf bifurcation, displaying complex oscillatory patterns indicative of biodiversity for low mobility, while in the case of high mobility diversity is wiped out [12, 13]. In this paper we study the effects of mobility -- spatial diffusion -- on the macroscopic collective dynamics of locally coupled, moving phase oscillators subjected to noise, in a one- dimensional ring. When oscillators are fixed in space, these systems can exhibit a series of steady states where local order is present [14 -- 16]. Such states have been called m-twist solutions [16], Fig. 1. Here we show that mobility can destabilize all m-twist solutions, enhancing the stability of the synchronized solution. We find that in finite systems there is a critical mobility above which either the synchronized or the disordered state is stable. 2 I. DIFFUSING PHASE OSCILLATORS We consider an ensemble of N identical phase oscillators that diffuse on a ring of perime- ter L. Oscillators are coupled to other oscillators in their local neighborhood, within an interaction range r. The dynamics of phase and position is described by 1 θi(t) = ω − γ  xi(t) = √2D ξx,i(t) , ni Xxi−xj<r sin(θi − θj)  + √2C ξθ,i(t) (1) (2) where i = 1, . . . , N is the oscillator label, θi(t) and xi(t) are the phase and position of the i-th oscillator at time t, ω is the autonomous frequency, and γ is the coupling strength -- whose inverse characterizes the typical relaxation time of the interaction. Each oscillator interacts with its ni neighbors in the range r through the coupling function in brackets, which defines an attractive interaction towards the local average of the phase. In steady state the spatial density is uniform and the number of neighbors is on average constant, n = ni = N r/L. With this definition of the coupling the thermodynamic limit is well defined, and the system reduces to the noisy Kuramoto model for r = L/2 [17]. The fluctuation terms ξθ,i and ξx,i represent two uncorrelated Gaussian noises such that hξθ,i(t)i = hξx,i(t)i = 0, and hξθ,i(t)ξθ,j(t′)i = hξx,i(t)ξx,j(t′)i = δi,jδ(t − t′). The strength of angular fluctuations is determined by the angular diffusion coefficient C, while the spatial diffusion coefficient D 2 π θ 0 0 (a) disorder (b) sync (c) 1-twist (d) 2-twist x L 0 x L 0 x L 0 x L FIG. 1: The system, Eqs. (1)-(2), exhibits different kinds of states: (a) disorder, (b) partial synchronization, and twisted states: e.g. (c) 1-twist and (d) 2-twist. Each dot (x, θ) represents an oscillator. Parameters in arbitrary units: N = 1000, L = 2π, r = L/100, ω = 0, γ = 0.1 and √2D = 0.01. In (a) √2C = 0.35 and in (b)-(d) √2C = 0.10. States (b)-(d) coexist. The snapshots were taken after 15000 time units, starting from random initial conditions. 3 (a) D / γ = 0.03 D / γ = 0.39 (b) C / γ = 0.11 C / γ = 0.25 (c) C / γ = 0.11 1.0 Z 0.5 1.0 Z 0.5 0.0 0 0.0 0 0.8 C / γ D / γ 0.4 100 10 1 0.1 ) Z ( P 100 10 1 0.1 ) Z ( P D / γ = 0.03 D / γ = 0.39 0 0.5 Z 1 FIG. 2: Twisted states affect the ensemble average of the global order parameter, hZi. (a) hZi vs. scaled angular diffusion C/γ for D/γ = 0.03 (green squares) and D/γ = 0.39 (blue dots). For small D/γ, hZi does not reach one, even for C = 0. Vertical dotted lines correspond to values of C/γ in (b). (b) hZi vs. scaled mobility D/γ, for C/γ = 0.11 (orange triangles) and C/γ = 0.25 (red diamonds). Notice that as D is reduced, hZi decays. Vertical dotted lines correspond to values of C/γ in (a). (c) The histogram of Z splits into two different peaks at low and high values of Z for low values of D/γ, but displays only one peak at high Z for high D/γ. Other parameters are N = 1000, L = 2π, r/L = 0.01, ω = 0 and γ = 0.1. determines the mobility of oscillators. The oscillators are point-like particles that can diffuse freely, i.e. their movement is not affected by the presence of other oscillators. Both the phase θi(t) and position xi(t) are periodic variables such that 0 ≤ θi(t) ≤ 2π and 0 ≤ xi(t) ≤ L. II. GLOBAL ORDER PARAMETER The system described by Eqs. (1)-(2) displays a range of states illustrated in Fig. 1. We can characterize global order by the order parameter Z, defined as the absolute value of the population average of the complex unit vectors of the oscillator phases Z(t) = N −1Pj eiθj (t) [18]. Mobility introduces a dramatic change in the ensemble average of the order parameter hZi, where h. . .i denotes an average over different initial conditions and realizations of the noise, Fig. 2. When mobility is large enough, the system displays global synchronization as C → 0, blue dots in Fig. 2(a). However, when mobility is reduced, global order is compromised, as indicated by the green squares in Fig. 2(a). Increasing mo- bility D results in an increase of the ensemble average hZi, Fig. 2(b). The cause for this 4 behavior can be traced back to the existence of twisted states, which display local order but have themselves a vanishing global order parameter, Fig. 1(c,d). Due to the presence of these states, hZi results from the average of a bimodal distribution P (Z), with a peak related to the twisted states and the other to global synchronization. Above a critical value of the mobility, P (Z) becomes unimodal, Fig. 2(c). Thus, although a global parameter is not suited to capture the complexity of a system with local interactions, its statistics reflect the existence of twisted states. III. STATISTICAL DESCRIPTION The role of twisted states can be studied using a statistical description that is valid when the number of oscillators is large. Given that the oscillators have identical autonomous frequencies ω, it is convenient to make the transformation θ → θ− ωt to a rotating reference frame. We coarse grain the microscopic model and describe the system in terms of ρ(x, θ, t), the density of oscillators at position x with phase θ, which obeys the Fokker-Planck equation ∂tρ(x, θ, t) = D∂xxρ(x, θ, t) + C∂θθρ(x, θ, t) (3) + γ n(x) ∂θ(cid:20)Z L dx′Z 2π 0 0 dθ′g(x − x′) sin(θ − θ′)ρ(x′, θ′, t)ρ(x, θ, t)(cid:21) , where g(x−x′) is a kernel accounting for the range and relative strength of local interactions, while n(x) = Z L 0 dx′Z 2π 0 dθ′g(x − x′)ρ(x′, θ′, t) denotes the effective number of oscillators in this range. In this paper we choose g(x−x′) = 1 for x − x′ < r and g(x − x′) = 0 otherwise, as in Eq. (1). The derivation of Eq. (3) relies on the assumption that ρ2(x, θ, t; x′, θ′, t) = ρ(x, θ, t)ρ(x′, θ′, t) [19]. Since the movement of the oscillators is purely diffusive, see Eq. (2), the spatial density 0 dθρ(x, θ, t) = N/L ≡ ρ0, and n(x) = 2rρ0. For small N, fluctuations in the spatial density can induce the formation of gaps in which the of oscillators is uniform in steady state, R 2π nearest oscillator is beyond the range of interaction. In this paper we consider large densities such that the lifetime of these gaps is much shorter than other typical time-scales. 5 A. Local order parameter The statistical description (3) can be cast in a more transparent form introducing a local mean field. Local order can be characterized by a local order parameter [14, 15] R(x, t)eiψ(x,t) =Z L dx′Z 2π 0 0 dθ′ g(x − x′) n(x) eiθ′ ρ(x′, θ′, t) , (4) where R(x, t) is a measure of local order and ψ(x, t) is the local average of the phase. Eq. (3) can be expressed in terms of this local order parameter as ∂tρ(x, θ, t) = D∂xxρ(x, θ, t) + C∂θθρ(x, θ, t) (5) + γR(x, t)∂θ (sin (θ − ψ(x, t)) ρ(x, θ, t)) , reflecting the fact that ψ(x, t) acts as a local mean field and R(x, t) is a local modulation to the coupling strength. IV. TRANSITION FROM DISORDER TO LOCAL ORDER Eq. (3) has a trivial steady state ρ(x, θ, t) = ρ0/2π ≡ ρd which corresponds to the disordered state of the system. We study the stability of ρd by inserting ρ(x, θ, t) = ρd + ǫf (x, t) cos(ℓθ) in Eq. (3) and keeping terms of order O(ǫ) [20]. Linear stability analysis reveals that the disordered solution ρd becomes unstable for ℓ = 1 when C < C ∗, with C ∗ = γ/2 . (6) This threshold is independent of ρ0 and D, and determines the value of C below which local order sets in. The critical C ∗ given by Eq. (6) coincides with the existence [17] and stability [20] threshold displayed by globally coupled noisy oscillators. V. LOCAL ORDER SOLUTIONS Once local order has set in, the system also supports twisted solutions. We specifically look for steady state solutions to Eq. (5) of the form ρs(x, θ) = f (θ − ψ(x)). Such wave- like solutions describe densities in which the angular distribution has the same shape, but is centered at position dependent phases ψ(x). Setting ∂tρ = 0 we obtain an ordinary 6 differential equation for f (cid:16)Dθ + Dx (ψ′(x))2(cid:17) f ′(ϕ) + γR (sin(ϕ) − Dxψ′′(x)) f (ϕ) = C(x) , (7) that we can solve together with periodic boundary conditions in phase and space to determine the arbitrary constant C(x) and phases ψ(x). Periodicity of the phase is consistent with solutions that fulfill ψ′′(x) = 0 and C(x) = 0, and periodicity of space sets the wave numbers k = 2πm/L with m integer. We obtain the m-twist steady state solutions ρs(x, θ) = N exp(cid:20) γR cos(θ − kx)(cid:21) , Deff where N is a normalization constant such thatR 2π 0 dθR L We have introduced the effective diffusion coefficient (8) 0 dxρs(x, θ) = N = Lρ0, see Fig. 3(a). Deff (k) = C + k2D , (9) which is a combination of angular diffusion C, and mobility D scaled by the square of the wave number k. Effective diffusion competes with the local coupling γR and controls the width of the angular distribution, which has a mean hϕi = kx and variance σ2 = 1 − I1(γR/Deff )/I0(γR/Deff ) = 1 − R/sinc(kr), where sinc(x) = sin(x)/x and In(z) = Z 2π 0 dθ cosn θ exp (z cos θ) is the modified n-order Bessel function of the first kind. The functional form of the steady state angular distribution Eq. (8) is known as the circular normal distribution [21]. A. Effective diffusion controls the local order parameter The m-twist solutions described by Eq. (8) exist if and only if the self-consistency condi- tion obtained by inserting Eq. (8) into Eq. (4) is fulfilled I1(γRm/Deff ) I0(γRm/Deff ) Rm = sinc(kr) , (10) see Fig. 3(b,c). A trivial solution to Eq. (10) is Rm = 0. Apart from this, an expansion of Eq. (10) for Rm ≪ 1 reveals that non-vanishing solutions Rm ≈ p8Dcγ−2 (Dc − Deff )1/2 7 (11) (a) D / γ = C / γ = 0.05 (b) D / γ = 0.05 (c) C / γ = 0.05 1.5 s ρ 1.0 0.5 0.0 0 m = 0 m = 1 m = 2 1.0 0.8 m R 0.6 m = 0 m = 1 m = 2 m R 0.4 0.2 0.0 0 θ 2π C / γ 0.5 1.0 0.8 0.6 0.4 0.2 0.0 0 m = 0 m = 1 m = 2 D / γ 0.5 FIG. 3: Twisted solutions and local order. (a) Angular distribution of steady state twisted solu- tions, Eq. (8). (b) The local order parameter Rm decreases with increasing angular fluctuations for all solutions, Eq. (10). (c) The local order parameter of twisted solutions also decreases with mobility, but is not affected for global order, Eq. (10). The range of interaction is r/L = 0.01. exist for Deff ≤ (γ/2) sinc(kr) ≡ Dc. Again, we see that Deff controls the growth of the local order parameter Rm characterizing the emergence of twisted solutions, through changes in phase fluctuations C and mobility D, Fig. 3(b,c). B. Existence of twisted solutions We can unfold the effects of spatial and angular fluctuations by writing Deff in terms of its components D and C. Setting Rm = 0 in Eq. (11) we get C + (2πm/L)2 D = (γ/2) sinc (2πmr/L) , (12) where we have expressed k in terms of L to stress system size dependence. For each value of m and range of interaction r, the surface in parameter space defined by Eq. (12) encloses the region where the m-twist solution exists. There are two ways in which fluctuations can destroy twisted solutions, by increasing either angular fluctuations or mobility. Mobility gets amplified for higher order modes as m2, causing twisted solutions to disappear sequentially as mobility increases, Fig. 4. For m = 0, Eq. (12) reduces to C = C ∗ = γ/2 and corresponds to global synchronization. Existence of global synchronization in steady state is not affected by mobility, as indicated by the dotted red line in Fig. 4(a,c). However, existence of twisted solutions in finite systems is controlled by angular diffusion C, mobility D, and range of interaction r as indicated by Eq. (12), Fig. 4. 8 (a) D = 0 disorder 0.5 γ / C (b) C / γ = 0.1 (c) r / L = 0.01 0.5 γ / D m=0 m=1 m=2 m=4 0.5 γ / D d i s o r d e r 0.0 0 r / L 0.0 0.5 0 r / L 0.0 0.5 0 C / γ 0.5 FIG. 4: Mobility controls existence of m-twist solutions. Three representative cuts of the phase diagram are shown for m = 0, 1, 2, 4. (a) D = 0, (b) C/γ = 0.1 and (c) r/L = 0.01. The dotted red line indicates the onset of local order, which is independent of D. Below the solid green line, the dashed blue line and the dot-dashed yellow line, the 1-twist, 2-twist, and 4-twist solution exists, respectively, see Eq. (12). An increasing number of co-existing twisted solutions can be found with decreasing mobility D and decreasing coupling range r. As L → ∞, the critical value of the spatial diffusion coefficient diverges for all m. There- fore, in the infinite system size limit all twisted solutions coexist with global synchronization for any finite D. These result is in agreement with [22], and indicates that identical noisy phase oscillators cannot exhibit a global synchronized state in 1D in this limit. C. Stability of twisted solutions and states While existence and stability thresholds coincide for the global order solution, Eqs. (6) and (12), this is not the case for m-twist solutions in finite systems, Fig. 5. We address the stability of the 1-twist solution by performing a numerical study of Eq. (3), using a finite difference scheme. To estimate the stability boundary, we continue a stable twist solution until it becomes unstable against small perturbations. We find that the instability is of modulational type. The m-twist solutions become stable only after the corresponding local order parameter Rm becomes larger than a certain value, i.e. twisted solutions become stable with a finite amplitude, Fig. 5. For vanishing spatial and angular diffusion C = D = 0, we encounter the system studied by Wiley et al. [16], see purple open circle in Fig. 5(a). The numerical solution seems to approach this point as C/γ → 0, but the numerics become 9 (a) D = 0 disorder (b) C / γ = 0.1 (c) r / L = 0.01 0.5 γ / C 0.5 γ / D 0.5 γ / D d i s o r d e r 0.0 0 r / L 0.0 0.5 0 r / L 0.0 0.5 0 C / γ 0.5 FIG. 5: The domains of existence and stability of m-twist solutions do not coincide. The 1-twist solution exists below the solid green line, Eq. (12), but it becomes stable below the dashed purple line, determined numerically. The purple dots show the stability of the 1-twist state of the Langevin Eqs. (1)-(2). The open circle in (a) corresponds to the limit case D = C = 0 studied in [16]. lengthy in this limit. We next compare the continuum Fokker-Planck description (3) with the discrete system, by means of Langevin simulations of Eqs. (1)-(2). To measure the stability of the 1-Twist state in Langevin simulations, we prepare the system in the 1-Twist state by randomly positioning the oscillators in space, and setting their phase to θ(x) = 2πx/L. We first let the system relax for 1000 units of time, so that the stationary shape of the angular distribution is reached. Then, we compute the ensemble average of the twist hmi, where h...i denotes an average over 100 realizations of the numerical simulation, which we performed for different sets of parameter values C, D, and r/L. To evaluate the stability of the 1-Twist solution we set a threshold on the average twist: when hmi falls below 0.05, we take the 1-twist state to be unstable. Simulations were performed with N = 1000 and L = 2π. Other parameters are indicated in Fig. 5. The stability threshold in Fig. 5(a) was found by varying r for various fixed values of C, in Fig. 5(b) by varying D for fixed values of r, and in Fig. 5(c) by varying C for fixed values of D. We find a good agreement between the stability of the 1-twist state of the discrete system, and the stability of the 1-twist solution found by numerical integration of the Fokker-Planck equation. 10 VI. ATTRACTION BASINS Twisted solutions co-exist with global order and among themselves, Fig. 4. As mobility is increased from low values, twisted solutions become unstable one by one, e.g. Fig. 4(c) and Fig. 5(c), and global order is enhanced resulting in an increasing value of the ensemble average of the global order parameter as displayed in Fig. 2(b). Angular fluctuations can also destabilize twisted solutions, Fig. 4. However, while increasing angular fluctuations can decrease the number of twisted solutions competing with global order, the width of the angular distribution corresponding to global order also grows as a consequence of increasing angular fluctuations. The interplay between these two competing effects may be the cause for the non-monotonic behavior observed in the ensemble average of the global order parameter as angular fluctuations C increase, see green squares in Fig. 2(a). For these reasons, it is interesting to explore the size of the attraction basins of the different states that the discrete system exhibits. The fraction of realizations B(m) in which the system ends up in a particular state after a short time, starting from random initial conditions, is a measure of the size of the basins of attraction of the state. Coexistence of twisted states with up to m = 5 is shown in Fig. 6(a) for low mobility and phase fluctuations, corresponding to the bottom left corner of Fig. 4(c). As C is decreased, the attraction basin corresponding to global order shrinks, while the attraction basins of the m-twist states expand, Fig. 6(b). Decreasing C does not necessarily yield global order at the ensemble level because the global ordered state shares the phase space with the m-twist solutions: the observed value of hZi results from the competition between an increase in local order for m = 0 and a decrease of B(0), see Fig. 2(a). An increase in the mobility D leads to a contraction of the attraction basin of the m-twist states, in favor of global order, Fig. 6(c). When all m-twist solutions are unstable, B(0) = 1 and the global order state is the only attractor below C ∗, see also Fig. 2(b). VII. DISCUSSION We have investigated the effects of mobility in a generic 1D model of locally coupled moving phase oscillators, and showed that oscillator mobility dramatically affects the collec- tive behavior of finite systems. More specifically, our results show that in low dimensional 11 (a) D / γ = C / γ = 5 x 10-4 (b) D / γ = 0.05 (c) C / γ = 0.05 0.3 0.2 ) m ( B 0.1 0 -6 -4 -2 0 2 4 6 m 1 ) m ( B 0 0 m=0 m=1 C / γ 1 ) m ( B m=0 m=1 m=2 0 0.5 0 D / γ 0.5 FIG. 6: The size of attraction basins of m-twist solutions depends on angular fluctuations and mobility. (a) Probability B(m) that starting from a random initial condition the system ends up in a m-twist solution. B(0) and B(1) as function of (b) C/γ and (c) D/γ. Parameters as in Fig. 2. systems global synchronization is compromised by the presence of multiple m-twist states exhibiting local order. At the onset of local order, the system can fall into the global syn- chronized state. However, the coexistence of local order m-twist states implies that the attraction basin of the global synchronization state is reduced. Strong mobility of the oscil- lators destabilizes these m-twist states, and thus promotes global synchronization. In this paper we have considered a high density limit such that the connectivity of the system is never interrupted by gaps. In the dilute limit, gaps in the connectivity play a crucial role in the synchronization dynamics. This problem was studied in the context of moving neighborhood networks, and under the assumption of a fast exchange of neighbors, a mean-field condition for the existence and stability of the global synchronization state has been derived [23]. According to this study, whenever the global synchronized state is stable the system reaches global synchronization, regardless of its spatial dimensionality. In other words, the study overlooks the possibility of coexistence of multiple solutions. Our findings reveal a different role for mobility, unrelated to the existence and stability of the global synchronization state: mobility disrupts all these multiple solutions except for the global synchronized state. Extensions of the current study towards dilute systems will be the subject of further investigations. Two-dimensional systems display a similar phenomenology, though the competing local order states can now take other forms, e.g. vortexes [24]. It has been recently reported that chaotic oscillators moving in a two-dimensional space can synchronize provided that 12 spatial dynamics is fast enough [25]. A related albeit different scenario occurs with chaotic advection mixing in two dimensional systems, where synchronization of excitable media is enhanced by strong mixing [26, 27]. These results indicate that mobility may also enhance synchronization in two-dimensional systems. We speculate that global synchronization may be achieved by destabilizing local deffects, as we show here for 1D systems. Further work is intended to clarify these issues. The theoretical framework introduced here may provide insight into other related prob- lems, as when movement is coupled to the oscillator phases. In this case synchronization can be interpreted as collective motion [28]. As a result of this coupling, strong spatial fluc- tuations and clustering effects dominate the system dynamics [29], and global order prevails even in the thermodynamic limit [30]. Finally, a compelling biological application of our framework may be found in the verte- brate segmentation clock, where global coupling is a good effective description of the system because of the high mobility of cells [31]: by precluding the appearance of local defects, mo- bility promotes global synchronization. Moreover, it has been recently shown that mobility decreases the relaxation times to achieve synchronization in a model of the segmentation clock that allows for flipping between neighboring cells [32]. However, this system also hosts spatial patterns [33], and mobility is not accounted for in current distributed models. It will be interesting to see how mobility affects the synchrony recovery times and pattern reorganization after perturbation in such models. Acknowledgements We thank S. Ares, H. Chat´e, M. Mat´ıas, A. C. Oates and D. Zanette for valuable com- ments and discussions. FP acknowledges financial support from the French ANR projects Morphoscale and Panurge. LGM acknowledges support from CONICET, ANPCyT PICT 876, and ERC grant 207634 SegClockDyn. [1] S. C. Manrubia, A. S. Mikhailov, and D. H. Zanette, Emergence of dynamical order: synchro- nization phenomena in complex systems (World Scientific, 2004), 1st ed. 13 [2] A. Pikovsky, M. Rosenblum, and J. Kurths, Synchronization: A Universal Concept in Non- linear Sciences (Cambridge University Press, 2008). [3] S. H. Strogatz, Sync: How Order Emerges From Chaos In the Universe, Nature, and Daily Life (Hyperion, 2004), 1st ed. [4] A. T. Winfree, The Geometry of Biological Time (Springer, 2001), 2nd ed. [5] L. Glass and M. C. Mackey, From Clocks to Chaos: The Rhythms of Life (Princeton University Press, 1988). [6] A. H. Cohen, P. J. Holmes, and R. H. Rand, J. Math. Biol. 13, 345 (1982). [7] A. F. Taylor, M. R. Tinsley, F. Wang, Z. Huang, and K. Showalter, Science 323, 614 (2009). [8] S. D. Monte, F. d'Ovidio, S. Danø, and P. G. Sørensen, Proc. Natl. Acad. Sci. USA 104, 18377 (2007). [9] A. Buscarino, L. Fortuna, M. Frasca, and A. Rizzo, Chaos 16, 015116 (2006). [10] M. R. Tinsley, A. F. Taylor, Z. Huang, and K. Showalter, Phys. Rev. Lett. 102, 158301 (2009). [11] B. Kerr, M. A. Riley, M. W. Feldman, and B. J. M. Bohannan, Nature (London) 418, 171 (2002). [12] T. Reichenbach, M. Mobilia, and E. Frey, Nature (London) 448, 1046 (2007). [13] T. Reichenbach, M. Mobilia, and E. Frey, Phys. Rev. Lett. 99, 238105 (2007). [14] Y. Shiogai and Y. Kuramoto, Prog. Theor. Phys. Suppl. 150, 435 (2003). [15] Y. Kuramoto, S. Shima, D. Battogtokh, and Y. Shiogai, Prog. Theor. Phys. Suppl. 161, 127 (2006). [16] D. A. Wiley, S. H. Strogatz, and M. Girvan, Chaos 16, 015103 (2006). [17] H. Sakaguchi, Prog. Theor. Phys. 79, 39 (1988). [18] Y. Kuramoto, Chemical Oscillations, Waves, and Turbulence. (Springer-Verlag, Berlin, 1984). [19] E. J. Hildebrand, M. A. Buice, and C. C. Chow, Phys. Rev. Lett. 98, 054101 (2007). [20] S. H. Strogatz and R. E. Mirollo, J. Stat. Phys. 63, 613 (1991). [21] R. von Mises, Phys. Z 19, 490 (1918). [22] N. D. Mermin and H. Wagner, Phys. Rev. Lett. 17, 1133 (1966). [23] J. D. Skufca and E. M. Bollt, Math. Biosci. Eng. 1, 347 (2004). [24] J. M. Kosterlitz and D. J. Thouless, J. Phys. C: Solid State Phys. 6, 1181 (1973). [25] M. Frasca, A. Buscarino, A. Rizzo, L. Fortuna, and S. Boccaletti, Phys. Rev. Lett. 100, 044102 (2008). 14 [26] C. R. Nugent, W. M. Quarles, and T. H. Solomon, Phys. Rev. Lett. 93, 218301 (2004). [27] C. Zhou and J. Kurths, New J. Phys. 7, 18 (2005). [28] F. Peruani, A. Deutsch, and M. Bar, EPJ-ST 157, 111 (2008). [29] D. H. Zanette and A. S. Mikhailov, Physica D 197, 203 (2004). [30] J. Toner, Y. Tu, and S. Ramaswamy, Ann. Phys. 318, 170 (2005). [31] I. H. Riedel-Kruse, C. Muller, and A. C. Oates, Science 317, 1911 (2007). [32] K. Uriu, Y. Morishita, and Y. Iwasa, Proc. Natl. Acad. Sci. USA 107, 4979 (2010). [33] L. G. Morelli, S. Ares, L. Herrgen, C. Schroter, F. Julicher, and A. C. Oates, HFSP J. 3, 55 (2009). 15
1905.04358
1
1905
2019-05-03T22:26:33
Three-Dimensional Microscopy by Milling with Ultraviolet Excitation
[ "physics.bio-ph", "physics.optics" ]
Analysis of three-dimensional biological samples is critical to understanding tissue function and the mechanisms of disease. Many chronic conditions, like neurodegenerative diseases and cancers, correlate with complex tissue changes that are difficult to explore using two-dimensional histology. While three-dimensional techniques such as confocal and light-sheet microscopy are well-established, they are time consuming, require expensive instrumentation, and are limited to small tissue volumes. Three-dimensional microscopy is therefore impractical in clinical settings and often limited to core facilities at major research institutions. There would be a tremendous benefit to providing clinicians and researchers with the ability to routinely image large three-dimensional tissue volumes at cellular resolution. In this paper, we propose an imaging methodology that enables fast and inexpensive three-dimensional imaging that can be readily integrated into current histology pipelines. This method relies on block-face imaging of paraffin-embedded samples using deep-ultraviolet excitation. The imaged surface is then ablated to reveal the next tissue section for imaging. The final image stack is then aligned and reconstructed to provide tissue models that exceed the depth and resolution achievable with modern three-dimensional imaging systems.
physics.bio-ph
physics
Three-Dimensional Microscopy by Milling with Ultraviolet Excitation Jiaming Guo1, Camille Artur2, Jason L. Eriksen3, and David Mayerich1,2* Preprint for Review, 2019 Abstract Analysis of three-dimensional biological samples is critical to understanding tissue function and the mechanisms of disease. Many chronic conditions, like neurodegenerative diseases and cancers, correlate with complex tissue changes that are difficult to explore using two-dimensional histology. While three-dimensional techniques such as confocal and light-sheet microscopy are well-established, they are time consuming, require expensive instrumentation, and are limited to small tissue volumes. Three-dimensional microscopy is therefore impractical in clinical settings and often limited to core facilities at major research institutions. There would be a tremendous benefit to providing clinicians and researchers with the ability to routinely image large three-dimensional tissue volumes at cellular resolution. In this paper, we propose an imaging methodology that enables fast and inexpensive three-dimensional imaging that can be readily integrated into current histology pipelines. This method relies on block-face imaging of paraffin-embedded samples using deep-ultraviolet excitation. The imaged surface is then ablated to reveal the next tissue section for imaging. The final image stack is then aligned and reconstructed to provide tissue models that exceed the depth and resolution achievable with modern three-dimensional imaging systems. Keywords Three-Dimensional Microscopy -- Ultraviolet Excitation -- Microtome 1University of Houston, Department of Electrical and Computer Engineering 2University of Houston, NSF BRAIN Center 3University of Houston, Department of Pharmacology *Corresponding author: [email protected] 9 1 0 2 y a M 3 ] h p - o i b . s c i s y h p [ 1 v 8 5 3 4 0 . 5 0 9 1 : v i X r a Current biomedical research and clinical diagnoses rely heavily on histological tissue sectioning to visualize phenotype and phe- notypic changes. However, two-dimensional sectioning provides a very limited representation of three-dimensional structure. These limitations are particularly difficult to reconcile for complex three- dimensional structures, such as neural and microvascular networks. Cancer and neurodegenerative diseases affect the surrounding tissue phenotype, introducing complex changes in tissue structure. For example, tumor-induced vascular endothelial growth factor (VEGF) stimulates angiogenesis, providing necessary substrates for tumor cell growth and spreading [1]. Neurodegenerative disorders such as Alzheimer's disease (AD) induce currently irreversible damage in vasculature and neural connectivity [2]. These changes are ex- tremely difficult to quantify with traditional histological sections, which are only 4 to 6µm thick. Three-dimensional (3D) microscopy, such as confocal, multi- photon, and light-sheet microscopy are common methods of 3D imaging. However, these techniques are extremely time-consuming, limited to small (≈1 mm thick) samples [3], and often expensive. Several attempts have been made to alleviate constraints on sample thickness, including optical modifications [4] and tissue clearing [5]. In addition to trade-offs in resolution, imaging depth is still limited by objective working distance and speed is limited by pho- tobleaching. Physical sectioning overcomes thickness constraints, with advanced developments including vaporizing imaged layers [6] and physical cutting [7, 8]. However, these methods are still either time-consuming or cost prohibitive, making them impractical for most research and clinical settings. In this paper, we propose a new methodology that offers (1) fast acquisition speeds comparable to 2D histology, (2) unlimited sam- ple thickness, (3) resolution that exceeds the diffraction limit along the axial direction, and (4) simple and low-cost construction. This imaging system is based on recent innovations in deep-ultraviolet histology [9] that allow block-face imaging of fresh samples to achieve histology-like 2D images. We describe the design of our prototype imaging system (Figure 1) and provide a demonstration and characterization of its performance for imaging high-resolution vascular and cellular components in formalin-fixed and paraffin- embedded (FFPE) samples. Theoretical Approach Deep Ultraviolet Optical Sectioning The proposed instrument is inspired by a new imaging technology known as microscopy with ultraviolet surface excitation (MUSE) [9] that allows slide-free histology on intact tissue using fluorescent dyes. The main advantage of UV excitation is that light penetration under direct illumination is limited to the sample surface (10 µm or less) [10]. Many common fluorophores are excited by deep UV, including 4(cid:48),6-diamidino-2-phenylindole (DAPI), Hoechst 33342 (HO342), and eosin (Figure 2). Since glass optics block UV light, no excitation/emission filters or dichroic mirrors are necessary, sig- nificantly reducing the cost of optics while allowing simultaneous multi-channel imaging using a color camera. In addition, the neces- sary deep-UV optics are inexpensive and readily available in the form of quartz lenses. Three-Dimensional Microscopy by Milling with Ultraviolet Excitation -- 2/7 resolution three-dimensional reconstruction of complex samples using both fluorescent (Figure 2) and absorbing (Figure 4) dyes. MUVE Instrumentation Our prototype MUVE imaging system (Figure 1) is based on an HM355S motorized microtome (Thermo Fisher Scientific) capable of automated 0.5 to 100µm sectioning. Our modifications include a FireEdge FE200 LED capable of up to 300 mW emission centered at 280 nm (Phoseon Technology, Hillsboro OR). Custom UV optics were designed to focus the UV source to a 1 mm spot at the sample block face. A custom microscope is mounted laterally to observe the block face. The light path of this microscope is simply com- posed of a 10X objective (Olympus Plan Fluorite objective, 0.3NA), a tube lens (Olympus U-TLU), and a 0.5X camera adapter (Olym- pus U-TV0.5XC-3). A Thorlabs CNS500 objective turret is used to support additional objectives. Emitted fluorescence was detected using a line-scan color camera (Thorlabs 1501C-GE) that provides a theoretical throughput of 1392×23 Hz≈32 kpixel/s at 3 colors per pixel, resulting in a throughput of approximately 96 kB/s. This microscope was rigidly mounted to a two-axis translation stage (Thorlabs XYT1) for positioning and focusing. Materials and Methods Tissue Collection and Labeling Mice were euthanized using CO2 based on guidelines provided by the American Veterinary Medical Association (AVMA). Mice were then perfused transcardially with 20 mL of room temperature phosphate-buffered saline (PBS) solution (pH 7.4), followed by 20 mL of room temperature 10% neutral-buffered formalin (pH 7.4). Perfusion with PBS and formalin removes blood from the circulatory system and fixes the tissue. Mice were then perfused with 10 mL of undiluted India-ink at a rate of ≈1 mL/s. We tested multiple vascular stains, includ- ing polyurethane resin (vasQtec PU4ii) and fluorescent tattoo ink (Skin Candy). Both fluorescent labels provided excellent contrast using block-face imaging. However, vasQtec resin was degraded by alcohol during dehydration prior to perfusion (both ethanol and isoproyl alcohol were tested). While the fluorescent tattoo inks sur- vived embedding, the dyes were composed of fluorescent particles ≈1 µm, resulting in blockages that prevented capillary labeling. We found that India ink (Higgins) [16] provided adequate perfusion and contrast for MUVE imaging. Organs were then removed and fixed in 10% neutral-buffered formalin for 24 h and finally stored in 70% ethanol (C2H5OH). Optionally, tissue samples were also stained using a variety of com- pounds to provide cellular contrast, including DAPI, Hoechst, and eosin [9] for fluorescent imaging and thionine [7, 17] for negative- contrast Nissl staining. Specimen Preparation and Embedding Organs were embedded in paraffin wax for imaging. UV pene- tration was controlled by doping molten paraffin with up to 14% UV27 dye (Epolin). Similar protocols were followed for all ranges of doped paraffin infiltration. Organ sections were dehydrated Figure 1. MUVE imaging. (a) Side view of the MUVE instrumentation showing (1) an automated microtome, (2) deep-ultraviolet (280 nm) source, and (3) standard microscope objective. (b) A planar view shows the optical train, where UV light incident on the sample excites fluorescent labels that are collected by the microscope objective. After imaging, a microtome ablates the tissue section and re-positions the sample for imaging the next section. Penetration depth can be further controlled by using a higher incident angle, however there is a trade-off with illumination in- tensity. Furthermore, recent research shows that water-immersion MUSE achieves approximately 50% reduction in imaging depth compared with air-immersion MUSE [11]. The proposed approach relies on doping the sample embedding medium with a soluble UV-absorbant dye (UV27, Epolin). Increas- ing the dye concentration within the embedding medium reduces the axial point-spread-function (PSF), providing optically thinner sections (Figure 3). Serial Ablation 3D Imaging Several approaches have been proposed for integrating block-face microscopy with serial ablation to enhance axial resolution and image depth. Early studies rely on all-optical imaging and ablation [6], which is time-consuming but applicable to a wide range of tissues. Serial block face scanning electron microscopy (SBF-SEM) [13] uses a microtome blade for ablation, providing nanometer- scale resolution of samples embedded in hard polymers. Alternative approaches achieve similar results using focused ion beams [14]. However, these methods are limited to extremely small micrometer- scale samples and lack molecular specificity. Integration of microtome sectioning with optical approaches has been proposed for large-scale imaging. However, these instruments are extremely expensive to construct and difficult to maintain. For example, knife-edge scanning microscopy (KESM) [7, 15] requires high-precision stages and time-consuming sample protocols, while two-photon tomography [8] requires expensive two-photon imaging systems. The proposed approach, which we refer to as MUSE milling or milling with ultraviolet excitation (MUVE) relies on block-face imaging, requiring the attachment of common microscope optics to an automated microtome. By leveraging the proposed UV-blocking approach described above, it is possible to image formalin-fixed paraffin embedded (FFPE) samples using MUSE, while milling away imaged sections using the microtome. This allows high- 10X Olympusobjectivemicrotomeblade3-channelimagingexponential decayoblique excitationsamplestage𝜃𝜃𝑠𝑠blade holdermicrotomeab𝜃𝜃𝑖𝑖 Three-Dimensional Microscopy by Milling with Ultraviolet Excitation -- 3/7 Figure 2. MUVE imaging of different mouse organs embedded in UV27-doped paraffin wax. (a) Singleplex imaging of mouse lung stained only with HO342. (b) Duplex imaging of mouse cerebellum perfused with India-ink and treated with DAPI. (c) Duplex imaging of mouse kidney stained with Eosin and HO342. (d) Singleplex imaging of mouse testicle stained only with HO342. (e) Duplex imaging of mouse liver perfused with India-ink and treated with HO342. (f) Duplex imaging of mouse spleen stained with Eosin and HO342. through a series of graded ethanols (70 to 100%) over the course of 8 h, followed by clearing with xylene substitute (SIGMA A5597) for 3 h. Standard paraffin wax (Tissue-Tek Paraffin) was selectively doped with UV27 at 60 ◦C, and samples were soaked in the selected mixture for 2 h to allow infiltration. The paraffinization process was performed with the aid of a tissue processor (Leica TP1020). Note that tissue shrinkage is always expected during paraffinization procedures and the degree of shrinkage can reach up to 40% in volume for brain tissue. This can be potential avoided using matri- ces that have low shrinkage artifacts, such as glycol-methacrylate resins (Electron Microscopy Sciences Technovit 7100) or urethane rubbers (Smooth On Clear Flex 95). In particular, we found that Technovit was highly UV opaque, but significantly more difficult to mill. Other nuclear stains, such as DAPI and Hoechst (HO342), are compatible with India ink perfusion. While these stains are sub- ject to bleaching during paraffin infiltration, we have found that paraffinized samples can be stained with DAPI and Hoechst, with penetration up to 1 mm after 3 days of in solution. For example, the Hoechst solution was prepared by diluting the HO342 stock solu- tion (Thermo Fisher Hoechst 33342) 1:2000 in 1X PBS. This also allows staining of 1 to 5µm embedded tissue (Figure 2). Staining was performed by covering the block face with solution for 2 to 3min prior to imaging. Image Collection Conventional microtome blades (DURAEDGE Low Profile) were used for cutting, with a cutting angle of 10◦ (Figure 1b). The single stroke operation mode of the microtome was used for semi- automated acquisition. Cutting velocities were randomized to pre- vent the reinforcement of artifacts such as knife chatter. However, the resting position of the microtome oscillates slightly around its sinusoidbinuclear hepatocytese25µmspermatidsdKupffercellhepatocytes10µmspermatocytesalveoliacapillarybrenal tubulesglomeruluscred pulpcentral arteriolewhite pulpfpneumocytesgranular cellsnephrocytes Three-Dimensional Microscopy by Milling with Ultraviolet Excitation -- 4/7 Figure 3. Monte-Carlo simulations of confocal and MUVE point-spread-functions using coupled-wave theory for absorbance in a layered homogeneous substrate [12]. All simulations show x-polarized coherent light propagating from left to right and intensities are normalized for each image. Contours indicate (from darkest to lightest) 1%, 10%, and 30% thresholds of maximum intensity. (red) Confocal PSFs for imaging in idealized (i.e. cleared) samples using 0.4NA (left), 0.8NA (center), and 1.0NA (right) objectives. In MUVE imaging, exponential absorbance of the excitation is the dominant factor describing the axial PSF. (purple) Incident deep-UV light is shown incident on a sample using a low-NA (≈0.25) objective. Varying the molar absorbance by doping the embedding compound reduces penetration, creating a smaller axial PSF. A UV-transparent sample (left) shows a significant contribution from back-scattered light. However, doping with UV27 (center, right) results in a significant improvement over high-end confocal imaging. (green) The excited region of the samples scales with the penetrating UV PSF, however ablation results in truncation of the left half. (bottom) A comparison is shown between confocal and MUVE axial PSFs. Lateral resolution is theoretically identical between MUVE and confocal. central position causing an offset along the cutting direction. We applied an automated alignment algorithm available in OpenCV [18] for compensation. Camera triggering used an image acqui- sition software package (Thorlabs ThorCam) controlled using an external TTL signal. Images were acquired using 100 ms exposure with a digital gain of 40, and image corrections were performed to adjust brightness and contrast using ImageJ. It took approximately 4 s for each slice: 2 s for cutting, 1 s for stage stabilization, and 1 s for image capturing. This process takes approximately 2 hours to collect 2,000 slices. While this prototype system uses a commercial microtome, a fully-automated system could achieve a data rate similar to 3D color histology for three-dimensional samples. Results Point Spread Function Characterization The lateral resolution of MUVE is diffraction limited, and similar to fluorescence microscopy is determined by the emission wavelength and objective numerical aperture (NA). We verified the lateral reso- lution using a USAF 1951 resolution test target (Edmund Optics). Images for this paper were acquired using a 40X Nikon objective (0.6NA). The horizontal construction of our MUVE prototype pro- hibited the use of immersion objectives, however previous work has already demonstrated MUSE compatibility with water-immersion optics [11]. MUVE axial resolution is dominated by the exponential ab- sorbance of the embedding medium (Figure 3). The presented prototype provides axial resolution beyond the diffraction limit due to limitations in the NA of air objectives. Further studies will be required to determine the practical PSF in other imaging media. MUVE resolution benefits were validated by imaging a phan- tom composed of 1 to 5µm fluorescent green beads (Cospheric Polyethylene Microspheres), (em. 515 nm) which were diluted 1000-fold into UV27-doped paraffin wax. We compared MUVE with wide-field fluorescence microscopy (Nikon Eclipse TI-E In- verted Microscope) using conventional excitation at 390 nm (DAPI excitation). The MUVE axial PSF shows a notable improvement over the traditional pattern of the wide-field microscope (Figure 5). The benefits of the MUVE PSF come in two forms: (1) physical 3210123z-axis (um)101x-axis (um)Confocal PSF at 500nm, 0.4NA3210123z-axis (um)101x-axis (um)Confocal PSF at 500nm, 0.8NA3210123z-axis (um)101x-axis (um)Confocal PSF at 500nm, 1.0NA3210123z-axis (um)101x-axis (um)MUVE Excitation PSF, A = 0.03210123z-axis (um)101x-axis (um)MUVE Emission PSF, A = 0.03210123z-axis (um)0.00.51.0normalized intensityPSF Profiles with MUVE A = 0.0confocal, 0.4NAconfocal, 0.8NAconfocal, 1.0NAMUVE3210123z-axis (um)101x-axis (um)MUVE Excitation PSF, A = 2.03210123z-axis (um)101x-axis (um)MUVE Emission PSF, A = 2.03210123z-axis (um)0.00.51.0normalized intensityPSF Profiles with MUVE A = 2.0confocal, 0.4NAconfocal, 0.8NAconfocal, 1.0NAMUVE3210123z-axis (um)101x-axis (um)MUVE Excitation PSF, A = 4.03210123z-axis (um)101x-axis (um)MUVE Emission PSF, A = 4.03210123z-axis (um)0.00.51.0normalized intensityPSF Profiles with MUVE A = 4.0confocal, 0.4NAconfocal, 0.8NAconfocal, 1.0NAMUVE Three-Dimensional Microscopy by Milling with Ultraviolet Excitation -- 5/7 Figure 4. Advantages of MUVE imaging as 3D microscopy. (a) Block-face imaging of paraffin-embedded brain (top) and UV27-doped paraffin-embedded brain (bottom) using a wide-field fluorescence microscope with DAPI excitation (390 nm). Near-visible penetration in tissue is large in both cases, making it impossible to reconstruct the 3D structure of microvessels. (b) Block-face imaging of paraffin-embedded brain (top) and UV27-doped paraffin-embedded brain (bottom) using MUVE (same regions as shown in (a)). Deep-UV penetration in tissue is significantly shorter than that of near-visible and UV27 infiltration further reduces the excitation volume. (c) Isosurface rendering of paraffin-embedded brain (top) shows uneven vessel surface reconstruction whereas isosurface rendering of UV27-doped paraffin-embedded brain (bottom) shows sharp and smooth vessel surface reconstruction. Figure 5. Central profiling of microspheres using a wide-field fluorescence microscope (blue) and MUVE (green). The axial measurements of micro-beads (≈4 µm) were collected at a 1.0 µm sectioning size for both optical and physical sectioning. Intensity plots were measured across the central line along the cutting direction indicated by a white arrow. 3D volume rendering of large beads (≈500 µm) showing imaging artifacts such as shadows (indicated by black arrows) involved in optical sectioning microscopy. ablation results in a truncated asymmetric emission spot, since pre- vious layers of the sample have been removed, and (2) absorbance of the doped embedding medium dominates the penetrating half of the PSF. This allows reconstruction of elements (i.e. lower parts of spheres) obstructed using optical sectioning. Block-Face Imaging of Fluorescent Samples Mouse organs, including brain, kidney, liver, lung, spleen, and testicle were embedded in UV27-doped paraffin wax and stained after embedding with DAPI. MUVE axial resolution sufficient to resolve individual cell bodies and their chromatin distributions. For example, two types of pneumocytes are distinguishable in the lung image (Figure 2a) and Kupffer cells and hepatocytes are also distinguishable in the liver (Figure 2d). Cerebellar neurons within the granular are also clear to determine, along with their chromatin structure (Figure 2c). The use of oblique UV illumination reveals tissue topographical information with enhanced contrast, consistent with previously published MUSE images [9]. For example, surface profiles of kidney renal tubules are visible using eosin (Figure 2e). The effectiveness of UV27 doping is shown using conventional microscopy (Figure 4a,d), MUSE (Figure 4b,e) and MUVE recon- structions (Figure 4c,f). Direct comparisons in light penetration and reconstruction behaviors between conventional paraffin em- bedding, which introduces some UV absorption, and UV27-doped paraffin-embedding tissues. For instance, the 3D reconstruction of paraffin-embedded tissue shows large and rough microvascular structures whereas the 3D reconstruction of UV27-doped paraffin- embedded tissue shows fine and smooth capillaries. Microvascular Imaging UV excitation is compatible with absorbing (negative) stains, where contrast is provided by exciting auto-fluorescence in the surround- ing tissue and embedding compound. This is particularly useful for microvascular reconstructions using India-ink (Figure 6), which mitigates the need for expensive fluorescent alternatives such as lectins [19]. This data set was imaged with a lateral sample spacing of 0.37 µm and a 2.0 µm cutting interval. Such spatial resolution is capable of reconstructing the smallest capillaries and their surface profiles. While the varying vessel thickness can introduce gaps in volume visualizations (Figure 6d), the images are particularly high contrast and simple to segment using existing algorithms [20]. This allowed us to create an explicit graph model with approximately 8,000 edges and 100,000 vertices of a cortical microvascular network (Figure 7), which was visualized using ParaView (Kitware). Note that networks of this size are challenging to reconstruct with optical sectioning due to increased light scattering with sample depth. 50µmabcdef2µm-10-50510Intensity (a.u.)Axial Position (µm)isocontourholes2µm Three-Dimensional Microscopy by Milling with Ultraviolet Excitation -- 6/7 Figure 6. Coronal MUVE imaging of mouse midbrain stained with India-ink. (a) Volume rendering of the entire data set (389×241×2134µm) showing the densely-connected microvascular network. (b) One complete cross section (z-axis position indicated by a yellow arrow) with a maximum intensity projection (MIP) overlapped over half. (c-d) Close-up view of small regions (128×128×128µm) showing that the sampling resolution of MUVE is large enough to resolve microvessels with different sizes. Microvascular and Nuclear Imaging Finally, we investigate combination staining of both microvascula- ture and nuclei in the brain using thionine with India ink perfusion (Figure 8). A region of the mouse thalamus was imaged at a 0.37 µm lateral resolution with 1.0 µm axial sections to resolve cell nuclei (Figure 8b). This demonstrates potential for studying cellular- vascular relationships linked to many neurodegenerative diseases. This data set was manually segmented using the multi-thresholding function in Amira (ThermoFisher) and visualized using volume rendering (Figure 8c). Discussion and Future Work We have introduced a high-throughput imaging methodology for multiplex imaging of large-scale samples at sub-micrometer reso- lution at low cost. MUSE milling is capable of imaging densely- interconnected microvascular networks, opening the door to simple acquisition and quantification of capillary changes common during disease progression [21] and guide the fabrication of in vitro disease models [22]. The proposed technique is compatible with a wide range of existing objectives, and can be integrated into immersion- based imaging systems to provide lateral resolution equivalent to existing fluorescence techniques. While MUSE milling eliminates constraints on sample depth, additional work on UV doping of embedding compounds will be necessary to further reduce and quantify axial resolution. Finally, the proposed method provides comparable speed to 2D fluorescence imaging, and was able to produce a deep microvascular network (≈2 mm) within 2 hours using an automated microtome. While the prototype is limited to a single FOV, custom microtome using 3-axis stages can provide a cost-efficient technology that is simple to build and maintain in most laboratories. Acknowledgments We would like to thank Pavel Govyadinov at University of Houston for his help on microvascular segmentation. This work was funded in part by the National Science Foundation I/UCRC BRAIN Cen- ter #1650566, Cancer Prevention and Research Institute of Texas (CPRIT) #RR140013, the University of Houston Division of Re- search, and the National Institutes of Health (NIH) / National Heart, Lung, and Blood Institute (NHLBI) #R01HL146745 and National Cancer Institute (NCI) #1R21CA214299. References [1] Goel, H. L. & Mercurio, A. M. VEGF targets the tumour cell. Nature Reviews Cancer 13, 871 -- 882 (2013). [2] Bell, R. D. & Zlokovic, B. V. Neurovascular mechanisms and blood -- brain barrier disorder in alzheimer's disease. Acta Neuropatho- logica 118, 103 -- 113 (2009). [3] Helmchen, F. & Denk, W. Deep tissue two-photon microscopy. Nature Methods 2, 932 -- 940 (2005). [4] Vettenburg, T. et al. Light-sheet microscopy using an airy beam. Nature Methods 11, 541 -- 544 (2014). [5] Rocha, M. D. et al. Tissue clearing and light sheet microscopy: Imaging the unsectioned adult zebra finch brain at cellular resolution. Frontiers in Neuroanatomy 13 (2019). [6] Tsai, P. S. et al. All-optical histology using ultrashort laser pulses. Neuron 39, 27 -- 41 (2003). 20µm25µmbcd100µm25µmacut Three-Dimensional Microscopy by Milling with Ultraviolet Excitation -- 7/7 Figure 8. Coronal MUVE imaging of mouse thalamus stained with India-ink and thionin. (a) Tissues are dark red while cell nuclei are dark brown (arrow) and vessels are black (dashed arrow) under UV illumination (left), providing enough contrast to segment both the cellular and vascular structures (right). (b) Volume rendering of the entire data set shows the density and organization of microvasculature with surrounding cellular details. (c) Volume rendering of a small region (100×100×100µm) shows series of connected microvessels (red) along with their associated cells (green), and separate channels show detailed cellular and vascular structures. [15] Li, A. et al. Micro-optical sectioning tomography to obtain a high- resolution atlas of the mouse brain. Science 330, 1404 -- 1408 (2010). [16] Mayerich, D. et al. Fast macro-scale transmission imaging of mi- crovascular networks using kesm. Biomedical optics express 2, 2888 -- 2896 (2011). [17] Choe, Y. et al. Specimen preparation, imaging, and analysis protocols for knife-edge scanning microscopy. JoVE (Journal of Visualized Experiments) e3248 (2011). [18] Evangelidis, G. D. & Psarakis, E. Z. Parametric image alignment using enhanced correlation coefficient maximization. IEEE Transactions on Pattern Analysis and Machine Intelligence 30, 1858 -- 1865 (2008). [19] Porter, G., Palade, G. & Milici, A. Differential binding of the lectins griffonia simplicifolia i and lycopersicon esculentum to microvascu- lar endothelium: organ-specific localization and partial glycoprotein characterization. European journal of cell biology 51, 85 -- 95 (1990). [20] Govyadinov, P. A., Womack, T., Eriksen, J. L., Chen, G. & Mayerich, D. Robust tracing and visualization of heterogeneous microvascu- IEEE Transactions on Visualization and Computer lar networks. Graphics 25, 1760 -- 1773 (2019). [21] Castillo-Carranza, D. L. et al. Cerebral microvascular accumulation of tau oligomers in alzheimer's disease and related tauopathies. Aging and Disease 8, 257 -- 266 (2017). [22] Guo, J. et al. Accurate flow in augmented networks (AFAN): an approach to generating three-dimensional biomimetic microfluidic networks with controlled flow. Analytical Methods 11, 8 -- 16 (2019). Figure 7. Mouse cerebral cortex microvasculature (top) and segmentation (bottom). The smallest microvessel (rendered in blue) acquired in this data set is ≈6 µm in diameter while the largest (rendered in purple) is ≈148 µm. [7] Mayerich, D., Abbott, L. & McCormick, B. Knife-edge scanning mi- croscopy for imaging and reconstruction of three-dimensional anatom- ical structures of the mouse brain. Journal of Microscopy 231, 134 -- 143 (2008). [8] Ragan, T. et al. Serial two-photon tomography for automated ex vivo mouse brain imaging. Nature methods 9, 255 (2012). [9] Fereidouni, F. et al. Microscopy with ultraviolet surface excitation for rapid slide-free histology. Nature Biomedical Engineering 1, 957 -- 966 (2017). [10] Meinhardt, M., Krebs, R., Anders, A., Heinrich, U. & Tronnier, H. Wavelength-dependent penetration depths of ultraviolet radiation in human skin. Journal of Biomedical Optics 13, 044030 (2008). [11] Yoshitake, T. et al. Rapid histopathological imaging of skin and breast cancer surgical specimens using immersion microscopy with ultraviolet surface excitation. Scientific Reports 8, 4476 (2018). [12] Davis, B. J., Carney, P. S. & Bhargava, R. Theory of midinfrared absorption microspectroscopy: I. homogeneous samples. Analytical chemistry 82, 3474 -- 3486 (2010). [13] Denk, W. & Horstmann, H. Serial block-face scanning electron mi- croscopy to reconstruct three-dimensional tissue nanostructure. PLoS biology 2, e329 (2004). [14] Knott, G., Marchman, H., Wall, D. & Lich, B. Serial section scanning electron microscopy of adult brain tissue using focused ion beam milling. Journal of Neuroscience 28, 2959 -- 2964 (2008). 100µmradius (µm)74.03.3310.070.030.020.050µmabccellsvessels10µm
1801.06844
2
1801
2019-08-13T18:37:08
Dynein catch bond as a mediator of codependent bidirectional cellular transport
[ "physics.bio-ph", "q-bio.SC" ]
Intracellular bidirectional transport of cargo on microtubule filaments is achieved by the collective action of oppositely directed dynein and kinesin motors. Experiments have found that in certain cases, inhibiting the activity of one type of motor results in an overall decline in the motility of the cellular cargo in both directions. This counter-intuitive observation, referred to as {\em paradox of codependence} is inconsistent with the existing paradigm of a mechanistic tug-of-war between oppositely directed motors. Unlike kinesin, dynein motors exhibit catchbonding, wherein the unbinding rates of these motors decrease with increasing force on them. Incorporating this catchbonding behavior of dynein in a theoretical model, we show that the functional divergence of the two motors species manifests itself as an internal regulatory mechanism, and leads to codependent transport behaviour in biologically relevant regimes. Using analytical methods and stochastic simulations, we analyse the processivity characteristics and probability distribution of run times and pause times of transported cellular cargoes. We show that catchbonding can drastically alter the transport characteristics and also provide a plausible resolution of the paradox of codependence.
physics.bio-ph
physics
Dynein catch bond as a mediator of codependent bidirectional cellular transport Palka Puri,1, 2 Nisha Gupta,3 Sameep Chandel,3 Supriyo Naskar,4 Anil Nair,5 Abhishek Chaudhuri,3, ∗ Mithun K. Mitra,1, † and Sudipto Muhuri5, ‡ 2Institute for Nonlinear Dynamics, Georg August University of Goettingen, Germany 1Department of Physics, IIT Bombay, Mumbai, India 3Indian Institute of Science Education and Research, Mohali, India 4Department of Physics, Indian Institute of Science, Bangalore, India 5Department of Physics, Savitribai Phule Pune University, Pune, India (Dated: August 15, 2019) Intracellular bidirectional transport of cargo on microtubule filaments is achieved by the collective action of oppositely directed dynein and kinesin motors. Experiments have found that in certain cases, inhibiting the activity of one type of motor results in an overall decline in the motility of the cellular cargo in both directions. This counter-intuitive observation, referred to as paradox of code- pendence is inconsistent with the existing paradigm of a mechanistic tug-of-war between oppositely directed motors. Unlike kinesin, dynein motors exhibit catchbonding, wherein the unbinding rates of these motors decrease with increasing force on them. Incorporating this catchbonding behavior of dynein in a theoretical model, we show that the functional divergence of the two motors species manifests itself as an internal regulatory mechanism, and leads to codependent transport behaviour in biologically relevant regimes. Using analytical methods and stochastic simulations, we analyse the processivity characteristics and probability distribution of run times and pause times of transported cellular cargoes. We show that catchbonding can drastically alter the transport characteristics and also provide a plausible resolution of the paradox of codependence. I. INTRODUCTION Bidirectional transport is ubiquitous in nature in the context of intracellular transport [1 -- 4]. Within the cell, oppositely directed motor proteins such as dynein and ki- nesin motors walk on microtubule (MT) filaments [1, 5] to transport diverse cellular cargo [1]. A theoretical framework proposed to explain the bidirectional trans- port is based on the tug-of war hypothesis [1, 3, 5 -- 10], which posits that the motors stochastically binds to and unbinds from the filament while mechanically in- teracting with each other through the cargo that they carry (Fig. 1a) [6 -- 8]. The resultant motion arises due to the competition between the oppositely directed motors [7, 8]. The tug-of-war model predicts that inhibiting the ac- tivity of one motor species would lead to an enhancement of motility in the other direction. While many experi- ments have provided support for this mechanical tug-of- war picture [5, 7, 11 -- 13], there remain a large class of experiments which are incompatible with the predictions of this model and show that there exists some coordi- nation mechanism due to which inhibition of one motor species results in an overall decline in the motility of the cargo [2, 6, 14 -- 17]. This apparently counterintuitive find- ing has been referred to as the paradox of codependence [1, 6]. The resolution of this paradox in terms of the underlying mechanisms which govern bidirectional trans- port remains an important open question. ∗ [email protected][email protected][email protected] Unlike kinesin, whose detachment rates from the fila- ment increases exponentially with increasing load force - a characteristic of slip bond [18 -- 20], dynein motors ex- hibit catchbonding : the propensity for the dynein motors to unbind decreases when subjected to increasing load forces in certain force regimes (Fig1b) [19, 21, 22]. While the effect of catchbonding has previously been incorporated in context of modeling of bidirectional transport of lipid droplets [21], their importance in me- diating codependent transport properties has not been realized and investigated. In this article we study the generic mechanism by which catchbonding in dynein may manifest as codependent transport behaviour for cellular cargoes and quantify the effects of the catchbond in terms of experimentally measurable cargo transport character- istics. In particular we explicitly show that catchbonding in dynein provides one plausible means of resolving the paradox of codependence. We use a threshold force bond deformation (TFBD) model to fit the experimentally observed unbinding rate of single dynein motors (Fig. 1(b)) [23]. With the TFBD model for dynein, and the usual slip bond model for ki- nesin [7, 20], we study the transport properties of bidirec- tional cargo motion by multiple motors, using experimen- tally relevant measures : (i) average processivity, defined as the mean distance a cargo travels along a filament be- fore detaching, (ii) probability distributions of runtime and pause times, and (iii) typical cargo trajectories as well as distributions of cargo velocities. Using these mea- sures we show that, in an experimentally viable param- eter space, the catchbonded response of dynein provides an internal regulatory mechanism that exhibits codepen- dent transport characteristics. 2 FIG. 1. (a) Schematic of bidirectional motion of cargo (C) attached to both kinesin (K) and dynein (D) motors on a microtubule (MT) filament; (b) Single dynein unbinding rate from experiments [21] (points) and the corresponding fit (solid line) from the TFBD model [23]. II. THEORY AND SIMULATION A. Model We consider transport of a cellular cargo with N+ kinesin motors and N− dynein motors. These motors stochastically bind to a MT filament with rates π± and unbind with rates ε±. At any instant of time, the state of the cargo is characterized by the number of attached Kinesin (n+) and Dynein motors (n−). The maximum of number of kinesin and dynein motors are N+ and N− re- spectively (0 < n+ < N+ and 0 < n− < N−). The time evolution of the system is then governed by the master equation [7] ∂p(n+, n−) ∂t = p(n+ + 1, n−)+(n+ + 1, n−) + p(n+, n− + 1)−(n+, n− + 1) +p(n+ − 1, n−)π+(n+ − 1, n−) + p(n+, n− − 1)π−(n+, n− − 1) −p(n+, n−) [+(n+, n−) + −(n+, n−) + π+(n+, n−) + π−(n+, n−)] (1) where, p(n+, n−) is the probability to find the cargo with n+ kinesin and n− dynein motors. The kinesin and dynein binding rates are assumed to be of the form π± = (N± − n±)π0±, where N+π0+ (N−π0−) is the rate for the first kinesin (dynein) motor to bind to the MT. Dynein motors exhibit catchbonding at forces larger than the stall force, Fs−, defined as the load force at which the cargo stalls [19, 21, 22]. This catchbonding regime is characterized by a decreasing detachment rate with increasing opposing load (see Fig. 1(b)). The load force is assumed to be shared equally among the attached motors. We use the phenomenological TFBD model for the unbinding rate of a dynein in an (n+, n−) state [23, 24], given by dissociation energy and the deformation energy respec- tively, while Fc is the cooperative force felt by the mo- tors due to the effect of the motors of the other species. Unlike dynein, the unbinding kinetics of kinesin ex- hibits usual slip behavior, and thus the unbinding rate for kinesin is given by the expression ε+(n+, n−) = n+ε0+ exp[Fc(n+, n−)/(n+Fd+)] [7]. The characteristic stall forces and detachment forces of kinesin are denoted by Fs+ and Fd+ respectively. The expression for the cooperative force felt by the motors is given by [8] Fc(n+, n−) = n+n−Fs+Fs− n−Fs−v0+ + n+Fs+v0− (v0+ + v0−) (4) ε− = n−ε0− exp[−Ed(Fc) + Fc/(n−Fd−)] where the deformation energy Ed sets in at F > Fs−, and is modeled by a phenomenological equation [23], Fc/n− − Fs− Ed(Fc) = Θ(Fc/n−−Fs−)α 1 − exp (cid:19)(cid:21) (cid:20) (cid:18) − (3) The parameter α sets the strength of the catch bond, while Fd− and F0 characterize the force scales for the v0+ = F0 Here, v0± denotes the velocity of kinesin (or dynein) mo- tors, (cid:26) vF + if vc > 0 vB+ if vc < 0 (cid:26) vF− if vc < 0 vB− if vc > 0 and v0− = and the cargo velocity is given by (2) vc(n+, n−) = n+Fs+ − n−Fs− n−Fs−/v0− + n+Fs+/v0+ (5) (a)(b)CMTDK 3 The average processivity reported in this manuscript is the average over all possible initial states of the motor conformations for a given maximum number of kinesins and dyneins, N+(cid:88) N−(cid:88) n+=0 n−=0 (cid:0)1 − δn++n−,0 (cid:1) , (cid:104)Ln+,n−(cid:105)n+n− = C Ln+,n− (8) where C is a normalization factor which depends on N+ and N−, with C−1 = (N+ + 1)(N− + 1) − 1. C. Simulations and Numerical techniques The Master equation is simulated using Stochastic Simulation Algorithm (SSA) [35, 36] to obtain individ- ual cargo trajectories. All possible initial configurations were generated for a (N+, N−) pair, and 1000 trajecto- ries were evolved for each initial configuration. A run finishes if the simulation continues until the maximum time TM AX ∼ 104s or if all motors detach from the MT. The runlength was then averaged over all initial configu- rations and all iterations. Probability distributions were also computed from the SSA trajectories after discarding initial transients. The simulated trajectories are then analysed to quantify the statistical properties of the sys- tem. Further we perform Brownian dynamics simulations and determine processivity of the cargo when the load is shared stochastically (see Appendix A). Further we also derive the associated Fokker-Planck equation (FPE) corresponding to the underlying Master equation, by treating the number of attached motors as continuum variables in the state space (see Appendix B). III. RESULTS A. Cargo Processivity Characteristics 2 (a)) and N+ (Fig. We show results for the average processivity (cid:104)Ln,m(cid:105)nm with varying N− (Fig. 2 (b)) for weak dynein (mammalian, Fs− = 1pN ) and strong dynein (yeast, Fs− = 7pN ). For all cases, the analytical value of the average processivity shows excellent agree- ment with SSA results. In Fig. 2 (a), we observe a sharp decrease in plus-end directed processivity with increase in N−, due to an increased propensity of catchbonded dynein motors to latch on to the filament. For weak dynein, the cargo in fact reverses direction, due to the activation of catchbond at lower forces. The Mean First Passage time (MFPT) in a particular bound motor state (n+, n−), Tn+,n−, is defined as the mean time for cargo starting with n bound kinesins and m bound dyneins to unbind, i.e; reach the (0, 0) state. This can be expressed in terms of mean residence time in that state τn+,n− and transition probabilities to other states, which leads to a recursion relation for the MFPT, of the form Tn+,n− = τn+,n− 1 + π+ n+,n−Tn++1,n− + π + ε+ − n+,n− Tn+−1,n− + ε n+,n− Tn+,n−−1 − n+,n− Tn+,n−+1 (cid:17) (6) n+,n− ). n+,n− + π− where the mean residence time in a (n+, n−) state is sim- ply the inverse of the sum of the transition probabili- ties to the other states, τn+,n− = 1/(π+ n+,n− + n+,n− + ε− ε+ We can similarly develop a recursion relation for the average cargo processivity (ACP) Ln+,n− , defined as the average distance a motor starting from the (n,m) state walks before it unbinds. In the state (n+, n−), the cargo walks with the cooperative velocity, vc(n+, n−), and the mean residence time in this state is τ (n+, n−). Hence the mean distance ηn+,n− that the cargo walks in the (n+, n−) state before transition to another state can be expressed as ηn+,n− = vc(n+, n−)τn+,n− . With this iden- tification, the recursion relation for the mean processivity becomes, Ln+,n− = ηn+,n− 1 + π+ n+,n− Ln++1,n− + π + ε+ n+,n− Ln−1,m + ε − n+,n− Ln+,n−−1 − n+,n− Ln+,n−+1 (cid:17) Parameter Kinesin Ref. Dynein Fs± Fd± π0± ε0± vF± vB± Ref. [26] [27] 6 pN [25] 1 pN (Weak) 7 pN (Strong) 3 pN 5/s 1/s [25] [28] [25] 0.65µm/s [31] [31] 1nm/s 0.67 pN [21, 23] 1/s (0.1 - 10)/s 0.65µm/s [29] [30] [32] 1nm/s [33, 34] TABLE I. Single motor parameter values used in the simula- tions. where, vF and vB are the forward and backward mo- tor velocities. Finally the stall forces for the two motor species are denoted by Fs±. The parameters used in the study are taken from the literature, and are summarized in Table I. B. First Passage Time and Processivity (cid:16) (cid:16) Together with the absorbing boundary conditions, T0,0 = 0 and L0,0 = 0, these define a linear system of equations which can be solved analytically to obtain the MFPT and the ACP. Diverse experiments have indicated that mutations of conventional kinesin in Drosophila can hamper motion of cellular cargo in both directions, by effectively reducing (7) B. Resolution of Paradox of Codependence 4 FIG. 2. Average processivity (a) as a function of N− for N+ = 4 , and (b) as a function of N+ for N− = 4. The colored points and lines correspond to the simulation results. Black crosses in all cases are obtained by the solutions of Eq. 8. Contour plots for processivity obtained from Eq. 8 in the N+ − N− plane for (c) Fs− = 1pN, α = 0, F0 = 7pN , and (d) Fs− = 1pN, α = 40kBT, F0 = 7pN . The color bar indicates the average processivity (in µm). The zero-force (un)binding rates for dynein are ε0− = π0− = 1/s FIG. 3. (a) Average processivity as a function of ε0− for different stall forces at α = 40kBT ; obtained using Eq. 8. (b) Contour plots of processivity in (Fs− − ε0−) plane for α = 40kBT and F0 = 7pN . Data shown is for N+ = 6, N− = 2, π0− = 1/s. the number of motors attached to the cargo [15, 16, 37 -- 39]. While the conventional tug-of-war model without the incorporation of catchbond does not exhibit code- pendent transport characteristics and fails to resolve the paradox of codependence observed in these experiments, the processivity characteristics reveals clear signature of −150−100−500500246810AverageProcessivity(µm)N−(a)−150−100−500501001500123456789AverageProcessivity(µm)N+(b)Fs−=1,α=0Fs−=1,α=40Fs−=7,α=40Fs−=7,α=0Fs−=1,α=0Fs−=1,α=40Fs−=7,α=40Fs−=7,α=00246810N+0246810N−<−150−100−50050100>150(c)0246810N+0246810N−<−150−100−50050100>150(d)−50−2502550012345678910(a)AverageProcessivity(µm)ε0−(/s)Fs−=1pNFs−=2pNFs−=4pNFs−=7pN012345678910Fs−(pN)12345ε0−(/s)<−50−40−2002040>50(b) plausible resolution of this Paradox by means catchbond mediated mechanism. To investigate this, in Fig. 2 (b), we look at the effect of variation of N+ on processivity, for a fixed value of N−. Remarkably, the average proces- sivity for weak dynein shows a non-monotonic behaviour In particular there is decrease of with increasing N+. processivity in the negative direction on decreasing the number of plus-end directed motors. This is a singular feature arising solely due to catchbonding in dynein, con- trary to usual tug-of-war predictions, and is reminiscent of the paradox of codependence. The robustness of this catchbond mediated phenomenon can be gauged from the observation that it persists for a wide range of biologi- cally relevant parameters even when the load is shared stochastically between the motors (see Appendix A and Fig. 8 for details). This codependent behaviour exemplified in processiv- ity characteristics may be understood in terms of the catchbond mechanism at play. In the absence of oppos- ing load, increasing N+ has the effect of increasing the mean first passage time (MFPT) for the kinesin motors. However in the presence of dynein, with larger number of kinesins, the load per dynein is higher, leading to engage- ment of the catchbond and thus fewer detachment events for dynein. The cargo is now in a tug-of-war state, lead- ing to higher detachment forces on the opposing kinesins, which detach with the usual slip kinetics. Thus, on av- erage, for some parameter regime, the kinesins detach at a higher rate than dyneins, leading to more configu- rations where there are no kinesins opposing the dynein team. Thus although the direct effect of the catch bond is a larger value of average unbinding time for dyneins, this leads to more configurations where the dyneins can walk towards the negative end leading to codependent transport. The corresponding contour plots of the processivity of the cargo, which provide an experimental testbed, in the (N+ − N−) plane are shown in Fig. 2(c-d), for weak dynein where the effect of dynein catch-bond is ro- bust. As expected, in the absence of catch-bond (α = 0) (Fig. 2(c)), there is a smooth transition from negative- directed runs to positive directed runs. In the presence of catch-bonded dynein (Fig. 2(d)), we observe a distinct regime where the processivity increases in the negative di- rection on increasing N+, reminiscent of anomalous code- pendent transport. Plus-end directed motion now occurs only for large N+ and low N−. This non-trivial effect of the catch bond is a robust feature that is observed for other values of kinesin and dynein motors (see Appendix C, Figs. 9 and Fig. 10) and can also be understood in terms of the average number of bound motors (see Ap- pendix D, Fig. 11). Experimental techniques to modulate cargo processiv- ity can also be achieved by modifying the (un)binding rates of the motor proteins. Dynactin mutations in Drosophila neurons affect the kinetics of dynein bind- ing to the filament, leading to cargo stalls [14]. Simi- larly, the tau protein has been observed to change the 5 unbinding rates of kinesin and dynein motors [40]. To investigate this, we look at the effect of variation of the bare unbinding rate of dynein motor on processivity of the cargo (ε0−) (Fig. 3a). Codependent transport be- haviour is again observed for a range of stall forces. For instance at Fs− = 2pN , we observe a non-monotonic behaviour of the processivity with increasing unbinding rate. At Fs− = 4pN , the run length in the positive direc- tion decreases on increase in ε0−. The contour plot of the processivity in the (Fs− − ε0−) plane (Fig. 3(b)) shows non-monotonic signatures of codependent transport - a feature akin to reentrant behaviour [41] - for a range of stall forces, and highlights the role of dynein stall force in determining the overall motion of the cellular cargo. The strength of catchbond (α) plays an important role in determining the nature of processivity of the cargo (Ap- pendix E, Fig. 12). A microscopic modeling of the catch bond in dynein based on the experimentally determined mechanism of the catch bond [42] can help identify bi- ologically relevant regimes for α and Fo and therefore constrain the predictions of the model. C. Probability distribution of runtime and cargo velocities In order to highlight the role of catchbond we provide quantitative measures which are biologically relevant for comparison with experimental data related to trajecto- ries of cellular cargo carried by molecular motors. We analyze the probability distribution of the time the cargo spends in the paused (tug-of-war) state versus the time it spends in the moving plus-end directed and minus-end directed state, as well as the probability distribution of the velocities of the cargo. Motivated by experiments on dictyostelium cell ex- tracts [13], we study the transport behaviour of a cargo with N+ = 2 and N− = 6 (Fig. 4 and Fig. 5). In the absence of catchbonding, cargoes predominantly move with positive velocity and the resultant motion is strongly plus-end directed (Fig. 4d). The probability distributions of runtime show that there are many more kinesin runs (Fig. 4c) than dynein runs (Fig. 4a), and the average run- time is also higher in the case of kinesins. The pauses in this case are also of extremely short duration (Fig. 4b). The corresponding probability distribution for the veloc- ities are shown in Fig. 5a In contrast, when dynein catch bond is switched on, the picture changes dramatically. While the cargo is in a paused state a significant fraction of time, around 35% of its runs are negative directed (Fig. 4f inset). Minus-ended runs become much more frequent than plus-ended runs, and the cargoes tend to move with a negative velocity (Fig. 5b) while the average pause time also increases by an order of magnitude compared to the non-catchbonded case, and becomes comparable to the average minus di- rected runtime. This is shown in Figs. 4(e)-(g). This prediction of minus-ended runs with intermittent pauses 6 FIG. 4. Probability distributions of runtime for N+ = 2, N− = 6. The top panels show the normalized histograms and sample trajectories for dynein in the absence of catch bond (α = 0). The bottom panels show the corresponding quantities for catch-bonded dynein (α = 40, F0 = 7pN ). (a) and (e) Distributions of runtime for minus directed runs (shown in red); (b) and (f) pausetime distributions (shown in blue)(c) and (g) distributions of runtime for plus directed runs (shown in green); and (d) and (h) sample trajectories. Insets in (a) and (g) show magnified views of the corresponding distributions. ended runs is about one order of magnitude lower than that of the plus-end directed run duration. Further there are now substantial duration of pauses (1 − 4 sec) dur- ing transport. These characteristics of the probability distributions result in typical cargo trajectories which exhibits bidirectional motion with pauses. The role of dynein catchbonding in altering the transport character- istics can also be seen for the simplest possible case of bidirectional transport of a cargo by a single kinesin and a single dynein motor (Appendix F, Fig. 13). D. Quantitative comparison with experiments In order to provide a quantitative comparison of our results with in-vivo experiments, we consider the specific case of kinesin inhibition in mouse neurons [17]. It was observed that inhibiting kinesin resulted in smaller retro- grade run lengths of prion protein vesicles, which is con- trary to expectations - a signature of codependent trans- port behaviour. In our model, kinesin inhibition is incor- porated by reducing the number of kinesins (N+) from 3 to 2 while the dynein number is held fixed (N− = 4). As shown in Fig. 7, this reduction in kinesin motors leads to smaller retrograde run lengths and larger anterograde run lengths when catch bond is switched on in dynein, as opposed to the situation when dynein unbinding ex- hibits slip behavior. This is the scenario of co-dependent FIG. 5. Probability distributions of velocities for N+ = 2, N− = 6.The left panels show the normalized histograms in the absence of catch bond (α = 0). The right panels show the corresponding histograms when dynein is catch-bonded (α = 40, F0 = 7pN ). qualitatively agrees with the experimental observation of transport of endosomes in Dictyostelium cells [13]. In a separate set of experiments on early endosomes in fungi, a team many kinesin motors (3-10) are involved in tug-of-war with 1 or 2 dynein motors during transport [5]. The results displayed in Fig. 6 for a cargo being trans- ported by six kinesins and two dyneins illustrates that while in the absence of catchbonding in dynein, the re- sultant motion would be strongly plus-end directed, with very small pause times, incorporation of catchbonding results in the frequency of minus-ended runs exceeding the frequency of plus-ended runs by almost one order of magnitude. However, the average duration of the minus- 01234500.511.5p(t−)t−(s)(a)Minusruns00.0250.0500.250.501234500.511.5p(t0)t0(s)(b)Pauseruns01234500.511.5p(t+)t+(s)(c)Plusruns02468101205101520x(µm)t(s)(d)Sampletrajectories01234500.511.5p(t−)t−(s)(e)Minusruns01234500.511.5p(t0)t0(s)(f)Pauseruns01234500.511.5p(t+)t+(s)(g)Plusruns00.00250.00500.250.5−5−4−3−2−1005101520x(µm)t(s)(h)Sampletrajectories00.20.40.60.81−0.650.00.65(a)P(vc)vc(µm/s)00.20.40.60.81−0.650.00.65(b)P(vc)vc(µm/s) 7 FIG. 6. Probability distributions of runtime for N+ = 6, N− = 2. The top panels show the normalized histograms and sample trajectories for dynein in the absence of catch bond (α = 0). The bottom panels show the corresponding quantities for catch-bonded dynein (α = 40, F0 = 7pN ). (a) and (e) Distributions of runtime for minus directed runs (shown in red); (b) and (f) pausetime distributions (shown in blue); (c) and (g) distributions of runtime for plus directed runs (shown in green); and (d) and (h) sample trajectories. Insets, where present, show a magnified view of the probability distributions. FIG. 7. Histograms showing scaled (a) retrograde and (b) anterograde run lengths with non-catchbonded (Expected, red, α = 0) and catchbonded (Observed, blue, α = 40kBT , F0 = 7pN ), when N+ is changed from 3 to 2, while N− = 4. The scaling is done with respect to the control ( without kinesin inhibition) and corresponds to N+ = 3. The zero-force (un)binding rates for dynein are ε0− = π0− = 1/s transport and compares well with the experimental ob- servations. Our assumption that kinesin inhibition leads to reduction in its number is a simplified view of the effect of the inhibition experiment in in-vivo conditions. Nonetheless, even with this assumption our results defini- tively points to the role of catchbond mediated mecha- nism in determining codependent transport behaviour. IV. CONCLUSION In summary, the findings of our model points to the crucial role played by catchbonding in dynein motors in internally regulating transport and providing a pos- sible resolution of the paradox of codependence. It also provides a framework to interpret diverse set of experi- ments where regulation of transport is achieved by differ- ent modes of modification of the motor properties. For 0123024p(t−)t−(s)(a)Minusruns00.00020.000400.050.10.150123024p(t0)t0(s)(b)Pauseruns0123024p(t+)t+(s)(c)Plusruns02468101205101520x(µm)t(s)(d)Sampletrajectories0123024p(t−)t−(s)(e)Minusruns0123024p(t0)t0(s)(f)Pauseruns0123024p(t+)t+(s)(g)Plusruns00.10.2012−101205101520x(µm)t(s)(h)Sampletrajectories00.511.522.5Retrograderunlength(a)00.511.522.5Anterograderunlength(b)ExpectedObservedExpectedObserved instance, while decreasing N− or increasing ε0− has the effect of weakening the dynein motor action, the manifes- tation of these two effects in the transport characteristics can in general be distinct. The results of these experi- ments can then qualitatively be understood in the light of Fig. 2 and Fig. 3, where weakening the dynein motor can lead to stalled motion of the cargo. Interestingly, while kinesin exhibits a conventional slip bond, the co- operative force exerted by the catch bonded dynein on kinesins, and vice-versa, introduces a complex interplay which results in signatures of codependent transport be- ing observed even on varying effective kinesin numbers. This effect is reflected in a preliminary comparison of processivity measurements for prion protein vesicles in mouse neurons [17] with our model predictions. These processivity measures also point to the sharp dif- ference in transport characteristics for strong and weak dynein. In the former case, regulatory role of catchbond- ing is very weak due to the high force scale at which catchbond is activated. This may provide a clue as to why the strong dynein in yeast is not involved in trans- port, while weak mammalian dynein are crucial to intra- cellular transport. Apart from the internal regulatory mechanism de- scribed here, external regulation by associated proteins is also expected to play an important role in determining the transport characteristics. Various candidate proteins such as Klar and JIP1 have been shown to modify trans- port behaviour [1, 3, 16, 43 -- 50]. Further, various other factors, such as memory effects during motor rebind- ing [19], interactions between multiple motors [51, 52], variable dynein step sizes [21, 53], and stochastic load sharing could also modify the transport behaviour of the cargo. However we show using simulations incorporating a stochastic sharing of load between attached motors, that the codependent behavior of cargo processivity is robust and is preserved even with additional inputs such as viscous friction and thermal noise (see Supplementary section I(E)). Various regulatory mechanisms are expected to achieve coordination through different means which may be re- flected in the transport characteristics of the cargo. For example, in the case of the catch-bonded tug-of-war me- chanical model, the pause state would in general be char- acterized by a slow velocity of the cargo. On the other hand, for mechanical inhibition [6, 54], microtubule teth- ering mechanism [6, 55] or steric disinhibition[6, 56], the motion of the cargo would either be diffusive or would show no movement. Increasing the binding rates of ei- ther motor species would result in shorter pause times if coordination is achieved through mediation by the catch bond, while it would have no effect on the pause times for some other mechanism. A careful examination of high resolution spatio-temporal measurement of cargo proces- sivity and pause durations obtained in various experi- ments is required to delineate the relative importance of these internal regulatory mechanisms. To conclude, we show that catchbonding in dynein dra- 8 matically alters the transport characteristics, and mani- fests as an internal regulatory mechanism that provides one possible resolution of the paradox of codependence. Acknowledgements. Financial support is acknowledged by MKM for Ramanujan Fellowship (13DST052), DST and IITB (14IRCCSG009); AC for SERB project No. EMR/2014/000791 and DST; SM and MKM for SERB project No. EMR /2017/001335; NG for UGC and SC for DST. MKM acknowledges hospitality of MPIPKS, Dresden. SM and MKM acknowledge helpful discussions with Roop Mallik, TIFR. P.P and N.G carried out simulations and analyzed data. S.C, S.N., and A.N. carried out simulations. A.C., M.K.M. and S.M. designed study, analyzed data and wrote manuscript. Appendix A: Stochastic load sharing FIG. 8. Average processivity as a function of N+, as the bead size σ is changed. Note that the friction constant ζ changes as a result. The blue curve shows the corresponding result under the equal load sharing assumption. Here N− = 4, α = 40 and Fo = 7pN . In order to ensure that the codependent transport characteristics obtained are not artifacts of the mean field assumption, where motors are assumed to share the load force equally, we also performed Brownian dynamic simu- lations where the load is shared stochastically, with each motor having a different extension, and hence facing a different opposing load. In the simulation, N motors are attached to the cargo. The motors are modeled as elastic springs with spring constant k = 0.32 pN/nm. The springs have a rest length l0 and generate a restoring force only when stretched be- yond the rest length. The rest length of the springs are chosen in accordance with earlier simulations, l0 = 100 02468101214N+-1000100Average processivity (µm)σ = 0.25σ = 0.30σ = 0.35MF modelN- = 4 nm for kinesin and l0 = 50 nm for dynein. In this one dimensional model, we start by putting the bead at the origin and all N motors attached irreversibly to the cargo at one end. The other end of the motors are allowed to bind to any point on the track within the rest length of the corresponding motor, on either side of the bead. At every time step, all the N motors are visited to determine if they are in the attached or detached state. Each motor position and their state are updated only once in a time step. If the motor is in the detached state, then it can re attach with a probability Pon = π±∆t, where π± are the binding rates of kinesin and dynein as defined earlier. The attachment happens within a dis- tance l0 on either side of the bead. If the ith motor is in an attached state, then the load force, Fi is cal- culated by multiplying the extension of the spring with the spring constant k. Depending on the load force, the motor could detach, with probability Poff = ε±(Fi)δt, where ε± are the unbinding rates of kinesin and dynein. Note that for dynein, Fi replaces Fc in Eq. 2 in the manuscript. If the motor does not detach, then we cal- culate the probability of taking a step, Pstep = kstep∆t, where kstep = (v0±/d)(1 − Fi/Fs±), where v0± is the un- loaded velocity of the single motor, Fs± is the stall force of the motor and d = 8 nm is the step length of the motor. Note that this form is used for backward loads Fi < Fs. For backward loads Fi > Fs, Pstep = 0. For forward loads, Fi = 0. If the motor steps, its position is updated from xi to xi + d. All motor states and their po- sitions are updated simultaneously in a given time step. Two sets of motors with their characteristic parameters as given in Table I, move in opposite directions. To update the position of the cargo (modeled as a bead of radius σ), we calculate the total force acting on the cargo due to both sets of molecular motors moving in opposite directions, Ftot =(cid:80) Fi. Note that the detached motors do not contribute to the total force, neither do the motors which lie within a rest length from the bead position. The bead is under the influence of both thermal and viscous forces with ξ = 0.001 pN-s/µ− m2 being the viscosity of the medium. The bead diffuses with diffusion constant D = kBT /ζ where ζ = 6πξσ is the friction constant. When the cargo is subjected to the force Ftot it moves with the velocity vd = Ftot/ζ. In the presence of thermal noise, the overdamped Brownian dynamics of the cargo is given by tain the associated Fokker Planck equation. We define, x = n+/N+ and y = n−/N−, and in terms of these vari- ables, 9 1 N+ p(x ± , y) = p(x, y) ± 1 p(x, y ± N− 1 N+ ) = p(x, y) ± , y) = +(x, y) + +(x + ∂xp(x, y) + ∂2 xp(x, y) ∂yp(x, y) + ∂2 yp(x, y) ∂x(x, y) + ∂2 x(x, y) ) = −(x, y) + ∂y(x, y) + ∂2 y(x, y) −(x, y + 1 N− 1 π+(x − N+ π−(x, y − , y) = π+(x, y) − 1 ) = π−(x, y) − N− ∂xπ(x, y) + ∂yπ(x, y) + 1 2N 2 + 1 2N 2− 1 2N 2 + 1 2N 2− 1 2N 2 + 1 2N 2− ∂2 xπ(x, y) ∂2 y(x, y) (B1) 1 N+ 1 N− 1 N+ 1 N− 1 N+ 1 N− Substituting in the Master equation, Eq. A1, and neglect- ing terms of order O(1/N 3±), we obtain, 2(cid:88) 2(cid:88) 1 ∂ ∂xi ∂2 p(x, y, t) = − 2(cid:88) i=1 j=1 ∂xi∂xj ∂ ∂t where, [vi(x, y)p(x, y, t)] + [Dij(x, y)p(x, y, t)] (B2) vx(x, y) = vy(x, y) = Dxx(x, y) = Dyy(x, y) = 1 N+ 1 N− 1 2N 2 + 1 2N 2− (π+(x, y) − +(x, y)) (π−(x, y) − −(x, y)) (π+(x, y) + +(x, y)) (π−(x, y) + −(x, y)) Dxy(x, y) = Dyx(x, y) = 0 (B3) The analysis of the FPE and the comparison of the steady state probabilities with those obtained by the nu- merical solution of the Master Equation will be presented in a separate manuscript. x(t + ∆t) = x(t) + vd∆t + η (A1) Appendix C: Average Processivity where η are drawn Gaussian distribution with (cid:104)η(t)(cid:105) = 0 and (cid:104)η(t)η(t(cid:48))(cid:105) = 2Dδ(t − t(cid:48)). Appendix B: Fokker Planck equation If the maximum number of kinesin (N+) and dynein (N−) is large, N+, N− (cid:29) 1, we can expand the proba- bilities in the Master equation in a Taylor series to ob- In Fig. 9, we look at the variation of the average proces- sivity as (a) a function of N− for fixed N+ = 9 (Fig. 9a), and (b) a function of N+ for fixed N− = 9 (Fig. 9b). In Fig. 9(a), we find that the average processivity decreases with N− for non-catchbonded dynein. Catch-bonded dynein with strong tenacity exhibits qualitatively similar behaviour as that of motor without catchbond with the processivity decreasing on increasing N− , while dyneins with weaker tenacity can stall the motion of the cargo. 10 FIG. 9. Average processivity (a) as a function of N− for N+ = 9 , and (b) as a function of N+ for N− = 9. The zero-force (un)binding rates for dynein are ε0− = π0− = 1/s, and Fo = 7pN . FIG. 10. Processivity contour plots in the N+ − N− plane for strong dynein (a) without catch bond (α = 0); and (b) with catch bonds (α = 40, Fo = 7pN ). The colorbar indicates the average processivity (in µm). Yellow regions denote strong plus ended runs, while dark blue regions indicate strong minus ended runs. The zero-force (un)binding rates for dynein are ε0− = π0− = 1/s This again arises because the catch bond is activated at smaller opposing loads for weak dynein, leads to drastic effects on the motion of the cargo. In Fig.9(b), we show that again, while strong dynein exhibits qualitatively similar behaviour to non-catch bonded dynein, weak dyneins show a counter-intuitive codependent behaviour as was seen in Fig. 2(b). As the number of kinesin motors increases initially, the cargo walks more in the negative direction, with the cargo walk- ing an average ∼ 100µm in the negative direction for 2 − 3 kinesin molecules, compared to around ∼ 40µm when no kinesins are present. Beyond 3 kinesin motors, on increasing the kinesin number, the processivity in the negative direction increases, as would be expected from the normal mechanical tug-of-war picture. looks similar to the one for weak dynein (Fig. 2c), with strong positive directed runs for N+ > 5. Strong nega- tive runs are achieved only for very high dynein number coupled to very low kinesin number. In the presence of the catch bond, dyneins are able to counteract the pos- itive directed load more efficiently, with strong positive runs occurring for higher number of kinesins than in the non-catch bonded case. The special cases corresponding to N+/N− = 4 are shown in Fig. 2(a) and (b), while the case corresponding to N+/N− = 9 are shown in Fig. 9(a) and (b). Appendix D: Average number of bound motors In Fig. 10 we show the contour plots of the proces- sivity of the cargo for strong dynein in the N+ − N− plane. In the absence of catch bond, the contour plots In Fig. 8, we look at the average processivity as the friction constant ζ is varied by varying σ. The rest of the parameters are as in fig. 2(b) of the main text with 01002003004005006007000246810AverageProcessivity(µm)N+−100−500501000246810AverageProcessivity(µm)N−Fs−=1,α=0Fs−=1,α=40Fs−=7,α=0Fs−=7,α=40Fs−=1,α=0Fs−=1,α=40Fs−=7,α=0Fs−=7,α=400246810N+0246810N−(a)-1001003005007000246810N+0246810N−(b)-100100300500700 α = 40 and Fo = 7pN . For all three values of σ, the non-monotonic behavior is reproduced indicative of the catchbonded codependent behavior. With increasing σ, the cusp is observed at higher values of N+ with average processivities close to values obtained in our mean field model. The effect of catchbonding on the processivity can also be understood in terms of the average number of bound motors. As illustrated in Fig. 11, in the absence of catch bond, the number of attached dyneins shows a very weak increase with increasing N−, saturating at a value of ∼ 0.3. The average number of kinesins is roughly around 3, leading to strong positive runs in the absence of catch bonds. For catch bonded weak dynein Fs− = 1pN , on increasing N−, the average number of bound dyneins increases sharply. For low N−, the aver- age number is almost the same as the maximum number, (cid:104)n−(cid:105) ∼ N−. The average number of attached kinesins also falls sharply to under 2. On increasing N− even fur- ther, the average number of bound dyneins keeps increas- ing, while the average number of bound kinesins roughly remains constant. The higher number of bound dyneins lead to overall minus directed runs in this regime. For strong dynein, both the increase in (cid:104)n−(cid:105) and (cid:104)n+(cid:105) are much less sharp, illustrating that catch bond plays a less drastic role here in contrast to weak dynein. The average number of bound kinesins and dyneins are comparable in this case, which effectively leads to no net motion for N− ≥ 2. The behaviour of the processivity as a function of N+ can also be understood in terms of the average number of bound motors. As shown in Fig. 11, in the absence of catch bonds, on increasing N+, the average number of dyneins fall drastically, approaching ∼ 0.1 for large val- ues of N+. In contrast, the average number of bound ki- nesins increases linearly, leading to stronger plus-end di- rected runs with increasing N+. For catch bonded weak dynein (Fs− = 1pN ), remarkably, the average number of attached dyneins increases with increasing N+. This is again a direct consequence of the catch bond, where the increasing opposing force due to more kinesin motors pushes dyneins into the catch bonded state, effectively increasing their numbers. This leads to the increased processivity of the cargo in the negative direction with increasing N+ as shown in Fig. 2(b). The effect for Fs− = 7pN dynein is much more muted, since it is diffi- cult to push these strong dyneins into the catch bonded regime. Appendix E: Catchbond Strength The strength of the dynein catch bond (α) is a phe- nomenological parameter in our model. Changing the strength of the catch bond can have dramatic conse- quences for the processivity characteristics. This is shown in Fig. 12 for three values of the catch bond strength. For α = 20kBT , on increasing the dynein 11 unbinding rate, the average processivity in the positive direction decreases monotonically. For α = 30kBT , on weakening the dynein, the processivity in the positive direction initially increases, as expected from standard tug-of-war. However, beyond a certain ε0−, weakening the dynein further, causes a net decreases in the proces- sivity in the positive direction. Finally, for α = 40kBT , on increasing ε0−, the runlength in the negative direction initially decreases, and beyond a certain point, saturates to almost zero, becoming insensitive to further changes in the unbinding rate, as has been discussed in the main text for Fig. 3(a). Appendix F: Probability distributions and sample trajectories Here, we analyse the case of bidirectional cargo trans- port by 1 kinesin motor and 1 dynein motor. In Fig. 13(a), (b) (c) we display the distribution of runtime along the negative direction, the distribution of times the cargo spends in pause state (which arises due to the si- multaneous attachment of dynein and kinesin motors to the filament leading to tug-of-war), and the distribution of runtime along the positive direction, in the absence of catchbonding in dynein. In this scenario, the frequency of the positive runs exceeds the negative runs by almost one order of magnitude as indicated by looking at the peaks of the probability distribution of the run time. This can be understood as a direct consequence of the relatively high tenacity of the kinesin with respect to dynein. Since stall force of the kinesin motor is around 5 times that of dynein, in a typical situation of tug-of-war, in the ab- sence of catchbonding, the unbinding rate of the dynein motors due to the opposing load of the kinesin motors rises far steeper than that of the unbinding rate of ki- nesin motor due to the opposing load of dynein motors. This leads to preponderance of the positive runs viz-a- viz negative runs. Further, the average runtime along the positive directions is more than the run lengths along the negative direction. This is simply a consequence of the fact that the kinesin binding rates are chosen to be higher than the dynein binding rates (see Table 1). Thus a plus-moving run, on average, continues for a longer time than a minus run, leading to larger average runtime along the positive direction. This trivially implies that the runlengths in the positive direction are also larger than those along the negative direction, the runlength being related linearly to the runtime through the forward velocity of the motor (vF + = vF− = 0.65µm/s). As seen in Fig. 13(b), in the absence of catchbond the average time the cargo spends in the paused state is an order of magnitude smaller than the time it spends in the plus moving state. Thus overall the motion of the cargo in the absence of catchbond for this case is strong plus ended motion with weak pauses and negligible runs along the negative direction ( Fig. 13d). Incorporation of the catch- bonding behaviour of dynein is demonstrated by compar- 12 FIG. 11. Average number of bound kinesins for (a) N+ = 4, (c) N+ = 9, (e) N− = 4, (g) N− = 9. Average number of bound dyneins for (b) N+ = 4, (d) N+ = 9, (f) N− = 4, (h) N− = 9. The data for both strong and weak dynein with and without catch bond. Here Fo = 7pN . ison of the probability distributions of runlengths and pauses as depicted in Fig. 13(e), (f) (g) with Fig. 13(a), (b), (c) . First of all by comparing Fig. 13(e) and 4(g) it can be seen that the frequency of the negative runs now exceed that of the positive directed runs. This is due to the fact that in the tug-of-war state, when the attached dynein experiences the load force due to the attached ki- nesin, the dynein enters a catchbonded state and thus its propensity to unbind from the filament diminishes, while that of the kinesin remains unaltered, resulting in a sit- uation where the pause state is more often transformed into a minus-end directed state of the cargo. While the characteristic pause times do not change substantially, they now become comparable to the runtime in the neg- ative direction, as can be seen by comparing Fig. 13(f) with Figs. 13(b) and (e). The average run length either along the plus or the minus end remains unaffected due to catchbonding with the average runs along the plus direc- tion being higher than that of the dynein due to higher kinesin binding rates compared to the dynein motors. The corresponding trajectories ( Fig. 13(g) ) then cor- respond to bidirectional motion, characterized by more frequency of negative runs but longer average plus ended runs, and more prominent pauses. 02468101.522.533.5<n+> (a) N+=4Fs-=1,,=0Fs-=1,,=40Fs-=7,,=0Fs-=7,,=4002468100123456<n-> (b) N+=4Fs-=1,,=0Fs-=1,,=40Fs-=7,,=0Fs-=7,,=400246810345678<n+> (c) N+=9Fs-=1,,=0Fs-=1,,=40Fs-=7,,=0Fs-=7,,=40024681002468<n-> (d) N+=9Fs-=1,,=0Fs-=1,,=40Fs-=7,,=0Fs-=7,,=4002468100246810<n+> (e) N-=4Fs-=1,,=0Fs-=1,,=40Fs-=7,,=0Fs-=7,,=40024681001234<n-> (f) N-=4Fs-=1,,=0Fs-=1,,=40Fs-=7,,=0Fs-=7,,=400246810N+0246810<n+> (g) N-=9Fs-=1,,=0Fs-=1,,=40Fs-=7,,=0Fs-=7,,=400246810N+02468<n-> (h) N-=9Fs-=1,,=0Fs-=1,,=40Fs-=7,,=0Fs-=7,,=40 13 FIG. 12. Processivity as a function of ε0− for different catch bond strengths (α) for weak dynein (Fs− = 1pN ). Data shown is for N+ = 6, N− = 2, π0− = 1/s and Fo = 7pN . FIG. 13. Probability distributions of runtime for N+ = 1, N− = 1. The top panel shows the normalized histograms and sample trajectories for dynein in the absence of catch bond (α = 0). The bottom panel shows the corresponding quantities for catch-bonded dynein (α = 40, Fo = 7pN ). (a) and (e) Distributions of runtime for minus directed runs (shown in red); (b) and (f) pausetime distributions (shown in blue); (c) and (g) distributions of runtime for plus directed runs (shown in green); and (d) and (h) sample trajectories. Insets, where present, show a magnified view of the probability distributions. [1] M. A. Welte, Current Biology 14, R525 (2004). [2] S. P. Gross, M. C. Tuma, S. W. Deacon, A. S. Serpin- skaya, A. R. Reilein, and V. I. Gelfand, The Journal of cell biology 156, 855 (2002). [3] M. A. Welte, S. P. Gross, M. Postner, S. M. Block, and E. F. Wieschaus, Cell 92, 547 (1998). [4] P. J. Hollenbeck, Front Biosci 1, d91 (1996). [5] M. Schuster, R. Lipowsky, M.-A. Assmann, P. Lenz, and G. Steinberg, Proceedings of the National Academy of 010020030040002468AverageProcessivity(µm)ε0−α=20α=30α=40012302p(t−)t−(s)(a)Minusruns012302p(t0)t0(s)(b)Pauseruns012302p(t+)t+(s)(c)Plusruns02468101205101520x(µm)t(s)(d)Sampletrajectories012302p(t−)t−(s)(e)Minusruns012302p(t0)t0(s)(f)Pauseruns012302p(t+)t+(s)(g)Plusruns−6−4−20205101520x(µm)t(s)(h)Sampletrajectories 14 Sciences 108, 3618 (2011). [6] W. O. Hancock, Nature reviews Molecular cell biology 15, 615 (2014). [7] M. J. Muller, S. Klumpp, and R. Lipowsky, Proceedings of the National Academy of Sciences 105, 4609 (2008). [8] M. J. Muller, S. Klumpp, and R. Lipowsky, Journal of P. Bassereau, and J. Prost, Proc. Natl. Acad. Sci. 101, 17096 (2004). [30] S. L. Reck-Peterson, A. Yildiz, A. P. Carter, A. Genner- ich, N. Zhang, and R. D. Vale, Cell 126, 335 (2006). [31] N. J. Carter and R. Cross, Nature 435, 308 (2005). [32] S. J. King and T. A. Schroer, Nature cell biology 2, 20 Statistical Physics 133, 1059 (2008). (2000). [9] M. J. Muller, S. Klumpp, and R. Lipowsky, Biophysical [33] H. Kojima, M. Kikumoto, H. Sakakibara, and K. Oiwa, journal 98, 2610 (2010). Journal of biological physics 28, 335 (2002). [10] D. Bhat and M. Gopalakrishnan, Physical Biology 9, [34] A. Gennerich, A. P. Carter, S. L. Reck-Peterson, and R. 046003 (2012). [11] A. G. Hendricks, E. Perlson, J. L. Ross, H. W. Schroeder III, M. Tokito, and E. L. Holzbaur, Current Biology 20, 697 (2010). D. Vale, Cell 131, 952 (2007). [35] D. T. Gillespie, Journal of computational physics 22, 403 (1976). [36] D. T. Gillespie, The journal of physical chemistry 81, [12] A. Gennerich and D. Schild, Physical biology 3, 45 2340 (1977). (2006). [13] V. Soppina, A. K. Rai, A. J. Ramaiya, P. Barak, and R. Mallik, Proceedings of the National Academy of Sciences 106, 19381 (2009). [14] M. Martin, S. J. Iyadurai, A. Gassman, J. G. Gindhart, T. S. Hays, and W. M. Saxton, Molecular biology of the cell 10, 3717 (1999). [37] W. M. Saxton, J. Hicks, L. S. Goldstein, and E. C. Raff, Cell 64, 1093 (1991). [38] D. D. Hurd and W. M. Saxton, Genetics 144, 1075 (1996). [39] J. G. Gindhart, C. J. Desai, S. Beushausen, K. Zinn, and L. S. Goldstein, The Journal of cell biology 141, 443 (1998). [15] S. Ally, A. G. Larson, K. Barlan, S. E. Rice, and V. I. [40] A. R. Chaudhary, F. Berger, C. L. Berger, and A. G. Gelfand, The Journal of cell biology 187, 1071 (2009). Hendricks, Traffic 19, 111(2018). [16] S. P. Gross, M. A. Welte, S. M. Block, and E. F. Wi- [41] T. Narayanan and A. Kumar, Physics Reports 249, 135 eschaus, The Journal of cell biology 156, 715 (2002). (1994). [17] S. E. Encalada, L. Szpankowski, C. H. Xia, and L. S. Goldstein, Cell 144, 551 (2011). [18] Y. V. Pereverzev and O. V. Prezhdo, Physical Review E 73, 050902 (2006). [19] C. Leidel, R. A. Longoria, F. M. Gutierrez, and G. T. Shubeita, Biophysical journal 103, 492 (2012). [20] S. Klumpp and R. Lipowsky, Proceedings of the National Academy of Sciences of the United States of America 102, 17284 (2005). [21] A. Kunwar, S. K. Tripathy, J. Xu, M. K. Mattson, P. Anand, R. Sigua, M. Vershinin, R. J. McKenney, C. Y. Clare, A. Mogilner, et al., Proceedings of the National Academy of Sciences 108, 18960 (2011). [22] R. Mallik, A. K. Rai, P. Barak, A. Rai, and A. Kunwar, Trends in cell biology 23, 575 (2013). [42] M. P. Nicholas, F. Berger, L. Rao, S. Brenner, C. Cho, and A. Gennerich, Proceedings of the National Academy of Sciences 112, 6371 (2015). [43] S. P. Gross, Current biology 13, R320 (2003). [44] S. P. Gross, Physical biology 1, R1 (2004). [45] S. P. Gross, Y. Guo, J. E. Martinez, and M. A. Welte, Current biology 13, 1660 (2003). [46] S. P. Gross, M. A. Welte, S. M. Block, and E. F. Wi- eschaus, The Journal of cell biology 148, 945 (2000). [47] G. T. Shubeita, S. L. Tran, J. Xu, M. Vershinin, S. Cer- melli, S. L. Cotton, M. A. Welte, and S. P. Gross, Cell 135, 1098 (2008). [48] M. A. Welte, S. Cermelli, J. Griner, A. Viera, Y. Guo, D.-H. Kim, J. G. Gindhart, and S. P. Gross, Current Biology 15, 1266 (2005). [23] A. Nair, S. Chandel, M. K. Mitra, S. Muhuri, and A. [49] R. J. McKenney, M. Vershinin, A. Kunwar, R. B. Vallee, Chaudhuri, Physical Review E 94, 032403 (2016). and S. P. Gross, Cell 141, 304 (2010). [24] S. Chakrabarti, M. Hinczewski, and D. Thirumalai, Jour- nal of Structural Biology 197, 50 (2017). [50] M. Hu and E. Holzbaur, J. Cell Biol. 202, 495 (2013). [51] F. Berger, C. Keller, S. Klumpp, and R. Lipowsky, Phys. [25] M. J. Schnitzer, K. Visscher, and S. M. Block, Nature Rev. Lett. 108, 208101 (2012) cell biology 2, 718 (2000). [26] R. Mallik, D. Petrov, S. Lex, S. King, and S. Gross, Cur- [52] M. Can Uar and R. Lipowsky, Soft Matter 13, 328 (2017) [53] R. Mallik, B. C. Carter, S. A. Lex, S. J. King, and S. P. rent Biology 15, 2075 (2005). Gross, Nature 427, 649 (2004). [27] S. Toba, T. M. Watanabe, L. Yamaguchi-Okimoto, Y. Y. Toyoshima, and H. Higuchi, Proc. Natl. Acad. Sci. 103, 5741 (2006). [54] S. Uemura, et al, Proceedings of the National Academy of Sciences 99, 5977 (2002) [55] J.R. Cooper and L. Wordeman, Curr. Opin. Cell Biol. [28] J. Beeg, S. Klumpp, R. Dimova, R. S. Gracia, E. Unger, 21, 68 (2009) and R. Lipowsky, Biophys. J 94, 532 (2008). [56] D.L. Coy, W.O. Hancock, M. Wagenbach, and J. Howard, [29] C. Leduc, O. Camp'as, K. B. Zeldovich, A. Roux, P. Jolimaitre, L. Bourel-Bonnet, B. Goud, J.-F. Joanny, Nature Cell Biol. 1, 288 (1999)
1607.05665
2
1607
2016-11-14T21:41:29
A 2-D suspension of active agents: the role of fluid mediated interactions
[ "physics.bio-ph", "cond-mat.soft" ]
Taking into account simultaneously the Vicsek short range ordering and also the far-field hydrodynamic interactions mediated by the ambient fluid, we investigate the role of long range interactions in the ordering phenomena in a quasi 2-dimensional active suspension. By studying the number fluctuations, the velocity correlation functions and cluster size distribution function, we observe that depending on the number density of swimmers and the strength of noise, the hydrodynamic interactions affect the ordering in a suspension. For a fixed value of noise, at large density of particles, long range interactions enhance the clustering in the system.
physics.bio-ph
physics
A 2-D suspension of active agents: the role of fluid mediated interactions Hojjat Behmadi,1 Zahra Fazli,2 and Ali Najafi2, 1, ∗ 2Department of Physics, Institute for Advanced Studies in Basic Sciences (IASBS), Zanjan 45137-66731, Iran. 1Physics Department, University of Zanjan, Zanjan 45371-38791, Iran. (Dated: May 21, 2018) Taking into account simultaneously the Vicsek short range ordering and also the far-field hydrody- namic interactions mediated by the ambient fluid, we investigate the role of long range interactions in the ordering phenomena in a quasi 2-dimensional active suspension. By studying the number fluctuations, the velocity correlation functions and cluster size distribution function, we observe that depending on the number density of swimmers and the strength of noise, the hydrodynamic inter- actions affect the ordering in a suspension. For a fixed value of noise, at large density of particles, long range interactions enhance the clustering in the system. PACS numbers: 47.63.Gd,05.65.+b,87.18.Gh,87.18.Nq Keywords: Active suspension, Hydrodynamic interac- tion, ordering Introduction. Active agents that can convert a non- mechanical form of energy to mechanical work, exhibit a wide range of interesting dynamical behaviors when they form a suspension [1–6]. Both at the scales of macro and micro, there are many examples of such systems that have attracted enormous interests recently. Schools of fishes and birds [7–10], bacterial suspensions [11, 12], gels of cytoplasmic polymers [13, 14], interacting active Janus particles [15] and swimmers in non newtonian fluid [16] are examples. One of the main questions that needs to be answered, is the nature of ordered phases in such systems. In this article, we concentrate on the dynamics of micron-scale active agents. A wide class of works includes numerical simulations of micro-suspensions, based on phenomeno- logical and simplified interaction terms between individ- ual particles[17–22]. The well known model of Vicsek that can correctly account the local ordering of elongated objects, built the core of such studies[23]. Such simula- tions reveal how a local ordering rule can lead the system to reach a state with large scale ordered phases [24–28]. Continuum thermodynamic description of active sus- pensions, is another line of approach that can address some macroscopic features of the systems[29–34]. Dy- namical equations derived by symmetry arguments or ob- tained from statistical averaging over microscopic forces, can capture the physics of ordered phases developed in such systems. Only a few studies have investigated the role of long-range hydrodynamic interactions (HD) be- tween the particles [35–38]. Effects of such interactions is essential, specially in the case of micro-scale exam- ples suspended in aqueous media. Fluid velocity pro- duced by a moving particle, propagates instantaneously (a property of small scale hydrodynamic) through the medium and affects the motion of other particles. Some researchers, using a simple dipolar flow interaction mech- anism, conclude that HD may prevent the emergence of long-range order [39]. The aim of this article is to im- FIG. 1. The geometry and internal structure of two interact- ing swimmers are shown. Each swimmer, with two internal degrees of freedom, has an intrinsic swimming direction de- noted by t. prove our understanding of the ordering phenomena and collective behavior in a suspension of active microscopic agents. To correctly account for such interactions, one needs to start from a hydrodynamic model that takes into account the internal structure of the swimmers. Starting from hydrodynamic interactions of a generic microscopic model, we numerically study the statistical parameters, that can reflect the nature of ordering in a suspension of such micron scale swimmers. Model. To study the dynamics of a two dimensional collection of N interacting self propelled objects, we as- sume that the interaction between swimmers can be ob- tained from two-particle interactions. Fig. 1, shows a schematic view of two swimmers that interact through both short and long range interactions. The position and the orientation of the i'th swimmer is shown by ri and ti, respectively. In addition to position and orientation, the internal structure of the swimmers is also important. Each swimmer has an internal structure, that allows it to swim. To model the internal structure of the swim- mers, we use a minimal model with two internal degrees 2 To overcome the complexity of short range hydrody- namic interactions, we approximate the short range part of the interactions by a very well known model of Vicsek interaction that is essentially a phenomenological short range interaction [23]. This interaction enforces an elon- gated object (like what we have shown by ellipsoids in Fig. 1) to change its direction according to the average orientations of its neighbors. The Vicsek model does not fully consider all features of the short rang hydrodynamic interaction, but as an approximation we neglect other de- tails of short range hydrodynamic interactions that are not included in Vicsek model. Vicsek's model mainly takes into account the steric interaction between the elon- gated objects. To simplify our study, we assume that there is a crossover length Rc that separate the short and long range forces. Two objects with separation smaller than this crossover length, interact with short range Vic- sek model and beyond this length, the long range hydro- dynamic interactions are present. Emergent collective motions of the Vicsek model are clearly known and our combined model here, will show how HD can affect such collective motions. In order to numerically study the dynamics of a sus- pension, we write the discrete dynamics of the i'th swim- mer as: ri(t + δt) = ri(t) + δt Θ(rij − Rc)Vij θi(t + δt) = θi(t) + θV i + ηξi(t) + δt Θ(rij − Rc)Ωij, j i = arg(cid:80)(cid:48) where Heaviside step function is Θ(x) = 1 for x ≥ 1 and it is 0 for x < 1. The short range contribution to the dynamics is given by: θV j eiθj (t), where the summation runs over all swimmers with rij < Rc. We assume that the fluctuations affect the dynamics of the swimmers, through a rotational noise represented by ηξi(t). ξ is random number with uniform probability in the interval [−π, π] and the strength of noise η, can take any positive value. Two swimmers scattering. Before studying the case of many swimmers, it is instructive to start with two par- ticles system. Rich behavior emerging from long range hydrodynamic interaction, promises a non trivial behav- ior for the trajectories of two interacting swimmers. A plethora of behavior, repulsive, attractive and oscillating trajectories can be observed. The details of such behavior has been studied extensively before [42, 43]. An important feature that one can learn from the two body system, is the stability of isotropic phase in a many swimer system. In an isotropic phase, all swimmers moves in random directions and no direction is preferred. Using a kinetics theory approach with two body scatter- ings, it is shown that for a dilute system, the nature of two body scattering is the essential mechanism that vti + (cid:88) (cid:88) j  , FIG. 2. Forward component of the change in total momen- tum, is plotted as a function of incoming angular separation ∆, for different values of noise strength η. At very low noise, the isotropic state is not stable. f (t) and (cid:96)i of freedom, namely the three beads connected by two arms[40, 41]. This is a generic model that can correctly explain the far field of both dipolar and quadrupolar swimmers. Denoting the lengths of front and back arms of a swimmer by (cid:96)i b(t) and the spheres radius by a, we can seek internal motions that are able to propel the swimmer. As it is verified experimentally, a simple har- monic undulating motion with a phase lag on both arm lengths, is able to propel the swimmer at low Reynolds condition [41]. For identical swimmers, we choose an in- ternal motion that is given by (cid:96)i f (t) = (cid:96) + u sin(ωt) and (cid:96)i b(t) = (cid:96)(1 + δ) + u sin(ωt + ϕi), where (cid:96) and (cid:96)(1 + δ) de- note the average arm lengths, u denotes the undulation amplitude, the frequency is shown by ω and the phase difference between the arms is denoted by ϕi. The orientation of a swimmer in a two dimensional reference frame can be represented by a single angle θi. In this case we have: ti = (cos θi, sin θi). Detail hy- drodynamic calculations (see appendix), show that the velocity of i'th swimmer, moving in the presences of j'th swimmer, can be written as [42, 43]: Vi = vti + Vij, θi = Ωij, where the intrinsic swimming velocity of the swimmer depends on the internal structure: v = v(a, u, ω, (cid:96), δ) and, the interaction terms are functions of the dis- tance rij = ri − rj and the orientation of swimmers: Vij = Vij(v, ti, tj, rij) and Ωij = Ωij(v, ti, tj, rij) (See appendix for details). The interaction terms, Vij and Ωij, obtained with this kind of modeling are valid only for very far swimmers: rij (cid:29) (cid:96). Complexity of hydrody- namic equations, does not allow us to achieve analytical results for the short range part of the interactions be- tween swimmers. 3 FIG. 3. Left: order parameter in terms of the number of swimmers. Right: density fluctuation as a function of average number is plotted for different strength of noise η. Results are compared for two cases where the hydrodynamic interaction is on or off (HD and NHD). as: ψ = 1N v(cid:80)N i=1 Vi fully polarized state (anisotropic phase) is given by ψ = 1 and the isotropic state corre- sponds to ψ = 0. Velocity autocorrelation and velocity- velocity correlation functions are defined as: N(cid:88) N (cid:104) 1 i=1 Ca(t) = Cvv(r) = 1 N (N − 1) (cid:104) Vi(0) · Vi(t) Vi(0)Vi(t)(cid:105), N(cid:88) N(cid:88) i=1 j(cid:54)=i Vi(t) · Vj(t) Vi(t)Vj(t)(cid:105). determines the stability of the isotropic state [44]. De- noting by P and δP, the initial total momentum and the change in total momentum after a binary scattering, we define the average forward component of the momentum change in a binary scattering by: µ = (cid:104)P· δP(cid:105)0. Averag- ing is done over all impact parameters and as shown in fig. 2(inset), the incoming angular separation is shown by ∆. Neglecting the self diffusion, for µ > 0, the isotropic state is unstable and the interactions will eventually lead the system to reach a polar state [44]. For µ < 0, the in- teraction between particles is not able to develop a polar state. Fig. 2, shows the forward component of averaged change in momentum as a function of incoming angular separation ∆. Here we have taken into account both the Vicsek and HD interactions as described in the previous section. As one can see from this figure, for small noises, µ is positive and it reflects the instability of the isotropic state. Following such instability and for small noises, or- dered state (anisotropis phase) emerges. In anisotropic phase, the rotational symmetric is spontaneously broken and all swimmers move in a preferred direction. Such an instability is a general feature of the Vicsek like interac- tion, and we see here that the long range hydrodynamic interaction does not affect the instability. As one can see from the figure, by increasing the noise, µ starts to have negative values that reflects the stability of isotropic state. This means that, still in the presence of HD, we expect to observe ordered (anisotropic) phase at small values of noise. In the following parts we numerically investigate the detail role of HD in the ordering of a suspension. Simulation. To investigate the role of hydrodynamic interaction in the long time behavior of a quasi 2-D sus- pension, we proceed by numerically simulating the sys- tem. Along this path we study a set of order parame- ter and correlation functions. To quantify the polar or- der of the system, we define the polar order parameter where (cid:104)···(cid:105) denotes the averaging over all particles. These two correlation functions contain information about correlation time and correlation length in fluctuat- ing system. Local spatial ordering and clustering in the system can be understood in terms of the radial distri- bution function g(r) that is defined by: g(r) = (cid:96)2 N (N − 1) (cid:104) δ(r − rij)(cid:105). N(cid:88) N(cid:88) i=1 j(cid:54)=i Defining all the required statistical parameters of our system, we will study the thermodynamic state of our system in the next section. In our numerical study, we consider a two dimensional suspension of N particles in a square box of length L with periodic boundary con- dition has been considered. To make the equations non dimensional, we use (cid:96) and v as characteristic length and velocity. In simulations, we choose a square box of size 50(cid:96) and change the particle numbers from 100 to 2000. The time step in dimensionless units is δt = 0.001 and a total number of ∼ 1.2 × 106 steps is necessary to reach steady state. Results. Swarming behavior in our model, results from interplay between hydrodynamic interactions, noise strength and number density of particles. As our main goal here is to investigate the role of hydrodynamic in- teractions, we repeat all simulations with and without 4 FIG. 4. Velocity auto-correlation function Ca(t) (a), velocity-velocity correlation function Cvv (b) and, radial distribution function g(r) (c) are plotted for a system with N = 2000 and η = 0.35. At large noises, both correlation time and correlation length and also the pairing strength are enhanced by HD interactions. hydrodynamic interaction. In the first set of simulations, we perform the simulations only with the Vicsek interac- tion then for the second set, we include the long range interactions as well. Comparison between the results of these two sets of results will provide an understanding of the role of hydrodynamic interaction. All results marked by NHD are obtained by taking into account the Vicsek interaction and the results marked by HD, denote the cases where both Vicsek and long range hydrodynamic interactions are present. Fig. 3(a) shows the polarization order parameter, ψ, as a function of number of the swimmers (for fixed box size). As we have expected from two-body scattering re- sults have been obtained at previous section, for a fixed noise, increasing the density will result instability in the isotropic phase and a stable anisotropic polarized phase will appear. Our numerical results show that the appear- ance of such polarized state with ψ (cid:54)= 0, is not sensitive to long range part of the interaction. This is consistent with the results of previous section where, as we dis- cussed, it is the short range part of the interaction that dominates in the instability mechanism for the isotropic phase. The number fluctuation, is the other quantity that we have studied in our simulations. Denoting the average number of particles by (cid:104)n(cid:105), we study its fluctu- ations: σ = (cid:104)n2(cid:105) − (cid:104)n(cid:105)2. This a statistical parameter that includes information about the non-equilibrium na- ture of a fluctuating system. The practical method which we have used to calculate the number fluctuations is as follows. For a given total number of swimmers N , we start by a small window in the middle of the simulation box and measure both average number of particles and its fluctuation inside this window. Then changing the size of this window will allow us to plot the number fluc- tuations as a function of average number. For a system that is in thermal equilibrium, we expect to see a relation like σ ∼ (cid:104)n(cid:105) 1 2 . Fig. 3(b) shows the results of numerical simulations, that as a result of particle's activity, devi- ate from equilibrium 1 2 power law [45]. The results, indi- cate that by increasing the noise strength the system will tend to approach equilibrium power law (for larger noise, the slopes are smaller). But for a large and fixed noise strength, the HD curve has a slope smaller than the NHD curve. It turns out that for this condition, the HD in- teraction diminishes the out of equilibrium nature of the system. This result critically depends on the strength of noise, our simulations shows that for smaller noise the HD does not have any critical role in the behavior of number fluctuation. Velocity auto-correlation function and velocity- velocity correlation function, are plotted in fig. 4(a) and (b). As one can see from the results, for large number density N = 2000 and large noise η = 0.35, the HD interaction increases both the correlation time and correlation length. For small noise η = 0.15 (results are given only for autocorrelation function), the increase in correlation time is very small. Fig. 4(c), shows the results for radial distribution function g(r). The strength of peak in this graph that corresponds to the pairing of the particles, strongly depends on the interaction between particles. HD interaction increases the pairing and clustering in the system. Size distribution function of clusters is another impor- tant quantity that can help us in correctly analyzing the swarming behavior in a suspension. In a many body sys- tem of active agents, clusters of different sizes intermit- tently form and break. As a results of particle exchange between different clusters, a power law distribution func- tion can be expected [25]. Intuitively, if two particles are within the alignment zone, they are considered to belong to a same cluster. Size of a cluster n, is defined as the number of particles that belong to a same cluster. Exam- ples of system snapshots, with and without HD interac- tions are shown fig. 5(up), left and right, respectively. In similar conditions, for the case where the HD interaction is on, the clusters are finely distinguishable. Denoting by P (n), the probability to have a cluster with size n, we study this function for different values of number density 5 FIG. 5. Up: two snapshots of the system for N = 1500 with (left) and without (right) hydrodynamic interactions. Different colors denote different cluster and one can distinguish that HD strongly enhances the clustering mechanism. Down: cluster size distribution function for various particle number and noise strengths. For large number of particles, having large clusters are most frequent in the case of HD. and a fixed strength of noise η = 0.35 in fig. 5(down). Having clusters with large sizes, depends strongly on the interaction and the number density. For larger densities, HD interaction increases the probability of finding large clusters, but by decreasing the density of particles, the HD may change its role. Consistence with the previous result, at small density, the hydrodynamic interactions do not show any observable effects in our simulations. In conclusion, we have numerically studied the effects of long range hydrodynamic interactions in the order- ing phenomena in an active suspension of micro-particles. We have shown that depending on the strength of noise and number density of particles, the interactions have critical effects on the number fluctuations, correlation functions and clustering phenomena. Along this work, we are studying the effects of interplay between inter- nal phases of the swimmers (here, we have assumed that all swimmer are in phase). Coherent effects observed in 6 We benefited from useful discussions with S. Ra- maswamy and R. Golestanian. We also acknowledge Farideh Khalili and Mohsen Yarifard for their helps at the early stages of this work. ∗ [email protected] [1] MC Marchetti, JF Joanny, S Ramaswamy, TB Liverpool, J Prost, Madan Rao, and R Aditi Simha. Hydrodynam- ics of soft active matter. Reviews of Modern Physics, 85(3):1143, 2013. [2] David Saintillan and Michael J Shelley. Active sus- pensions and their nonlinear models. Comptes Rendus Physique, 14(6):497–517, 2013. [3] Julia K Parrish and Leah Edelstein-Keshet. Complexity, pattern, and evolutionary trade-offs in animal aggrega- tion. Science, 284(5411):99–101, 1999. [4] Moslem Moradi and Ali Najafi. Rheological properties of a dilute suspension of self-propelled particles. EPL (Europhysics Letters), 109(2):24001, 2015. [5] Fernando Peruani, Andreas Deutsch, and Markus Bar. Nonequilibrium clustering of self-propelled rods. Physical Review E, 74(3):030904, 2006. [6] Vishwajeet Mehandia and Prabhu R Nott. The collective dynamics of self-propelled particles. Journal of Fluid Me- chanics, 595:239–264, 2008. [7] D Weihs. Hydromechanics of fish schooling. Nature, 241:290–291, 1973. [8] Julia K Parrish, Steven V Viscido, and Daniel Grunbaum. Self-organized fish schools: an examina- tion of emergent properties. The biological bulletin, 202(3):296–305, 2002. [9] SJ Hall, CS Wardle, and DN MacLennan. Predator eva- sion in a fish school: test of a model for the fountain effect. Marine biology, 91(1):143–148, 1986. [10] Ronald P Larkin and Barbara A Frase. Circular paths of birds flying near a broadcasting tower in cloud. Journal of Comparative Psychology, 102(1):90, 1988. [11] Eshel Ben-Jacob, Ofer Schochet, Adam Tenenbaum, Inon Cohen, Andras Czirok, Tamas Vicsek, et al. Generic modelling of cooperative growth patterns in bacterial colonies. Nature, 368(6466):46–49, 1994. [12] Naohiko Shimoyama, Ken Tsuyoshi Mizuguchi, Yoshinori Hayakawa, and Masaki Sano. Collective motion in a system of motile elements. Physical Review Letters, 76(20):3870, 1996. Sugawara, [13] Karsten Kruse, Jean-Fran¸cois Joanny, Frank Julicher, Jacques Prost, and Ken Sekimoto. Asters, vortices, and rotating spirals in active gels of polar filaments. Physical review letters, 92(7):078101, 2004. [14] Ha Youn Lee and Mehran Kardar. Macroscopic equa- tions for pattern formation in mixtures of microtubules and molecular motors. Physical Review E, 64(5):056113, 2001. [15] Parvin Bayati and Ali Najafi. Dynamics of two inter- acting active janus particles. The Journal of chemical physics, 144(13):134901, 2016. [16] Gaojin Li and Arezoo M. Ardekani. Collective motion of microorganisms in a viscoelastic fluid. Phys. Rev. Lett., 117:118001, Sep 2016. [17] Shawn D Ryan, Andrey Sokolov, Leonid Berlyand, and 2 [Mmntjmtjn] tj + 3 T = −3[Mmntjmtjn] rij D = − 3 2 [Mmnktjmtjntjk] rij E = −3[Mmntjmtjn] ti + 3[Mmnktjmtjntik] rij F = 3[Mmnk tjm tjn tik] (rij × ti), G = [Mmnktjmtjntik] (tj × ti) − 5[Mmnkltjmtjntjktil] (rij × ti), H = − 15 2 [Mmnkltjmtjntiktil] (rij × ti), where ti represents the director of i'th swimmer and tik stands for its k'th component. In above relations, sum- mation over indices m, n, k, l are assumed. The sym- metric and traceless tensors used at the above equations, are given by: Mij(r) = rirj − 1 Mijk(r) = 4rirj rk − (δjk ri + δik rj + δij rk), 2 δij, small systems [46], promise us to see interesting effects in suspensions. Appendix. Following the method given in references [42, 43], the intrinsic velocity of a single swimmer and it's hydrodynamic interaction with second swimmer can be written as: v = 7 24 ( a (cid:96)2 ) (1 + δ) (u2 ω) sin ϕi, Vij = α1( Ωij = β1 ( (cid:96) rij (cid:96) rij )2 T + ( 3 ) F + ( 3 ) 4 ) (cid:96) rij (cid:96) rij [α2 D + α3 E] , [β2 G + β3 H] . The above results are obtained by averaging over a com- plete period of internal motion (2π/ω). In terms of  = a by: (cid:96)2 u2ω and ϕij = ϕi − ϕj, the parameters are given a δ (cid:96) ) sin ϕj, α2 = −  4 (δ + 2) sin ϕj, [(3 + δ) sin ϕi + (3 + 2δ) (sin ϕj + sin ϕij)] , ) sin ϕj (2 − δ), a δ ( (cid:96)2 ) sin ϕj, ) [(3 − δ) sin ϕi + (3 − 2δ)(sin ϕj + sin ϕij)] . β2 = 3 8 (  (cid:96) ( α1 = 29 64  α3 = 24 β1 = − 29 64 7 48 ( β3 = and  (cid:96) Mijkl(r) = 6rirj rk rl + 1 −(δij rk rl + δklrirj + δik rj rl + δjlrirk + δjk rirl + δilrj rk). 4 (δijδkl + δikδjl + δilδjk) The above equations are valid for two swimmers that are moving inside a 2-D plane. For numerical calculations we have set all phases to ϕi = π/2 and δ = 0.1. Igor S Aranson. Correlation properties of collective mo- tion in bacterial suspensions. New journal of physics, 15(10):105021, 2013. [18] Juan P Hernandez-Ortiz, Christopher G Stoltz, and Michael D Graham. Transport and collective dynamics in suspensions of confined swimming particles. Physical Review Letters, 95(20):204501, 2005. [19] Patrick T Underhill, Juan P Hernandez-Ortiz, and Michael D Graham. Diffusion and spatial correlations in suspensions of swimming particles. Physical Review Letters, 100(24):248101, 2008. [20] Shawn D Ryan, Brian M Haines, Leonid Berlyand, Falko Ziebert, and Igor S Aranson. Viscosity of bacterial sus- pensions: Hydrodynamic interactions and self-induced noise. Physical Review E, 83(5):050904, 2011. [21] Takuji Ishikawa and TJ Pedley. Diffusion of swimming model micro-organisms in a semi-dilute suspension. Jour- nal of Fluid Mechanics, 588:437–462, 2007. [22] A Decoene, A Lorz, S Martin, B Maury, and M Tang. Simulation of self-propelled chemotactic bacteria in a stokes flow. In ESAIM: proceedings, volume 30, pages 104–123. EDP Sciences, 2010. [23] Tam´as Vicsek, Andr´as Czir´ok, Eshel Ben-Jacob, Inon Cohen, and Ofer Shochet. Novel type of phase transi- tion in a system of self-driven particles. Physical review letters, 75(6):1226, 1995. [24] Guillaume Gr´egoire and Hugues Chat´e. Onset of col- lective and cohesive motion. Physical review letters, 92(2):025702, 2004. [25] Cristi´an Huepe and Maximino Aldana. Intermittency and clustering in a system of self-driven particles. Phys- ical review letters, 92(16):168701, 2004. [26] Hugues Chat´e, Francesco Ginelli, Guillaume Gr´egoire, and Franck Raynaud. Collective motion of self-propelled particles interacting without cohesion. Physical Review E, 77(4):046113, 2008. [27] Gabriel Baglietto and Ezequiel V Albano. Finite-size scaling analysis and dynamic study of the critical be- havior of a model for the collective displacement of self- driven individuals. Physical Review E, 78(2):021125, 2008. [28] Gabriel Baglietto and Ezequiel V Albano. Nature of the order-disorder transition in the vicsek model for the col- lective motion of self-propelled particles. Physical Review E, 80(5):050103, 2009. [29] Luis H Cisneros, John O Kessler, Sujoy Ganguly, and Raymond E Goldstein. Dynamics of swimming bacte- ria: Transition to directional order at high concentration. Physical Review E, 83(6):061907, 2011. [30] David Saintillan and Michael J Shelley. Orientational order and instabilities in suspensions of self-locomoting rods. Physical review letters, 99(5):058102, 2007. [31] Christel Hohenegger and Michael J Shelley. Stability of active suspensions. Physical Review E, 81(4):046311, 2010. [32] Ganesh Subramanian and Donald L Koch. Critical bac- 7 terial concentration for the onset of collective swimming. Journal of Fluid Mechanics, 632:359–400, 2009. [33] Henricus H Wensink, Jorn Dunkel, Sebastian Heiden- reich, Knut Drescher, Raymond E Goldstein, Hartmut Lowen, and Julia M Yeomans. Meso-scale turbulence in living fluids. Proceedings of the National Academy of Sci- ences, 109(36):14308–14313, 2012. [34] Knut Drescher, Jorn Dunkel, Luis H Cisneros, Sujoy Ganguly, and Raymond E Goldstein. Fluid dynamics and noise in bacterial cell–cell and cell–surface scatter- ing. Proceedings of the National Academy of Sciences, 108(27):10940–10945, 2011. [35] Aparna Baskaran and M Cristina Marchetti. Statisti- cal mechanics and hydrodynamics of bacterial suspen- sions. Proceedings of the National Academy of Sciences, 106(37):15567–15572, 2009. [36] Andrey Sokolov, Igor S. Aranson, John O. Kessler, and Raymond E. Goldstein. Concentration dependence of the collective dynamics of swimming bacteria. Phys. Rev. Lett., 98:158102, Apr 2007. [37] Ricard Matas-Navarro, Ramin Golestanian, Tan- niemola B Liverpool, and Suzanne M Fielding. Hydrody- namic suppression of phase separation in active suspen- sions. Physical Review E, 90(3):032304, 2014. [38] Andreas Zottl and Holger Stark. Hydrodynamics deter- mines collective motion and phase behavior of active col- loids in quasi-two-dimensional confinement. Physical re- view letters, 112(11):118101, 2014. [39] Hugues Chat´e, Francesco Ginelli, Guillaume Gr´egoire, Fernando Peruani, and Franck Raynaud. Modeling col- lective motion: variations on the vicsek model. The Eu- ropean Physical Journal B, 64(3-4):451–456, 2008. [40] Ali Najafi and Ramin Golestanian. Propulsion at low reynolds number. Journal of Physics: Condensed Matter, 17(14):S1203, 2005. [41] Marco Leoni, Jurij Kotar, Bruno Bassetti, Pietro Cicuta, and Marco Cosentino Lagomarsino. A basic swimmer at low reynolds number. Soft Matter, 5(2):472–476, 2009. [42] Majid Farzin, Kiyandokht Ronasi, and Ali Najafi. Gen- eral aspects of hydrodynamic interactions between three- sphere low-reynolds-number swimmers. Physical Review E, 85(6):061914, 2012. [43] CM Pooley, GP Alexander, and JM Yeomans. Hydrody- namic interaction between two swimmers at low reynolds number. Physical review letters, 99(22):228103, 2007. [44] Khanh-Dang Nguyen Thu Lam, Michael Schindler, and Olivier Dauchot. Polar active liquids: a universal classi- fication rooted in nonconservation of momentum. Jour- nal of Statistical Mechanics: Theory and Experiment, 2015(10):P10017, 2015. [45] Sriram Ramaswamy, R Aditi Simha, and John Toner. Active nematics on a substrate: Giant number fluctu- ations and long-time tails. EPL (Europhysics Letters), 62(2):196, 2003. [46] A. Najafi and R. Golestanian. Coherent hydrodynamic coupling for stochastic swimmers. EPL (Europhysics Let- ters), 90(6):68003, 2010.
1208.5126
3
1208
2013-05-16T17:12:13
Statistical Mechanics of DNA unzipping under periodic force: Scaling behavior of hysteresis loop
[ "physics.bio-ph", "cond-mat.soft", "q-bio.BM" ]
A simple model of DNA based on two interacting polymers has been used to study the unzipping of a double stranded DNA subjected to a periodic force. We propose a dynamical transition, where without changing the physiological condition, it is possible to bring DNA from the zipped/unzipped state to a new dynamic (hysteretic) state by varying the frequency of the applied force. Our studies reveal that the area of the hystersis loop grows with the same exponents as of the isotropic spin systems. These exponents are amenable to verification in the force spectroscopic experiments.
physics.bio-ph
physics
Statistical Mechanics of DNA unzipping under periodic force: Scaling behavior of hysteresis loop Department of Physics, Banaras Hindu University, Varanasi 221 005, India Sanjay Kumar and Garima Mishra A simple model of DNA based on two interacting polymers has been used to study the unzipping of a double stranded DNA subjected to a periodic force. We propose a dynamical transition, where without changing the physiological condition, it is possible to bring DNA from the zipped/unzipped state to a new dynamic (hysteretic) state by varying the frequency of the applied force. Our studies reveal that the area of the hystersis loop grows with the same exponents as of the isotropic spin systems. These exponents are amenable to verification in the force spectroscopic experiments. PACS numbers: 05.10.-a, 87.15.H-, 82.37.Rs, 89.75.Da The mechanism involved in the separation of a double stranded DNA (dsDNA) into two single stranded DNA (ssDNA) is a prerequisite for understanding processes like replication and transcription. In vitro, opening of DNA is achieved either by increasing the temperature (85-90◦C) termed as thermal melting or by changing the pH value of the solvent (< 3 or > 9) called DNA denaturation [1]. However, such a drastic change in the physiological condition is not possible in living systems. The mecha- nism of opening of dsDNA in vivo is quite complex and is initiated by helicases, DNA and RNA polymerase, ssb proteins etc, which exert force of the order of pN and as a result DNA unwinds. It is now possible to unzip the two strands of a DNA using techniques like optical tweezers, atomic force microscopy, magnetic tweezers etc [2, 3]. Theoretical understanding of DNA unzipping is mostly based on equilibrium conditions [4 -- 7]. However, living systems are open systems and never at equilibrium. Understanding the separation of DNA in the equilibrium is one approach, but another route is to perform the analysis in a situation which closely resembles the living systems i.e. in non-equilibrium con- ditions. Moreover, helicases are ATP driven molecular motors. The periodic hydrolysis of ATP to ADP can gen- erate a continuous push and pull kind of motion which spans a wide range of length and time scales. As a di- rect consequence of these chemo-mechanical cycles, bi- ological machines act like a repetitive force generators, and it is believed that forces with periodic signatures are experienced by biomolecules in many physiological con- texts. For instance, it has been postulated that DNA- B, a ring like hexameric helicase, pushes through the DNA like a wedge and produces unidirectional motion and strand separation [8]. Active rolling model and inch- worm model are two mechanisms, which suggest that PcrA goes through cycle of pulling the ds part of the DNA and then moving on the ss part during ATP hy- drolysis [9]. Williams and Jankowsy [10] showed that viral RNA helicase NPH-II hops cyclically from the ds to the ss part of DNA and back during the ATP hydrolysis cycle. Apart from these examples, there are several stud- ies [11 -- 14], which suggest that the force acting on DNA (at the junction of the Y fork i.e. ssDNA and dsDNA) N/2 N/2+1 2 1 N−1 N f (t) CLOSE y OPEN f (t) FIG. 1. DNA in zipped and unzipped state. One end is fixed and the other end is subjected to a periodic force. is periodic in nature rather than constant. Surprisingly, in most of the studies (experiments, theories and simula- tions), the applied force or loading rate is kept constant [15], and hence the results provide a limited picture of the DNA opening. Application of periodic force would intro- duce new aspects, which are not possible in the steady force case. In DNA unzipping, the equilibrium response of the re- action coordinate (extension (y)) to the constant force is well understood in terms of simple models amenable to statistical mechanics [15 -- 17]. However, when a ds- DNA is driven by an oscillatory force, y will also oscil- late and lag behind the force due to the relaxation de- lay. This relaxation delay induces hysteresis in the force- extension (f −y) curve, which has been recently observed in simulations and experiments [2, 19, 20]. The nature of hysteresis and its dependence on the amplitude (F ) and frequency (ν) of the applied force is well studied in the context of spin systems [21 -- 24]. It is found that the area under the hysteresis loop Aloop scales as F ανβ. The values of α and β differ from system to system [24]. How- ever, for DNA unzipping, the non-equilibrium response of y to the oscillatory force remains elusive. In this Letter, we show that under a certain physiologi- cal condition, a dsDNA remains in the steady and stable (zipped or open) state for an extended period of time. Furthermore, without any change in temperature (T ) or pH of the solvent, by varying ν alone, a dsDNA may be brought from the time averaged open or zipped state to 0.15 0.20 0.25 0.30 f 0.35 0.40 1000 1000 Q 30 20 10 0 0 30 20 10 0 0 30 20 10 0 0 30 20 10 0 0 500 F=0.5 500 F=0.6 500 F=1.0 500 2 ν=4.0Ε−4 F=0.4 500 1000 F=0.5 500 1000 F=0.6 500 1000 F=1.0 500 1000 30 20 10 0 0 30 20 10 0 0 30 20 10 0 0 30 20 10 0 0 ν=1.24Ε−3 ν=1.6Ε−4 ν=8.0Ε−5 ν=4.0Ε−5 ν=2.0Ε−5 ν=1.0Ε−5 ν=5.0Ε−6 ν=1.24Ε−6 30 25 20 15 10 5 > ) f ( y < (b) (a) 40 30 > ) f ( y < 20 10 0 0 0.2 0.4 f 0.6 0.8 1 FIG. 2. hyi of DNA as a function of the cyclic force of ampli- tudes (a) 0.4 and (b) 1.0 at different ν. (a) At high ν, DNA remains in the zipped state with a small hysteresis loop. As ν decreases, the system extends from the zipped state to the open state with a bigger loop. For ν → 0, the hysteresis loop vanishes and the system approaches to the equilibrium path. (b) DNA remains in the open state at high ν and approaches the equilibrium path from below as ν → 0. a new dynamic state (hysteretic), oscillating between the zipped and unzipped states, which is dynamical in origin and vanishes in the quasi static limit [24]. We evalu- ate the scaling exponents α and β associated with Aloop, which are amenable to verification in the force spectro- scopic experiments. We also show that using the work theorem [25], it is possible to extract the equilibrium f −y curve from the non-equilibrium pathways. We consider a simple model of DNA (Fig.1), where a string of beads (complementary nucleotides) corresponds to a single strand. The beads are connected by harmonic springs. The excluded volume and and base-pairing inter- actions are modeled by the Lennard Jones potential. It was recently shown that the model captures some of the essential properties of DNA and the equilibrium force- temperature diagram is in good agreement with the two state model in the entire range of the f and T [2]. In order to study the dynamical stability of DNA under the peri- odic force f (t), we add an energy −f (t)y(t) to the total energy of the system and perform Langevin Dynamics simulation to monitor the separation y of the terminal base pairs [7, 26]. The random force Γ [7, 26], has also been super imposed on the periodic force to take account of stochastic fluctuations of the system. Here, one may fix ν and vary F or keep F constant and vary ν. The value of f increased to its maximum value F in ms steps at interval ∆f (= 0.01) and then it is taken to 0 in the same way [7]. Since we are interested in the non-equilibrium regime, we allow only n LD time steps (much below the equilibrium time) in each increment of ∆f . We keep sum of the time spent τ (= 2nms) in each force cycle constant to keep ν(= 1/τ ) constant. In the following, we keep T = 0.1 and F > 0.32 [28, 29]. In Fig.2, we plot the value of y(f ) (averaged over C =1000 cycles) with f for different values of ν. It is inter- esting to note that the average value of y(f ) for different initial conformations remains almost the same, showing ν=2.57Ε−2 F=0.4 ν=3.2Ε−3 F=0.4 1000 1000 30 20 10 0 0 30 20 10 0 0 30 20 10 0 0 30 20 10 0 0 500 F=0.5 500 F=0.6 500 F=1.0 500 C 1000 1000 1000 1000 FIG. 3. The time sequence of Q for different ν and F . that the system is in the steady state [32]. All plots show hysteresis. The area under the loop is the measure of the energy dissipated over a cycle and is defined as a dynamic order parameter [24] Aloop = I y.df, (1) which depends upon F and ν. If y(f ) is less than 5, we call the system to be in the zipped state, where as if y(f ) > 5, it is in the unzipped state [7]. At high ν, it is evident that for small F , dsDNA remains in the zipped state (Fig. 2a), whereas at high F , it is in the unzipped state (Fig. 2b), irrespective of initial confor- mations. Moreover, the path of y(f ) for the force 0 to F is different from that of the path for F to 0, which constitutes a hysteresis loop. A decrease in ν leads to a bigger path of the hysteresis loop [26]. Depending on the amplitude, the system starts from the zipped conforma- tion as shown in Fig. 2a (or open conformations shown in Fig. 2b) and then gradually approaches the open state (or the zipped state) and back to the initial state. One may note that even though f decreases from F to 0 (Fig. 2a), y(f ) increases and there is some lag, af- ter which it decreases. Recall that the relaxation time is much higher compare to the time spent at each interval of ∆f . Therefore, an increase in y(f ) with decreasing f indicates that the system gets more time to relax. As a result y(f ) approaches a path, which is close to the equilibrium. Once the system gets enough time, the lag disappears. A similar lag is expected, when the system starts from the open state at high ν. However, in this case as ν decreases, y(f ) decreases with increasing f (Fig. 2b). In both cases , whether dsDNA starts from the zipped or open state, as ν → 0, the system approaches the equi- librium f − y curve and Aloop vanishes [26]. Moreover, at high ν, Aloop also vanishes (Fig. 2a & 2b), but the system goes away from the equilibrium. The other dynamic order parameter Q = 1 τ I y(t)dt, p o o l A (2) studied in the context of magnetic systems [24], has re- cently been applied to obtain the the force-frequency di- agram of DNA hairpin [7]. In Fig. 3, we plot Q with cycles for different ν and F . The distribution shows that the path remains in the zipped state or in the open state or in the dynamic (hysteretic) state, depending on F and ν. In contrast to the DNA hairpin, which shows the co- existence of different states, a dsDNA shows a continuous transition from the zipped state to the new dynamic state as the frequency decreases. We now focus our study on the scaling of the area of the hysteresis loop. In Fig. 4a, we have plotted Aloop as a function of (F ν)0.5. For low ν, we observe that all plots for different F collapse on a straight line. This gives the value of α = 0.5 = β. At high ν, depending on the amplitude, the system remains either in the zipped state (low F ) or in the open state (high F ). In contrast to the spin system, where the average applied field is zero over a cycle, in this case, the average applied force is finite over a cycle because the two states are asymmetric. In fact, at low F , we find that Aloop scales as ν−1(F − fc)2.0±0.1, where fc is the equilibrium critical force at that temper- ature. The proposed scaling is consistent with the mean field values for a time dependent hysteretic response to periodic force in case of the isotropic spin [23] and found to be independent of length [33]. In single-molecule experiments, measurements have been done at non-equilibrium conditions. It is possible to infer the equilibrium properties of the system from these data to have a better understanding of the system. In or- der to do so, generally measurements have been taken in the quasi static limit [3] so that the techniques involved in thermodynamics can be employed. In recent years, there has been considerable work to extract equilibrium properties from the non-equilibrium data e.g. Jarzynski equality, which relates the free energy differences between two equilibrium states through non-equilibrium processes [34]. In another approach a dominant reaction pathway algorithm is developed to compute the most probable re- action pathways between two equilibrium states [35]. Here, we use the work theorem to derive the equilib- rium path between the two states [25]. Instead of repeat- ing the force cycle C times, we now randomly choose C initial conformations, which belong to equilibrium con- formations at that T and f (=0). We follow a simi- lar protocol as described above to reach the final state (f=F) from the initial state (f=0). No attempt is made to achieve equilibrium during this process. The total work performed on the system going from the zipped state i=1 yi. When the applied force decreases (backward path) from F to 0, we start with C initial conformations, which be- to open state (forward path) is wms = −∆f Pms F=0.40 F=0.50 F=0.60 F=1.00 15 10 5 0 0 0.02 0.01 ν 0.5 F 0.5 3 (a) 12 (b) p o o l A 8 4 0 0 F=0.40 F=0.45 F=0.50 F=0.55 F=0.60 F=0.65 0.1 ν −1 0.2 B(F) 0.3 0.4 FIG. 4. The scaling of the loop area of hysteresis (Aloop) with respect to ν 0.5F 0.5 in the low frequency limit (a) and with respect to ν −1B(F ) in the high frequency limit. Here, B(F ) ∼ (F − fc)2±0.1 (b). zipped state can be written as w1 = ∆f P1 long to equilibrium conformations at that T and f (=F). The work done by the system from the open state to the i=ms yi. The equilibrium distance yk for the force fk for the forward path can be obtained by assigning the weight exp(−βwk) to all forward C paths [25] at that instant k, which can be written as yk = PC PC i ykexp(−βwk) i exp(−βwk) . (3) Similarly, yk for the reverse path can also be obtained. In Fig.5a, we have plotted the simple average of ex- tension over many (C = 1000) forward paths as well as backward paths (n = 104 LD time steps). One can see the existence of hysteresis. In this figure, the solid line corresponds to the equilibrium path, which is the same, whether we start from the zipped or open confor- mation. We have used 2 × 109 time steps out of which the first 5 × 108 steps have not been taken in the aver- aging. The results are averaged over many trajectories, which are almost the same within the standard devia- tion. The weighted average of y(f ) for the forward and the backward paths obtained from Eq. 3 have also been depicted in this plot. One can see from these plots that the weighted average even for n = 104 LD steps is quite close to the equilibrium value (1.5 × 109 LD steps). We further note that the weighted average of the backward path almost overlaps with the equilibrium path. This may be because of the fact that in a reverse path, two strands of DNA are in the open state and the system can access more configurational space. This gives the higher probability of choosing rare conformations, which have dominant contributions in Eq. 3. It may be noted that the underlying assumption behind the work theorem re- lies on the fact that the initial state of the system should be in the thermal equilibrium. Whereas for the scal- ing, the system needs not be in the equilibrium, but in the steady state. Moreover, scaling involves frequency, whereas the equilibrium path obtained from the work theorem is independent of frequency. In another study, Chattopadhyay and Marenduzzo [36] 4 riodic force on DNA unzipping. We showed the ex- istence of a dynamic transition, where by varying the frequency of the applied force, a dsDNA can go from the zipped/unzipped state to a new dynamic (hysteretic) state. The area of the hysteresis loop found to scale with the same exponents as of spin systems. The scaling ex- ponents are found to be quite robust and independent of length [33] and friction coefficient [26]. We also showed that by using the work theorem, it is possible to ex- tract the equilibrium properties of the system from the non-equilibrium data, which have potential application in single molecule force measurements. At this stage, additional theoretical and numerical investigations are needed to establish a connection between dynamical tran- sition in the spin systems and a polymer under periodic force. Since, the role of hysteresis in biological processes remains unexplored territory, our work calls for further experiments on periodically driven DNA to explore such hitherto unknown dynamical phase transitions and re- lated scaling. n o i s n e t x e e g a r e v A 30 25 20 15 10 5 0 (a) 0.24 0.28 0.32 f 0.36 0.4 0.36 0.34 (b) fu fz f 0.32 Regime I Regime II 0.30 0.28 107 108 Time period 109 FIG. 5. (a) The variation of the average extension as a func- tion of force f . Dotted and dashed lines correspond to simple average of backward and forward paths, respectively. The weighted average of backward and forward paths are shown by up and down triangles, respectively. The solid line rep- resents the equilibrium f − y curve. (b) Simulated values of unzipping force (fu) and rezipping force (fz) as a func- tion of time. Regime I corresponds to the non-equilibrium, where hysteresis has been seen [19]. Regime II corresponds to the equilibrium, where equilibrium response of DNA has been studied [3, 39]. studied the dynamics of a polymer chain, whose ends are anchored. An oscillatory force was applied at the intermediate bead. They also observed hysteresis for the flexible polymer chain. However, they showed a crossover from a periodic limit cycle (hysteresis) to an aperiodic dynamics as the polymer gets stiffer. Since the unzipping experiments [3, 19] usually are performed on a long chain (few Kbps) much greater than the persistence length of DNA, their model studies [36] also imply the existence of hysteresis under periodic force [37]. Are the dynamic transition and the scaling proposed here observable in single molecule experiments? To an- swer this, in Fig. 5b, we show how the system approaches the equilibrium (regime II) from the non-equilibrium (regime I). This is in accordance with experiment fol- lowed by simulation [19]. For a two state model, time needed to cross the energy barrier ∆E(10 − 20kBT ) de- pending upon the length and sequence of DNA lies in between 4s to 15 min [38]. The equilibrium response of DNA unzipping (regime II) has been studied in experi- ments belong to this time scale [3, 39] as we obtained in our simulation, but in (µs) range. There is a mismatch of the time scale because of the coarse grained description of the model. One of the possible ways to check the feasibil- ity of the experiment from our simulation is to compare the ratio of time needed for the equilibrium (shown in Fig.2 by turquoise color) and the non-equilibrium regime (say black in Fig. 2). From our simulation, this ratio turns out to be ∼ 1000. If the experimental equilibrium time is 900 seconds [3] then the lower limit of time is 900/1000 ∼ 1 second. Hence, by manipulating the am- plitude and the frequency in the intermediate time scale (1s-15 minutes), it is possible to perform experiments where the dynamical transition may take place. In conclusion, we have studied the response of a pe- We thank S. M. Bhattacharjee, Deepak Dhar, Sriram Ramaswamy and Madan Rao for many helpful discus- sions on the subject. We acknowledge the financial sup- ports from the DST and CSIR, India. Generous com- puter supports from MPIPKS Dresden are gratefully ac- knowledged by authors. 5 [1] R. M. Wartell and A. S. Benight, Physics Reports 126, 67 (1985). [2] U. Bockelmann, B. Essevaz-Roulet, and F. Heslot, Phys. Rev. Lett. 79, 4489 (1997). [3] C. Danilowicz et al., Phys. Rev. Lett. 93, 078101 (2004). [4] S. M. Bhattacharjee, J. Phys. A 33, L423 (2000). [5] D. K. Lubensky and D. R. Nelson, Phys. Rev. Lett. 85, 1572 (2000). [6] Y. Kafri, D. Mukamel, and L. Peliti, Phys. Rev. Lett. 85, 4988 (2000). [7] D. Marenduzzo, S. M. Bhattacharjee, A. Maritan, E. Or- landini, and F. Seno, Phys. Rev. Lett. 88, 028102 (2002). 268, 319 (2007). g/mol and σ = 5.17 A, we get t ≈ 3.0 t∗ ps. However, this conversion does not work in the entire range of f and T [7]. [30] M. P. Allen and D. J. Tildesley, Computer simulations of liquids (Oxford Science, 1987). [31] W. A. Kibbe, Nucl. Acids Res. 35, W43 (2007). [32] We have analyzed 80 conformations, but in Fig. 2, only 10 conformations are shown. [33] R. K. Mishra et al. to be published. [34] C. Jarzynski, Phys. Rev. Lett. 78, 2690 (1997). [35] P. Faccioli, A. Lonardi, and H. Orland, J. Chem. Phys. 133, 045104 (2010). [36] A. K. Chattopadhyay and D. Marenduzzo, Phys. Rev. [8] I. Donmez and S. S. Patel, Nucl. Acids Res. 34, 4216 Lett. 98, 088101 (2007). [37] Even in a realistic simulation which includes persistence length of dsDNA and ssDNA as well as helical structure in its description, one would expect hysteresis curve with no major change in the time scale (emerging from large energy gap) as reduction in entropy due to persistence length will be roughly compensated by the mobility DNA. [38] S. Cocco, Eur. Phys. J. E 10, 153 (2003). [39] M. T. Woodside et al., PNAS, 103, 6190 (2006). (2006). [9] S. S. Velankar, P. Soultanas, M. S. Dillingham, H. S. Sub- ramanya, and D. B. Wigley, Cell 97, 75 (1999). [10] M. E. Fairman-Williams and E. Jankowsky, J. Mol. Biol 415, 819 (2012). [11] D. S. Johnson et al Cell 129 1299 (2007). [12] A Basu, A J Schoeffler, J M Berger and Z Bryant Nat. Struc. & Mol. Bio. 19, 538 (2012). [13] N. Fili et al. NAR 38 4448 92010). [14] P Szymczak and H Janovjak, J. Mol.Biol. 390 443 (2009). [15] S. Kumar and M. S. Li, Phys. Rep. 486, 1 (2010). [16] J. F. Marko and E. Siggia, Macromolecules 28, 8759 (1995). [17] S. Kumar et al., Phys. Rev. Lett. 98, 128101 (2007). [18] G. Mishra, D. Giri, M. S. Li, and S. Kumar, J. Chem. Phys 135, 035102 (2011). [19] K. Hatch, C. Danilowicz, V. Coljee, and M. Prentiss, Phys. Rev E. 75, 051908 (2007). [20] R. Kapri, arXiv: 1201.3709 (2012). [21] M. Rao and R. Pandit, Phys. Rev. B 43, 3373 (1991). [22] M. Rao, H. R. Krishnamurthy, and R. Pandit, Phys. Rev. B 42, 856 (1990). [23] D. Dhar and P. Thomas, J. Phys. A 25, 4967 (1992). [24] B. K. Chakrabarti and M. Acharyya, Rev. Mod. Phys. 71, 847 (1999). [25] P. Sadhukhan and S. M. Bhattacharjee, J. Phys. A. 43, 245001 (2010). [26] See supplementary material for detail. [27] G. Mishra, P. Sadhukhan, S. M. Bhattacharjee, and S. Kumar, arxiv: 1204.2913 (2012). , t = ( ǫ ǫ mσ2 )1/2t∗, r = r∗ [28] For T = 0.1, critical force fc was found to be ∼ 0.32 [2]. [29] Following relations may be used to convert dimensionless units to real units: T = kB T ∗ σ [4], where T ∗, t∗, r∗ and ǫ are temperature, time, distance and energy in real units. σ is the inter particle distance at which potential goes to zero. For example, if we set effective base pairing energy ǫ ∼ 0.1 eV, we get T ∗ m ≈ 85 0C, which corresponds to Tm = 0.23 in the reduced unit as reported in Ref [2]. It is consistent with the one obtained by OligoCalc [6] for the same sequence. Similarly by setting, average molecular weight of each bead ∼ 308 Supplementary material Here, we briefly describe the simulation details adopted in this study. The model includes a string of beads, each bead represents a base associated with sugar and phos- phate groups. We consider a sequence in such a way that the first half of the chain is complementary to the other half. This gives the possibility of the formation of a ds- dNA at low temperature [1]. The energy of the model system, we adopted, is defined as E = N −1 Xi=1 k(di,i+1 − d0)2 + N −2 N Xi=1 Xj>i+1 4( B d12 i,j − Aij d6 i,j ), (4) where N (= 32) is the number of beads. The distance between beads dij , is defined as ~ri − ~rj, where ~ri and ~rj denote the position of bead i and j, respectively. In the Hamiltonian (Eq. 4), we use dimensionless distances and energy parameters. The harmonic (first) term with spring constant k (=100) couples the adjacent beads along the chain. The remaining terms correspond to Lennard-Jones (LJ) potential.The first term of LJ po- tential takes care of the "excluded volume effect", where we set B = 1. We assign the base pairing interaction Aij = 1 for native contacts and 0 for non-native ones [2]. This choice corresponds to the Go model [3]. By na- tive, we mean that the first base forms pair with the N th (last one) base only and second base with (N − 1)th base and so on as shown in Fig.1. The parameter d0(= 1.12) corresponds to the equilibrium distance in the harmonic potential, which is close to the equilibrium position of the average L-J potential. In the Hamiltonian (Eq. 4), we use dimensionless distances and energy parameters. We obtained the dynamics by using the following Langevin equation [4, 5] m d2ri dt2 = −ζ dri dt + Fc + Γ (5) where m(= 1) and ζ(= 0.4) are the mass of a bead and friction coefficient, respectively. Here, Fc is defined as − dE dr and the random force Γ is a white noise [5] i.e., < Γ(t)Γ(t′) >= 2ζT δ(t − t′). This keeps temperature constant throughout the simulation. The equation of motion is integrated using 6th order predictor corrector algorithm with a time step δt = 0.025 [2]. ǫ , t = ( ǫ Following relations may be used to convert dimension- less units to real units: T = kB T ∗ mσ2 )1/2t∗, r = r∗ σ [4], where T ∗, t∗, r∗ and ǫ are temperature, time, dis- tance and energy in real units. σ is the inter particle distance at which potential goes to zero. For example, if we set effective base pairing energy ǫ ∼ 0.1 eV, we get T ∗ m ≈ 85 0C, which corresponds to Tm = 0.23 in the re- duced unit as reported in Ref [2]. It is consistent with the one obtained by OligoCalc [6] for the same sequence. Similarly, by setting the average molecular weight of each 6 bead ∼ 308 g/mol and σ = 5.17 A, we get t ≈ 3.0 t∗ ps. However, this conversion is not valid for the entire range of force and temperature [7]. I. DEPENDENCE OF LOOP AREA ON THE APPLIED FREQUENCY AND AMPLITUDE OF FORCE The area under the loop is the measure of the energy dissipated over a cycle. At fixed F , as ν decreases, Aloop increases (Fig. 6a). At some value of ν, Aloop is found to be maximum and then it approaches to zero. In Fig. 6b, we have depicted the plots of Aloop with F for different ν. In this case also, Aloop increases with F and above a certain amplitude, it starts decreasing. However, one can observe here that Aloop is the slow variant of F and never approaches to zero for all F . p o o l A 20 15 10 5 0 (a) (b) ν=1.24Ε−3 ν=6.2Ε−4 ν=3.12Ε−4 F=0.4 F=0.5 F=0.6 F=1.0 20 15 10 5 p o o l A 0 0.0005 ν 0.001 0.0015 0 0.0 1.0 F 2.0 3.0 FIG. 6. Variation of loop area Aloop (a) with ν at different F , (b) with F at different ν. II. DEPENDENCE OF SCALING EXPONENTS ON FRICTION COEFFICIENT ζ We have checked our results for different values of ζ. Here, we have shown results for 0.4 and 1.2. We found that the scaling exponents in high (Fig. 7) and low (Fig. 8) frequency regime, are independent of friction coeffi- cient. p o o l A 12 8 4 0 0 (a) ξ= 0.4 0.1 0.2 B(F) ν −1 (b) ξ= 1.2 12 8 p o o l A 4 F=0.40 F=0.45 F=0.50 F=0.55 F=0.60 F=0.65 F=0.40 F=0.45 F=0.50 F=0.55 F=0.60 F=0.65 0.3 0.4 0 0 0.1 0.3 0.4 0.2 ν −1 B(F) FIG. 7. Loop area scales as ν −1B(F ) at high frequencies. Here, B(F ) ∼ (F − fc)2±0.1. 7 20 15 p o o l A 10 5 0 0 (a) ξ= 0.4 F=0.40 F=0.50 F=0.60 F=1.00 0.01 ν 0.5 F 0.5 0.02 20 15 p o o l A 10 5 0 0 (b) ξ= 1.2 F=0.40 F=0.50 F=0.60 F=1.0 0.01 ν 0.5 F 0.5 0.02 FIG. 8. Loop area scales as ν 0.5F 0.5 at low frequencies. [1] S. Kumar and G. Mishra, Phys. Rev. E 78, 011907 (2008). [2] G. Mishra, D. Giri, M. S. Li, and S. Kumar, J. Chem. [5] D. Frenkel and B. Smit, Understanding molecular simula- tion (Academic Press, London, 2002). Phys 135, 035102 (2011). [3] N. Go and H. Abe, Biopolymers 20, 991 (1981). [4] M. P. Allen and D. J. Tildesley, Computer simulations of liquids (Oxford Science, 1987). [6] W. A. Kibbe, Nucl. Acids Res. 35, W43 (2007). [7] G. Mishra, P. Sadhukhan, S. M. Bhattacharjee, and S. Kumar, arxiv: 1204.2913 (2012).
1302.5730
1
1302
2013-02-22T22:36:21
Orientation Determination in Single Particle X-ray Coherent Diffraction Imaging Experiments
[ "physics.bio-ph", "cond-mat.soft", "physics.optics" ]
Single particle diffraction imaging experiments at free-electron lasers (FEL) have a great potential for structure determination of reproducible biological specimens that can not be crystallized. One of the challenges in processing the data from such an experiment is to determine correct orientation of each diffraction pattern from samples randomly injected in the FEL beam. We propose an algorithm (see also O. Yefanov et al., Photon Science - HASYLAB Annual Report 2010) that can solve this problem and can be applied to samples from tens of nanometers to microns in size, measured with sub-nanometer resolution in the presence of noise. This is achieved by the simultaneous analysis of a large number of diffraction patterns corresponding to different orientations of the particles. The algorithms efficiency is demonstrated for two biological samples, an artificial protein structure without any symmetry and a virus with icosahedral symmetry. Both structures are few tens of nanometers in size and consist of more than 100 000 non-hydrogen atoms. More than 10 000 diffraction patterns with Poisson noise were simulated and analyzed for each structure. Our simulations indicate the possibility to achieve resolution of about 3.3 {\AA} at 3 {\AA} wavelength and incoming flux of 10^{12} photons per pulse focused to 100\times 100 nm^2.
physics.bio-ph
physics
Orientation Determination in Single Particle X-ray Coherent Diffraction Imaging Experiments O.M. Yefanov∗ Deutsches Elektronen-Synchrotron (DESY), Notkestrasse 85, D-22607 Hamburg, Germany I.A. Vartanyants Deutsches Elektronen-Synchrotron (DESY), Notkestrasse 85, D-22607 Hamburg, Germany and National Research Nuclear University "MEPhI", 115409 Moscow, Russia (Dated: November 4, 2018) Abstract Single particle diffraction imaging experiments at free-electron lasers (FEL) have a great potential for structure determination of reproducible biological specimens that can not be crystallized. One of the challenges in processing the data from such an experiment is to determine correct orientation of each diffraction pattern from samples randomly injected in the FEL beam. We propose an algorithm [1] that can solve this problem and can be applied to samples from tens of nanometers to microns in size, measured with sub-nanometer resolution in the presence of noise. This is achieved by the simultaneous analysis of a large number of diffraction patterns corresponding to different orientations of the particles. The algorithms efficiency is demonstrated for two biological samples, an artificial protein structure without any symmetry and a virus with icosahedral symmetry. Both structures are few tens of nanometers in size and consist of more than 100 000 non-hydrogen atoms. More than 10 000 diffraction patterns with Poisson noise were simulated and analyzed for each structure. Our simulations indicate the possibility to achieve resolution of about 3.3 A at 3 A wavelength and incoming flux of 1012 photons per pulse focused to 100×100 nm2. 3 1 0 2 b e F 2 2 ] h p - o i b . s c i s y h p [ 1 v 0 3 7 5 . 2 0 3 1 : v i X r a ∗ [email protected]; present address: Center for Free-Electron Laser Science, Notkestrasse 85, D- 22607 Hamburg, Germany. 1 I. INTRODUCTION The problem of solving the structure of individual biological specimens to high resolution is critical for many branches of modern life- and bio-science. Two widely used techniques for high resolution structure determination are x-ray crystallography and electron microscopy. X-ray crystallography can only be used for molecules that form crystals [2], whereas trans- mission electron microscopy is limited to structures with a thickness well below one micron [3]. Therefore the majority of samples must be sliced [4] and the minimum thickness of the slices limits the resolution of this method. Single particle coherent diffraction imaging [5 -- 7] is one of the promising new techniques for the investigation of biological samples to subnanometer resolution. It has become possi- ble only recently due to the development of x-ray free-electron lasers [8 -- 11] , which produce ultra-short (less than 100 fs), coherent x-ray pulses with high intensity (more than 1012 photons in a single pulse). Short and intense pulses are required to overcome the radia- tion damage of biological particles during the pulse propagation [12 -- 14] and to produce a high number of elastically scattered photons [5, 15]. The coherence of the incident beam is important for a successful reconstruction of the electron density of the sample [16 -- 18]. However, after the pulse propagation the particles will explode, and only one projection of the sample can be measured. This problem can be overcome by injecting particles one after another with random orientations and collecting a set of diffraction patterns [6]. Each mea- sured diffraction pattern corresponds then to an unknown particle orientation. A method to determine the orientation of the particle, corresponding to each diffraction pattern, is the main subject of this paper. When the relative angular orientation of all diffraction patterns is determined the full three-dimensional (3D) intensity distribution in reciprocal space can be obtained. The structural information, or electron density of the sample is determined then by the phase retrieval [19, 20]. During the last few years there was a significant progress in the practical implementation of these ideas at hard x-ray FELs (see, for example, [21 -- 23]). There were few attempts to determine three-dimensional (3D) structure in single particle imaging experiments [24], however, the methods are still under development. Several approaches have been proposed so far to find an unknown particle orientation in these experiments. One is based on the common arc algorithm [25] originally developed for electron microscopy [26, 27]. This al- 2 gorithm exploits the fact that all two-dimensional (2D) diffraction patterns of reproducible particles in random orientations represent sections by the Ewald sphere of the 3D intensity distribution in reciprocal space. As such, all diffraction patterns have one common point, the origin of reciprocal space, and intersect along common arcs. The intensities along these arcs must be equal, and using this information the relative orientation of all diffraction pat- terns can be determined. The main problem of this method is its demand for a high signal to noise ratio, which is difficult to satisfy even with the present high power FEL sources. It was suggested to overcome this limitation by an additional classification step [25, 28], in which diffraction patterns with similar particle orientations are averaged prior to orientation determination. This step improves the statistics of each averaged diffraction pattern, but at the same time reduces its contrast. As a result, the classification step decreases the achiev- able resolution and can produce artifacts in the final stage of electron density reconstruction. Another method is based on generative topographic mapping and neural networks [29, 30]. This approach works well for a low signal to noise ratio but scales poorly with the number of resolution elements in terms of computational time and memory. The same is valid for a method based on an expectation maximization technique [31]. Here, we propose an orientation determination method based on an improved common arc algorithm [1]. Instead of a classification step we perform a simultaneous analysis of common arcs between many diffraction patterns. To improve the quality of the orientation determination, a 3D angular refinement procedure is applied at the final step. This algorithm works well even with a low photon signal down to 0.05 photons per pixel for sampling rate of three at the edge of the detector. It scales linearly with the number of resolution elements and number of measured diffraction patterns. Memory requirements are relaxed because most of the data can be processed in parts. Finally, the algorithm is highly parallelizable since most of the analysis is done between pairs of diffraction patterns. The paper is organized in the following way. In section two we describe our implemen- tation of the common arc algorithm. Section three describes our approach to treat poor signal to noise ratio data as well as orientation refinement procedure. Tests of the proposed algorithm on simulated data from two different biological structures are presented in sec- tion four. The paper is completed by the conclusion section. The details of the algorithm implementation are presented in the Appendix. 3 II. COMMON ARC ALGORITHM In a typical single particle diffraction imaging experiment, a sample with unknown orien- tation is injected into the focused coherent x-ray beam of an FEL (Fig. 1,a). The scattered radiation is measured in the far field by a 2D detector. This diffraction pattern can be mapped on an Ewald sphere [32], and represents a 2D cut of the 3D intensity distribution in reciprocal space (Fig. 1,b). Alternatively, the diffraction pattern can be considered as a perspective projection of an Ewald sphere sector onto the 2D detector plane as viewed from the sample position (Fig. 2). Our previous studies suggest [7] that, in order to increase the scattered signal, it is favorable to use longer x-ray wavelengths, since the x-ray scattering cross-sections are larger at these wavelengths. At the same time the energy of the incident x-rays should be sufficient to penetrate the sample. To achieve high resolution the detector should also cover high scattering angles. Under these conditions a large sector of an Ewald sphere is covered, which is beneficial for orientation determination. When two independent measurements of identical particles with different orientations are considered, the orientation of the first particle can be fixed as known. The orientation of the second particle can be uniquely described relative to the first one. Alternatively, two measured diffraction patterns could be considered to originate from the same particle in different experimental geometries. In this case the particle orientation is fixed, but the di- rection of the incident beam and the detector orientation are different for each measurement as shown in Fig. 2. For the first measurement the incident beam direction, given by its wavevector Ki1, can be taken along the qz axis in the reciprocal space coordinate system shown in Fig. 1,b. The direction of the incident beam for the second measurement is given by its wavevector Ki2 (Fig. 2). The relative orientation of the second geometry with respect to the first one can be described by three Euler angles φ, θ, ψ [33]. The choice of Euler angles is convenient, since rotations around the angles φ and ψ in reciprocal space are equivalent to rotations by the same angles of detectors one and two in real space, respectively. For monochromatic x-rays the Ewald sphere has the radius K = 2π/λ, where λ is the wavelength of the incident radiation. The Ewald spheres corresponding to the two mea- surements pass through the origin of the reciprocal space coordinate system (Fig. 2). The origin of the first Ewald sphere (see point A in Fig. 2), for the incident vector Ki1 is at 4 (0, 0,−K) and the origin of the second sphere (point B in Fig. 2), for the incident vector Ki2, is at (qx0, qy0, qz0) . The coordinates qx0, qy0, qz0 are determined by a rotation of the point (0, 0,−K) around the reciprocal space origin (0, 0, 0) by the Euler angles φ and θ. The intersection of the two spheres is a common arc that also passes through the origin of reciprocal space (see Fig. 3,a). This common arc is projected on the two detectors (curves a and b in Fig. 2). It is clear from this construction that the intensity along these arcs must be the same at both detectors. By analyzing the intensity correlations along all possible common arcs, the unique relative orientation of the two measurements can be determined. It should be noted that a common arc can fix the relative orientation of two patterns only for experimental geometries with large scattering angle. Otherwise, the measured sector of the Ewald sphere can be considered as flat, and the common arc reduces to a straight line. This common line fixes only the angles φ and ψ, but not the angle θ, therefore a simultaneous analysis of at least three diffraction patterns is needed in this case [27]. The projection of the common arc on the first detector (curve a in Fig. 2) can be expressed in the detector 2D coordinate system (x, y) by the following equation (see for details Appendix A) x0 − (qz0 + K)2)x2 + (q2 (q2 2qy0qx0xy + 2dqx0(qz0 + K)x + 2dqy0(qz0 + K)y = 0 , y0 − (qz0 + K)2)y2 + (1) where d is the sample-detector distance and x, y are the coordinates of the common arc projection on the first detector. Similar projection of the common arc on the second detector (curve b in Fig. 2) is also described by equation (1) by substituting x, y coordinates to x(cid:48), y(cid:48) = −y. As it follows from equation (1), the curvature of the common arcs a and b at the detectors one and two is determined only by the angle θ and sample-detector distance d. Practically, the coordinates of the projections of common arcs at the detector planes are obtained by solving equation (1) for each value of θ and d with the fixed angles φ = ψ = 0. A set of curves corresponding to the fixed value of the angle θ and all other values of angles φ and ψ is determined by rotation of the curve obtained on the previous step. This is implemented by rotation of the coordinate system (x, y) corresponding to the first detector by angle φ and the coordinate system (x(cid:48), y(cid:48)) of the second detector by an angle ψ (see Fig. 2 and Appendix B for details). 5 For each set of Euler angles and fixed sample-detector distance d the coordinates of both arcs (a and b in Fig. 2) are determined, and the intensities along these lines are compared by calculating the cross-correlation coefficient (CCC) cab(φ, θ, ψ) (cid:80) i (θ, φ)]2(cid:112)(cid:80) i (θ, φ)J b i J a i [J a (cid:112)(cid:80) i (θ, ψ) i [J b i (θ, ψ)]2 cab(φ, θ, ψ) = , (2) where J a,b i (θ, φ) = ln[I a,b i (θ, φ) along the first (a) and second (b) arc. The correct orientation of the second measurement with i (θ, φ)+1] are logarithms of the intensities I a i (θ, φ) and I b respect to the first one is given by the set of angles φB, θB, ψB that maximize the CCC in Eq. (2). To determine this orientation all three Euler angles (φ, θ, ψ) are varied sequentially with some angular step and CCCs for every orientation are calculated and compared. This procedure is applied to all diffraction patterns until their orientation relative to the first one is determined (Fig. 3,b). It should be noted here that not only orientation but also the position of a particle in space should be taken into account explicitly in the analysis. The transverse position relative to the incoming x-ray beam adds only a phase to the scattered amplitude and scales its intensity. If each diffraction pattern is properly centered this does not cause any problems in the analysis since the phase is not recorded by the detector. The intensity of each diffraction pattern can be rescaled at the stage of composing the 3D intensity distribution in reciprocal space, as described later. At the same time, the particle-detector distance d, must be taken into account explicitly, due to its strong influence on the diffraction pattern. If two measurements are performed at different sample-detector distances d1 and d2, equation (1) must be solved separately for both detectors taking into account the corresponding distances. This is especially important in real experimental conditions, when particles are injected in the beam, due to variations of the distance d from shot to shot. We also assume in our analysis that the particle size is much smaller than any variations of beam intensity. More details on our practical implementation of the common arc algorithm are presented in Appendix B. The common arc algorithm described in this section performs well for data sets with high signal to noise ratio [25]. However, it often fails in practical applications for a low number of scattered photons. One way to overcome this problem is presented in the following section. 6 III. ADVANCED ALGORITHM FOR ORIENTATION DETERMINATION IN THE PRESENCE OF NOISE In the previous section, the common arcs between one diffraction pattern (that we define as a base pattern) and all other diffraction patterns were analyzed. At the same time we should note that each diffraction pattern has a common arc with other diffraction patterns (see Fig. 3). Therefore, common arcs between all patterns could, in principle, be ana- lyzed simultaneously. Such analysis can significantly improve the fidelity of the orientation, however, its practical implementation requires an increase in computational resources. A compromise can be found by implementing the following strategy. As a first step a set of base diffraction patterns Nbase is analyzed with respect to each other to determine the cor- rect orientations of these chosen patterns. In the next step all other patterns are oriented with respect to each of these base patterns. This implementation requires Nbase times more calculations compared to a single base pattern. In the final step all intensities are mapped to a 3D array of voxels in reciprocal space by 3D gridding and averaging procedure. The benefit of this approach is the possibility to solve the orientation problem for noisy data, as will be demonstrated in the following section (see also for a detailed discussion Appendix C). In a real experimental situation all diffraction patterns have different intensities due to shot to shot intensity jitter of the FEL and the fact that each injected particle is hit by a different part of a focused beam. As a consequence all measured diffraction patterns have to be rescaled. This is implemented in the algorithm by utilizing the fact that each of two patterns have a common arc and that the intensities along this arc must be equal. The scaling factor for the intensities can be determined by taking the ratio of intensities of two diffraction patterns along the common arc. Having all information about the experimental geometry, orientation, and scaling factor for each pattern the 3D intensity distribution in reciprocal space can be constructed. The orientation determination can be significantly improved by an additional refinement that is based on the correlations between an individual pattern and the whole 3D intensity distribution. It can be implemented in the following way. First, the 3D intensity distribution is obtained from all but one selected diffraction pattern. Then the orientation of the selected pattern is varied in a small angular range and the correlation between this 2D pattern and the 7 whole 3D intensity distribution is analyzed. The orientation corresponding to the highest correlation value is considered to be the correct one. Then, the rescaled intensity of the selected pattern with the refined orientation is included in the new 3D intensity distribution in which the next diffraction pattern is excluded and the refining procedure is repeated. By applying this approach to all diffraction patterns the final 3D intensity distribution is obtained. This procedure can also be applied to identify diffraction patterns from "wrong" particles (the ones that do not belong to a set of samples under investigation). Clearly, correlation coefficients of diffraction patterns originating from these particles and the whole 3D intensity distribution will be quite low, which can be used as a criteria for rejection of these diffraction patterns from the future analysis. If the structure has a known symmetry, this can be used as an additional constraint for orientation determination [27, 34]. Using symmetry conditions each diffraction pattern can be oriented individually with respect to the selected symmetry axis. This is contrary to the structures without symmetry when at least two patterns are required for orientation determination. Applying symmetry conditions it is possible to get sufficient number of diffraction patterns for 3D representation of the scattered intensity in reciprocal space even with a limited data set or a large area of missing data due to a big beamstop. This approach was successfully used for simulated data for a sample with icosahedral symmetry discussed in the next section as well as for experimentally measured diffraction patterns of a Mimi virus obtained in coherent diffraction imaging experiment at FLASH [35]. It is interesting to note that presented implementation of the common arc algorithm also allows to determine the unknown symmetry of the object. This can be obtained by the analysis of angular orientations appearing with the highest probability. Such orientations can be found in a 3D angular map (φ, θ, ψ) of all possible orientations and reveal themselves as regions with high density (see for details Appendix C). For example, for structures with icosahedral symmetry it will correspond to 120 most likely orientations in reciprocal space that are related to the icosahedral symmetry transformation matrix. A sampling rate of at least two in each direction in diffraction pattern is required for a successful implementation of the algorithm described here. The same requirement is valid for the phase retrieval algorithms applied for the reconstruction of electron density of the samples. A higher sampling rate is beneficial for orientation determination because each speckle consists of more pixels. At the same time data with a lower sampling rate have 8 a higher signal in each pixel, which could become important for orientation determination of data with a low signal to noise ratio [36]. By testing different sampling conditions we found that an optimal sampling rate is in the region from two to three. In practice, to increase signal the experimental data could be binned for orientation determination, while the reconstruction is performed on the original unbinned data set. Applying on a final step phase retrieval algorithms [19, 20] to the 3D data set of the intensity distribution, the electron density of the sample can be obtained. IV. NUMERICAL TEST OF THE ALGORITHM The algorithm was tested with two different biological structures. The first one was an artificial protein structure without any symmetry combined from the 2BTV and 8RUC macromolecular structures [37] (see Fig. 4,a). It has a size of 13 × 19 × 28 nm3 and consists of about 124 000 non-hydrogen atoms. The second one was a human adenovirus penton base 2 12 chimera 2c6s [37]. It has icosahedral symmetry with the diameter of 27 nm and consists of about 200 000 non-hydrogen atoms (Fig. 5,a). Diffraction patterns for both structures where calculated [38] at 3 A wavelength, with a detector size of 100 mm and a sample-detector distance of 50 mm, providing the maximum scattering angle of 45◦. The achievable resolution in this geometry was 3.92 A at the detector edge and 3.3 A at its corner. The number of detector pixels was 512×512 for the first sample (providing minimum sampling rate of 2.5) and 360 × 360 for the second (with a sampling rate of two). The incoming flux was 1012 photons focused uniformly on a 100 × 100 nm2, and Poison noise was added to each diffraction pattern. The average flux at the edge of the detector was 0.05 and 0.15 photons per pixel corresponding to 0.45 and 0.6 photons per Shannon angle for the first and second structure, respectively. For the first structure 36×36× 18 =23 328 patterns were simulated with a 10◦ increment for each Euler angle. For the second structure 12 000 randomly oriented patterns were simulated. A beamstop with diameter of about 2 mm covering 1.5 speckles was introduced in all simulated diffraction patterns, and this region was excluded from the calculation of correlation coefficients. Typical diffraction patterns for a single FEL pulse simulated in the experimental conditions described above are shown in Fig. 4,b and Fig. 5,b for the first and second structure, respectively. The correct orientation of each diffraction pattern was determined using the algorithm 9 described in the previous sections. This allowed us to obtain the full 3D intensity distribution in reciprocal space for each sample. A central slice through this distribution constructed from the oriented diffraction patterns corresponding to the first structure is presented in Figure 4,d. For comparison, the same slice through the 3D intensity distribution obtained from the known orientation of each diffraction pattern is shown in Fig. 4,c. It is well seen that the slice obtained as a result of orientation determination well reproduces all features of an "ideal" intensity distribution, small deviations can be attributed to angular misalignment. This misalignment between the angles obtained from the common arc algorithm and the correct angles for the first structure is presented as a plot in Fig. 6. The accuracy of the orientation determination correlates strongly with the angular step size for the Euler angles (φ, θ, ψ). Clearly, a finer angular step size requires more computational time that scales as a third power of the step size. It is well seen in Fig. 6 that a three degree angular step being five times slower still gives higher accuracy in orientation determination comparing to a five degree step. An additional improvement in the angular determination is obtained by the final orientational refinement of each diffraction pattern with respect to 3D intensity distribution, as described in the previous section (see Fig. 6). It is interesting to observe how the signal is increased by the number of diffraction patterns used in the analysis. In Fig. 5,d a central slice through the 3D array representing the number of patterns contributing to each voxel of the constructed 3D intensity distribution is presented. It can be seen from this figure that at least 100 patterns from the analyzed 12 000 contribute to each voxel inside a resolution ring of 4 A. One more intriguing feature can be observed in this figure. Though the initial diffraction patterns were simulated up to 3.92 A resolution at the edge of the detector the 3D intensity distribution obtained from the algorithm has distinguishable features up to 3.3 A resolution (see dark outer ring in Fig. 5,d). This additional signal comes from the corners of the diffraction patterns. V. SUMMARY In summary, we proposed a method for the angular orientation determination in single particle coherent imaging experiments based on the common arc algorithm. We obtained a significant improvement of this approach by introducing a simultaneous analysis of the common arcs for several diffraction patterns. This gives the possibility to apply the method 10 to data with a low level of signal to noise ratio as well as to skip classification step which can reduce achievable resolution. Additionally, we proposed an orientational refinement of diffraction patterns that can improve the quality of the final 3D intensity distribution in reciprocal space. The algorithm proposed here has several advantages compared to other approaches [29, 31]. It scales linearly with the number of measured patterns and total number of pixels in the diffraction patterns. The algorithm is easy to parallelize, because most of the cross- correlation analysis is performed between pairs of independent diffraction patterns. It has minimum memory requirements, because the data can be processed in parts. We foresee that this approach has the potential to be the key for the success in the analysis of single particle diffraction imaging experiments and will allow to reach sub-nanometer resolution in three-dimensional imaging of biological specimens. ACKNOWLEDGMENTS We are thankful to E. Weckert for a permanent interest and support of this project, as well as for the use of program moltrans for simulation of diffraction patterns from biological structures, and to A. Singer and U. Lorenz for a careful reading of the manuscript. Part of this work was supported by BMBF Proposal 05K10CHG ''Coherent Diffraction Imaging and Scattering of Ultrashort Coherent Pulses with Matter'' in the framework of the German- Russian collaboration ''Development and Use of Accelerator-Based Photon Sources'' and the Virtual Institute VH-VI-403 of the Helmholtz association. Appendix A: Derivation of the main equations Equation (1) was derived using the following considerations. Both intersecting Ewald spheres (Fig. 7) have radii equal to the wave vector K = 2π/λ (Ki2 = Kf2 = K 2), where λ is wavelength of the incident radiation. Therefore the coordinates (qx, qy, qz) of the intersection curve must satisfy the equations q2 x + q2 y + (qz + K)2 = K 2, (qx − qx0)2 + (qy − qy0)2 + (qz − qz0)2 = K 2. (A1) (A2) 11 The center of the second Ewald sphere (qx0, qy0, qz0) lies on the distance K from the center of reciprocal space (Ki22 = K 2), therefore q2 x0 + q2 y0 + q2 z0 = K 2. (A3) From equations (A1 - A3) the formula describing intersection of two Ewald spheres can be derived: (q2 y0 + q2 x0)q2 y + 2qy0qz0(qz − K)qy + (q2 z0 + q2 x0)q2 z + K 2(q2 z0 − q2 x0) − 2Kq2 z0qz = 0. (A4) As soon as the diffracted vector Kf 1 (Fig. 7) has the same direction in both real and reciprocal spaces (angles coincide) the following relation between coordinates of a pixel on the detector (x, y, z) in real space and corresponding coordinates of the end of Kf (qx, qy, qz) in reciprocal space can be written: (A5) As soon as the distance from the sample to the detector (d) is fixed, z ≡ d. Equation (1) = = . x qx y qy z qz can be easily derived from equations (A4, A5). Appendix B: Common arc algorithm Due to the properties of Euler angles, angle φ (0 ≤ φ < 2π) can be attributed to the rotation of the reciprocal space coordinate system around the incident beam (Ki1), angle θ (0 ≤ θ < π) corresponds to the rotation around the new position of vector qy and angle ψ (0 ≤ ψ < 2π) is the final rotation of the coordinate system around the new position of the vector qz - vector Ki2 in Fig. 2. In practice, the curvature of the common arcs a and b in Fig. 2 is determined only by the angle θ and distance d. The coordinates of the projections of common arcs on detector planes can be obtained for each value of θ and d, with the fixed angles φ = ψ = 0, by solving equation (1). Other curves (for all values of angles φ and ψ) at the fixed value of angle θ are determined by rotation of the curve obtained at the previous step. Coordinate system (x, y), corresponding to the first detector, is rotated by an angle φ and the coordinate system (x(cid:48), y(cid:48)), corresponding to the second detector, by an angle ψ (Fig. 2). The common arc algorithm described in this paper was implemented using the following scheme (supplementary Fig. 8). Calculation starts with fixed angles φ = 0 and ψ = 0. Then 12 coordinates qx0, qy0, qz0 are calculated by Euler rotation on θ angle of the vector Ki1 (point (0, 0,−K)). After this equation (1) is solved and coordinates (x,y) for the common arc are found. This curve is rotated on angle φ for the first pattern (each pair of (x, y) is multiplied by corresponding rotation matrix) and on the angle ψ for the second pattern (with exchange y → −y). After full determination of the curves for both patterns corresponding values of intensities (along the curve) are extracted using the interpolation described below. The intensities along the curves are then correlated. This process continues for all angles ψ in the region 0 ≤ ψ < 2π and for all angles φ in the region 0 ≤ φ < 2π. Then the whole process is repeated for different θ value. The common arcs approach has difficulties when angle θ is large. In this case the in- tersection between two spheres reduces to a closed circle. When angle θ approaches π this circle shrinks to a point at the origin of reciprocal space coordinates (0, 0, 0). Therefore at large θ (close to π) the projection of the common arc to the detector plane, described by the equation (1), degenerates to an ellipse. Therefore the length along an arc and a circle can be different. Moreover the correlation coefficients found for the arcs with different curvature can hardly be compared, because the ends of such curves correspond to different q-range and so some curves will have good signal at the ends and some - mostly noise. From this considerations it is clear that it is difficult to compare curves obtained for small and big θ angles. To solve this problem we limited the range of acceptable θ angles to the range 0 ≤ θ ≤ π/2. To cover the range of angles π/2 < θ < π we used the fact that reciprocal space is centro-symmetric for scattering on non-absorbing objects (Friedel's law). To take this into account for angles π/2 < θ < π we invert direction of the vector Ki2 (Fig. 2) to its opposite −Ki2, that corresponds to the following transformation: rotation of φ by π (φ → φ + π), rotation of θ by θ → π − θ and final rotation of ψ by π (ψ → ψ + π). To finish inversion transformation we change y → −y (equation 1) in detector plane. To find the intensities corresponding to each point of the curve described by equation (1), some sort of gridding must be performed. In our calculations we used interpolation in the form of an average of four nearest neighbor pixels. We checked also other schemes of interpolation: nearest neighbor and bilinear interpolation [39]. The first one is faster but lacks accuracy, the second is computationally much slower without noticeable increase in quality. All common arcs for different θ values (different curvature) should cover the same q- 13 distance in reciprocal space. Therefore the step between points on the curve should remain constant. For this reason equation (1) was differentiated analytically and starting from the center of detector (x = y = 0) each next point on the curve is calculated keeping the distance (dx2 + dy2) = const. The accuracy of calculations of cross-correlation coefficient (equation 2) for all patterns can be increased by replacing intensities I(q) with their logarithms, more precisely by ln(1 + I(q)). Also for noisy data the following form of CCC is more beneficial: (cid:80) (J a(qi)− < J a(qi) >)(J b(qi)− < J b(qi) >) (cid:112)(cid:80) (J a(qi)− < J a(qi) >)2(cid:112)(cid:80) (J b(qi)− < J b(qi) >)2 , cab(φ, θ, ψ) = (B1) where J(q) = ln(1 + I(q)) and < J(qi) > is a radial averaged value of intensity corre- sponding to a ring with the radius qi on a diffraction pattern. Appendix C: Advanced algorithm for processing noisy data Fig. 9 shows typical correlations, calculated with equation (2), between two noisy diffrac- tion patterns for different angles φ with fixed angles θ and ψ (bottom curve). For comparison, correlation coefficients corresponding to a perfect data set are plotted in the same Fig. 9 (top curve). The best correlation coefficient between two noisy patterns can correspond to completely wrong orientation (like point B in Fig. 9). Therefore several orientations (Nangl) corresponding to the best set of correlations must be stored. To avoid storage of almost identical angles, some tolerance Atol in the best orientation angles determination is neces- sary. Only one angle in the range ±Atol with the highest correlation coefficient is stored. This leads to storage of only one angle per tolerance region (rectangles in Fig. 9). Therefore only one point per marked rectangle in Fig. 9 is stored in the list of "best" angles (for example in Fig. 9 Nangl = 10 and Atol = 5◦). The algorithm of orientation determination is following. In the first step all base patterns are correlated to each other using the algorithm described in the section II. As a result, Nangl "best" angles between each pattern and all other Nbase − 1 base patterns are stored. Therefore each pattern has (Nbase − 1)Nangl "best" angles with respect to all other base patterns. Then all these angles are recalculated with respect to one selected pattern (which attributed all angles equal to 0) - we shall call it the zeroth pattern. This is done in the following way: each pattern's base angles with respect to other bases are recalculated to 14 angles with respect to zeroth pattern taking into account that each base has Nangl angles with respect to zeroth. In this way (Nbase − 2)N 2 respect to zeroth are determined. angl + Nangl angles for each pattern with Let's explain this step on an example. Consider one of the base patterns Pi. This pattern has Nangl angles with respect to the zeroth pattern P0. The pattern Pi has also Nangl angles with respect to the first pattern P1. But the first pattern P1 itself has Nangl angles with respect to the P0. So, pattern Pi has already Nangl + N 2 angl angles to the P0. Then it also has N 2 angl angles to the P0 through the second pattern P2. This process is continued for all base patterns. Finally all (Nbase− 2)N 2 angl + Nangl angles determined for one base pattern (Pi) are plotted in 3D space of angles (φ, θ, ψ) and the angular region with the maximum density of points is selected as the best angle. In practice it is done in the following way: a number of points (in the tolerance region ±Atol for each Euler's angle) near each point is calculated. The point with the biggest number of neighbors is considered to be the best estimate. Then the correct angle is determined by averaging of all positions of the neighbors. In Fig. 10 an example of such operation in 2D case (for angles θ and φ with fixed ψ) is presented. The best angle corresponds to the middle of the rectangle in Fig. 10. As soon as all correct angles for the base patterns are determined, all other patterns can be oriented with respect to known bases. At this step only NbaseNangl angles need to be considered for each pattern under analysis in the algorithm described above. For better orientation determination the base patterns should be selected carefully among all measured. All base patterns should have different orientations, more precisely different θ angles. Because initially all angles are unknown this requirement can be satisfied by the analysis of 2D correlations between different patterns. This is performed by calculating 2D correlation between pairs of patterns for different angles φ and the best correlation coefficient is stored. Then patterns with the worst 2D cross-correlations between each other are selected as bases. Also for experimental data analysis, base patterns should be selected according to the recorded signal quality. The transformation to a Cartesian coordinate system in reciprocal space is performed in the following way. The whole reciprocal space is divided into elementary set of 3D voxels with the size corresponding to the pixel size of the detector. Each voxel could contain few values of the measured intensities that effectively increase signal in the 3D intensity intensity distribution (see Fig. 5,d). All intensities in each voxels are averaged and the full 3D dataset 15 is obtained. Orientation determination problem for the symmetrical structure (2c6s) without intro- ducing the symmetry is more difficult. This is due to the fact that instead of one dense spot in Fig. 10 there will be 60 (for structure with icosahedral symmetry) less dense spots in 3D angular space. So it is harder for algorithm to select the right orientation. The accuracy (or time) of orientations determination can be greatly improved if the symmetry of the sample is known. But we want to underline the important feature of the algorithm, that it can be used even for symmetrical data with unknown symmetry and also for the structures with pseudo-symmetry. There is one more important issue for speed and accuracy optimization while processing low flux data with noise. For the initial orientations determination there is no need to process high-Q region of diffraction pattern where the radially averaged photon counts is less then approximately 1-2 photons per Shannon pixel (a pixel with sampling rate equal to one). The region with smaller photon counts just lowers the accuracy of orientation determination based on common arcs. Of course this argument cannot be applied to objects with highly anisotropic scattering in different directions, like, for example, pyramids [40]. At the same time the final 3D reciprocal space and 3D angular refinement can be performed for the full datasets thus the whole procedure does not lower the resolution. The described in this chapter parameters for the two analyzed structures were: Nbase = 64 base patterns, angular step for each Euler's angle was 3◦, Atol = 5◦ and Nangl = 30. The initial orientation determination of the simulated diffraction patterns for both model structures was performed for the circular low-Q region, 150 pixels in diameter, whereas the 3D refinement for 512 (8RUC) and 360 (2BTV) pixels with angular step of 0.5◦ in the range of 5◦ near the previously found position. The initial orientation determination of 23 328 patterns with a 3◦ angular step and Nbase = 64 base patterns took about one week on a single 8-core computer. The refinement of the found orientation took about one day. [1] O. Yefanov, I. Vartanyants, and E. Weckert, Photon Science - HASYLAB Annual Report 2010, http://hasylab.desy.de/annual report/files/2010/20101462.pdf. [2] J. Drenth, Principles of Protein X-Ray Crystallography (Springer Science, 2007). 16 [3] M. Beck, V. Luccic, F. Foerster, W. Baumeister, and O. Medalia1, Nature 449, 611 (2007). [4] J. Ayache, L. Beaunier, J. Boumendil, G. Ehret, and D. Laub, Sample Preparation Handbook for Transmition Electron Microscopy (Springer Science, 2010). [5] R. Neutze, R. Wouts, D. van der Spoel, E. Weckert, and J. Hajdu, Nature 406, 752 (2000). [6] K. J. Gaffney and H. N. Chapman, Science 316, 1444 (2007). [7] A. P. Mancuso, O. M. Yefanov, and I. A. Vartanyants, J. Biotechnology 149, 229 (2010). [8] W. Ackermann et al., Nature Photonics 1, 336 (2007). [9] P. Emma et al., Nature Photonics 4, 641 (2010). [10] T. Ishikawa et al., Nature Photonics 6, 540 (2012). [11] M. Altarelli et al., Technical Design Report DESY, XFEL The European X-ray Free-Electron Laser., Tech. Rep. (XFEL, DESY, 2006) http://xfel.desy.de/technical information/tdr/tdr. [12] M. Howells et al., J. Elect. Spectrosc. and Relat. Phenom. 170, 4 (2009). [13] H. M. Quiney and K. A. Nugent, Nature Physics 7, 142 (2011). [14] U. Lorenz, N. M. Kabachnik, E. Weckert, and I. A. Vartanyants, Phys. Rev. E 86, 051911 (2012). [15] M. Bergh, G. Huldt, N. Timneanu, F. R. N. C. Maia, and J. Hajdu, Q Rev Biophys. 41, 181 (2008). [16] I. A. Vartanyants and I. K. Robinson, J. Phys.: Condens. Matter 13, 10593 (2001). [17] I. A. Vartanyants and I. K. Robinson, Journal of Synchrotron Radiation 10, 409 (2003). [18] G. J. Williams, H. M. Quiney, A. G. Peele, and K. A. Nugent, Phys. Rev. B 75, 104102 (2007). [19] J. R. Fienup, Appl. Opt. 21, 2758 (1982). [20] S. Marchesini, Rev. Sci. Instrum. 78, 011301 (2007). [21] M. M. Seibert et al., Nature 470, 78 (2011). [22] S. Kassemeyer et al., Optics Express 20, 4149 (2012). [23] N. D. Loh et al., Nature (London) 486, 513 (2012). [24] N. D. Loh, M. J. Bogan, V. Elser, A. Barty, S. Boutet, S. Bajt, J. Hajdu, T. Ekeberg, F. R. N. C. Maia, J. Schulz, M. M. Seibert, B. Iwan, N. Timneanu, S. Marchesini, I. Schlichting, R. L. Shoeman, L. Lomb, M. Frank, M. Liang, and H. N. Chapman, Phys. Rev. Lett. 104, 225501 (2010). [25] G. Huldt, A. Szoke, and J. Hajdu, J.Struct.Biol. 144, 219 (2003). 17 [26] B. Vainshtein and A. Goncharov, Proc. llth Intern. Congr. on Elec. Mirco. , 459 (1986). [27] M. van Heel, Ultramicroscopy 21, 111 (1987). [28] G. Bortel, G. Faigel, and M. Tegze, J.Struct.Biol. 166, 226 (2009). [29] R. Fung, V. Shneerson, D. Saldin, and A. Ourmazd, Nature Physics 5, 64 (2009). [30] P. Schwander, R. Fung, G. Phillips, and A. Ourmazd, New Journal of Physics 12, 134 (2010). [31] N.-T. D. Loh and V. Elser, Phys.Rev.E 80, 026705 (2009). [32] J. Als-Nielsen and D. McMorrow, Elements ao Modern X-Ray Physics (Willey, 2001). [33] L. D. Landau and E. M. Lifshitz, Mechanics (3rd ed.) (Oxford: Butterworth-Heinemann, 1996). [34] B. K. Vainstein and A. B. Goncharov, Dokl. Akad. Nauk SSSR (in Russian) 287, 1131 (1986). [35] O. Yefanov, R. Dronyak, A. Barty, H. Chapman, J. Hajdu, and I. Vartanyants, (2011), (unpublished). [36] A. Mancuso, A. Schropp, B. Reime, et al., Phys. Rev. Lett. 102, 035502 (2009). [37] RCSB Protein Data Bank, http://www.rcsb.org. [38] All diffraction patterns were simulated using the program moltrans. [39] Wolfram Research, http://www.wolfram.com. [40] O. M. Yefanov, A. V. Zozulya, I. A. Vartanyants, J. Stangl, C. Mocuta, T. H. Metzger, G. Bauer, T. Boeck, and M. Schmidbauer, Applied Physics Letters 94, 123104+ (2009). 18 FIG. 1. (Color online) Schematic view of the experimental geometry. (a) In real space, diffraction pattern from a sample in random orientation is measured by a single FEL pulse. (b) In reciprocal space the measured diffraction pattern correspond to a cut of the 3D intensity distribution by an Ewald sphere sector. Vectors Ki and Kf denote the incident and diffracted wavevectors. FIG. 2. (Color online) Measurements of two reproducible samples at random orientation can be considered as two measurements of the same sample with two different incident beam directions indicated by vectors Ki1 and Ki2. Angles φ, θ, ψ are Euler's rotation angles. Points A and B are the centers of the corresponding Ewald's spheres. Coordinates on the first and second detector are indicated as x, y and x(cid:48), y(cid:48), respectively. 19 FIG. 3. (Color online) Few Ewald sphere sectors intersecting the 3D intensity distribution of the sample in reciprocal space. Yellow lines indicate common arcs between different patterns. FIG. 4. (Color online) Artificial protein structure without any symmetry combined from the 2BTV and 8RUC macromolecular structures. (a) Iso-surface of the electron density, (b) single diffraction pattern (edge resolution 3.92A), (c,d) 2D central cuts (edge resolution 3.3A) through the constructed 3D intensity distribution in reciprocal space for the patterns with a known orientation (c), and the patterns with the orientations determined using the proposed algorithm (d). All diffraction patterns are presented in logarithmic scale. 20 FIG. 5. (Color online) Human adenovirus penton base 2 12 chimera 2c6s structure with the icosahedral symmetry. (a),(b),(c) The same as in Fig. 4, (d) number of diffraction patterns contributing to each voxel of (c) (see section III for details). All diffraction patterns are presented in logarithmic scale. FIG. 6. (Color online) Distribution of the angular error of the determined orientations for the structure without symmetry. Blue line corresponds to 5◦ angular step, green line 3◦, and red line 3◦ after the 3D refinement (see text for details). 21 FIG. 7. (Color online) Schematic view of the intersection of two Ewald spheres. FIG. 8. Flowchart for efficient data analysis with common arc algorithm. 22 FIG. 9. (Color online) Correlation coefficient between two patterns for different φ angles. Upper curve for ideal data, the lower one for the noisy data. FIG. 10. (Color online) Two-dimensional (φ, θ) distribution of angles corresponding to good corre- lation between a pattern and all base patterns. A blue box correspond to the region with highest density of good orientations. 23
1004.4327
3
1004
2011-03-15T12:14:18
Stochastic kinetics of a single headed motor protein: dwell time distribution of KIF1A
[ "physics.bio-ph", "cond-mat.stat-mech", "q-bio.SC" ]
KIF1A, a processive single headed kinesin superfamily motor, hydrolyzes Adenosine triphosphate (ATP) to move along a filamentous track called microtubule. The stochastic movement of KIF1A on the track is characterized by an alternating sequence of pause and translocation. The sum of the durations of pause and the following translocation defines the dwell time. Using the NOSC model (Nishinari et. al. PRL, {\bf 95}, 118101 (2005)) of individual KIF1A, we systematically derive an analytical expression for the dwell time distribution. More detailed information is contained in the probability densities of the "conditional dwell times" $\tau_{\pm\pm}$ in between two consecutive steps each of which could be forward (+) or backward (-). We calculate the probability densities $\Xi_{\pm\pm}$ of these four conditional dwell times. However, for the convenience of comparison with experimental data, we also present the two distributions $\Xi_{\pm}^{*}$ of the times of dwell before a forward (+) and a backward (-) step. In principle, our theoretical prediction can be tested by carrying out single-molecule experiments with adequate spatio-temporal resolution.
physics.bio-ph
physics
Stochastic kinetics of a single headed motor protein: dwell time distribution of KIF1A Ashok Garai1 and Debashish Chowdhury1 1Department of Physics, Indian Institute of Technology, Kanpur 208016, India. KIF1A, a processive single headed kinesin superfamily motor, hydrolyzes Adenosine triphosphate (ATP) to move along a filamentous track called microtubule. The stochastic movement of KIF1A on the track is characterized by an alternating sequence of pause and translocation. The sum of the durations of pause and the following translocation defines the dwell time. Using the NOSC model (Nishinari et. al. PRL, 95, 118101 (2005)) of individual KIF1A, we systematically derive an analytical expression for the dwell time distribution. More detailed information is contained in the probability densities of the "conditional dwell times" τ±± in between two consecutive steps each of which could be forward (+) or backward (-). We calculate the probability densities Ξ±± of these four conditional dwell times. However, for the convenience of comparison with experimental data, we also present the two distributions Ξ∗ ± of the times of dwell before a forward (+) and a backward (-) step. In principle, our theoretical prediction can be tested by carrying out single-molecule experiments with adequate spatio-temporal resolution. PACS numbers: 87.16.ad, 87.16.Nn, 87.10.Mn I. INTRODUCTION Molecular motors are nano-devices which perform me- chanical work by converting part of the input energy; for the motors of our interest in this paper, the input is de- rived from the hydrolysis of ATP molecules [1]. In reality, these motors are also enzymes that hydrolyze ATP and utilize the input chemical energy to perform mechanical work. In this paper we specifically consider the mem- bers of a particular superfamily of motors, called Kinesin, which are involved in intracellular transport processes in living cells. This family is designated as KIF1A and the members of this family move along filamentous tracks called microtubule (MT) [2]. One unique feature of a KIF1A is that, at least under the conditions of in-vitro experiments, it functions as a single-headed motor. The average properties, e.g., the average velocity, of these motors have been calculated analytically by using a theoretical model developed by Nishinari, Okada, Schadschneider and Chowdhury (from now onwards, referred to as the NOSC model) [3, 4] which is an extension of the general approach pioneered by Fisher and Kolomeisky [5]. In single molecule experiments, individual motor pro- teins are observed to move in an alternating sequence of pause and translocation. The sum of the pause at a bind- ing site and the subsequent translocation can be defined as the corresponding "dwell time". Because of the in- trinsic irreversibility of mechano-chemical kinetics of the system, the inverse of the mean dwell time is the aver- age velocity of a motor. The dwell time distribution g(t) contains more detailed information on the stochastic ki- netics of a motor than that revealed its average velocity. For example, the randomness parameter < t2 > − < t >2 r = < t >2 (1) provides an estimate of the lower bound on the number of rate-limiting kinetic steps in each cycle of the motor [6]. For some other motors, which move on nucleic acid strand, the analytical forms of the distributions of the dwell times have been reported recently [7, 8]. In this paper we report the exact analytical expression for the distribution of the dwell times of a KIF1A mo- tor in the NOSC model during a single processive run in between its attachment to the track and the next detach- ment. What makes the calculation more difficult in the case of KIF1A, compared to those of those reported in ref. [7, 8], is the occurrence of branched pathways in its mechano-chemical cycle. For motors which can step both forward and backward, one can define conditional dwell times which may be more easily extracted from the data obtained from single molecule experiments [9, 10]. There- fore, in this paper we also report analytical expressions for the probability densities of these conditional dwell times as well as that of a few other closely related ran- dom variables. II. MODEL AND STOCHASTIC KINETICS A. The Model A MT track of the motor is modelled as a one- dimensional finite lattice having L number of discrete sites. Each site corresponds to a KIF1A binding site on the MT and the lattice spacing is the separation be- tween the successive binding sites on a MT. A KIF1A motor is represented by a particle with two possible chemical states labeled by the indices 1 and 2. The states 1 and 2 correspond to the strongly bound and weakly bound states, respectively. fig. 1 illustrates the detailed mechano-chemical cycle of KIF1A. The transi- tions 1j ↔ 2j are purely chemical whereas the transi- tions 2j ↔ 2j±1, which correspond to the Brownian mo- 2 R ∞ ¯S(q, t) and the Fourier ¯W (q, t) are given by W (q, s) = R ∞ The Laplace transforms of transforms S(q, s) = of 0 dte−st ¯S(q, t) and 0 dte−st ¯W (q, t). Therefore, using S(q, s) and W (q, s) as the two compo- nents of a vector P(q, s), the matrix equation (4) can now be written in a more compact notation as P(q, s) = R(q, s)−1P(0), withR(q, s) ≡ sI − M(q) and P(0) is a vector whose elements are determined by the initial conditions for S and W . The initial conditions are S(j, 0) = δjk and W (j, 0) = 0 where k is an arbitrarily selected site. The dwell time at the site k is the total duration for which the motor stays at that site, starting from the initial condition mentioned above, irrespective of the chemical state, before its next departure from the same site. (6) Thus the determinant of R(q, s) is a 2nd order poly- nomial in s; that is [11], R(q, s) = s2 + α(q)s + γ(q) where, α(q) = ωh + 2ωb + ωs + ωf − ωb(e−ıq + eıq) (7) (8) γ(q) = ωh(2ωb + ωf ) − ωbωh(e−ıq + eıq) − ωhωf e−ıq (9) Note that γ(q) is the determinant of the transition matrix M (q). Because of conservation of probability, in the q → 0 limit, all of the columns of M sum to zero. Therefore, we obtain M(0) = γ(0) = 0. We define [11] the position probability density ¯P (q, t) = ¯S(q, t) + ¯W (q, t), and hence P (q, s) = s + α(0) s2 + α(q)s + γ(q) . (10) C. Velocity and Diffusion constant Following ref.[11], the average velocity v of KIF1A and the diffusion constant D are found to be v = −ı γ(0) α(0) = ωf ωh (ωf + ωh + ωs) . (11) and , = D = 2α(0) γ(0) − 2ıv α(0) − 2β(0)v2 2ωbωh + ωf ωh − 2v2 2(ωf + ωh + ωs) (12) where γ(0) and γ(0) are the first and second derivatives, respectively, of γ(q) with respect to q evaluated at q = 0 while β(0) is the coefficient of s2 at q = 0; in this case β(0) = 1. Interestingly, the velocity v depends only on the coefficients γ and α, of the lowest two orders of the polynomial obtained from the determinant of R(q, s). On the other hand, the diffusion coefficient D depends on the three lowest order coefficients β, α, and γ of the determinant of R(q, s). FIG. 1: Two state model for KIF1A. The indices ..., j−1, j, j+ 1, ... label the equispaced sites for the binding of the motor to its track. The states 1 and 2 correspond to the "chemical" states in which the motor is bound strongly and weakly, re- spectively, to the microtubule track. The allowed transitions are shown by the arrows along with the corresponding rate constants (transition probability per unit time). tion, are purely mechanical. In contrast, the transition [3] for a more 2j → 1j+1 is mechano-chemical (see ref. detailed description). B. Kinetics and the master equations We define S(j, t) and W (j, t) as the probabilities of finding KIF1A in state 1 and state 2 respectively, at site j at time t. The master equations for these probabilities are dS(j, t) dt = −ωhS(j, t) + ωf W (j − 1, t) + ωsW (j, t) (2) dW (j, t) dt (3) = ωhS(j, t) − (ωf + ωs + 2ωb)W (j, t) + ωb(W (j − 1, t) + W (j + 1, t)) with ωh = ω0 h[AT P ], where [AT P ] is the concentra- tion of ATP. Now we introduce Fourier transforms of j=−∞ S(j, t)e−ıqj and j=−∞ W (j, t)e−ıqj , where ı = √−1 and the ¯W (q, t)# = M (q)(cid:20) ¯S(q, t) " ¯S(q, t) ¯W (q, t)(cid:21) S(j, t) and W (j, t) by ¯S(q, t) = P∞ ¯W (q, t) = P∞ lattice spacing, d = 1. Thus, (4) where the transition matrix M(q) is given by, M(q) =(cid:18) −ωh ωh − {(2ωb + ωs + ωf ) − ωb(e−ıq + eıq)} (cid:19) . (ωf e−ıq + ωs) (5) n o i t u b i r t s i d y t i l i b a b o r P 30 25 20 15 10 5 0 ω b=0 s-1 ω b=30 s-1 ω b=100 s-1 ω b=1125 s-1 ω h=125 s-1 ω s=145 s-1 ω f=55 s-1 Thus, = p+ψ+(s) + p ′ ′ ψ (s) = − a+(s) a0 ωbs + ωh(ωb + ωf ) s2 + (2ωb + ωh + ωf + ωs)s + ωh(2ωb + ωf ) Inverse Laplace transform yields, p+ψ+(t) + p ′ ′ ψ (t) = c1e−r1t/2 − c2e−r2t/2 2r0 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 Dwell time (s) where, FIG. 2: Dwell time distribution is plotted from equation (30) for few different values of ωb. III. DWELL TIME DISTRIBUTION We define the individual branching probabilities as p+, p sponding dwell time distributions are ψ+, ψ respectively. Thus ′ forward and backward and p− and the corre- and ψ−, ′ p+ = ωb 2ωb + ωf ′ p = ωf 2ωb + ωf p− = ωb 2ωb + ωf , , . (13) (14) (15) and, r0 =q(2ωb + ωf + ωh + ωs)2 − 4(2ωb + ωf )ωh c1 = (2ω2 b + ωbωf − ωbωh − 2ωf ωh + ωbωs + ωbr0) (22) c2 = (2ω2 b + ωbωf − ωbωh − 2ωf ωh + ωbωs − ωbr0) (23) r1 = 2ωb + ωf + ωh + ωs + r0 r2 = 2ωb + ωf + ωh + ωs − r0 Similarly, 1 s P (q, s){ρ+(q)=0} = a0 + a−(s)ρ−(q) where a0 is given by eq. (18) and a−(s) = − ωbs + ωbωh s2 + (ωf + ωh + ωs)s 3 (19) (20) (21) (24) (25) (26) (27) .(28) We define the Fourier weights ρ+(q) = e−ıq and ρ−(q) = eıq for the forward and backward steps, respectively (step size d = 1 in our units). Hence [11], = p−ψ−(s) = − a−(s) a0 ωbs + ωbωh s2 + (ωh + 2ωb + ωs + ωf )s + ωh(2ωb + ωf ) , Inverse Laplace transform yields, 1 s P (q, s)nρ ′ (q)=0o = k6=k 1 − ζ(s)ρ+(q) − p−ψ−(s)ρ−(q) 1 − ζ(s) − p−ψ−(s) ψ (16) (s)) and the symbol ′ ′ where ζ(s) = (p+ψ+(s) + p ′ is zero; in our case, can represent either the forward (+) or the backward nρk6=k′ (q) = 0o expresses the condition that the Fourier weight for all possible steps, except k k (−) steps. For forward branching we obtain, 1 − p+ψ+(s)ρ+(q) − p (s)ρ+(q) ψ 1 ′ ′ ′ s P (q, s) {ρ−(q)=0} = 1 − p+ψ+(s) − p′ ψ′(s) = a0 + a+(s)ρ+(q) (17) where a0 = s2 + (2ωb + ωh + ωf + ωs)s + ωh(2ωb + ωf ) s2 + (ωh + ωf + ωs)s a+ = − ωbs + ωh(ωb + ωf ) s2 + (ωh + ωf + ωs)s (18) p−ψ−(t) = ωb(cid:8)(r1 − 2ωh)e−r1t/2 − (r2 − 2ωh)e−r2t/2(cid:9) 2r0 (29) Now total dwell time distribution can be written as fol- lows: g(t) = p+ψ+(t) + p ′ ′ ψ (t) + p−ψ−(t) = {ωb(r2 − 2ωh) − ωf ωh}(e−r1t/2 − e−r2t/2) (30) r0 The total dwell distribution is plotted in the fig. 2. We observe that with the increase of ωb the most probable of dwell time shifts towards a smaller value. This is because rate constant ωb is related to diffusion of KIF1A in state 2. For smaller values of ωb dwell time is dominated by the rate ωf , but with the increase of rate ωb the dwell time depends on both ωb and ωf and eventually the mean dwell time decreases. IV. PROBABILITIES OF SPLITTING AND step, are given by 4 CONDITIONAL DWELL TIMES A typical trajectory of a KIF1A motor consists of a random sequence of forward and backward steps. There- fore, we can define four different conditional dwell times τ±±. Here τ++ is the dwell time in between two consec- utive forward steps whereas τ−− is the dwell time in be- tween two consecutive backward steps. Similarly, τ+− is the dwell time in between two consecutive steps of which the first is forward and the second is backward whereas the opposite if true in case of τ−+. We denote the probability density functions for the conditional dwell times by the symbols Ξ±±(t). The inte- grated probabilities obtained from these probability den- sities are given by Q+ µ (0) = ωf δµ1 + ωbδµ2 ωb + ωf and Q− µ (0) = δµ2 (35) where µ = 1, 2, Q1 = S and Q2 = W as described above. Using the expressions for the probability currents as- sociated with the allowed transitions in our model, we get Π±+P±+(t) = Z t Π±−P±−(t) = Z t 0 0 dt′{ωbQ2(t′) + ωf Q2(t′)} dt′{ωbQ2(t′)} (36) Following Lind´en and Wallin [9], by using the Ansatz 0 Ξ±±(t′)dt′ P±±(t) =Z t Obviously, limt→∞ P±±(t) = R t 0 Ξ±±(t′)dt′ = 1. More- over, we introduce the "pairwise splitting" probabilities Π±± where Π++ and Π+− represent the probability that a forward step is followed by a forward step or a backward step, respectively. Similarly, Π−+ and Π−− denote the probability that a backward step is followed by a forward step or a backward step, respectively. For the analysis of the experimental data and comparison with theoreti- cal predictions, it is sometimes more convenient to divide the dwell times into two groups depending on the direc- tion of the following step. We use the symbols Ξ∗ +(t) and Ξ∗ −(t) to denote the probability density functions for the dwell times before a forward (+) and a backward (−) steps, respectively. Obviously, Ξ∗ ±(t) are given by [9] (31) P++(t) = P−−(t) = 1 + εµeλµt (37) 2 Xµ=1 together with eqs. (35)-(36) to compute ∂tΠ±+P±+(t) and ∂tΠ±−P±−(t) we obtain the following systems of lin- ear equations: 1 λ1 λ2 (cid:19)(cid:18) ε++ (cid:18) 1 1 ε++ 2 = −1 −1 Π++ ω2 b ωb Π+−(ωb+ωf ) ε+− 1 ε+− 2 ε−+ 1 ε−+ 2 −1 ωb+ωf Π−+ ε−− 2 (cid:19) 1 ε−− Π−− ! . −1 ωb (38) Solving this and using Ξ±±(t) = ∂tP±±(t) we obtain the following distributions of the conditional dwell times: Ξ∗ Ξ∗ +(t) = Π++Ξ++(t) + Π−+Ξ+−(t) −(t) = Π+−Ξ−+(t) + Π−−Ξ−−(t) Ξ+−(t) = u2 (32) In this section we derive analytical expressions for Π±±, Ξ±±(t), and hence, Ξ∗ ±(t) following the procedure adopted in ref.[9, 10]. In our model, immediately after a forward step the KIF1A motor can be found in either of the two states (Qµ(t); Q1(t) = S(t) i.e. strongly bound state, and Q2(t) = W (t) i.e. weakly bound state) whereas it can exist only in state 2 immediately after a backward step. The first escape problem for our model is governed by a reduced Master equation (as described in refs. [9] and [17]), with the transition matrix Υ =(cid:18) −ωh ωh −(ωf + ωs + 2ωb)(cid:19) . ωs (33) The eigen values of Υ are λ1,2 = −u1 ±p(ωh − ωf − ωs − 2ωb)2 + 4ωhωs 2 (34) where, u1 = ωh + ωf + ωs + 2ωb and define u2 = λ1λ2 = ωf ωh + 2ωbωh. The initial conditions, describing the dis- tribution of states just after a ± (forward and backward) Ξ++(t) = u2 Ξ−+(t) = u2 Ξ−−(t) = u2 Hence (eλ1t − eλ2t) λ1 − λ2 (eλ1t − eλ2t) λ1 − λ2 (eλ1t − eλ2t) λ1 − λ2 (eλ1t − eλ2t) λ1 − λ2 + + + + (λ1 − λ2) (λ1eλ1t − λ2eλ2t) ωb Π++ ω2 b (λ1eλ1t − λ2eλ2t) Π+−(ωb + ωf )(λ1 − λ2) (ωb + ωf ) (λ1eλ1t − λ2eλ2t) Π−+ (λ1 − λ2) (λ1eλ1t − λ2eλ2t) (39) ωb Π−− (λ1 − λ2) ΥT (cid:18) Π1− Π1+ Π2− Π2+ (cid:19) = −(cid:18) 0 ωb (ωf + ωb)(cid:19) . 0 (40) To derive the pairwise splitting probabilities we first solve for the Πµ± and then weight them according to the initial conditions as of eq. (35): Π±+ =Xµ Πµ+Q(±) µ (0), Π±− =Xµ Πµ−Q(±) µ (0) (41) We finally get the following splitting probabilities Π++ = Π−+ = Π−− = Π+− = ωh(ωf + ωb) ωh(ωf + 2ωb) ωhωb u2 (42) ) ) t ( - / + * Ξ ( y t i s n e d y t i l i b a b o r P 200 180 160 140 120 100 80 60 40 20 0 +(t):ω Ξ* -(t):ω Ξ* +(t):ω Ξ* -(t):ω Ξ* +(t):ω Ξ* -(t):ω Ξ* ω ω ω b=0.01 b=0.01 b=10 b=10 b=100 b=100 h=125 s-1 s=145 s-1 f=55 s-1 3 2.5 2 1.5 1 r e t e m a r a p s s e n m o d n a R 5 ω s=145 s-1 ω f=55 s-1 45 ω b=0s-1 ω b=30s-1 ω b=50s-1 ω b=100s-1 b=1125 s-1 ω f=5s-1 ω ω f=20s-1 ω f=55s-1 30 15 0 1e-08 1e-06 0.0001 0.01 0 0.01 0.02 0.03 0.04 0.05 0.06 Dwell time (sec) 0.5 1e-08 1e-07 1e-06 1e-05 0.0001 0.001 0.01 [ATP] (M) FIG. 3: The distribution Ξ∗ values of ωb. ±(t) is plotted for few different Thus, probability of a forward step is [ωh(ωf + ωb)]/[ωh(ωf + 2ωb)] irrespective of the direction of the preceeding step. Similarly, the probability of a backward step is (ωhωb)/[ωh(ωf +2ωb)] irrespective of the direction of the preceeding step. Both these are consistent with the kinetic pathways shown in fig. 1 as well as with the fact that Π++ + Π+− = 1 and Π−− + Π−+ = 1. Substituting the expressions (42) and (39) into (32) we get the analytical expressions for Ξ∗ ±(t). This distri- bution is plotted in fig. 3 for a few different values of ωb. The most probable dwell time before a forward step decreases with increasing ωb. V. RANDOMNESS PARAMETER Using eq. (30) we obtain < t >= 4α (r2 + r1)(r2 − r1) (r1r2)2 and < t2 >= 16α (r2 − r1)(r2 2 + r1r2 + r2 1) (r1r2)3 (43) (44) where r1, r2 are given by eqs. (24), (25) and α is given by 2ω2 α = b + ωbωf + ωbωs − ωbωh − ωf ωh − ωbr0 . (45) r0 FIG. 4: The randomness parameter r, defined by eq. (46), is plotted against ATP concentration for a few different values of ωb. The inset shows the same for few different values of ωf . ω f=55.0 s-1 ω s=145 s-1 ) - * r ( r e t e m a r a p s s e n m o d n a R 8 7 6 5 4 3 2 1 ω b=20s-1 ω b=50s-1 ω b=100s-1 ω b=20s-1 ω b=50s-1 ω b=100s-1 1e-06 0.0001 0.01 [ATP] (M) 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 ) + * r ( r e t e m a r a p s s e n m o d n a R 0 1e-08 1e-07 1e-06 1e-05 0.0001 0.001 0.01 [ATP] (M) FIG. 5: The randomness parameter r∗ +, defined by eq. (47), is plotted against ATP concentration for a few different val- ues of ωb. The inset depicts the dependence of randomness parameter r∗ −, defined by eq. (47), against ATP concentration for the same set of ωb values. values of ωf . With increasing concentration of ATP, r decreases and finally sturates near to unity. At low ATP concentrations r is greater than 1 which may be the effect of multi-exponentiality in dwell time distribution. It also depicts that increasing concentration of ATP reduces the fluctuations in dwell time. For the conditional dwell times, the randomness pa- rameters are defined by r0 is given by eq. (21). Using equations (1), (43) and (44) we obtain randomness parameter as follows r∗ ± = < (t∗ ±)2 > − < t∗ ± >2 < t∗ ± >2 (47) r = r1r2(r2 2 + r1r2 + r2 1) − α(r2 − r1)(r2 + r1)2 α(r2 − r1)(r1 + r2)2 (46) Using eq. (32) we obtain Randomness parameter (see eq. (46)) is plotted in fig. 4 against ATP concentration for a few different values of ωb and same is plotted in the inset for a few different < t∗ + >= −2Π++ u2(λ1 + λ2) (λ1λ2)2 − Π++ω2 b Π+−(ωb + ωf )λ1λ2− ωb λ1λ2 (48) ) + * r ( r e t e m a r a p s s e n m o d n a R 1.8 1.7 1.6 1.5 1.4 1.3 1.2 1.1 1 0.9 ω s=145 s-1 ω b=1125 s-1 ) - * r ( r e t e m a r a p s s e n m o d n a R 1.3 1.25 1.2 1.15 1.1 1.05 1 0.95 ω f=10s-1 ω f=30s-1 ω f=55s-1 ω f=10s-1 ω f=30s-1 ω f=55s-1 0.0001 0.001 0.01 [ATP] (M) 0.0001 0.001 [ATP] (M) 0.01 FIG. 6: The randomness parameter r∗ +, defined by eq. (47), is plotted against ATP concentration for a few different val- ues of ωf . The inset depicts the dependence of randomness parameter r∗ −, defined by eq. (47), against ATP concentration for the same set of ωf values. and < (t∗ +)2 > = 4Π++ u2(λ2 + 2ωb(λ1 + λ2) (λ1λ2)2 (49) 1 + λ1λ2 + λ2 2) (λ1λ2)3 ω2 b (λ1 + λ2) +2Π++ Π+−(ωb + ωf )(λ1λ2)2 where Π++, Π−+, Π−−, and Π+− are given by eq. (42). λ1 and λ2 are given by equation (34). Similarly, < t∗ − >= −2Π−− u2(λ1 + λ2) (λ1λ2)2 − Π+−(ωb + ωf ) Π−+λ1λ2 − ωb λ1λ2 (50) and < (t∗ +)2 > = 4Π−− u2(λ2 1 + λ1λ2 + λ2 2) (λ1λ2)3 + 2ωb(λ1 + λ2) (λ1λ2)2 +2Π+− (ωb + ωf )(λ1 + λ2) Π−+(λ1λ2)2 (51) The conditional randomness paremeters (eq. (47)) are plotted in the figs. 5 and 6 against ATP concentration 6 for different values of ωb and ωf respectively. The non monotonic variation of conditional randomness parame- ters with ATP concentration changes to a monotonic de- crease (see fig. 6) when the magnitude of ωb is sufficiently high. Variation of the randomness parameter indicates the changes in the number of rate-limiting steps. Any value of the randomness parameter higher than unity may, at first sight, appear counter-intuitive. But, this is quite common for systems with branched mechano- chemical kinetics. VI. SUMMARY AND CONCLUSION Theoretical calculation of the dwell time distribution of two-headed conventional kinesin motors has been re- ported earlier [12]. In this paper we have derived an exact analytical expression for the distribution of dwell times of single-headed kinesin motors KIF1A at each binding site during a processive run on a microtubule. We have used the NOSC model [3] for single-headed KIF1A mo- tors to derive our results. Since both forward and back- ward steps of this motor are possible, we have also de- fined conditional dwell times and calculated their distri- butions analytically. The experimentally measured dwell time distributions for some of the other processive mo- tors, like conventional kinesin [13], myosin-V [14], dynein [15], ribosome [16], etc., have been reported in the lit- erature. To the best of our knowledge, the dwell time distribution of KIF1A has not been reported so far; we hope our theoretical prediction will stimulate experimen- tal investigations. Acknowledgments This work is suppoted by a research grant from CSIR (India). AG thanks UGC (India) for a senior research fellowship. [1] Howard J. Mechanics of motor proteins and the cytoskele- [7] Garai A., Chowdhury D., Chowdhury D. and Ramakr- ton (Sinauer Associates, Sunderland, MA, 2001). ishnan T. V., Phys. Rev. E. 80, 011908 (2009). [2] Hirokawa N., Nitta R. and Okada Y., Nat. Rev. Mol. Cell [8] Tripathi T., Schutz G. M. and Chowdhury D., J. Stat. Biol. 10, 877 (2009), and references therein. Mech.: Theory and Experiment, P08018 (2009). [3] Nishinari K., Okada Y., Schadschneider A. and Chowd- [9] Lind´en M. and Wallin M., Biophysical Journal 92, 3804 hury D., Phys. Rev. Lett. 95, 118101 (2005). (2007). [4] Greulich P., Garai A., Nishinari K., Schadschneider A. [10] Tsygankov D., Lind´en M., and Fisher M. E., Phys. Rev. and Chowdhury D., Phys. Rev. E 75, 041905 (2007). E 75, 021909 (2007). [5] Kolomeisky A. B. and Fisher M. E., Annu. Rev. Phys. [11] Chemla Y. R., Moffitt J. R., and Bustamante C., J. Chem.58, 675 (2007) and references therein. Phys. Chem. B 112, 6025 (2008). [6] Schnitzer M. J. and Block S. M., Cold Spring Harbor [12] Valleriani A., Liepelt S. and Lipowsky R., EPL, 82, Symp. Quant. Biol.LX, 793 (1995). 28011 (2008). [13] Asbury C. L., Fehr A. N. and Block S. M., Science 302, 2130 (2003). [14] Pierobon P., Achouri S., Courty S., Dunn A. R., Spudich J. A., Dahan M. and Cappello G., Biophys. J. 96, 4268 (2009). [15] Reck-Peterson S. L., Yildiz A., Carter A. P., Gennerich A., Zhang N., and Vale R. D., Cell 126, 335 (2006). [16] Wen J. -D., Lancaster L., Hodges C., Zeri A. -C., Yoshimura S. H., Noller H. F., Bustamante C., and I. Tinoco, Nature 452, 598 (2008). [17] N. G. van Kampen Stochastic Processes in Physics and Chemistry (Elsevier, Amsterdam, The Netherlands, 2nd Ed., 1992). 7
1702.03725
1
1702
2017-02-13T11:46:38
Time-domain THz spectroscopy reveals coupled protein-hydration dielectric response in solutions of native and fibrils of human lyso-zyme
[ "physics.bio-ph" ]
Here we reveal details of the interaction between human lysozyme proteins, both native and fibrils, and their water environment by intense terahertz time domain spectroscopy. With the aid of a rigorous dielectric model, we determine the amplitude and phase of the oscillating dipole induced by the THz field in the volume containing the protein and its hydration water. At low concentrations, the amplitude of this induced dipolar response decreases with increasing concentration. Beyond a certain threshold, marking the onset of the interactions between the extended hydration shells, the amplitude remains fixed but the phase of the induced dipolar response, which is initially in phase with the applied THz field, begins to change. The changes observed in the THz response reveal protein-protein interactions me-diated by extended hydration layers, which may control fibril formation and may have an important role in chemical recognition phenomena.
physics.bio-ph
physics
Time-domain THz spectroscopy reveals coupled protein-hydration dielectric response in solutions of native and fibrils of human lyso- zyme Fabio Novelli1, Saeideh Ostovar Pour2, Jonathan Tollerud1, Ashkan Roozbeh1, Dominique R. T. Appadoo3, Ewan W. Blanch2, Jeffrey A. Davis1,* 1 Centre for Quantum and Optical Science, Swinburne University of Technology, Victoria 3122, Australia 2 School of Science, RMIT University, GPO Box 2476, Melbourne, Victoria 3001, Australia 3 Australian Synchrotron, 800 Blackburn Road, Clayton, 3168 VIC, Australia ABSTRACT: Here we reveal details of the interaction between human lysozyme proteins, both native and fibrils, and their water environment by intense terahertz time domain spectroscopy. With the aid of a rigorous dielectric model, we deter- mine the amplitude and phase of the oscillating dipole induced by the THz field in the volume containing the protein and its hydration water. At low concentrations, the amplitude of this induced dipolar response decreases with increasing con- centration. Beyond a certain threshold, marking the onset of the interactions between the extended hydration shells, the amplitude remains fixed but the phase of the induced dipolar response, which is initially in phase with the applied THz field, begins to change. The changes observed in the THz response reveal protein-protein interactions mediated by extended hydration layers, which may control fibril formation and may have an important role in chemical recognition phenomena. Proteins are the most versatile macromolecules of the cell and the ability to maintain their native, functional state is fundamental to biological activity. A great number of mammalian disorders have some linkage with protein misfolding and aggregation. Amyloid fibril related disor- ders[1-4], for example, are associated with the unwanted filamentous aggregation of particular proteins or peptides. Human lysozyme is a good model system that has been used for decades to study amyloid fibrillogenisis. The com- mon features of amyloidogenic proteins are β-sheet for- mation and fibrillar morphology[5], however, the detailed mechanism of β-sheet formation and the molecular level properties of fibrils are not yet understood despite being subjects of great interest[6-8]. The importance of the dielectric solvated environment, particularly of water molecules in mediating the secondary and tertiary structure of proteins, interactions between proteins[9,10], and conformational changes that lead to ag- gregation is widely appreciated[11-21], but poorly charac- terized. Light at terahertz (THz) frequencies (0.1-10 THz or 3,000-30 𝜇m) is strongly absorbed by water and provides a very sensitive probe of any perturbations to the hydrogen bonding network[22,23]. At the interface of a macromole- cule in solution several things typically happen: the water molecules close to the surface will be strongly influenced and their dynamics retarded by charges on the surface of the macromolecule, leading to changes in the THz absorp- tion spectrum[24,25]. The influence of protein-water inter- actions also extends beyond these tightly bound water molecules[26-29], causing changes to the hydrogen bond- ing network and thus to the contribution of both single wa- ter molecules and the larger domains to the THz absorp- tion spectrum. Details of these changes, and how far these effects propagate from the surface of the macromolecule, remain open questions. Techniques based on NMR[30] and neutron or X-ray scattering[31] probe only the tightly bound water molecules, indicating that a sub-monolayer of water molecules is formed around each protein at near physiological conditions. Similar conclusions are drawn from terahertz spectroscopy when films[32-34] or relatively large protein concentrations are studied (≥ 60 mg/ml for hen egg white lysozyme[35]). However, when lower con- centrations are used, terahertz techniques suggest that the population of hydration water is much larger and extends over multiple layers. Extended hydration shells, where the effect of the protein on the water molecules is evident, have been reported to vary between 15 Å and 25 Å from the so- lute surface for different proteins[25-29]. While these works are based on the simplistic assumption that the ab- sorption coefficient of a complex system can always be written as the weighted sum of the absorption of its com- ponents[35], here we show how determining the complex dielectric response provides a much more informative and reliable assessment of the impact of protein-water interac- tions on the water environment. Another recent work in- vestigated the lysozyme films[34]. However, results on films[33,34] cannot be easily full dielectric response of compared with results on solutions because additional col- lective modes appear upon film formation[36], and be- cause films typically have much lower water content (com- pare e.g. the dielectric functions shown in Ref.[34] with the one of bare water in Ref.[23]). Here we report high-resolution terahertz time-domain measurements of solutions containing native human lyso- zyme or fibrils down to concentrations as low as 5 mg/ml (corresponding to ~0.36% volume fraction) at room tem- perature. By determination of the complex dielectric func- tion, we are then able to determine the amplitude and phase of the induced dipole in the volume containing the protein and the extended hydration water. Further details of the hydration layer and interactions between proteins are then revealed by tracking the concentration depend- ence of these values. We measured the amplitude and the phase of the THz field transmitted through 0.5 mm thick lysozyme-water so- lutions at room temperature. We used large amplitude THz fields generated by tilted-front optical rectifica- tion[37-41] and detected by electro-optical sampling[42- 44], as described in the Supplementary Information. The associated pulse fluence is lower than 2 𝜇J/𝑐𝑚2 and any thermal effects can be disregarded[45]. Human lysozyme was obtained from Sigma Aldrich (Australia). The samples were prepared in milli-Q water with pH value of 4.80 for solutions containing native proteins, and with pH 2.0 for solutions containing fibrils. In order to form fibrils of hu- man lysozyme the samples were incubated at 96 °C for 16 hours in a solution with pH 2.0 [46]. The formation of fi- brils following this protocol has previously been verified by AFM, TEM and Thiflavin T assay[46]. Additional details of sample preparation are discussed in the Supplementary In- formation. The distilled water and the protein-free solu- tions at pH 2.0 and 4.8 all have, within our resolution, the same optical absorption and refraction properties at the te- rahertz frequencies investigated in this work. In the follow- ing we simply take distilled water as the background for zero protein concentrations. The samples are well shaken before the measurements, and the fields transmitted by a reference and by the filled quartz cell are acquired alter- nately 10 times for a total acquisition window shorter than 15 minutes per sample. In Figure 1 we report typical time-domain traces of the THz pulses transmitted by a reference (Figure 1a) and by 100 mg/ml solutions of native human lysozyme (blue curve in Figure 1b) and fibrils (red curve in Figure 1b). By com- paring the blue and red curves in Figure 1b it can be seen that the terahertz response of fibrils and native solutions is distinctly different: the solution containing the fibrils dis- plays lower transmission and is delayed in time. The inset of Figure 1b shows the spectral amplitude of the transmis- sion, compared to the incident spectrum (the dips in the incident THz spectrum at and above ∼ 1.1 THz are due to absorption by water molecules in the air[47]). The large drop (>20x) in the THz spectral amplitude transmitted by the protein solutions is primarily due to absorption by bulk water. However, by normalizing the transmission of the protein solutions to the transmission through pure water (see Supplementary Figure1(a)), it can be seen that the spectral response of both native lysozyme and fibrils is flat and that the fibrils transmit less than the native lysozyme. These results are confirmed by additional measurements performed at the FAR-IR beamline of the Australian Syn- chrotron (see Supplementary Figure 1(b,c)). Figure 1. a) Amplitude of the pulsed THz field transmit- ted through the empty sample holder as a function of the sampling pulse delay. b) Typical THz pulses transmitted by water-based solutions containing 100 mg/ml native (blue) and fibrils (red). In the inset the magnitude of the Fourier transformations are shown as a function of frequency. c) The index of refraction for the lysozyme solutions and d) The corresponding absorption coefficient. From the time-domain THz experiments we calculate the full dielectric properties of the solutions, both the in- dex of refraction (Figure 1c) and the absorption coefficient (Figure 1d), assuming only the first Fresnel transmission coefficient is relevant (see Supplementary Information for details). It is important to note that scattering cannot ex- plain the differences observed between fibril and native protein solutions (see Supplementary Information). 2 For the following analysis we focus on the average values for the optical properties over the 0.4-0.9 THz range. The average absorption coefficient is shown in Figure 2a for na- tive (blue solid squares) and fibril (red solid squares) lyso- zyme solutions, together with the expected effect of re- moving a volume of bulk water equal to the protein volume (black line in Figure 2a). The optical properties depend on the volume fraction, 𝜂, of the proteins in solution. For na- tive lysozyme we measured a variety of solutions with 𝜂 ranging from about 0.36% to 42.7% (or, equivalently, from 5 to 600 mg/ml, taking the volume of each protein as a sim- ple sphere with radius ~ 15.9 A[48] and assuming ideal be- havior [49]). Three distinct regions can be identified in Fig. 2(a) for the native human lysozyme (blue squares): above the solution saturation threshold of 𝜂 ∼ 17.8% (250 mg/ml) a strong deviation from water-removal effects is seen, as expected[35]; between 𝜂 = 7.1% and 17.8% (100 and 250 mg/ml) a linear decrease is seen, showing reasonable agreement with bulk water-removal effects[35]; and at vol- ume fractions in the range 0.36% – 3.6% (5-50 mg/ml) a flatter region is evident (zoomed in Figure 2b), which can- not be explained as removal of bulk water[29], but can be described in part by contributions from the water mole- cules that are affected by the presence of the pro- tein[24,29]. These results are consistent with previous measurement of THz absorption of lysozyme over different parts of this concentration range[29,35]. The solutions containing fibrils of lysozyme have been studied below the saturation point, between 𝜂 ∼ 0.36% (5 mg/ml) and 14.2% (200 mg/ml). It is evident that the solu- tions with fibrils display a larger absorption and index of refraction compared to the native lysozyme (see Figure 1 and Figure 2). In the following we focus on the results obtained below 𝜂 ∼ 3.6% (50 mg/ml) where the protein volume is known not to change for different concentrations[49]. To understand the volume-fraction dependent response of the fibril and native human lysozyme solutions shown in Figure 2 we begin by following the approach of Heyden et al.[24]. The starting point is to consider the protein vol- ume as a void. The first order effect is simply the removal of bulk water, which leads to a reduction in the absorption linearly proportional to the volume fraction of the protein in solution (this is the solid black line in Figure 2a and Fig- ure 2b). However, in a dielectric medium the applied field will also induce a polarization at the interface between the void and the dielectric. If we consider the protein to be a spherical void and water to be a static dielectric medium, Maxwell's equations tell us that the induced field will be in the opposite direction to the applied field, with magnitude given by 𝑴𝑀 = −𝑉𝑷 3𝜖𝑆 where 𝑉 is the void volume, 𝑷 is 2𝜖𝑆+1 the transverse bulk polarization, and 𝜖𝑆 is the dielectric constant of the solvent[24]. Under these assumptions the resultant average dipole moment induced over the solu- tion, and hence the THz absorption, is further reduced. Figure 2. a) The average absorption coefficient in the 0.38-0.92 THz range of the lysozyme solutions is shown versus the volume fraction of lysozyme. b) Same as a) but on an expanded scale for small concentrations or volume fractions. Error bars are ± 1 standard deviations calculated from 10 measurements (see Supplementary Information). c) Average index of refraction at low concentrations. Previous measurements of proteins (including lysozyme) in water solutions have, however, shown that rather than the THz absorption being lower than expected when con- sidering only water removal, the THz absorption is actually higher[24-29], consistent with the results reported here (Fig. 2). The absorption is enhanced because the protein is not a void and will typically have some absorption and surface charges, and because the water environment of the protein is not a static dielectric but is able to move and rotate to compensate for an electromagnetic perturbation. It is clear then that the pure Maxwell picture described above does not provide a complete description. Heyden et al.[24] thus defined the scalar quantity "𝑎" relating the actual dipole moment induced in the volume of solution containing one protein and its hydration layers, 𝑴𝑖𝑛𝑡, to the dipole mo- ment predicted in the case of one spherical void in a static dielectric 𝑴𝑀: 𝑎 = 𝒙^ ∙ 𝑴𝑖𝑛𝑡 𝑴𝑀 Eq. 1 where 𝒙^ is an unitary vector along the direction of the applied electric field. This can be combined with the con- tribution from bulk water and its removal to give the com- plex, frequency-dependent dielectric response of the solu- tion, 𝜖(𝜔): 3 ] Eq.2 𝜖(𝜔) = 1 + [𝜖𝑆(𝜔) − 1][1 − 𝜂] − 𝜂 [𝑎(𝜔) [𝜖𝑆(𝜔)−1]2 2𝜖𝑆(𝜔)+1 where 𝜂 is the volume fraction of the protein in solution and 𝜖𝑆(𝜔) is the dielectric function of the solvent alone[24]. In the case that 𝑎 = 1 the induced dipole will follow the behaviour predicted from Maxwell's electrostat- ics 𝑴𝑖𝑛𝑡 = 𝑴𝑀; for 𝑎 = 0 the solvent rearrangement will perfectly compensate for the Maxwell dielectric field and the behaviour of the solution can be described simply as removal of bulk water. In several previous works[28,50] the THz absorption of protein solutions decreases slower than expected for sol- vent removal, or in some cases even increases as a function of protein concentration[24-27,29]. This corresponds to a negative value of a and an induced dipole, in the volume containing protein and hydration water molecules (i.e. 𝑴𝑖𝑛𝑡), that is parallel with the applied field direction. This is consistent with the absorption behavior observed here; however, we also have the refractive index, which allows a reassessment of Eq.1 and the determination of both ampli- tude and phase of the induced dipole. The analysis proposed by Heyden et al.[24], based on Eq. 1, indicates that only the component of the induced field in phase (negative 𝑎) or 𝜋 out of phase (positive 𝑎) with the applied field is relevant. In general, however, the induced dipole can have any phase relative to the applied field. De- termination of the full complex dielectric function, as we obtain here, provides additional constraints that allow us to determine a complex value for 𝒂 which relates to the amplitude and phase of 𝑴𝑖𝑛𝑡, not simply the projection of the induced-dipole onto the applied field direction. In this context, the scalar product in Eq. 1 is between the induced dipole and the axis of the applied field rather than the os- cillating field direction. We determine the complex valued 𝒂(𝝎) using Eq. 2, the 𝛼 and 𝑛 values obtained previously (see Fig.1 and Fig.2), and the measured values for the complex dielectric func- tion of the pure solvent. The amplitude and phase of 𝒂 are plotted for both the fibrils and native lysozyme in Figure 3 (the lines represent different THz frequencies while the points correspond to the average values). For native lysozyme the average amplitude of a, a, is initially 1.4 ± 0.3 and the phase, (a), is −𝜋. This corre- sponds to an induced field, 𝑴𝑖𝑛𝑡, oscillating in phase with the applied field and having a magnitude consistent with previous reports on a variety of different macromolecules in water solutions (see e.g. Ref.[24] and references therein). A subsequent rapid drop of a as the concentration is in- creased takes the amplitude down to 0.6 ± 0.1, where it re- mains flat for the rest of the concentration range. From this point on, however, the average phase increases from −𝜋 at the lowest concentration to −0.4𝜋 at the volume fraction of 3.6%. Figure 3. a) Relative amplitude and b) phase of the in- duced dipole at the protein-solvent boundary versus vol- ume fraction of lysozyme for solutions containing native human lysozyme (blue) and fibrils (red). Squares represent the average response in the 0.46-0.77 THz range, each thin line corresponds to a different frequency from 0.46 THz to 0.77 THz by ~0.07 THz steps (legend inside panel b). Precise quantitative details of the induced dipole mo- ment at the low concentration limit are determined by the nanoscopic interactions within and between the protein/s and the surrounding water molecules, and requires de- tailed molecular modelling of the dynamic response, which is beyond the scope of this work. However, as a phenome- nological explanation of our results we propose that the two regimes observed correspond to the two cases shown in Figure 4(b). At low concentrations a reduces as the probability of the solvation shells overlapping slightly in- creases, as proposed by Heyden et al[24]. In this case, most of the proteins and their solvation waters are not interact- ing with the solvation shells of other proteins. Above the critical concentration, where a remains constant and 4 (a) changes, there is significant overlap between the solv- ation shells of neighboring proteins. In this regime the solvation water is the dominant "type" of water and the strength of the interaction between the solvation waters and the proteins increases as the average distance between them decreases. This then affects the ability of the induced dipole to follow the field, which results in the phase shift observed. Figure 4. a) Cartoon of a human lysozyme protein (red sphere) in water together with the populations of: i) water molecules tightly-bound to the protein surface (white), ii) extended hydration layers (blue), and iii) unperturbed bulk water (light blue). b) Sketch of the evolution of the modu- lus and the phase of the induced dipole in a unit volume of solution versus protein concentration. The effect of phase gain at larger concentrations is pictorially represented by darker colors on the bottom right panel. Simply assuming that all bulk water has been replaced at the point where the trends of both a and (a) change, (i.e. at η ~ 1% or 15 mg/ml for native solutions), and con- sidering a uniform cube lattice of proteins, we estimate[48] an upper limit for the extent of the hydration layer around each native protein. This indicates the hydration layer may extend up to 42 Å from the protein surface for dilute solu- tions. This value is on the same order of magnitude as pre- vious reports[26-29], and is very similar to the size of the extended hydration layer calculated by molecular dynam- ics calculations for other protein systems[24]. For the solutions containing fibrils, the plots in Figure 3 for the amplitude (red squares) and phase (red triangles) of 𝒂 are quite different: the amplitude slowly decreases while the phase remains flat at –π. This behavior of the fi- brils up to 𝜂 ∼ 3.6% (50 mg/ml) is equivalent to the behav- ior of native lysozyme for 𝜂 < 1% (< 15 mg/ml) and is con- sistent with the phenomenological model proposed above, whereby the dominant trend is determined by the average distance between macromolecular assemblies. For a given concentration of native lysozyme the separation between fibrils, each originating from many closely spaced lyso- zyme molecules, is much larger, less hydration water is present, and the hydration layers do not overlap. A precise quantitative comparison, however, is prevented by the large heterogeneity in the fibril size (see Methods in SI). The other major difference observed in the fibrils is that the initial value for 𝒂 is 2.8 ± 0.6, significantly larger than the initial value for native lysozyme. This indicates that the induced dipole, 𝑴𝑖𝑛𝑡, for the fibrils is larger than for the native. Previous work has revealed that in fibrils the lyso- zyme tertiary structure changes, such that there is a high density of 𝛽-sheets on their surface[51]. This may explain the greater value of 𝑴𝑖𝑛𝑡 because the higher surface charge density from the 𝛽-sheets can cause a greater disruption to the hydrogen-bonding network of water, however, detailed modelling is required to quantitatively compare the two. Regardless, the demonstrated ability to reveal differ- ences in the THz response of native proteins and fibrils in- dicates that terahertz techniques can help to understand the nature of protein-environment interactions and the role they play in stabilizing tertiary structure and driving changes to it[52]. In conclusion, determination of the complex dielectric function of proteins in solution, enabled by time domain THz spectroscopy, allows detailed analysis beyond what is possible from simple absorption measurements. Following this approach we are able to determine both the phase and amplitude of the induced dipole in the volume containing the protein and solvating water. Separating the amplitude and phase of this response reveals that not only can the amplitude of the induced dipole vary as the concentration is changed, but above a concentration threshold so too does the phase of the induced dipole oscillations. While the precise details of the THz response are determined by dynamic atomistic interactions, we propose a phenomeno- logical model that explains the available data, whereby the phase of the induced dipole in the protein-solvent interac- tion region begins to vary when there is significant overlap between solvation layers of neighboring proteins. This re- sult suggests that indirect electromagnetic protein-protein interactions could take place if mediated by the extended hydration layers surrounding each protein which might also be of great relevance for chemical recognition[52,53]. AUTHOR INFORMATION NOTES Corresponding Author * [email protected] ACKNOWLEDGMENT We thank the Australian Research Council for funding (DP130101690). FN acknowledges funding from Swinburne University via the Early/Interrupted research career scheme 2014. This work has been partially done at the Australian Syn- chrotron during beamtime AS153/FIR/9692 at the FAR-IR beamline. REFERENCES [1] Baumeister, R.; Eimer, S. Amyloid Aggregates Presenilins, and Alzheimers Disease. Angewandte Chemie International Edi- tion 1998, 37 (21), 2978–82. 5 [2] Rhoades, E.; Agarwal, J.; Gafni, A. Aggregation of an Amy- loidogenic Fragment of Human Islet Amyloid Polypeptide. Bio- chimica Et Biophysica Acta (BBA) - Protein Structure and Molecu- lar Enzymology 2000, 1476 (2), 230–38. [3] DiFiglia, M.; Sapp, E.; Chase, K.O.; Davies, S.W.; Bates, G.P.; Vonsattel, J.P.; Aronin, N. Aggregation of Huntingtin in Neuronal Intranuclear Inclusions and Dystrophic Neurites in Brain. Science 1997, 277 (5334), 1990–93. [4] Prusiner, S.B. Molecular Biology of Prion Diseases. Science 1991, 252 (5012), 1515–22. [5] Campioni, S.; Mannini, B.; Zampagni, M.; Pensalfini, A.; Par- rini, C.; Evangelisti, E.; Relini, A.; Stefani, M.; Dobson, C.M.; Cec- chi, C.; Chiti, F. A Causative Link between the Structure of Aber- rant Protein Oligomers and Their Toxicity. Nature Chemical Biol- ogy 2010, 6 (2), 140–47. [6] Liu, R.; He, M.; Su, R.; Yu, Y.; Qi, W.; He, Z. Insulin Amyloid Fibrillation Studied by Terahertz Spectroscopy and Other Bio- physical Methods. Biochemical and Biophysical Research Commu- nications 2010, 391 (1), 862–67. [7] Png, G.M.; Falconer, R.J.; Abbott, D. Tracking Aggregation and Fibrillation of Globular Proteins Using Terahertz and Far-In- frared Spectroscopies. IEEE Transactions on Terahertz Science and Technology 2016, 6 (1), 45–53. [8] Zakaria, H.A.; Fischer, B.M.; Bradley, A.P.; Jones, I.; Abbott, D.; Middelberg, A.P.J.; Falconer, R.J. Low-Frequency Spectro- scopic Analysis of Monomeric and Fibrillar Lysozyme. Applied Spectroscopy 2011, 65 (3), 260–64. [9] Ebbinghaus, S.; Meister, K.; Born, B.; DeVries, A.L.; Grue- bele, M.; Havenith, M. Antifreeze Glycoprotein Activity Correlates with Long-Range Protein-Water Dynamics J. Am. Chem. Soc. 2010, 132 (9), 12210–12211. [10] Xu, Y.; Havenith, M. Watching low-frequency vibrations of water in biomolecular recognition by THz spectroscopy The Jour- nal of Chemical Physics 2015, 143, 170901. [11] Cho, C.H.; Singh, S.; Robinson, G.W. Liquid Water and Bi- ological Systems: the Most Important Problem in Science That Hardly Anyone Wants to See Solved. Faraday Disc.1996, 103, 19- 27. [12] Roberts, C.J. Kinetics of Irreversible Protein Aggregation: Analysis of Extended Lumry-Eyring Models and Implications for Predicting Protein Shelf Life. The Journal of Physical Chemistry B 2003, 107 (5), 1194–1207. [13] Tsai, C.J.; Maizel, J.V.; Nussinov, R. The Hydrophobic Ef- fect: A New Insight from Cold Denaturation and a Two-State Wa- ter Structure. Critical Reviews in Biochemistry and Molecular Biol- ogy 2002, 37 (2), 55–69. [14] Ben-Amotz, D.; Underwood, R. Unraveling Water's En- tropic Mysteries: A Unified View of Nonpolar Polar, and Ionic Hy- dration. Accounts of Chemical Research 2008, 41 (8), 957–67. [15] Disalvo, E.A.; Lairion, F.; Martini, F.; Tymczyszyn, E.; Frías, M.; Almaleck, H.; Gordillo, G.J. Structural and Functional Proper- ties of Hydration and Confined Water in Membrane Interfaces. Biochimica Et Biophysica Acta (BBA) - Biomembranes 2008, 1778 (12), 2655–70. [16] Gallo, P.; Rovere, M. Structural Properties and Liquid Spi- nodal of Water Confined in a Hydrophobic Environment. Physical Review E 2007, 76 (6), 061202. [17] Szyperski, T.; Mills, J.L. NMR-Based Structural Biology of Proteins in Supercooled Water. J Struct Funct Genomics 2011, 12 (1), 1–7. [18] Prabhu, N.; Sharp, K. ProteinSolvent Interactions. ChemInform 2006, 37 (30), 200630298. Review of Biophysics and Biomolecular Structure 2005, 34 (1), 173– 99. [20] Teeter, M.M. Water-Protein Interactions: Theory and Ex- periment. Annu. Rev. Biophys. Biophys. Chem. 1991, 20 (1), 577– 600. [21] Schnepf, M.I. Protein-Water Interactions. In Biochemistry of Food Proteins, Hudson, B.J.P., Eds.; Springer: 1992, 1–33. [22] Falconer, R.J.; Markelz, A.G. Terahertz Spectroscopic Anal- ysis of Peptides and Proteins. Journal of Infrared Millimeter, and Terahertz Waves 2012, 33 (10), 973–88. [23] Vinh, N.Q.; Sherwin, M.S.; Allen, S.J.; George, D.K.; Rah- mani, A.J.; Plaxco, K.W. High-precision gigahertz-to-terahertz spectroscopy of aqueous salt solutions as a probe of the femtosec- ond-to-picosecond dynamics of liquid water. The Journal of Chemical Physics 2015, 142, 164502. [24] Heyden, M.; Tobias, D.J.; Matyushov, D.V. Terahertz Ab- sorption of Dilute Aqueous Solutions. The Journal of Chemical Physics 2012, 137 (23), 235103. [25] Xu, J.; Plaxco, K.W.; Allen, S.J. Collective Dynamics of Ly- sozyme in Water: Terahertz Absorption Spectroscopy and Com- parison with Theory. The Journal of Physical Chemistry B 2006, 110, 24255. [26] Ebbinghaus, S.; Kim, S.J.; Heyden, M.; Yu, X.; Heugen, U.; Gruebele, M.; Leitner, D.M.; Havenith, M. An Extended Dynam- ical Hydration Shell around Proteins. Proceedings of the National Academy of Sciences 2007, 104 (52); 20749–52. [27] Born, B.; Kim, S.J.; Ebbinghaus, S.; Gruebele, M.; Havenith, M. The terahertz dance of water with the proteins: the effect of protein flexibility on the dynamical hydration shell of ubiquitin. Faraday Discuss. 2009, 141, 161-173. [28] Bye, J.W.; Meliga, S.; Ferachou, D.; Cinque, G.; Zeitler, J.A.; Falconer, R.J. Analysis of the Hydration Water around Bovine Se- rum Albumin Using Terahertz Coherent Synchrotron Radiation. J. Phys. Chem. A 2014, 118 (1), 83–88. [29] Sushko, O.; Dubrovka, R.; Donnan, R.S. Sub-Terahertz Spectroscopy Reveals That Proteins Influence the Properties of Water at Greater Distances than Previously Detected. The Journal of Chemical Physics 2015, 142 (5), 055101. [30] Kuntz, I.D. Hydration of Macromolecules. III. Hydration of Polypeptides. J. Am. Chem. Soc. 1971, 93 (2), 514–16. [31] Svergun, D.I.; Richard, S.; Koch, M.H.J.; Sayers, Z.; Kuprin, S.; Zaccai, G. Protein Hydration in Solution: Experimental Obser- vation by x-Ray and Neutron Scattering. Proceedings of the Na- tional Academy of Sciences 1998, 95 (5), 2267–72. [32] Knab, J.; Chen, J.Y.; Markelz, A. Hydration Dependence of Conformational Dielectric Relaxation of Lysozyme. Biophysical Journal 2006, 90 (7), 2576–81. [33] Woods, K.N. Solvent-induced backbone fluctuations and the collective librational dynamics of lysozyme studied by te- rahertz spectroscopy. Physical Review E 2010, 81, 031915. [34] Yamamoto, N.; Ohta, K.; Tamura, A.; Tominaga, K. Broad- band Dielectric Spectroscopy on Lysozyme in the Sub-Gigahertz to Terahertz Frequency Regions: E ff ects of Hydration and Ther- mal Excitation. The Journal of Physical Chemistry B 2016, 120, 4743. [35] Vinh, N.Q.; Allen, S.J.; Plaxco, K.W. Dielectric Spectros- copy of Proteins as a Quantitative Experimental Test of Compu- tational Models of Their Low-Frequency Harmonic Motions. J. Am. Chem. Soc. 2011, 133 (23), 8942–47. [36] Laman, N.; Harsha, S.S.; Grischkowsky, D.; Melinger, J.S. High-Resolution Waveguide THz Spectroscopy of Biological Mol- ecules. Biophysical Journal 2008, 94, 1010. [19] Dill, K.A.; Truskett, T.M.; Vlachy, V.; Hribar-Lee, B. Mod- eling Water the Hydrophobic Effect, and Ion Solvation. Annual [37] Hebling, J.; Yeh, K.L.; Hoffmann, M.C.; Bartal, B.; Nelson, K.A. Generation of High-Power Terahertz Pulses by Tilted-Pulse- 6 Front Excitation and Their Application Possibilities. Journal of the Optical Society of America B 2008, 25 (7), B6. [38] Jewariya, M.; Nagai, M.; Tanaka, K. Enhancement of Te- rahertz Wave Generation by Cascaded  Processes in LiNbO3. Journal of the Optical Society of America B 2009, 26 (9), A101. [39] Blanchard, F.; Sharma, G.; Razzari, L.; Ropagnol, X.; Ban- dulet, H.C.; Vidal, F.; Morandotti, R.; Kieffer, J.C.; Ozaki, T.; Tie- dje, H.; Haugen, H.; Reid, M.; Hegmann, F. Generation of Intense Terahertz Radiation via Optical Methods. IEEE J. Select. Topics Quantum Electron. 2011, 17 (1), 5–16. [40] Fülöp, J.A.; Pálfalvi, L.; Almási, G.; Hebling, J. Design of High-Energy Terahertz Sources Based on Optical Rectification. Opt. Express 2010, 18 (12), 12311. [41] Hirori, H.; Tanaka, K. Single-Cycle Terahertz Pulses with Amplitudes Exceeding 1 MV/Cm Generated by Optical Rectifica- tion in LiNbO3 and Applications to Nonlinear Optics. Proc. SPIE 2012, 8240. [42] Gallot, G.; Zhang, J.; McGowan, R.W.; Jeon, T.I.; Grischkowsky, D. Measurements of the THz Absorption and Dis- persion of ZnTe and Their Relevance to the Electro-Optic Detec- tion of THz Radiation. Appl. Phys. Lett. 1999, 74 (23), 3450. [43] Wu, Q.; Zhang, X.C. Ultrafast Electro-Optic Field Sensors. Appl. Phys. Lett. 1996, 68 (12), 1604. [44] Casalbuoni, S.; Schlarb, H.; Schmidt, B.; Steffen, B.; Schmuser, P.; Winter, A. Numerical Studies on the Electro-Optic Sampling of Relativistic Electron Bunches. Proceedings of the 2005 Particle Accelerator Conference 2005, 1591367. [45] Kristensen, T.T.; Withayachumnankul, W.; Jepsen, P.U.; Abbott, D. Modeling Terahertz Heating Effects on Water. Opt. Ex- press 2010, 18 (5), 4727. [46] Krebs, M.R.H.; Wilkins, D.K.; Chung, E.W.; Pitkeathly, M.C.; Chamberlain, A.K.; Zurdo, J.; Robinson, C.V.; Dobson, C.M. Formation and seeding of amyloid fibrils from wild-type hen lyso- zyme and a peptide fragment from the β-domain Journal of Mo- lecular Biology 2000, 300 (3), 541-549. [47] Slocum, D.M.; Slingerland, E.J.; Giles, R.H.; Goyette, T.M. Atmospheric Absorption of Terahertz Radiation and Water Vapor Continuum Effects. Journal of Quantitative Spectroscopy and Ra- diative Transfer 2013, 127, 49–63. [48] Lee, B. Calculation of Volume Fluctuation for Globular Protein Models. Proceedings of the National Academy of Sciences 1983, 80 (2), 622–26. [49] Fredericks, W.J.; Hammonds, M.C.; Howard, S.B.; Rosen- berg, F. Density, thermal expansivity, viscosity and refractive in- dex of lysozyme solutions at crystal growth concentrations. Jour- nal of Crystal Growth 1994, 141, 183. [50] Nieuhaus, G.; Heyden, M.; Schmidt, D.A.; Havenith, M. Ex- ploring Hydrophobicity by THz Absorption Spectroscopy of Solv- ated Amino Acids. Faraday Discuss.2011, 150, 193. [51] Greenwald, J.; Riek, R. Biology of amyloid: structure, func- tion, and regulation. Structure 2010, 18 (10), 1244–1260. [52] Fichou, Y.; Schirò, G.; Gallat, F.X.; Laguri, C.; Moulin, M.; Combet, J.; Zamponi, M.; Härtlein, M.; Picart, C.; Mossou, E.; Lor- tat-Jacob, H.; Colletier, J.P.; Tobias, D.J.; Weik, M. Hydration Wa- ter Mobility Is Enhanced around Tau Amyloid Fibers. Proceedings of the National Academy of Sciences 2015, 112 (20), 6365–70. [53] Heyden, M.; Bründermann, E.; Heugen, U.; Nie, G.; Leitner, D.M.; Havenith, M. Long-Range Influence of Carbohydrates on the Solvation Dynamics of Water: Answers from Terahertz Ab- sorption Measurements and Molecular Modeling Simulations. J. Am. Chem. Soc. 2008, 130 (17), 5773–79. 7
1510.05171
2
1510
2015-11-11T13:01:35
Interplay of curvature-induced micro- and nanodomain structures in multicomponent lipid bilayers
[ "physics.bio-ph", "cond-mat.soft" ]
We discuss different mechanisms for curvature-induced domain formation in multicomponent lipid membranes and present a theoretical model that allows us to study the interplay between the domains. The model represents the membrane by two coupled monolayers, which each carry an additional order parameter field describing the local lipid composition. The spontaneous curvature of each monolayer is coupled to the local composition, moreover, the lipid compositions on opposing monolayers are coupled to each other. Using this model, we calculate the phase behavior of the bilayer in mean-field approximation. The resulting phase diagrams are surprisingly complex and reveal a variety of phases and phase transitions, including a decorated microdomain phase where nanodomains are aligned along the microdomain boundaries. Our results suggest that external membrane tension can be used to control the lateral organization of nanodomains (which might be associated with lipid "rafts") in a multicomponent lipid bilayer.
physics.bio-ph
physics
Interplay of curvature-induced micro- and nanodomain structures in multicomponent lipid bilayers Leonie Brodbek · Friederike Schmid 5 1 0 2 v o N 1 1 ] h p - o i b . s c i s y h p [ 2 v 1 7 1 5 0 . 0 1 5 1 : v i X r a Received: date / Accepted: date Abstract We discuss different mechanisms for curvature- induced domain formation in multicomponent lipid mem- branes and present a theoretical model that allows us to study the interplay between the domains. The model represents the membrane by two coupled monolayers, which each carry an additional order parameter field describing the local lipid composition. The spontaneous curvature of each monolayer is coupled to the local com- position; moreover, the lipid compositions on oppos- ing monolayers are coupled to each other. Using this model, we calculate the phase behavior of the bilayer in mean-field approximation. The resulting phase dia- grams are surprisingly complex and reveal a variety of phases and phase transitions, including a decorated mi- crodomain phase where nanodomains are aligned along the microdomain boundaries. Our results suggest that external membrane tension can be used to control the lateral organization of nanodomains (which might be associated with lipid "rafts") in a multicomponent lipid bilayer. Keywords Lipid bilayers · Multicomponent mem- branes · Lipid Rafts · Ginzburg-Landau theory · Elasticity · Curvature 1 Introduction Biomembranes are not homogeneous [1]. They consist of a self-assembled lipid bilayer which acts as a support for L. Brodbek Institut fur Physik, Johannes Gutenberg-Universitat Mainz, DE F. Schmid Institut fur Physik, Johannes Gutenberg-Universitat Mainz, DE E-mail: [email protected] proteins and other biomolecules [2]. In the last decades, there is increasing evidence that biomembranes are lat- erally structured [1], and this structure is believed to be important for their biological function. For example, the so-called lipid raft hypothesis states that biomembranes are often filled with nanodomains, which typically have sizes between 10 to 100 nm, a higher cholesterol con- tent and higher local order [3,4,5,6,7,8,9]. Indirect evi- dence for the existence of such rafts is provided, e.g., by superresolution images showing nanoscopic clusters of raft-associated proteins in membranes [10,11,12]. How- ever, the main forces driving this internal organization of membranes have not yet been identified unambigu- ously. In particular, the role of the lipids and the ques- tion whether and how they contribute to the structuring of membranes is still discussed controversially. On the one hand, it is argued that the formation of nanoscopic or larger protein clusters in membranes could be driven by the proteins alone. On the other hand, nano- and microstructures have also been ob- served in pure lipid membranes. Already one-component phospholipid membranes exhibit a modulated "ripple phase" in the transition region between the high tem- perature fluid phase and the low-temperature gel phase [13,14,15,16,17,18]. Experimental evidence for nanodomains with properties similar to those attributed to rafts has been provided by atomic force microscopy [19] and neu- tron scattering [20,21,22,23]. On larger scales, multi- component giant vesicles were found to feature ordered patterns of micron-size domains under certain circum- stances [24]. Lateral phase separation was observed in model multicomponent vesicles [25,26,27] and in plasma vesicles extracted from living rat cells [28], and large up to micron-size critical clusters could be visualized in the vicinity of critical points [19,28,29,30,31]. 2 Leonie Brodbek, Friederike Schmid Several physical mechanism have been proposed that may generate micro- or nanostructures in pure lipid membranes [32]: One, mentioned above, is critical clus- ter formation in the vicinity of a demixing phase transi- tion [19,29,31]. A second suggestion is that line-active molecules in multicomponent mixtures may reduce the line tension and eventually turn an originally phase- separated mixture into a microemulsion [33,34,35,36, 37,38,39,40]. In the present paper, we focus on a third generic class of domain-stabilizing mechanism in mem- branes: The formation of microemulsions or modulated structures due to curvature-induced elastic interactions. Two variants of such a mechanism have been pro- posed in the literature. In the first (termed I hereafter), it is assumed that the two opposing leaflets tend to have different lipid composition, and that this imposes a spontaneous curvature on the bilayer as a whole [41]. A tensionless membrane responds by bending around [42]. If tension is applied, it is forced to be planar, and this creates elastic stress. The membrane reacts by forming staggered domains with curvatures of opposite sign. This effect is illustrated in Fig. 1a). It was first proposed by Andelman and coworkers and further in- vestigated by a number of authors [41,42,43,44,45,46, 47,48]. Schick and coworkers argued that it might ac- count for raft formation in membranes [45,46,47]. The characteristic length scales of the domains are in the range of 100 nm to micrometers and diverge for ten- sionless membranes. The second mechanism (termed II) was recently pro- posed in our group [18,49]. It assumes that the lipids tend to demix laterally, but the lipid composition on opposing leaflets are preferably in registry. Different lipid compositions are associated with different spon- taneous curvature parameters in the monolayer. Since the membrane as a whole remains planar, this also cre- ates elastic stress, which is relieved by keeping domain sizes finite. The effect is illustrated in Fig. 1b). Charac- teristic length scales according to this mechanism are in the range of 10 nm and depend on the membrane ma- terial. We have argued that the same effect could also account for the formation of modulated ripple states, which also have characteristic wave lengths in the same order of magnitude, hence ripples and nanodomains in lipid bilayers could be closely related phenomena. Both mechanisms have in common that they are driven by a coupling between lipid composition and spontaneous curvature. The main difference lies in the nature of the local correlation between the lipid com- positions on the two leaflets. Mechanism I assumes that domains are anticorrelated, which favors microdomains. Mechanism II assumes that domains are in registry, which favors nanodomains. Experimentally, it has been (a) (b) (c) microdomains Dµ (mechanism I) nanodomains nD (mechanism II) (d) (e) disordered asymmetric DIS ASY demixed 2Φ Fig. 1 Curvature-induced domains in mixed bilayers: (a) microdomains formed by mechanism I [41, 43, 45] (b) nan- odomains formed by mechanism II [18, 49]. Panels (c-e) show homogeneous membrane states that are also considered in this paper: (c) symmetric disordered state, (d) asymmetric state (e) lateral phase separation reported that domains on opposing leaflets tend to be in registry [50]. However, the correlation is never perfect, and we shall see below that microdomains may be ener- getically favorable under certain circumstances even in systems where the coupling between lipid compositions on opposing leaflets is weakly positive. The purpose of the present paper is to provide a uni- fied picture of curvature-induced phenomena in mixed lipid bilayers. We propose an elastic theory which re- produces both the mechanisms I and II of domain for- mation. Using mean-field approximation, we can assess the influence of membrane tension, curvature coupling, and composition coupling across leaflets on the bilayer structure, and calculate a representative set of phase diagrams which demonstrate the complex interplay be- tween microdomain and nanodomain formation in these systems. Our paper is organized as follows. In the next sec- tion, we introduce the framework of our theoretical ap- proach and derive a Ginzburg-Landau model which de- scribes both nanodomain and microdomain formation. In Section 3 we present and analyze the possible phase diagrams of this system, which we have calculated ana- lytically within mean-field approximation and a varia- tional single-mode Ansatz. The results are summarized and discussed in Section 4. 2 Theoretical framework Our theory is based on an elastic model for membranes that describes the lipid bilayer by a system of two cou- pled elastic monolayers [51,52,53,54]. Each monolayer is represented by a two-dimensional manifold that de- scribes the position of the monolayer-water interface in Interplay of curvature-induced micro- and nanodomain structures in multicomponent lipid bilayers 3 a coarse-grained sense. We assume that the membrane is planar on average and aligned in the (x, y) plane, and that bubbles and overhangs can be neglected. Each monolayer surface can then be parametrized by a func- tion zi(x, y) (where i = 1, 2 and z1 < z2). We consider a symmetric situation where both monolayers have the same elastic properties, except that their spontaneous curvatures ci may vary locally and differ from each other. We adopt a sign convention according to which ci is positive if the outer surface of the monolayer has a tendency to bend inwards, towards the membrane in- terior. Furthermore, we use a Gaussian approximation, i.e., we expand the free energy up to second order of zi about a fully planar reference state. The total elastic energy of the coupled monolayer system is then given by [53,54,55,56] Fel = Z d2r n kA 8t2 0 + kc ζ 2t0 kc 4 (cid:2)(∆z1)2 + (∆z2)2(cid:3) (1) (z1 − z2)2 + (z1 − z2)(cid:2)∆z1 − ∆z2(cid:3) 2 (cid:2) Γ 1 + kc(cid:2)c1∆z1 − c2∆z2(cid:3) + 2∇(z1 + z2)(cid:3)2o Here we have not included the contribution of the Gaussian curvature, which is a constant for planar mem- branes with fixed topology [57]. The parameter t0 is the mean monolayer thickness, kA and kc are the (bilayer) compression and bending modulus, respectively, ζ is a curvature-related parameter [53,54,55], and Γ the ten- sion of the membrane [56]. We must briefly comment on the interpretation of the parameter Γ . It has been noted [58] and confirmed by simulations [59,60,61] that the externally applied tension (the "frame tension") and the tension experi- enced by the lipids inside the membrane (the "intrinsic" tension) differ slightly from each other in fluctuating membranes. However, fundamental symmetry consid- erations [62,63] suggest that the "tension" parameter governing the amplitude of membrane undulations co- incides with the frame tension in the thermodynamic limit, except for a small multiplicative correction fac- tor that accounts for the difference between the actual area and the area projected on the (x, y)-plane [64]. This implies that in a mean-field theory based on a quadratic approximation such as (1), the parameter Γ is most appropriately interpreted as a frame tension. Several simulations have confirmed that frame tension and fluctuation tension are equal [60,63,65,66,67], but deviations have also been reported, especially for low tensions [59,68,69,70]. If the membrane tension is close to zero, the existence of the thermodynamic limit for which the theoretical arguments [62,63] would apply becomes questionable. One consequence is that the am- plitude of undulations may depend on the statistical ensemble under consideration [60]. However, these ef- fects are small and we can neglect them in the context of this work. Hence we identify Γ with a frame tension, which can be applied and controlled externally. We will also neglect the fact that the material pa- rameters of the membrane may change under tension [56,71,72]. Our elastic model does not explicitly ac- count for the effect of lipid orientation and possible lipid tilt [73,74,75,76,77]. A detailed consideration of such factors might help to establish a molecular basis for the relation between the elastic parameters and the lipid structure [78,79]. Here we wish to keep our model as simple as possible. We emphasize that Eq. (1) contains all terms up to second order of zi and the spatial deriva- tives that are allowed by symmetry in a system of two coupled identical monolayers described by two surfaces zi [80] (apart from the contribution of the Gaussian cur- vature). We have used the elastic model, Eq. (1), to fit deformation profiles in the vicinity of inclusions, with good results down to molecular length scales [55,81]. To describe lateral phase separation and domain for- mation within the monolayers, we supply each mono- layer with an additional order parameter field ϕi(x, y). Here ϕ is a collective variable designed to characterize the local lipid composition in a multicomponent system -- not necessarily a binary system -- with a propensity to locally phase separate. The associated free energy is described by a Ginzburg-Landau functional FGL = Z d2rn 2 Xi=1 (cid:2) g 2 (∇ϕi)2+ t 4 ϕ2 i + v 8 ϕ4 i(cid:3)− s 2 ϕ1ϕ2o(2) with v > 0. The last term couples the compositions on the two leaflets and is constructed such that it favors equal composition for s > 0 (positive coupling) and different composition for s < 0 (negative coupling). In the absence of any coupling, lateral phase separation occurs for t < 0 and the monolayers remain homoge- neous (ϕi ≡ 0) for t > 0. Hence the parameter t is temperature-like and describes the distance from the critical demixing transition in parameter space. Thus far, we have combined standard theories for fluid elastic sheets and phase separating order param- eter fields. The key additional feature that we must introduce to describe curvature-induced domain forma- tion is a coupling between the order parameters ϕi(x, y) and the local spontaneous curvature of the monolayer ci(x, y). Kollmitzer et al. [79] recently discussed how the monolayer curvature might depend on the mem- brane composition for raft-forming lipids. Specifically, 4 Leonie Brodbek, Friederike Schmid we assume that the relation between ci and ϕi is roughly linear, ci(x, y) = c0 + c ϕi(x, y), (3) where c0 is the spontaneous curvature in the mixed ho- mogeneous system. Eqs. (1) - (3) define our theoretical model. To pro- ceed, we express the degrees of freedom in terms of a set of new fields that make the symmetries in the system more transparent: The position of the membrane mid- plane, h(x, y) = 1 2 (z1 + z2), the variations of the local mean monolayer thickness, u(x, y) = 1 2 (z1 − z2) − t0, the mean local order parameter Φ = 1 2 (ϕ1 + ϕ2), and the local order parameter difference between monolayer leaflets Ψ = 1 2 (ϕ1 − ϕ2). Rewritten as a functional of these new fields, the total free energy reads u2 + kc 2 (∆u)2 + 2kc ζ t0 u∆u (4) 0 + 2t2 F = Z d2r n kA Γ (∇h)2 + 2 g (∇Φ)2 + 2 g (∇Ψ )2 + 2 + + (∆h)2 kc 2 1 (t − s)Φ2 + 2 1 (t + s)Ψ 2 + 2 v 4 v 4 Φ4 Ψ 4 + 2kcc [Φ∆u + Ψ ∆h] + 3 2 v Φ2Ψ 2o. . In this representation, it becomes clear that the two sets of fields (u, Φ) and (h, Ψ ) are largely independent of each other. The only coupling between them is intro- duced by the last term in Eq. (4), a nonlinear fourth order term. The domain forming mechanism I is asso- ciated with modulations in the fields (h, Ψ ) and mech- anism II is associated with modulations in (u, Φ) (see Fig. 1). Since the free energy (4) is quadratic in h and u, these degrees of freedom can be integrated out. Alterna- tively, we may adopt a mean-field approximation from the outset and minimize the free energy, Eq. (4), with respect to h and u. Apart from uninteresting constants, the result is the same and best written in a mixed Fourier- and real space representation (omitting con- stants): with KΦ(x) = t − s + KΨ (x) = t + s + 0 g ξ2 g ξ2 Γ x2 x − 4kcc2 x − 4kcc2 x (1 − x)2 + Λx 1 + x . (6) (7) scales ξ0 = (t2 Here we have introduced the characteristic length 0kc/kA)1/4 and ξΓ = (kc/Γ )1/2 and the dimensionless parameter Λ = 2−4ζpkc/kA. The length scale ξ0 and the parameter Λ are intrinsic properties of the membrane. Experimentally, one finds that the bend- ing rigidity, kc, is roughly proportional to kAt2 0, hence ξ0 should be of the order of the membrane thickness, 2t0. The parameter Λ characterizes the bilayer thick- ness profiles in the vicinity of a local perturbation [53, 55]: For Λ ≥ 4, they decay towards the equilibrium thickness in a purely exponential manner; for Λ < 4, an oscillatory component emerges, and the membrane becomes unstable towards deformations at Λ < 0. In- serting numbers from experiments [82], all-atom sim- ulations [54,83,84], or coarse-grained simulations [55] for the fluid phase of DPPC bilayers (dipalmitoyl phos- phatidylcholine, of one of the most common lipids in natural biomembranes), one consistently obtains val- ues around ξ0 ∼ (0.9 − 1.4) nm, and Λ ∼ 0.7 [49]. With typical experimental values for the tension [85,86] in the range of Γ ∼ 10−5N/m, one finds that ξΓ is of the order ∼ 100 nm. It diverges for tensionless membranes. -- The parameter Λ, which characterizes the membrane material. It must be positive, otherwise the mem- brane is not stable -- The rescaled curvature coupling C = 2c ξ0pkc/g -- The rescaled composition coupling s = s ξ2 0 /g -- The rescaled temperature parameter t = t ξ2 -- The rescaled tension Γ = Γ t0/√kc kA In this notation, the characteristic length scale ξΓ is 0 /g simply given by ξΓ = ξ0 /p Γ . Our task is to minimize the free energy (5) with respect to the fields Φ and Ψ , and to calculate the re- sulting phase diagrams. This is done in the next section. 3 Phase behavior of mixed bilayers 3.1 Stability analysis F = (2π)2 2 A Xq n 1 KΦ(cid:0)(ξ0 q)2(cid:1)Φq2 KΨ(cid:0)(ξΓ q)2(cid:1)Ψq2o 4 Z d2r n(cid:0)Φ4 + Ψ 4(cid:1) + 6 Φ2Ψ 2o 1 2 + v + (5) At high "temperature" t, the system is disordered, Φ = Ψ ≡ 0. The disordered state is unstable towards or- dering or phase separation if either KΦ(x) or KΨ (x) become negative for at least one x > 0. Hence the in- stability limit is determined by the maximum value of t at which the minimum or KΦ or KΨ becomes zero. Four types of instabilities are possible: Interplay of curvature-induced micro- and nanodomain structures in multicomponent lipid bilayers 5 (i) Instability of KΦ(x) at x = 0. This corresponds to an instability with respect to macroscopic demixing (state 2Φ in Fig. 1e) (ii) Instability of KΦ(x) at some x > 0. This corre- sponds to an instability with respect to a modu- lated state characterized by a periodic array of nan- odomains (state nD in Fig. 1b) (iii) Instability of KΨ (x) at x = 0. This corresponds to an instability with respect to the formation of a globally asymmetric membrane (state ASY in Fig. 1d) (iv) Instability of KΨ (x) at some x > 0. This corre- sponds to an instability with respect to a modulated phase with a periodic array of microdomains (state µD in Fig. 1a) The instabilities of KΨ ((iii) and (iv)) have been discussed by Shlomovitz and Schick [46]. At low cur- vature coupling C, the instability (iii) dominates and the membrane becomes asymmetric at t = −s. At high C, the instability (iv) takes over, and modulated mi- Γ −1/4( C − crodomains with a wavelength λΨ = 2πξ0 p Γ )−1/2 emerge for t < tΨ := −s + ( C −p Γ )2. (8) can expand both equations in powers of δ, which leads to the expansion tΦ − s = ǫ(cid:2)1 + Λ 4 (ǫ − ǫ2) + Λ 16 (Λ + 4)ǫ3 + O(ǫ4)(cid:3). (11) Far from the bicritical point, for large ǫ, one obtains the asymptotic behavior Λ for Λ > 2. tΦ − s → ǫ (cid:26) 1 + Λ/(4 − Λ) for Λ < 2 Hence tΦ − s is roughly proportional to ǫ in the whole range of C. Below, we will use the approximate expres- sion derived from the leading order of the series expan- sion, (12) 1 Λ (13) ( C 2 − Λ). tΦ ≈ s + ǫ = s + In the total system of coupled order parameters, both the instabilities in Φ and Ψ can destroy the homoge- neous membrane, and the membrane enters the phase associated with the instability at highest "temperature" t. Within the ordered or metastable phase, the nonlin- ear terms in the free energy expression, Eq. (5), become important, and we must include them to calculate the full phase diagram. This is done in the next section. The two regimes are separated by a Lifshitz critical point at C = p Γ , at which point the wavelength λΨ of the modulations diverges. The picture that one obtains after analyzing the in- stabilities of KΦ ((i) and (ii)) is similar, but the result- ing scenario differs from that described above in one important aspect. At low curvature coupling C, the in- stability (i) dominates and the membrane phase sepa- rates at t = s. At high C, the instability (ii) dominates, and a modulated nanodomain phase emerges. However, the wavelength of this phase remains finite. The two regimes are connected by a bicritical point at C = √Λ connecting a line of Ising-type transitions (regime (i)) and Brazovskii-type transitions [87] (regime (ii)), and the wavelength of the modulated phase at this point is λΦ = 2πξ0 . The transition point tΦ between the mod- ulated and the disordered phase at C > √Λ is defined through a set of implicit equations 3.2 Phase diagrams In the present work, we are not interested in the exact numerical solution of the free energy model, (4), but rather in a qualitative picture of the interplay of the different ordering mechanisms in the membrane. There- fore, we calculate the mean-field phase diagram within a single mode approximation which provides us with analytical expressions for the phase boundaries. Specif- ically, we make the following Ansatz for the shape of the order parameter field: Φ(r) = Φm cos(kΦz)+Φs, Ψ (r) = Ψ m cos(kΨ z)+Ψ s, (14) where kΦ = 2π/λΦ and kΨ = 2π/λΨ are the most un- stable wavevectors for given C and Γ calculated in Sec. 3.1. Inserting this Ansatz in the functional (5), we ob- tain the free energy per area ǫ := C 2 Λ − 1 = δ4 + 2δ(1 + δ)2Λ Λ(1 + δ)(Λ(1 + δ) − 2δ) tΦ = s + δ(1 + δ)(2 + δ) Λ(1 + δ) − 2δ , where δ is related to the characteristic wavelength λΦ at the transition via λΦ = 2πξ0 /√1 + δ and vanishes at the bicritical point. Close to the bicritical point, we (9) (10) F/A = 2 v 4nΦ 2 s(t − s) s(t + s) 2 2 m(t − tΦ) + 2Φ m(t − tΨ ) + 2Ψ +Ψ 3 (Φ m) + Φ 8 3 2 m + 3(Φ m + Ψ mΨ + + Φ 4 2 2 2 4 4 4 s + Ψ s + 3(Φ mΦ 2 2 2 2 2 mΨ s + Φ sΨ m) + 6Φ sΨ 2 s)o, (15) 2 s + Ψ 2 mΨ 2 s) 6 Leonie Brodbek, Friederike Schmid (a) tŽ 4 3 2 1 0 -1 DIS 2Φ nD (b) tŽ 3 2 1 0 -1 -2 DIS 2Φ µD nD (c) tŽ 3 2 1 0 -1 -2 0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5 Ž C µ D nD Ž C µ D (e) tŽ 3 2 1 0 -1 -2 DIS 2Φ (d) tŽ 4 3 2 1 0 -1 DIS 2Φ 0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5 Ž C Ž C DIS nD µ D ASY 0.5 0.0 1.0 1.5 2.0 2.5 Ž C DIS µD (f) tŽ 3 2 1 0 -1 -2 ASY 0.5 0.0 1.0 1.5 2.0 2.5 Ž C Fig. 2 Typical phase diagrams in the plane of "temperature" t and curvature coupling C for strong, weak, and negative monolayer composition coupling (s = 1, 0.1, −0.1) and two choices of the membrane parameter Λ, one with Λ < 1 and one with Λ > 1: (a) s = 1, Λ = 0.7; (b) s = 0.1, Λ = 0.7; (c) s = −0.1, Λ = 0.7; (d) s = 1, Λ = 1.4; (e) s = 0.1, Λ = 1.4; (f) s = −0.1, Λ = 1.4. The dimensionless tension is Γ = 10−4 in all cases, corresponding to a characteristic length scale for microdomains of ξΓ = 100ξ0 . The membrane structures in the different phases correspond to those shown in Fig. 1. Solid gray lines correspond to continuous transitions, black lines to first order transitions, and dashed grey lines transition regions between a a pure µD phase (at high t) and a decorated µD phase which also contains nanodomains at the microdomain boundaries (at lower t). The red arrow indicates the position of a Lifshitz point. which we can now minimize with respect to the amplitudes Φm, Φs, Ψ m, Ψ s in a straightforward man- ner. We find that states with minimal free energy can only sustain one type of order, i.e., all amplitudes are zero except, possibly, one. Transitions between different states are first order. These predictions are consistent with the known be- havior near Lifshitz critical points and seem reasonable in many other aspects as well. However, they clearly do not capture the interplay of microdomains and nan- odomains in the limit ξΓ /ξ0 → 0 or Γ ≪ 1 when the characteristic lengths of microdomains and nan- odomains are very different. If the order parameter Ψ (r) varies very slowly compared to Φ(r), it acts almost like a constant field on Φ, and it is clear that nanodomains might emerge in regions with Ψ ≈ 0 even if the global amplitude of Ψ modulations does not vanish. Hence an adiabatic approximation seems more appropriate for this case, where Φm is allowed to vary in space and to depend on the local value of Ψ 2. Let us first assume that we can impose a given con- stant asymmetry Ψ on a mixed membrane with Φ ≡ 0. 3 (tΦ− t), the free energy (15) can then As long as Ψ 2 < 1 be lowered by introducing a periodic modulation with 3 (t − tΦ + 3Ψ 2). The asso- squared amplitude Φ ciated free energy gain per area is Fm/A = − 3v 32 Φ m. If we now consider a modulated microdomain phase µD with very slowly varying order parameter Ψ (r), we can m = − 4 4 2 v lower the free energy by allowing for the formation of nanodomains at the interfaces between microdomains, and the associated adiabatic free energy can be approx- imated as Fad/A = − However, the membrane is not entirely filled with nan- odomains in this state, the nanodomains only build up in narrow regions close to the microdomain boundaries. The transition between the decorated µD and the nD phase remains discontinuous. 6 Z d2r h min(cid:0)t − tΦ + 3Ψ (r)2, 0(cid:1)i2 (16) . Combining all these results, we can now finally cal- culate the phase behavior of the coupled membrane system. Some representative phase diagrams for mem- branes with low tension are shown in Fig. 2. They fea- ture the disordered phase at high "temperatures" t and a variety of ordered and modulated states at low t. At low curvature coupling C, the low-temperature state is a homogeneous membrane, which is either phase sep- arated (2Φ) for positive composition coupling s across leaflets (Fig. 2a,b,d,e), or asymmetric (ASY) for nega- tive coupling s (Fig. 2c,f). At high curvature coupling C, both the modulated microdomain phase µD and the nanodomain phase nD may appear. Not surprisingly, the nanodomain phase tends to be favored if the com- position coupling s between leaflets is strongly positive (Fig. 2a,d). Another, less obvious factor that selects between modulated phases is the value of the mem- brane parameter Λ. This can be related to the different Interplay of curvature-induced micro- and nanodomain structures in multicomponent lipid bilayers 7 (a) DIS (d) DIS nD nD (b) DIS µD nD 2Φ (e) DIS µ D nD ASY (c) (f) DIS µ D 2Φ µD DIS ASY Fig. 3 Typical phase diagrams in the plane of "temperature" t and dimensionless tension Γ for Λ = 0.7 and different choices of monolayer composition coupling s curvature coupling C. (a) s = 0.1, C = 1.5; (b) s = 0.1, C = 1.; (c) s = 0.1, C = 0.75; (d) s = −0.1, C = 2.; (e) s = −0.1, C = 1.3; (f) s = −0.1, C = 0.75; Solid gray lines correspond to continuous transitions, black lines to first order transitions, and red arrow indicates the position of a Lifshitz point. slopes of tΦ( C) and tΨ ( C). For small tensions, Γ ≪ 1, and close to the bicritical point of Φ at C = p Γ , the slopes are given by dtΦ/d C = 2 C/ Γ and dtΨ /d C ≈ 2 C. Hence the nanodomain state can supersede a mi- crodomain state as C increases if Λ < 1 (Fig. 2a-c), the microdomain state will remain dominant for Λ > 1 (Fig. 2d-f). We recall that phospholipid bilayers have values of Λ around 0.7. If this finding is representative for lipid bilayers, nanodomains would tend to be favored over microdomains at large curvature mismatch. Fig. 2b-f) also indicates regions (in the µD phase below the dashed line), where we believe that nan- odomains may exist within microdomains, based on the adiabatic approximation described above. Such states should be particularly interesting since they are filled with lines of nanodomains, either regularly ordered or -- in case fluctuations destroy the global order -- in a foam-like structure. Next we study the influence of tension on the phase behavior. The results for selected parameter sets are shown in Fig. 3. As a rule, applying tension destabilizes the microdomain phase. This is also reported in Ref. [46]. With increasing tension, the temperature range where the microdomain phase is stable decreases and it is eventually replaced by the disordered structure, a nanodomain state, or a homogeneous membrane struc- ture. The phase transitions between the other states are not affected by membrane tension. We should note, however, that the tensions needed to bring about a phase transition are rather extreme. The tension unit used to rescale the dimensionless ten- sion parameter Γ is Γ/ Γ = √kckA/t0, which corre- sponds to (50-100) mN/m for typical experimental val- ues of the elastic parameters kc and kA/t2 0 of phospho- lipid bilayers[82]. The tension strength where lipid bi- layers rupture is typically around 2-10 mN/m [88]. On short time scales, higher tensions up to 20-30 mN/m can be sustained [89]. Still, the membranes will rupture for most of the tensions considered in Fig. 3. Also, it should be noted that membranes undergo significant structural rearrangement under high tension [56], which will likely interfere with domain formation. Hence, the phase di- agrams shown in Fig. 3 are of rather academic interest with probably little practical relevance. The main con- trol that one can exert by applying tension, according to our model, is to modulate the characteristic length scale of the microdomains. This could again be inter- esting in the mixed microdomain / nanodomain state, because it implies that tension can be used to manipu- late the structural arrangements of microdomains. 4 Summary and discussion In sum, we have presented a theoretical framework that allows us to describe two curvature-driven mechanisms for the formation of modulated structures in multi- component lipid bilayers in a unified manner: A mi- crodomain structure associated with a staggered com- position profile on the two monolayers, and a nanodomain structure associated with domains that oppose each other on both monolayers. We have seen that both microdomain and nanodomain structures can be ob- 8 Leonie Brodbek, Friederike Schmid served for both negative and positive couplings between the lipid compositions on opposing monolayer leaflets. However, a strong positive coupling favors nanodomain formation and a strong negative coupling favors mi- crodomain formation. We have also studied the influ- ence of membrane tension and found, in agreement with Ref. [46], that microdomains are destabilized at high tensions. In our opinion, however, the most important effect of tension is that it can be used to control the characteristic wave length of the microdomains, Our mean-field analysis revealed an unexpectedly complex phase behavior. It includes two types of Bra- zovskii-transitions and Ising transitions, bicritical points where Brazovskii lines meet with Ising lines or with each other, and possibly one Lifshitz point. We have consid- ered the simplest possible system, a perfectly symmet- ric bilayer made of identical monolayers and a Ginzburg Landau free energy that is perfectly symmetric in the order parameter. These are the systems most often stud- ied in laboratory experiments. The possible scenarios will be even more complicated if we consider asymmet- ric systems where, e.g., one order parameter is favored -- which will create a competition between modulated phases with different symmetry [43] -- or the lipid com- position is different on both sides of the membranes [46]. Furthermore, we have assumed that only the spon- taneous curvature is coupled to the local order param- eter. In reality, one would expect all elastic parame- ters to depend on the local lipid composition [90]. In many cases, this will probably change the picture only in a quantitative sense, phase boundaries and length scales will be shifted. However, qualitatively new be- havior may emerge if the Gaussian modulus becomes composition dependent. In that case, the contribution of the Gaussian curvature to the elastic free energy can no longer be neglected, and a new type of modulated phase may emerge that might be interesting in the con- text of membrane fusion. In our mean-field approach, we have neglected fluc- tuations. They should have a severe impact on the phase diagrams. It is known that thermal fluctuations turn a continuous Brazovskii transition into a weak first order transition [87,91], and they destabilize Lifshitz points, such that a microemulsion channel may open up and/or part of the Ising critical line may become first order [47, 48]. Additionally, we have presented a simple scaling ar- gument in Ref. [49] according to which nanodomain for- mation should preempt homogeneous demixing for all values of the curvature coupling C. The implications of this effect for the topology of the phase diagrams is still unclear. Most likely the Ising-type demixing transition at low C will persist. As mentioned in the introduction, Ising-type demixing transitions have been observed in multicomponent membranes and the Ising critical ex- ponents have been verified with great care [29,30,92]. However, the coexisting phases may be structured and contain a certain amount of nanodomains. The Bra- zovskii phases will become disordered in the vicinity of the mean-field Brazovskii line and be replaced by a rel- atively strongly structured microemulsion over a wide parameter range. Thus the lateral structure of the membrane is pre- sumably characterized by a combination of disordered micro- and nanodomains, where the nanodomains tend to accumulate along the borders of microdomains. Given their nanoscale size, the nanodomains can possibly be related to the lipid "rafts" discussed in the introduction. Our results indicate that microdomains can be used to manipulate rafts, but not vice versa. Since the size of the microdomains can be tuned by varying the applied tension, this reveals a way how the larger lateral or- ganization of nanodomains or rafts could be controlled externally by a physical process. According to the raft hypothesis, raft proteins use small lipid domains as templates, but they sometimes assemble to larger units to be functional [7]. We have seen how membrane tension could be used to assist such a process. However, our results rely on a mean-field theory and a number of further simplifying approxima- tions. Studying the multicomponent system in detail by numerical simulations that also include thermal fluctu- ations will be the subject of future work. Acknowledgements The ideas presented in this paper are based on previous work, mostly simulations, that were carried out by Stefan Dolezel, Gerhard Jung, Olaf Lenz, Sebastian Meinhardt, Jorg Neder, and Beate West. These simulations have given us trust in the coupled monolayer model which on which the present model is built. We also wish to thank Frank Brown, Laura Toppozini, Maikel Rheinstadter, and Richard Vink for collaborations that have helped to shape our view on lipid bilayers. We thank in particular Michael Schick for helpful comments on the manuscript and for pointing out Refs. [85, 86, 90]. References 1. Vereb, G., Szollosi, J., Matko, J., Nagy, P., Farkas, T., Vigh, L., Matyus, L., Waldmann, T.A., Damjanovich, S.: Dynamic, yet structured: The cell membranes three decades after the singer-nicolson model. PNAS 100, 8053 -- 8058 (2003) 2. Singer, S.J., Nicolson, G.K.: Fluid Mosaic Model of Struc- ture of Cell-Membranes. Science 175(4023), 720 -- 731 (1972) 3. Ahmed, S.N., Brown, D.A., London, E.: On the ori- gin of sphingolipid/cholesterol-rich detergent-insoluble cell membranes: Physiological concentrations of choles- terol and sphingolipid induce formation of a detergent- Interplay of curvature-induced micro- and nanodomain structures in multicomponent lipid bilayers 9 insoluble, liquid-ordered lipid phase in model membranes. Biochemistry 36(36), 10,944 -- 10,953 (1997) 4. Simons, K., Ikonen, E.: Functional rafts in cell mem- branes. Nature 387, 569 -- 572 (1997) 5. Brown, D.A., London, E.: Structure and origin of ordered lipid domains in biological membranes. J. Membr. Biol- ogy 164(2), 103 -- 114 (1998) 6. Pike, L.: Lipid rafts: Bringing order to chaos. J. Lip. Res. 44, 655 -- 667 (2003) 7. Pike, L.: Rafts defined: A report on the keystone sym- posium on lipid rafts and cell function. J. Lip. Res. 47, 1597 -- 1598 (2006) 8. Leslie, M.: Do lipid rafts exist? Science 334, 1046 -- 1047 (2011) 9. Lingwood, D., Simons, K.: Lipid rafts as a membrane- organizing principle. Science 327, 46 -- 50 (2010) 10. Eggeling, C., Ringemann, C., Medda, R., Schwarzmann, G., Sandhoff, K., Polyakova, S., Belov, V.N., Hein, B., von Middendorf, C., Schonle, A., Hell, S.W.: Direct ob- servation of the nanoscale dynamics of membrane lipids in a living cell. Nature 457, 1159 -- 1162 (2009) 11. Mizuno, H., Abe, M., Dedecker, P., Makino, A., Rocha, S., Ohno-Iwashita, Y., Hofkens, J., Kobayashi, T., Miyawaki, A.: Fluorescent probes for superresolution imaging of lipid domains on the plasma membrane. Chemical Science 2, 1548 -- 1553 (2011) 12. Owen, D.M., Magenau, A., Williamson, D., Gaus, K.: The lipid raft hypothesis revisited -- new insights on raft composition and function from super-resolution fluores- cence microscopy. Bioessays 34, 739 -- 747 (2012) 13. Koynova, R., Caffrey, M.: Phases and phase transitions of the phosphatidylcholines. Biochimica et Biophysica Acta - Reviews on Biomembranes 1376, 91 -- 145 (1998) 14. Koynova, R., Caffrey, M.: An index of lipid phase dia- grams. Chemistry and Physics of Lipids 115, 107 -- 219 (2002) 15. Katsaras, J., Tristram-Nagle, S., Liu, Y., Headrick, R.L., Fontes, E., Mason, P.C., Nagle, J.F.: Clarification of the ripple phase of lecithin bilayers using fully hydrated, aligned samples. Phys. Rev. E 61, 5668 -- 5677 (2000) 16. Leidy, C., Kaasgaard, T., Crowe, J.H., Mouritsen, O.G., Jorgensen, K.: Ripples and the formation of anisotropic lipid domains: imaging two-component supported dou- ble bilayers by atomic force microscopy. Biophys. J. 83, 2625 -- 2633 (2002) 17. Lenz, O., Schmid, F.: Structure of symmetric and asym- metric ripple phases in lipid bilayers. Phys. Rev. Lett. 98, 058,104 (2007) 18. Schmid, F., Dolezel, S., Loenz, O., Meinhard, S.: On rip- ples and rafts: Curvature induced nanoscale structures in lipid membranes. J. Phys.: Conf. Series 487, 012,004 (2014) 19. Connell, S.D., Heath, G., Olmsted, P.D., Kisil, A.: Crit- ical point fluctuations in supported lipid membranes. Faraday Disc. 161, 91 -- 111 (2013) 20. Armstrong, C.L., Marquardt, D., Dies, H., Kucerka, N., Yamani, Z., Harroun, T.A., Katsaras, J., Shi, A.C., Rhe- instadter, M.C.: The observation of highly ordered do- mains in membranes with cholesterol. PLoS One 8, E66,162 (2013) 21. Rheinstadter, M.C., Mouritsen, O.G.: Small-scale struc- ture in fluid cholesterol-lipid bilayers. Current Opinion in Colloid & Interface Science 18, 440 -- 447 (2013). DOI 10.1016/j.cocis.2013.07.001 22. Toppozini, L., Meinhardt, S., Armstrong, C.L., Yamani, Z., Kuvcerka, N., Schmid, F., Rheinstadter, M.: The structure of cholesterol in lipid rafts. Phys. Rev. Lett. 113, 228,101 (2014) 23. Nickels, J.D., Cheng, X., Mostofian, B., Stanley, C., Lindner, B., Heberle, F.A., Perticaroli, S., Feygenson, M., Egami, T., Standaert, R.F., Smith, J.C., Myles, D.A.A., Ohl, M., Katsaras, J.: Mechanical properties of nanoscopic lipid domains. J. Am. Chem. Soc. p. just accepted (2015). DOI 10.1021/jacs.5b08894 24. Baumgart, T., Hess, S.T., Webb, W.W.: Imaging coexist- ing fluid domains in biomembrane models coupling cur- vature and line tension. Nature 425, 821 -- 824 (2003) 25. Veatch, S.L., Keller, S.L.: Separation of liquid phases in giant vesicles of ternary mixtures of phospholipids and cholesterol. Biophys. J. 85(5), 3074 -- 3083 (2003) 26. Veatch, S.L., Keller, S.L.: Seeing spots: Complex phase behavior in simple membranes. Biochimica et Biophysica Acta 1746, 172 -- 185 (2005) 27. Veatch, S.L., Keller, S.L.: Miscibility phase diagrams of giant vesicles containing sphingomyelin. Phys. Rev. Lett. 94, 148,101 (2005) 28. Veatch, S.L., Cicuta, P., Sengupta, P., Honerkamp-Smith, A.R., Holowka, D., Baird, B.: Critical fluctuations in plasma membrane vesicles. ACS Chem. Biol. 3, 287 -- 293 (2008) 29. Veatch, S.L., Soubias, O., Keller, S.L., Gawrisch, K.: Crit- ical fluctuations in domain-forming lipid mixtures. PNAS 104, 17,650 -- 17,655 (2007) 30. Honerkamp-Smith, A.R., Cicuta, P., Collins, M.D., Veatch, S.L., den Nijs, M., Schick, M., Keller, S.J.: Line tensions, correlation lengths, and critical exponents in lipid membranes near critical points. Biophys. J. 95, 236 -- 246 (2008) 31. Honerkamp-Smith, A.R., Veatch, S.L., Keller, S.J.: An introduction to critical points for biophysicists: Observa- tions of compositional heterogeneity in lipid membranes. Biochimica et Biophysica Acta 1788, 53 -- 63 (2009) 32. Komura, S., Andelman, D.: Physical aspects of hetero- geneities in multi-component lipid membranes. Adv. Coll. Interf. Sci. 208, 34 -- 46 (2014) 33. Simons, K., Vaz, W.L.C.: Model systems, lipid rafts, and cell membranes. Annu. Rev. Biophys. Biomol. Struct. 33, 269 -- 295 (2004) 34. Brewster, R., Pincus P. A. Safran, S.A.: Hybrid lipids as a biological surface-active component. Biophys. J. 97, 1087 -- 1094 (2009) 35. Hirose, Y., Komura, S., Andelman, D.: Coupled bilayers: A phenomenological model. modulated ChemPhysChem 10, 2839 -- 2846 (2009) 36. Yamamoto, T., Brewster, R., Safran, S.A.: Chain or- dering of hybrid lipids can stabilize domains in sat- urated/hybrid/cholesterol lipid membranes. EPL 91, 28,002 (2010) 37. Yamamoto, T., Safran, S.A.: Line tension between do- mains in multicomponent membranes is sensitive to the degree of unsaturation of hybrid lipids. Soft Matter 7, 7021 -- 7033 (2011) 38. Hirose, Y., Komura, S., Andelman, D.: Concentration fluctuations and phase transitions in coupled modulated bilayers. Phys. Rev. E 86, 021,916 (2012) 39. Palmieri, B., Safran, S.A.: Hybrid lipids increase the probability of fluctuating nanodomains in mixed mem- branes. Langmuir 29, 5246 -- 5261 (2013) 40. Palmieri, B., Yamamoto, T., Brewster, R.C., Safran, S.A.: Line active molecules promote inhomogeneous structures in membranes: Theory, simulations and exper- iments. Adv. Coll. Interf. Sci. 208, 58 -- 65 (2014) 41. Leibler, S., Andelman, D.: Ordered and curved meso- structures in membranes and amphiphilic films. J. de Physique 48, 2013 -- 2018 (1987) 10 Leonie Brodbek, Friederike Schmid 42. Safran, S.A., Pincus, P., Andelman, D.: Theory of spon- taneous vesicle formation in surfactant mixtures. Science 248, 354 -- 356 (1990) 43. Kumar, P.B.S., Gompper, G., Lipowsky, R.: Modulated phases in multicomponent fluid membranes. Phys. Rev. E 60, 4610 -- 4618 (1999) 44. Harden, J.L., MacKintosh, F.C.: Shape transformations of domains in mixed-fluid films and bilayer membranes. Europh. Lett. 28, 495 -- 500 (1994) 45. Schick, M.: Membrane heterogeneity: Manifestation of a curvature-induced microemulsion. Phys. Rev. E 85, 031,902 (2012) 46. Shlomovitz, R., Schick, M.: Model of a raft in both leaves of an asymmetric lipid bilayer. Biophys. J. 105, 1400 -- 1413 (2013) 47. Shlomovitz, R., Maibaum, L., Schick, M.: Macroscopic phase separation, modulated phases, and microemul- sions: A unified picture of rafts. Biophys. J. 106, 1979 -- 1985 (2014) 48. Sadeghi, S., Muller, M., Vink, R.L.C.: Raft formation in lipid bilayers coupled to curvature. Biophys. J. 107, 1591 -- 1600 (2014) 49. Meinhardt, S., Vink, R.L.C., Schmid, F.: Monolayer cur- vature stabilizes nanoscale raft domains in mixed lipid bilayers. PNAS 12, 4476 -- 4481 (2013) 50. Collins, M.D., Keller, S.L.: Tuning lipid mixtures to in- duce or suppress domain formation across leaflets of un- supported asymmetric bilayers. PNAS 105, 124 -- 128 (2008) 51. Dan, N., Pincus, P., Safran, S.A.: Membrane-induced in- teractions between inclusions. Langmuir 9, 2768 -- 2771 (1993) 52. Dan, N., Berman, A., Pincus, P., Safran, S.A.: Membrane-induced interactions between inclusions. J. de Physique II 4, 1713 -- 1725 (1994) 53. Aranda-Espinoza, H., Berman, A., Dan, N., Pincus, P., Safran, S.: Interaction between inclusions embedded in membranes. Biophys. J. 71, 648 -- 656 (1996) 54. Brannigan, G., Brown, F.: A consistent model for thermal fluctuations and protein-induced deformations in lipid bi- layer. Biophys. J. 90, 1501 -- 1520 (2006) 55. West, B., Brown, F.L.H., Schmid, F.: Membrane-protein interactions in a generic coarse-grained model for lipid bilayers. Biophys. J. 96, 101 -- 115 (2009) 56. Neder, J., West, B., Nielaba, P., Schmid, F.: Coarse- J. grained simulations of membranes under tension. Chem. Phys. 132, 115,101 (2010) 57. Safran, S.A.: Statistical Thermodynamics of Surfaces, In- terfaces, and Membranes. Perseus Books, Cambridge, Massachusetts (1994) 58. Farago, O., Pincus, P.: The effect of thermal fluctuations on schulman area elasticity. Eur. Phys. J. E 11, 399 -- 408 (2003) 59. Fournier, J.B., Barbetta, C.: Direct calculation from the stress tensor of the lateral surface tension of fluctuating fluid membranes. Phys. Rev. Lett. 100, 078,103 (2008) 60. Schmid, F.: Are stress-free membranes really "tension- less"? EPL 95, 28,008 (2011) 61. Shiba, H., Noguchi, H., Fournier, J.B.: Monte carlo study of the frame, fluctuation and internal tensions of fluctuat- ing membranes with fixed area. arxiv:1507.08722 (2015) 62. Cai, W., Lubensky, T.C., Nelson, P., Powers., T.: Mea- sure factors, tension, and correlations of fluid membranes. J. de Physique II 4, 931 -- 949 (1994) 63. Farago, O., Pincus, P.: Statistical mechanics of bilayer membrane with a fixed projected area. J. Chem. Phys. 120, 2934 -- 2950 (2004) 64. Diamant, H.: Model-free thermodynamics of fluid vesi- cles. Phys. Rev. E 84, 0611,203 (2011) 65. Wang, Z.J., Frenkel, D.: Modeling flexible amphiphilic bilayers: A solvent-free off-lattice Monte Carlo study. J. Chem. Phys. 122, 234,711 (2005) 66. Noguchi, H., Gompper, G.: Meshless membrane model based on the moving least-squares method. Phys. Rev. E 73, 021,903 (2006) 67. Farago, O.: Mechanical surface tension governs mem- brane thermal fluctuations. Phys. Rev. E 84, 051,944 (2011) 68. Imparato, A.: Surface tension in bilayer membranes with fixed projected area. J. Chem. Phys. 124, 154,714 (2006) 69. Stecki, J.: Balance of forces in simulated bilayers. J. Phys. Chem. B 112(14), 4246 -- 4252 (2008) 70. Tarazona, P., Chacon, E., Bresme, F.: Thermal fluctu- ations and bending rigidity of bilayer membranes. J. Chem. Phys. 139, 094,902 (2013) 71. Akimov, S.A., Kuzmin, P.I., Zimmerberg, J., Cohen, F.S.: Lateral tension increases the line tension between two domains in a lipid bilayer membrane. Phys. Rev. E 75, 011,919 (2005) 72. Watson, M.C., Morriss-Andrews, A., Welch, P.M., Brown, F.L.H.: Thermal fluctuations in shape, thickness, and molecular orientation in lipid bilayers ii: Finite sur- face tensions. J. Chem. Phys. 139, 084,706 (2013) 73. Fournier, J.B.: Coupling between membrane tilt- difference and dilation: A new "ripple" instability and multiple crystalline inclusions phases. Europh. Lett. 43, 725 -- 730 (1998) 74. Fournier, J.B.: Microscopic membrane elasticity and in- teractions among membrane inclusions: Interplay be- tween the shape, dilation, tilt and tilt-difference modes. Eur. Phys. J. B 11, 261 -- 272 (1999) 75. Bohinc, K., Kralj-Iglic, V., May, S.: Interaction between two cylindrical inclusions in a symmetric lipid bilayer. J. Chem. Phys. 119, 7435 -- 7444 (2003) 76. Fosnaric, M., Iglic, A., May, S.: Influence of rigid inclu- sions on the bending elasticity of a lipid membrane. Phys. Rev. E 74, 051,503 (2006) 77. Watson, M.C., Penev, E.S., Welch, P.M., Brown, F.L.H.: Thermal fluctuations in shape, thickness, and molecular orientation in lipid bilayers. J. Chem. Phys. 135, 244,701 (2011) 78. Kuzmin, P.I., Akimov, S.A., Chizmadzhev, Y.A., Zim- merberg, J., Cohen, F.S.: Line tension and interaction energies of membrane rafts calculated from lipid splay and tilt. Biophys. J. 88, 1120 -- 1133 (2005) 79. Kollmitzer, B., Heftberger, P., Rappolt, M., Pabst, G.: Monolayer spontaneous curvature of raft-forming mem- brane lipids. Soft Matter 9, 10,877 -- 10,884 (2013) 80. Schmid, F.: Fluctuations in lipid bilayers: Are they un- derstood? Biophysical Reviews and Letters 8, 1 -- 20 (2013). DOI 10.1142/S1793048012300113 81. Neder, J., Nielaba, P., West, B., Schmid, F.: Interactions of membranes with coarse-grain proteins: A comparison. New J. of Physics 14, 125,017 (2012) 82. Marsh, D.: Elastic curvature constants of lipid monolay- ers and bilayers. Chemistry and Physics of Lipids 144, 146 -- 159 (2006) 83. Lindahl, E., Edholm, O.: Mesoscopic undulations and thickness fluctuations in lipid bilayers from molecular dy- namics simulations. Biophys. J. 79, 426 -- 433 (2000) 84. Marrink, S.J., Risselada, H.J., Yefimov, S., Tielemann, D.P., de Vries, A.H.: The martini force field: Coarse grained model for biomolecular simulations. J. Phys. Chem. B 111, 7812 -- 7823 (2007) Interplay of curvature-induced micro- and nanodomain structures in multicomponent lipid bilayers 11 85. Gauthier, N.C., Masters, T.A., Sheetz, M.: Mechani- cal feedback between membrane tension and dynamics. Trends Cell Biol. 22, 527 -- 535 (2012) 86. Pontes, B., Ayala, Y., Fonseca, A., Romao, L., Amaral, R., Salgado, L., Lima, F.R., Farina, M., Viana, N.B., Moura-NEto, V., Nussenzveig, H.M.: Membrane elastic properties and cell function. PLOS one 8, 67,708 (2013) 87. Brazovskii, S.A.: Phase transitions of an isotropic system to a nonuniform state. Soviet Physics JETP 41, 85 -- 89 (1975) 88. Needham, D., Nunn, R.S.: Elastic deformation and fail- ure of lipid bilayer membranes containing cholesterol. Biophys. J. 58, 997 -- 1009 (1990) 89. Evans, E., Heinrich, V., Ludwig, F., Rawicz, W.: Dy- namic tension spectroscopy and strength of biomem- branes. Biophys. J. 85, 2342 -- 2350 (2003) 90. Amazon, J.J., Goh, S.L., Feigenson, G.W.: Competition between line tension and curvature stabilizes modulated phase patterns on the surface of giant unilamellar vesi- cles: A simulation study. Phys. Rev. E 87, 022,708 (2013) in transitions: Nucleation fluctuation-driven first-order of lamellar phases. Phys. Rev. E 52, 1828 -- 1845 (1995) 91. Hohenberg, P.C., Swift, J.B.: Metastability 92. Honerkamp-Smith, A.R., Machta, B.B., Keller, S.J.: Ex- perimental observations of dynamic critical phenomena in a lipid membrane. Phys. Rev. Lett. 108, 26,507 (2012)
1806.08179
1
1806
2018-06-21T11:44:16
Classification of red blood cell shapes in flow using outlier tolerant machine learning
[ "physics.bio-ph", "physics.flu-dyn" ]
The manual evaluation, classification and counting of biological objects demands for an enormous expenditure of time and subjective human input may be a source of error. Investigating the shape of red blood cells (RBCs) in microcapillary Poiseuille flow, we overcome this drawback by introducing a convolutional neural regression network for an automatic, outlier tolerant shape classification. From our experiments we expect two stable geometries: the so-called `slipper' and `croissant' shapes depending on the prevailing flow conditions and the cell-intrinsic parameters. Whereas croissants mostly occur at low shear rates, slippers evolve at higher flow velocities. With our method, we are able to find the transition point between both `phases' of stable shapes which is of high interest to ensuing theoretical studies and numerical simulations. Using statistically based thresholds, from our data, we obtain so-called phase diagrams which are compared to manual evaluations. Prospectively, our concept allows us to perform objective analyses of measurements for a variety of flow conditions and to receive comparable results. Moreover, the proposed procedure enables unbiased studies on the influence of drugs on flow properties of single RBCs and the resulting macroscopic change of the flow behavior of whole blood.
physics.bio-ph
physics
Classification of red blood cell shapes in flow using outlier tolerant machine learning Alexander Kihm1, Lars Kaestner1,2 Christian Wagner1,3,* Stephan Quint1 1 Department of Experimental Physics, Saarland University, Campus E2 6, 66123 Saarbrucken, Germany 2 Theoretical Medicine and Biosciences, Saarland University, Campus University Hospital, Building 61.4, 66421 Homburg, Germany 3 Physics and Materials Science Research Unit, University of Luxembourg, Luxembourg *[email protected] Abstract The manual evaluation, classification and counting of biological objects demands for an enormous expenditure of time and subjective human input may be a source of error. Investigating the shape of red blood cells (RBCs) in microcapillary Poiseuille flow, we overcome this drawback by introducing a convolutional neural regression network for an automatic, outlier tolerant shape classification. From our experiments we expect two stable geometries: the so-called 'slipper' and 'croissant' shapes depending on the prevailing flow conditions and the cell-intrinsic parameters. Whereas croissants mostly occur at low shear rates, slippers evolve at higher flow velocities. With our method, we are able to find the transition point between both 'phases' of stable shapes which is of high interest to ensuing theoretical studies and numerical simulations. Using statistically based thresholds, from our data, we obtain so-called phase diagrams which are compared to manual evaluations. Prospectively, our concept allows us to perform objective analyses of measurements for a variety of flow conditions and to receive comparable results. Moreover, the proposed procedure enables unbiased studies on the influence of drugs on flow properties of single RBCs and the resulting macroscopic change of the flow behavior of whole blood. Author summary Artificial neural networks represent a state-of-the art technique in many branches of natural sciences due to their ability to fastly detect and categorize image features with high throughput. We use a special type of neural network, the so-called convolutional neural network (CNN) for the classification of human red blood cell shapes in microcapillary Poiseuille flow. Following this approach, phase diagrams of two distinct classes (slippers, croissants) are generated and, by comparison with a manually obtained phase diagram, optimized threshold ranges for categorizing the output values are established. This allows us to better understand the complex fluid behavior of blood depending on the intrinsic properties of single red blood cells. For future studies, we aim to predict phase diagrams under the influence of certain drugs. 1/15 Introduction Amongst all human organs, blood is the most delocalized one, delivering oxygen from the respiratory system to the tissues in the body and transporting carbon dioxide back. On a microscopic scale, this is performed by red blood cells (RBCs) which form the largest fraction of cells in whole blood (≈ 99 %). At rest, RBCs are biconcave discocytes with an average diameter of 8 µm and a height of 2 µm. Due to their flexible membrane, RBCs alter their shape under external stress prevalent in the microvascular network [1, 2]. This feature is one of the key properties of RBCs, which allows them to squeeze through geometrical constrictions much smaller than their stress-free shape [3], which is partly an intrinsic property of RBC morphology [4, 5] and partly an active adaptation process [6, 7]. Although data on the mechanical properties of RBC suspensions is widely known from rheological measurements [8], the linkage to individual cell behavior is limited. Further, the comparison between capillary Poiseuille flow and pure shear flow prevalent in rheometers is difficult. Consequently, mimicking flow under physiological conditions in vitro demands for experimental setups such as PDMS-based microchannels, ubiquitous in lab-on-a-chip devices [9]. In this work, we focus on experiments of individual flowing RBCs, providing a holistic insight to individual cell mechanics. Hereby, the experimental data originate from a previous study on RBC shape geometry [10] and the data is reused for the introduction of a fully automated data analysis approach based on a deep learning convolutional neural network (CNN) [11]. As described in the preceding work, two stable RBC shapes are expected from the measurements: The so-called 'croissant' and the 'slipper' shape [12–14]. Their frequency of occurrence highly depends on the imposed flow conditions. Whereas axisymmetric croissants mostly appear at lower flow velocities, non-axisymmetric slippers are observed at higher shear rates [15, 16]. However, besides these stable geometries, also a large number of indefinite shapes occur, especially in the so called phase transition range. These outliers or 'other' shapes make a considerable amount of the whole statistics. Due to their large shape variance, these cells cannot be assigned to a mutual exclusive class and can therefore not easily be involved in the machine learning processes. Overcoming this drawback, we present a regression based CNN aiming to distinguish between croissants, slippers and others by applying statistically derived thresholds to the net response. As a result, we obtain an automatically generated RBC phase diagram which relinquishes any subjective user input. Prospectively, we will be able to run comparable studies on RBC shape geometries in microchannels of variable size and to generate unbiased phase diagrams. Above all, data evaluation can then be performed in a highly time-effective manner. Materials and Methods Ethics Statement Human blood withdrawal from healthy donors as well as blood preparation and manipulation were performed according to regulations and protocols that were approved by the ethic commission of the "A rztekammer des Saarlandes" (reference No 24/12). We obtained informed consent from the donors after the nature and possible consequences of the studies were explained. Architecture of convolutional neural networks CNNs are digital image processing systems [17] with the ability to resolve and evaluate the details of an input image. Usually, they are used to e.g. recognize and classify 2/15 fully connected output node Fig 1. Layer structure of the used CNN. The input layer accepts cell images of 90 × 90 px2. Avoiding border effects (top and bottom) caused by irregular light refraction at the channel edges, input images are weighted by a Tukey window (α = 0.25) causing a fading effect towards the upper and lower edge (Eq. 3). In a first processing stage, images are convoluted by 25 different convolution kernels of size 21 × 21. This results in 25 intermediate images of size 70 × 70 px2, which undergo a non-linear rectification (reLU layer) before getting down-sampled by a max-pooling procedure (2 × 2, stride 2) to a size of 35 × 35 px2. The combination of convolution, rectification, and max-pooling are repeated twice using different sets of convolution kernels (see Tab. 1). The output node then intertwines all resulting subimages by a full interconnection of all available pixel values and maps them to a linear output range. Sizes of subimages (blue/grey) as well as the indications of convolution kernels (black) are chosen to scale, illustrating that kernels obey the characteristic features of input and subimages. particular objects or humanoid faces [18] within pictures. Here, a CNN is exploited to distinguish between the shape characteristics of RBCs in flow and to detect undefined outliers which occur due to channel imperfections, membrane damages, cell-cell-interactions, shape transitions or transients, and optical ambiguities. Independent of the particular use case, the architectures of CNNs usually follow the same design rules. They consist of an image input layer followed by a certain number of subsequent convolution stages (Fig. 1), and provide so-called interconnected layers forming an artificial neural network (ANN) to combine the convolution data in a final stage before the information is fed to the output layer nodes. The main stages of a CNN usually consist of several sublayers [19] including the actual convolutional layer, a non-linear rectification 'reLU' layer [20], and a pooling layer [21]. Even though more sophisticated designs [22, 23] exist, for RBC shape recognition we restrict ourselves to these layer types keeping the system as simple as possible. Convolutional layers make use of a number of convolutional kernels (feature maps) of particular size and are optimized to find the major characteristics of a set of various input images being subject to a sophisticated training process (cf. section Training). For instance, this can include horizontal or vertical edges but also more detailed characteristics, such as the 'tail' of a cell, see Fig. 2. Convolving an image with a set of specially optimized convolution kernels then results in a number of output images showing clear differences in between the RBC shape classes. 3/15 layer kernel size [px2] subimage size [px2] input layer conv. layer 1 reLU layer - 21 × 21 - max pooling layer 2 × 2, stride 2 conv. layer 2 reLU layer 14 × 14 - max pooling layer 2 × 2, stride 2 conv. layer 3 reLU layer 6 × 6 - max pooling layer 2 × 2, stride 2 output layer, regression type - 90 × 90 70 × 70 70 × 70 35 × 35 22 × 22 22 × 22 11 × 11 6 × 6 6 × 6 3 × 3 1 × 1 Table 1. Overview of the used layers in the indicated deep learning CNN. In the center column, the kernel size of the corresponding layer is given. The resulting image size after layer passage is given in the rightmost column. In a mathematical sense, convolutions are linear operations. Thus, CNNs could easily be contracted to a simple linear signal processing system if there were no non-linearities involved. Indeed, the usage of non-linearities resolving higher order dependencies in between the characteristics of a set of distinct input images is one of the key ideas behind neural networks. For CNNs, it is common practice to use a reLU layer to set the negative values of the subimages to zero. This renders image transformations non-unitary and allows to introduce further non-linear evalution branches within the tree-like structure of the system. After the convolution and rectification stage, the amount of data is fairly increased and usually corresponds to a multiple of the input data. Thus, it is reasonable to reduce data with the aid of pooling layers [21, 24]. Within such layer, subsets of each intermediate image are pooled according to their mean or maximum values. Results are then assigned to a target image of smaller size. For our CNN, subsets of size 2 × 2 px2 with a stride of 2 px are evaluated according to the maximum value aiming to halve the sizes of intermediate images at the end of each convolution stage. A further advantage of such pooling strategy is to obtain certain tolerance with respect to spatial translations of objects. As special feature, the here employed CNN makes use of a regression output layer providing a linear transfer function. In contrast to standard classification networks offering a purely binary output, realized e.g. by a logistic transfer function or softmax approach [25], the output node of our CNN is able to take on a range of floating point values. Taking the number space of input images into account (8-bit space), we define our output to move in the same range thus the same order of magnitude. Taking values at this scale, we observe a fast convergence of the training process. Consequently, perfect slippers are defined at value −127 whereas croissants are located at 127. Input images leading to an output value which significantly differs from these targets are assumed to be indefinite and therefore discriminated to be an outlier of type other (cf. section Training). 4/15 Fig 2. Resulting subimages (bottom) of two contrary RBC shapes (croissant, upper left; slipper, upper right) passing the first convolutional layer of a CNN. The convolution kernels as well as the subimages are represented by a false color mapping for the sake of better visibility. Boxes in the input images indicate typical features of both cell shape classes and the respective enhancement of these after convolution (indicated by arrows). Training of convolutional neural networks CNNs are not ready to use if neurons are not 'trained' to solve certain regression or classification problems. This means that initially randomly chosen convolution kernels, weights and bias terms (concluded in the vector σ of free parameters) first need to be reasonably updated. For this purpose, we employ a supervised learning approach which requires three major ingredients: A labeled set of training data, a suitable loss function that has to be minimized, as well as an appropriate optimization strategy. As loss function, the root mean square error (RMSE) is chosen (Eq. 1), expressing the cumulated differences between all target and actual output values of the training data set. 5/15 (1) ( Here, ai expresses the actual output of a given input image i and ti the respective, predefined target value. Finding the global minimum of the RMSE, σ is optimized using a stochastic gradient descent solver with momentum (SGDM), see [26] and [27]. The SGDM algorithm updates the vector σl in discrete steps, where l denotes the actual iteration: σl+1 = σl − α∇E (σl) + γ (σl+1 − σl) . (2) E (σl) corresponds to the loss function, respectively the RMSE, α = 0.001 to the constant learning rate, and γ = 0.9 to the momentum term. Achieving a faster convergence, a single optimization iteration takes into account a so-called mini-batch [25, 28, 29] consisting of a subset (128 images) of the randomly shuffled training data. In this context, a full pass of the training data is called an 'epoch'. A typical temporal evolution of training states is shown in Fig. 3 where the loss function is plotted with respect to the number of training epochs. We set a maximum of ten training epochs, since a prolongation to more training epochs rather causes overtraining instead of a gain in performance. This overtraining of the neural network can be monitored in the validation loss: A termination criterion of the training status is implemented by a consecutive increase in validation loss five times, since a divergence between training and validation loss is an indicator for not being in the global optimum. Further, we choose a validation frequency (update frequency of validation loss) of 50 mini-batches, to not meet the termination criterion by random fluctuations of the training process. In general, the validation data set consists of data disjoint from the ones for training to guarantee an independent validation of the networks' performance. For our purposes, 5% of the training data set was chosen randomly for validation, whereas the rest was the true training data. For training, besides slippers and croissants, we additionally define an auxiliary class called 'sheared croissant' (Fig. 4). This is due to the fact that the characteristic features of pure croissants and sheared croissants show a larger mutual similarity than pure croissants and slippers. Assuming a linear scale of output values, this subtype is biased towards a value of 64. Contrasting the nomenclature, they are not any type of croissants, but most probably resemble a slipper flipped by 90◦ perpendicular to the optical axis. However, due to the lack of information of our 2D images, we cannot make a clear statement on the actual shape. Our training data set consists of 4, 000 manually classified cells (1, 500 each for slippers, and croissants, resp. and 1, 000 sheared croissants), which is augmented to the doubled number (8, 000 in total) by mirroring each image along the centerline in flow direction. Each of those 4,000 initial training cell images represents one distinct cell, i.e. no cell was considered more than once in the training data set. We intentionally have fewer sheared croissants in the training data set than croissants and slippers since it is only an auxiliary cell class not reflected in the phase diagram but only to increase the precision of croissant classification. The three subsets (one per cell class) were taken at different pressure drops. Slippers were recorded at 700 − 1, 000 mbar, croissants at 100 − 200 mbar and sheared croissants at 300 − 500 mbar. Within these ranges, we applied several different pressure drops to ensure a certain variability of our training data set. All cells of a class are then mixed since they should be identified independent from the applied pressure. Further, we intentionally recorded cells for different lighting conditions by varying the power of the light source randomly (within a range of 50 − 80 %). Together with a bilinear contrast adjustment of the resulting images (cf. experimental setup), this ensures high contrast dynamics of each image and corrects for minor changes in illumination. Due to constant manual observation during the recording process of the input data, we ensure collecting focused cell images. However, to ensure an invariance of our CNN approach with respect to optical misalignment and to reach a robust algorithm in terms of optical imaging, we also train for defocused cells by recording training images at different positions within a slightly tilted channel. The outcome will then be a neural network with a self-contained training set and thus is perfectly suited for a non-biased analysis. 6/15 Fig 3. Diagram of training (red) and validation (black) status: Evolving convergence loss with growing number of training epochs. We set a maximum of ten epochs since a prolongation to more training epochs yields no gain in performance but rather causes overtraining. As training method, a gradient descent solver with momentum (SGDM) is used. The red line indicates the training loss, whereas the black dots represent the loss of the validation data set (validation loss). The progression of the validation loss serves as an indicator whether the CNN is overtrained, since the training and validation losses would diverge. As loss function, a root-mean-square error is chosen, being a standard approach for regression problems. Experimental setup For our measurements, dilute suspensions of RBCs are required to observe single cells flowing through microcapillaries. Therefore, capillary blood is drawn with consent from a healthy individual and resuspended in a buffer solution (phosphate buffered saline, PBS, gibco by life technologies). After centrifugation at 1,500 g for 5 minutes, RBCs sediment on the bottom of the sample tube, whereas leukocytes and platelets form a buffy coat on top of these (for further details concerning the blood sampling, see [30]). Since the presence of platelets can alter the dynamic properties of RBCs due to shear-induced platelet activation (SIPA, see [31]), this buffy coat is removed together with the supernatant and the residual pellet of RBCs is again resuspended in PBS. This procedure is repeated three times until the final pellet of cells will be adjusted to a hematocrit of Ht ≤ 1% in a base solution of both PBS and bovine albumin (BSA, Sigma-Aldrich) at a concentration of 1 mg/ml, to avoid the well-known glass slide effect, turning discocytous RBCs to echinocytes. With the aid of a high-precision pressure device (Elveflow OB 1 Mk II), this highly diluted suspension of RBCs is then driven through microcapillaries for a discrete set of 12 pressure drops ∆p ∈ {20, 50, 100, 200, 300, 400, 500, 600, 700, 800, 900, 1000} mbar, yielding various flow velocities, and recorded with a high-speed camera at a framerate of 400 Hz. Over the full pressure range, we record in total 3, 090 RBCs and obtain 7/15 experimental data, independent of the training data set. The microcapillaries are formed in PDMS (polydimethylsiloxane), have a length of Lx = 4 cm and a rectangular cross-section with dimensions Ly = 11.9 ± 0.3 µm and Lz = 9.7 ± 0.3 µm. The y-direction is perpendicular to the camera-axis, whereas the z-direction points inward the camera axis. Several channels are branched from a common reservoir to the fluid inlet, which is connected to the pressure device via flexible tubing. Post-processing the recorded media involves single particle tracking of cells, including distinction and sorting out non-isolated RBCs (i.e. RBCs being too close to each other, such that hydrodynamic interactions are not negligible). Due to the used optical setup (Nikon CFI Plan Fluor 60× oil-immersion objective, NA = 1.25) camera (Fastec HiSpec 2G) and capture settings, typical cell dimensions are in the range of 80 px. However, to allow for minor (physiological) variations, we crop the individual cell images to a format of 90 × 90 px2. Since the channel width is smaller than 90 px, we apply a Tukey window w(y) (3) to the images, causing a smooth fade-out towards the channel walls (3) where y denotes the relative position (0 ≤ y ≤ 1) in vertical direction of an arbitrary cell image. This reduces the influence of markable edges probably influencing the training and output of the CNN (Fig. 2). For image preparation, we additionally map the image intensities to the full 8-bit range, such that the bottom 1 % and the top 1 % of all pixel values are saturated. This transformation yields higher signal dynamics and equal intensity profiles of all cell pictures and renders our approach more stable regarding slight illumination variations. Results and discussion Neural networks are usually benchmarked by opposing their output with manually classified data. For our presented analysis, we compared the CNN-classified cell shapes with a manually ascertained phase diagram, presented first in [10]. This manual result serves as reference for all presented system benchmarks. As can be seen in Fig. 4, the positions of the main populations of slippers (−117), croissants (115), and sheared croissants (≈ 40) differ significantly from the target values of −127, respectively 127 and 64. Deviations mostly occur due to different microscopy settings such as illumination and focusing. However, since thresholds are applied by statistical means, this shifting is not crucial for the final result. The center values for the occurring populations as well as the background are determined by applying four Gaussian fit functions. Based on the standard deviations of the Gaussian distributions of the main classes, thresholds can be defined to assigning cells at certain position the respective population. Cells which do not fall in a designated range are classified as others rendering our regression based CNN an outlier tolerant classification system. In Fig. 5, both the output values for the whole image data as well as the corresponding phase diagrams for different thresholds are shown. According to three predefined training classes (slippers, sheared croissants, croissants), we again fit the results of the CNN by four independent Gaussians, one of them accounting for the background. In Fig. 5 (a), we labelled all cells within a confidence interval of 1σ (1σcroissant, and 1σslipper, resp.) to assign the cells to the associated distinct class. In Fig. 8/15 slippers croissa nts sheared Fig 4. We estimate perfect slippers to be around the peak of the distribution at ≈ −117, whereas croissants occur around ≈ 115. By fitting the whole spectrum by four Gaussians, we are able to separate the respective contributions of each cell shape class and thus can determine a respective confidence interval. In the lower part, typical cell shapes are depicted for different output value ranges. Starting from the leftmost cell image, we undergo a shape change from slippers (image 1-3) to others (image 4-5) and finally to sheared (image 6-7) and pure croissants (image 8-9). 5 (b), manually evaluated data from [10] are depicted as areas, whereas the automatically evaluated data are marked as dash-dotted lines. The slipper population is represented in light blue, whereas croissant-labelled cells appear in light red. Although the qualitative behavior of the resulting phase diagram can be reproduced, for a threshold of 1σ we significantly underestimate the fraction of slippers and croissants compared to the result from manual classification. By a change of these confidence bounds from 1σ to 2σ, we perceive a better accordance of the manually classified croissants and the CNN-evaluated ones. Slippers still tend to be underestimated by the neural network, however, the deviation to the manually obtained graph is smaller throughout the whole velocity regime (Fig. 5 (c),(d)), causative of predicting a wrong phase transition point (intersection of both graphs). In Figs. 5 (e),(f), we therefore apply an adapted threshold range for each population, proposing a phase diagram in very good accordance to the benchmark of manually classified cell shapes. Even though a slight deviation to the manually generated phase diagram is still noticeable, both for slippers and croissants, the prediction of the phase transition point is in very good agreement with the original one, as well as the peaks of both cell fractions. The underlying adapted confidence interval is derived from an iterative process, finding a minimum of a predefined cost function. As cost function, we calculate the number of false negatives and false positives for each cell shape class compared to the manual selection. By scanning the center values as well as threshold intervals of each population separately, both values (false positives and false negatives) are summarized to receive a cost value. Consequently, by minimizing this cost value, we obtain the optimal threshold range setting for each population. Finally, we yield adapted ranges at −116.6 ± 28.5 for slippers and 116.8 ± 35.5 for croissants. All deviating individuals 9/15 Fig 5. CNN output values for all cell images. The gray solid line is the network's output for the whole dataset, whereas the black solid line represents a fit with four Gaussians, one for each distinct class (croissants, slippers and sheared croissants, resp.) and one to account for indistinguishable cell shapes. The thresholds are shown in light blue and light red, respectively. In the right column, the obtained classification is compared with the manually ascertained phase diagram (solid lines). We stress the fact that the solid line is a guide to the eye, since we have a discrete number of flow veolcities due to the given number of applied pressure drops. In figure (b), a threshold of 1σ was used as a confidence interval to classify the cells into one of the two categories. Figures (d) and (f) show the resulting phase diagrams for a threshold of 2σ and an adapted σ, resp. 10/15 false negative croissants false positive croissants Fig 6. Image montage of all false negative (left) and false positive (right) classified croissants with respect to manual classification. On the left, false negative croissants are shown, i.e. all cells being classified as croissant manually, but not by the neural network. In contrast, all cells classified as croissant shapes by automated analysis but not by hand are depicted in the right montage (false positive croissants). Numerical values given in the yellow box of each picture correspond to the respective output value of the CNN. either for croissants and slippers are shown in Figs. 6 and 7. In total, the associated confusion matrix of our classification is depicted in Fig. 8. false negative slippers false positive slippers Fig 7. Image montage of all false negative (left) and false positive (right) classified slippers with respect to manually obtained classification. All cells being classified as croissants manually but not by the CNN (false negative slippers) are depicted in the left image, whereas all false positive slippers are shown in the right montage (cells classified as slipper shapes by automated analysis but not by hand). Additionally, each cell image contains a yellow box with the according CNN output values. However, Figs. 6 and 7 illustrate that either the CNN as well as the manual evaluation of particular cells include many indifferent decisions especially for the false 11/15 croissant 551 (85.6%) 0 (0.0%) 93 (14.4%) t n a s s i o r c r e p p i l s r e h t o s s a l c d e t c i d e r p other 91 (8.2%) 118 (10.6%) 901 (81.2%) actual class slipper 0 (0.0%) 1227 (91.8%) 109 (8.2%) Fig 8. Confusion matrix with absolute values and relative percentages to evaluate the performance of the CNN approach. The rows hereby indicate the predicted, i.e. real class, whereas the columns indicate the actual class, corresponding to the CNN output. Thus, all values on the diagonal represent the correctly classified cells. negative cases. Adjusted for these contrary decisions, it can be assumed that the false negative count rate influenced by subjective input is significantly lower. In contrast, with few exceptions, the false positive rate is in good accordance with the indicated value. Referring to the final RMSE (Fig. 3) which corresponds to a value of ≈ 10 regarding a target value of ± 127, the false detection rates of slippers are in a very good agreement with the final error of the training process (≈ 8%). For the croissants, we observe much higher deviations due to the occurrence of sheared individuals which are closely neighbored to the pure croissant population. This especially affects the false positive counts, whereas the false negative count contains a noticeably high amount of indifferent cases. Future perspectives Using artificial intelligence for classification issues is a well-known, yet rare approach in biological systems. Although a wide variety of accessible, pre-trained, highly sophisticated neural networks (e.g. AlexNet, ResNet) exist, they tend to be over-engineered for most purposes in this field. Due to the complexity of their architecture with respect to the degrees of freedom, they require millions of input images. Furthermore, they are not designed for regression problems as required for the present study, but rather have a binary classification output. The presented CNN is a step forward to a fully automated analysis of recorded microscopy images associated with a big gain in evaluation efficiency. It moreover constitutes an unbiased classification system for RBC shapes in flow. Any restriction to cell classes is purely artificial in the sense that we train the CNN solely with RBC shapes in flow. Similarly, one could tailor a set of training data for other distinct cell classes or even cell types. We aim to conduct further experiments with different channel geometries and flow conditions, amending the existing phase diagram.The phase diagram and especially the phase transition point between croissants and slippers can be altered adding drugs to the RBC suspension. Since certain drugs, e.g. acetylsalicylic acid, show a strong effect on the membrane structure [32], we conjecture to resolve this feature in phase diagrams of future studies, leading to insights towards mechanical properties of individual RBCs in flow. Following this scope, the change in flow behavior of whole blood under drug influence can be predicted due to the influence on its main constituent, the red blood 12/15 cells. This might play a key role in finding appropriate drugs to avoid pathogenic incidents, e.g. stenosis. Analogously, we clearly see evidence to detect maladies causing an alteration of cell's intrinsic mechanical properties, e.g. in sickle cell disease, thalassemias and numerous rare anaemias. By slight technical adaptations (e.g. the usage of high-performance graphics cards), it is also possible to classify cells on-the-fly, i.e. in real-time rather than with the presented frame-based approach. Prospectively, the here proposed CNN will be a useful quantitative tool in hematology aiming to investigate cell membrane characteristics. References 1. Helfrich W. Elastic Properties of Lipid Bilayers: Theory and Possible Experiments. Z Naturforsch. 1973;28 c:693–703. 2. Pries AR, Secomb TW. Blood flow in microvascular networks. In: Microcirculation. Elsevier; 2008. p. 3–36. 3. Skalak R, Branemark PI. Deformation of Red Blood Cells in Capillaries. Science. 1969;164(3880):717–719. doi:10.1126/science.164.3880.717. 4. Freund JB, Orescanin MM. Cellular Flow in a Small Blood Vessel. J Fluid Mech. 2011;671:466–490. doi:10.1017/S0022112010005835. 5. Freund JB. The Flow of Red Blood Cells through a Narrow Spleen-like Slit. Phys Fluids. 2013;25(11):110807. doi:10.1063/1.4819341. 6. Danielczok JG, Terriac E, Hertz L, Petkova-Kirova P, Lautenschlager F, Laschke MW, et al. Red Blood Cell Passage of Small Capillaries Is Associated with Transient Ca2+-mediated Adaptations. Frontiers in Physiology. 2017;8. doi:10.3389/fphys.2017.00979. 7. Cahalan SM, Lukacs V, Ranade SS, Chien S, Bandell M, Patapoutian A. Piezo1 links mechanical forces to red blood cell volume. eLife. 2015;4. doi:10.7554/elife.07370. 8. Lazaro GR, Hernandez-Machado A, Pagonabarraga I. Rheology of red blood cells under flow in highly confined microchannels: I. effect of elasticity. Soft Matter. 2014;10:7195–7206. doi:10.1039/C4SM00894D. 9. Brust M, Aouane O, Thi´ebaud M, Flormann D, Verdier C, Kaestner L, et al. The plasma protein fibrinogen stabilizes clusters of red blood cells in microcapillary flows. Scientific Reports. 2014;4(1). doi:10.1038/srep04348. 10. Guckenberger A, Kihm A, John T, Wagner C, Gekle S. Numerical-experimental observation of shape bistability of red blood cells flowing in a microchannel. arXiV. 2017;doi:arXiv:1711.06986. 11. Xu M, Papageorgiou DP, Abidi SZ, Dao M, Zhao H, Karniadakis GE. A deep convolutional neural network for classification of red blood cells in sickle cell anemia. PLOS Computational Biology. 2017;13(10):e1005746. doi:10.1371/journal.pcbi.1005746. 12. Noguchi H, Gompper G. Shape transitions of fluid vesicles and red blood cells in capillary flows. Proceedings of the National Academy of Sciences. 2005;102(40):14159–14164. doi:10.1073/pnas.0504243102. 13/15 13. Tahiri N, Biben T, Ez-Zahraouy H, Benyoussef A, Misbah C. On the Problem of Slipper Shapes of Red Blood Cells in the Microvasculature. Microvasc Res. 2013;85:40–45. doi:10.1016/j.mvr.2012.10.001. 14. Quint S, Christ AF, Guckenberger A, Himbert S, Kaestner L, Gekle S, et al. 3D tomography of cells in micro-channels. Applied Physics Letters. 2017;111(10):103701. doi:10.1063/1.4986392. 15. Secomb TW, Skalak R, O zkaya N, Gross JF. Flow of axisymmetric red blood cells in narrow capillaries. Journal of Fluid Mechanics. 1986;163(-1):405. doi:10.1017/s0022112086002355. 16. Guido S, Tomaiuolo G. Microconfined flow behavior of red blood cells in vitro. Comptes Rendus Physique. 2009;10(8):751 – 763. doi:https://doi.org/10.1016/j.crhy.2009.10.002. 17. Lecun Y, Bottou L, Bengio Y, Haffner P. Gradient-based learning applied to document recognition. Proceedings of the IEEE. 1998;86(11):2278–2324. doi:10.1109/5.726791. 18. Matsugu M, Mori K, Mitari Y, Kaneda Y. Subject independent facial expression recognition with robust face detection using a convolutional neural network. Neural Networks. 2003;16(5-6):555–559. doi:10.1016/s0893-6080(03)00115-1. 19. LeCun Y, Bengio Y, Hinton G. Deep learning. Nature. 2015;521(7553):436–444. doi:10.1038/nature14539. 20. Nair V, Hinton GE. Rectified Linear Units Improve Restricted Boltzmann Machines. In: Proceedings of the 27th International Conference on International Conference on Machine Learning. ICML'10. USA: Omnipress; 2010. p. 807–814. 21. Scherer D, Muller A, Behnke S. Evaluation of Pooling Operations in Convolutional Architectures for Object Recognition. In: Proceedings of the 20th International Conference on Artificial Neural Networks: Part III. ICANN'10. Berlin, Heidelberg: Springer-Verlag; 2010. p. 92–101. 22. Krizhevsky A, Sutskever I, Hinton GE. ImageNet Classification with Deep Convolutional Neural Networks. In: Pereira F, Burges CJC, Bottou L, Weinberger KQ, editors. Advances in Neural Information Processing Systems 25. Curran Associates, Inc.; 2012. p. 1097–1105. 23. Szegedy C, Liu W, Jia Y, Sermanet P, Reed SE, Anguelov D, et al. Going Deeper with Convolutions. CoRR. 2014;abs/1409.4842. 24. Boureau YL, Roux NL, Bach F, Ponce J, LeCun Y. Ask the locals: Multi-way local pooling for image recognition. In: 2011 International Conference on Computer Vision. IEEE; 2011. 25. Bishop C. Pattern Recognition and Machine Learning. Springer; 2006. 26. Rumelhart DE, Hinton GE, Williams RJ. Learning representations by back-propagating errors. Nature. 1986;323(6088):533–536. doi:10.1038/323533a0. 27. Qian N. On the momentum term in gradient descent learning algorithms. Neural Networks. 1999;12(1):145–151. doi:10.1016/s0893-6080(98)00116-6. 28. Kubat M. An Introduction to Machine Learning. Springer; 2017. 14/15 29. Hastie T, Tibshirani R, Friedman J. The Elements of Statistical Learning. 2nd ed. Springer; 2008. 30. Baskurt OK, Boynard M, Cokelet GC, Connes P, Cooke BM. New guidelines for hemorheological laboratory techniques. Clinical Hemorheology and Microcirculation. 2009;42(2):75–97. doi:10.3233/CH-2009-1202. 31. Holme PA, Orvim U, Hamers MJAG, Solum NO, Brosstad FR, Barstad RM, et al. Shear-Induced Platelet Activation and Platelet Microparticle Formation at Blood Flow Conditions as in Arteries With a Severe Stenosis. Arteriosclerosis, Thrombosis, and Vascular Biology. 1997;17(4):646–653. doi:10.1161/01.atv.17.4.646. 32. Himbert S, Alsop RJ, Rose M, Hertz L, Dhaliwal A, Moran-Mirabal JM, et al. The Molecular Structure of Human Red Blood Cell Membranes from Highly Oriented, Solid Supported Multi-Lamellar Membranes. Scientific Reports. 2017;7:39661. doi:10.1038/srep39661. 15/15
1804.06442
1
1804
2018-04-17T19:24:39
Evidence for a Solenoid Phase of Supercoiled DNA
[ "physics.bio-ph", "cond-mat.soft", "q-bio.BM" ]
In mechanical manipulation experiments, a single DNA molecule overwound at constant force undergoes a discontinuous drop in extension as it buckles and forms a superhelical loop (a plectoneme). Further overwinding the DNA, we observe an unanticipated cascade of highly regular discontinuous extension changes associated with stepwise plectoneme lengthening. This phenomenon is consistent with a model in which the force-extended DNA forms barriers to plectoneme lengthening caused by topological writhe. Furthermore, accounting for writhe in a fluctuating solenoid gives an improved description of the measured force-dependent effective torsional modulus of DNA, providing a reliable formula to estimate DNA torque. Our data and model thus provide context for further measurements and theories that capture the structures and mechanics of supercoiled biopolymers.
physics.bio-ph
physics
Evidence for a Solenoid Phase of Supercoiled DNA Andrew Dittmore and Keir C. Neuman Laboratory of Single Molecule Biophysics, National Heart, Lung, and Blood Institute, National Institutes of Health, Bethesda, Maryland 20892, USA (Dated: April 23, 2018) In mechanical manipulation experiments, a single DNA molecule overwound at constant force un- dergoes a discontinuous drop in extension as it buckles and forms a superhelical loop (a plectoneme). Further overwinding the DNA, we observe an unanticipated cascade of highly regular discontinuous extension changes associated with stepwise plectoneme lengthening. This phenomenon is consistent with a model in which the force-extended DNA forms barriers to plectoneme lengthening caused by topological writhe. Furthermore, accounting for writhe in a fluctuating solenoid gives an im- proved description of the measured force-dependent effective torsional modulus of DNA, providing a reliable formula to estimate DNA torque. Our data and model thus provide context for further measurements and theories that capture the structures and mechanics of supercoiled biopolymers. Supercoiling of DNA permits its compaction in a cell and is crucial to the biological activities of DNA inter- acting proteins. Understanding the structures and me- chanics of DNA supercoiling is therefore a topic of fun- damental interest [1]. Experimentally, the force and link- ing number of a topologically closed DNA double helix can be specified in a single-molecule manipulation ex- periment with torsional control [2]. Following a theorem that states that the imposed torsion or excess linking number (Lk) partitions into interconverting components of twist (Tw) and writhe (Wr) [3–5], the DNA molecule can either twist about its backbone or writhe into more complex shapes that depend on force and solution condi- tions. These DNA shapes produce mechanical signatures observable by experiment. At a critical torque, twisted DNA buckles: It loops upon itself and wraps into a superhelical writhed struc- ture called a plectoneme (Fig. 1). The plectoneme length- ens with additional imposed turns. Although this is widely assumed to be a smooth and continuous process, here we present experimental evidence that a DNA plec- toneme lengthens in discrete steps that occur with a peri- odicity that exceeds the characteristic plectoneme repeat length (Fig. 1a). This behavior indicates supercoiling involves unexpected periodic free-energy barriers of un- known origin. The observed plectoneme lengthening steps can be in- terpreted as discontinuous exchanges of twist for writhe. At fixed force and torsion, the conjugate variables, ex- tension and torque, are free to fluctuate. Since torque increases with increasing twist, the discrete plectoneme lengthening phenomenon corresponds to a progressive buildup, then release, of DNA torque (Fig. 1b). To explain this phenomenon, we propose a model in which barriers to plectoneme lengthening arise from the force-extended DNA adopting a solenoid structure (Fig. 2). If the force-extended region were straight with- out writhe, it would have continuous translational sym- metry and adding length to the plectoneme would in- crease energy smoothly. In contrast, a solenoid shape is periodic and has discrete translational symmetry, result- ing in a periodic energy associated with adding length to the plectoneme: To preserve boundary matching condi- tions in the joining region, the solenoid discontinuously feeds into the plectoneme in steps of one solenoid wave- length (Fig. 2). To compare this model with measure- ments, we first develop a heuristic computational frame- work describing equilibrium DNA mechanics (eqns. 1-6) [7, 8], and then use this to calculate steps predicted by a solenoid model (Fig. 3); we find that measurements are broadly consistent with predictions. In further sup- port of this model, we find that changes in torsional elasticity with force can be accounted for by estimating the writhe of the solenoid (Fig. 4). Thus, the solenoid model we propose to explain our primary observation of an unanticipated discontinuous supercoiling phenomenon potentially also resolves an existing discrepancy between theory and experiment regarding the decrease in the ef- fective torsional persistence length with decreasing force (eqn. 7). Solenoid shapes were anticipated by theorists [13], and expected to range from a circularly writhed coil at zero force to a straight-twisted rod at high force (Fig. 4 inset). In their treatment of DNA supercoiling, this intuition was expressed by Marko and Siggia [13]: "... whatever extension exists must be bridged by a helical random coil or solenoid." Though this structure has not been seen in DNA, solenoids directly visualized in actin filaments may have biological functions [14]. However, a straight- shape hypothesis has prevailed in the DNA literature, in part because thermally fluctuating shapes are difficult to visualize, and also because calculations assuming an initially straight elastic rod show that the lowest energy structure is a straight-twisted rod and not a solenoid [15]. We can conclude from this that if solenoids occur in DNA, as they do on a larger scale in microscopic actin filaments, these solenoids are likely different than the solenoids of classical elastic rod theory [16]. Reconciliation with theory comes by relaxing the as- sumption that the DNA is initially straight. In the pres- ence of intrinsic sequence-dependent bends as well as entropic bending fluctuations [17], overwound DNA is never straight, and can reach a lower energy by trad- ing twist for writhe [9]. Consistent with this, Thompson 8 1 0 2 r p A 7 1 ] h p - o i b . s c i s y h p [ 1 v 2 4 4 6 0 . 4 0 8 1 : v i X r a 2 FIG. 2. Schematic structural model of plectoneme lengthen- ing in discrete steps. (a) Plectoneme joining a solenoid. (b) Planar representation of solenoid of wavelength λ. The plec- toneme projects into the plane at the black dot, where up and down arrows are internal forces in balance at this point. (c) Translational symmetry breaking. Translating either end of the solenoid by λ/4 to lengthen the plectoneme (total trans- lation length λ/2) would incur a maximum torsional energy penalty due to the force-couple, 2rf (with r ∼ (βf )−1), inte- grated through rotation by π/2. This estimate is used only to establish the existence of periodic energy barriers exceeding β−1 and ignores deformation or other energy contributions to the unknown transition state structure. Colored points correspond to initial contour positions in a. (d) Torsional energy penalty vs translation length. The symmetry break- ing depicted in c results in periodic barriers to continuous plectoneme lengthening. The symmetry in b is reflected in positions with zero energy penalty that repeat periodically in incremental translation lengths of one solenoid wavelength, predicting a step size [7]. with data, we first calculate the continuous or average elastic response curves, and then include the predicted discrete states that correspond to twist-writhe fluctua- tions and plectoneme lengthening steps that arise due to the solenoid structure adopted by the extended DNA. Using three phenomenological constants, c1, c2, and c3, we develop a heuristic quantitative model for the elas- tic response of DNA under tension and torsion, favoring empirical relationships where current theory is lacking. We limit our attention to long molecules of positively su- percoiled DNA, subject to small forces that allow ther- mal bending fluctuations, and small torques that pre- serve the rigidly stacked double-helical structure [7, 8]. Under these conditions, DNA is the prototypical exam- ple of a semiflexible and twist-storing polymer, well de- scribed as an elastic rod with bend persistence length A = 50 nm and twist persistence length C = 100 nm much larger than the constituent chemical monomers. These length scales are related to the distances over which bend and twist fluctuations induced by the ther- mal energy, β−1 = 4.1 pN nm, become uncorrelated. For a DNA molecule of contour length L0, imposed torsion FIG. 1. Single-molecule data on 2 µm DNA at 3.6 pN force and 0.5 M salt. (a) End-to-end extension of the DNA vs excess linking number [2]. The schematic depicts the surface- immobilized DNA molecule overwound by Lk rotations of the tethered magnetic bead at constant upward force f . The hor- izontal buckled structure is a plectoneme. Adjacent points recorded at 5-ms intervals are separated by 10−4 in Lk. The extension histogram (right; 10-nm bins) reveals a series of discrete post-buckling states colored in varying hues. The scale is from 0 to 8000. (b) Torque vs excess linking number [6]. The torsional response is linearly elastic up to the maxi- mum torque Γ† at the point of buckling (Lk†, near Lk = 24). The torque then varies about an average value Γb (Lk > Lk†), increasing linearly in Lk within each discrete state while fluc- tuating discontinuously among states. The torque overshoot, (Γ† − Γb) assumed to occur from Lk∗ to Lk†, is comparable in scale. and Champneys showed that torsion causes initially bent elastic rods to develop a nonclassical, one-turn-per-wave solenoid [18]. This previously unrecognized solenoid is not a new solution to the elastic rod equations, but is instead the more general form of the straight-twisted rod solution [19], in agreement with intuition [13]. Motivated to investigate whether signatures of a one- turn-per-wave solenoid can be found in the mechanical characteristics of overwound DNA, we develop a heuristic computational approach to reproduce the observed be- havior of supercoiled DNA [7, 8]. To compare our model 0204060500100015002000ExcessLinkingNumber1020304050607015202530ExcessLinkingNumberTorque (pNnm)Extension (nm)(a)(b)Lkf bLk†Lk*Lk†Lk*Γ†Γ(a)(b)2rTranslation Length(c)λ(d)λ2λTorsionalEnergy Penalty0fπrf 3 FIG. 4. Force dependence of DNA torsional elasticity and partitioning of twist and writhe (see eqn. 1). The effective tor- sional persistence length Ceff increases with force and asymp- totically approaches C. Points are data from Mosconi et al. [10]. The dashed curve is the theoretical prediction of Mo- roz and Nelson [9], Ceff (f ) ≈ ( 1 βAf )−1. The solid C + curve (eqn. 7), Ceff (f ) ≈ ( 1 βAf )−1 − C(1 − xWLC), √ 1 gives an improved description of data by including an esti- mate of writhe per link in a solenoid, Wr/Lk ≈ 1 − xWLC [11]. Both curves are plotted with bending persistence length A = 50 nm and bare twist persistence length C = 100 nm. This comparison indicates that the total writhe per link (Wr/Lk = 1 − Tw/Lk; right axis) is consistent with the sum of a solenoid (depicted schematically) plus fluctuations about this shape [12]. C + 4A √ 1 4A subtle, approximately quadratic, decrease in the exten- sion vs linking number curve prior to buckling (Fig. 1). We observe that a fraction 1/c1 of the torsional energy reduction relative to a straight-twisted rod goes into the work of decreasing the extension at constant force. With c1 ≈ 3 a constant, (cid:104)Xu(f, L0, Lk)(cid:105) ≈ L0xWLC(f ) − 2(πLk)2(C − Ceff (f )) c1βL0f (2) is a simple approximation for the extension vs excess link- ing number behavior of twisted and writhed DNA prior to buckling. Once the DNA buckles, its end-to-end exten- sion decreases linearly, on average, due increases in link- ing number being taken up as writhe in the plectoneme. Each additional turn lengthens the plectoneme and short- ens the force-extended portion by q ≈ 3πΓbxWLC/(2f ) [21], where Γb(f, c) ≈ c2f 0.72/ ln(c) (3) is the average torque in buckled DNA. With monova- lent salt concentration c in mM and force f in pN, the constant c2 ≈ 54. The ln(c)−1 scaling is expected from electrostatic models [22], whereas the f 0.72 power-law is empirical [10], and not yet explained by theory. Next, comparing eqns. 1 and 3, we locate a useful feature point of DNA supercoiling curves: The torque in ∗ unbuckled DNA is equal to Γb at Lk = βL0Γb/(2πCeff ). This is assumed to be the point at which the post- buckling slope intersects the (cid:104)Xu(cid:105) vs Lk curve (eqn. 2). By extrapolating the slope q from Lk∗, we can write the FIG. 3. Model and comparison with data as a function of force and salt concentration. (a) End-to-end extension vs ex- cess linking number calculated under conditions identical to Fig. 1a. The average extension curve (eqn. 5) is plotted in black. The step size and difference in linking number be- tween steps calculated from the solenoid model are indicated on the discontinuous gray curve. (b) Torque vs excess link- ing number calculated under conditions identical to Fig. 1b. The average torque curve (eqn. 6) is plotted in black. The torque increases linearly in each state in the discontinuous gray curve. (c) Step size vs force and (d) change in linking number between steps vs force. (e) Step size vs salt concen- tration and (f) change in linking number between steps vs salt concentration. The black curves correspond to a unit wave- length of solenoid, equal to (cid:104)Xu(f, L0, Lk†)(cid:105)/Lk†. The gray curves correct for (cid:112)A/(βf ) bending fluctuations [9], which are expected to shorten each adjacent half-wave of solenoid above and below the plectoneme. Error bars are standard er- ror of the mean across multiple post-buckling steps (Fig. 1). Force was varied at fixed 1 M monovalent salt; and salt con- centration was varied at fixed 3.6 pN force. increases the interlinking of the two strands of the dou- ble helix by excess linking number Lk. The force f and concentration c of monovalent salt are held constant as the DNA is overwound. The torque Γ increases in direct proportion to torsion Lk according to the linear elasticity equation, (cid:104)Γu(f, L0, Lk)(cid:105) = 2πCeff (f )/(βL0)Lk = 2πC/(βL0)(cid:104)Tw(cid:105), (1a) (1b) where braces denote the thermal average. With Lk = Tw + Wr, it is clear that the conventional use of the effective torsional modulus (force-dependent Ceff /β in eqn. 1a) instead of the microscopic twist modulus (force- independent C/β in eqn. 1b) is a concise statement that overwound DNA develops writhe (see Fig. 4). The nonzero writhe is evidence for bends adopting chi- rality in the direction of torque [9], leading to a reduc- tion of the extension of unbuckled DNA. We start with the Marko-Siggia wormlike chain (WLC) model to obtain the normalized extension xWLC(f ) of a length L0 DNA molecule at zero torque [20], and note that an increase in writhe due to torque-enhanced fluctuations explains the 500100015002000Extension (nm)0204060102030ExcessLinkingNumberTorque (pNnm)01234501001502000123425507510001234123456ΔLkbetweensteps00.511234[Na+] (M)Stepsize (nm)Force (pN)[Na+] (M)Force (pN)ΔLkbetweenstepsStepsize (nm)(a)(b)(c)(e)(d)(f)Lk†Lk*ΔLkbetweenstepsstep sizeCeff(nm)Force (pN)2550751000.750.50.250Wr/Lk01324 post-buckling extension as (cid:104)Xb(f, c, L0, Lk)(cid:105) = (cid:104)Xu(f, L0, Lk ∗ )(cid:105)− q(Lk− Lk ∗ ). (4) Finally, ∗ ∗ † to Lk to complete the average elastic response curves, we estimate details near the buckling transition. The DNA extension at buckling is marked by a discontin- uous transition between unbuckled (eqn. 2) and buckled (eqn. 4) states. These are occupied with equal proba- † bility at the transition midpoint Lk , where unbuckled DNA and buckled DNA featuring a small plectoneme are equal in energy. We assume the torque increase from Lk is proportional to this linking number dif- ) ≈ c3f /Γb(f, c) and ference (eqn. 1), with (Lk constant c3 ≈ 20 nm used to approximate the scale of this feature for DNA [23], which is known as the torque overshoot (Fig. 1). Near the discontinuity, the micro- scopic nature of this transition (energy scale comparable to β−1) permits coexistence fluctuations that give rise to smoothly varying average extension and torque curves. Buckling is dictated by the exchange of free energy and described using Boltzmann statistics, with probabilities Pu = (1 + exp [4π2(Lk of the unbuckled state and Pb = 1− Pu of the buckled state [7, 24]. Joining eqns. 2 and 4, )(Lk − Lk † − Lk ∗ † − Lk )Ceff /L0]) † −1 (cid:104)X(f, c, L0, Lk)(cid:105) = Pu(cid:104)Xu(cid:105) + Pb(cid:104)Xb(cid:105). Similarly, joining eqns. 1 and 3, (cid:104)Γ(f, c, L0, Lk)(cid:105) = Pu(cid:104)Γu(cid:105) + Pb(cid:104)Γb(cid:105). (5) (6) Eqns. 5 and 6 describe the average elastic response to supercoiling a DNA molecule whose total length is par- titioned into a plectonemic fraction and a force-extended fraction. In general, lengths of DNA can exchange be- tween these fractions, interconverting twist and writhe, leading to variance and discontinuities in X and Γ at a given value of Lk (Fig. 1). The pronounced discontinuity at the buckling transition is attributed to plectoneme for- mation. We find that as Lk is increased past the critical point Lk at buckling, a series of discontinuities ensues (Fig. 1). † We investigate whether a structural model might ex- plain this. In the most general sense, the phenomenon we observe might be explained by the energy of locally bent DNA (the tails) feeding into the plectoneme. Linking- number-dependent variations in the tails could occur due to superstructure or sequence causing variations in bend- ing energy. Since we observe modulations on what is otherwise a regular, periodic signal (Fig. 1), we attribute the periodicity to superstructure and the modulations to sequence. Since the simplest periodic superstructure is a solenoid, we consider whether the existence of a force- extended solenoid is consistent with the observed elas- ticity and supercoiling behaviors of DNA. We do this through calculations of a solenoid model, which we com- pare to data without fitting. 4 Marko and Siggia anticipated a "plecto-noid" coex- istence state of torsionally buckled DNA [13], consis- tent with our proposed structural model (Fig. 2). Our phenomenological framework for constructing the av- erage elastic response curves (continuous black curves in Fig. 3a) provides all necessary parameters for pre- dicting discrete states in the context of this model [7], where writhe in extended DNA causes discontinuous plectoneme lengthening corresponding to an extension change, or step size, of one solenoid wavelength (Fig. 2). In the simplest case of a one-turn-per-wave solenoid, the expected number of waves is equal to the imposed number of turns, Lk. Therefore, absent fluctuations in a simple solenoid shape (Fig. 2), the expected step size is equal to the extension per wave, or wavelength, (cid:104)Xu(f, L0, Lk , at buckling (Fig. 3). A more refined estimate accounts for bending fluctuations [9], which are expected to facilitate barrier crossing and re- duce the predicted step size by 2(cid:112)A/(βf ) (Fig. 3). The )(cid:105)/Lk † † corresponding linking number change ∆Lk calculated be- tween steps is equal to the step size divided by q, and can be visually compared with the periodic recurrence in linking number of discrete torque fluctuation states (Fig. 1b and Fig. 3b). We find that this model recapit- ulates the essential force- and ionic strength-dependent features of discontinuous states observed experimentally (Fig. 3), lending credence to the idea that torsionally buckled DNA has solenoidal superstructure in the force- extended region. √ To investigate this further, a solenoid model predicts additional writhe not accounted for in a straight-rod model. This should have consequences for the torque in DNA prior to buckling. As seen in eqn. 1, the torque, which is directly proportional to twist, is reduced by writhe and calculated using the effective torsional persis- tence length Ceff . In particular, Ceff reflects the writhe Wr = Lk(1− Ceff /C) that develops as unbuckled DNA is overwound. The force dependence of this relationship is examined in Fig. 4: We compare the theory of torsional directed walks [9] with the data of Mosconi et al. [10]. In their insightful theory, Moroz and Nelson derived the ex- pression Ceff (f ) ≈ C(1+C/(4A βAf ))−1 by considering writhe caused by fluctuations about a straight-twisted rod [9]. The straight-rod assumption may explain why, particularly at low forces, their expression overestimates the experimental values of Ceff (Fig. 4). This topic is highlighted by two alternative models recently proposed to explain the apparent discrepancy between the Moroz- Nelson theory and data [25, 26]. Both models fit experi- mental data better than the Moroz-Nelson formula, but do so by including additional fitting parameters. Our model of solenoidal superstructure recasts the Moroz- Nelson result without additional parameters. Under the assumption that the total writhe includes linearly addi- tive terms, we approximate a static component of writhe [11], and find that a simple term, −C(1−xWLC(f )), aug- ments the Moroz-Nelson theory and corrects the overes- timate of Ceff remarkably well (Fig. 4). This suggests a 5 modified formula, improved approximation for Ceff . Ceff (f ) ≈ ( 1 C + √ 1 βAf 4A )−1 − C(1 − xWLC(f )), (7) can be used as an improved estimate, which, when input to eqn. 1, gives a reasonable approximation for the torque in overwound DNA [7, 8]. The Moroz-Nelson theory is entirely self-consistent and the bare twist persistence length C is correctly estimated using their procedure of fitting to X vs Lk data [9]. This works because the approximately quadratic change in ex- tension used for fitting is caused by torque-coupled fluc- tuations. However, the discrepancy between the Moroz- Nelson formula for Ceff and data suggest there is an ad- ditional source of writhe that is decoupled from torque. We found that the missing writhe is consistent with a fluctuating solenoid – rather than straight-rod – shape (Fig. 4), and therefore recommend use of eqn. 7 as an We presented experimental evidence that DNA super- coiling features unexpected discontinuities or steps oc- curring at highly regular intervals of extension or link- ing number, indicating sawtooth-like variations in torque associated with extending the plectoneme of buckled DNA. We found that both the observed steps and force- dependent torsional elasticity are quantitatively consis- tent with twisted and force-extended DNA storing writhe in the form of a solenoid structure. To show this, we compared our data to model predictions by developing a framework for computing the elastic response of DNA overwound at constant tension. To promote further mea- surements and theories relating shape to mechanics in DNA or other supercoiled biopolymers, we provide this as an open-source tool for general use [8]. Building from this template, our model might be refined and extended by exact calculations of fluctuating solenoid structure. [1] C. J. Benham and S. P. Mielke. Ann Rev Biomed Eng, of Elasticity, 4th ed, (Dover, New York, 1944). 7:21–53, 2005. [17] T. M. Okonogi, A. W. Reese, S. C. Alley, P. B. Hopkins, [2] T. R. Strick, J. F. Allemand, D. Bensimon, and V. Cro- B. H. Robinson. Biophys J 77(6):3256-3276, 1999. quette. Biophys J, 74(4):2016–2028, 1998. [18] J. M. T. Thompson and A. R. Champneys. Proc R Soc [3] G. Calugareanu. Czech Math J 11:588-625, 1961. [4] J. H. White. Am J Math, 91:693, 1969. [5] F. B. Fuller. Proc Natl Acad Sci U S A, 68:815, 1971. [6] Data in Fig. 1b are data from Fig. 1a transformed by setting the average torque Γb in the plectoneme equal to 2f q/(3πxWLC) [21], with small changes in Γ assumed linearly proportional to small changes in Lk. Lond A, 452(1944):117–138, 1996. [19] A. R. Champneys, G. H. M. van der Heijden, and J. M. T. Thompson. Phil Trans R Soc Lond A, 355:2151–2174, 1997. [20] J. F. Marko and E. D. Siggia. Macromolecules, 28: 8759- 8770, 1995. [21] N. Clauvelin, B. Audoly, and S. Neukirch. Macro- [7] Supplemental Material includes annotated equations, ad- molecules, 41(12):4479–4483, 2008. ditional figures, methods, and refs. [27, 28]. [8] Our computational framework is further described and demonstrated in annotated Python code, which is ap- pended here and will be available on GitHub. [9] J. D. Moroz and P. Nelson. Macromolecules, 31(18):6333- 6347, 1998. [10] F. Mosconi, J. F. Allemand, D. Bensimon, and V. Cro- quette. Phys Rev Lett, 102(7), 2009. [11] The writhe in a static one-turn-per-wave solenoid is Wr = Lk(1−sin γ) [13], with solenoid opening angle γ. In our simple approximation, sin γ equals the wavelength di- vided by the contour length per wave, or fractional DNA extension, xWLC. This estimate avoids double counting of fluctuations already accounted for in the Moroz-Nelson theory [12]. [12] Torque-coupled fluctuations further increase writhe and reduce the normalized extension to Xu/L0 ≤ xWLC [9]. [13] J. F. Marko and E. D. Siggia. Science, 265(5171):506– 508, 1994. [14] N. Leijnse, L. B. Oddershede, and P. M. Bendix. Proc Natl Acad Sci U S A, 112(1):136–141, 2015. [15] B. Fain, J. Rudnick, and S. Ostlund. Phys Rev E 55(6):7364–7368, 1997. 1923. [22] P. Debye and E. Huckel. Phys Zeitschrift, 24: 185-206, [23] We posit (Lk† − Lk∗) ≈ c3f /Γb(f, c), noting that it pro- vides reasonable force dependence as a sublinear power- law, a logarithmic dependence on salt concentration, and works well over our specified parameter range [7, 8]. In our model, (Lk† − Lk∗) also determines the torque differ- ence (Γ† − Γb) and is reflected in sigmoidal shape of the average extension vs excess linking number curve near buckling [24]. [24] In these probabilities, the argument of the exponential is −β times the change in torsional free energy upon buckling, estimated by assuming a constant local torque drop, (Γ† − Γb) = 2πCeff /(βL0)(Lk† − Lk∗), for Lk near Lk†. This is consistent with the sigmoidal shape of the buckling transition in average elastic response curves, as well as experimental estimates of (Γ† − Γb) = −(2πβ)−1∂ln Pu /∂Lk from two-state fluctuation data. Pb [25] S. K. Nomidis, F. Kriegel, W. Vanderlinden, J. Lipfert, E. Carlon. Phys Rev Lett 118(21):7801–7806, 2017. [26] J. M. Schurr J Phys Chem B 119(21):6389–6400, 2015. [27] H. Brutzer, N. Luzzietti, D. Klaue, and R. Seidel. Bio- phys J, 98(7):1267–1276, 2010. [16] A. E. H. Love. A Treatise on the Mathematical Theory [28] Y. Seol and K. C. Neuman. Methods Mol Biol, 783:265– 93, 2011.
1510.05474
2
1510
2015-11-12T21:55:26
Brownian dynamics simulations of an idealized chemical reaction network under spatial confinement and crowding conditions
[ "physics.bio-ph", "physics.chem-ph" ]
We investigate, via Brownian dynamics simulations, the reaction dynamics of a simple, non-linear chemical network (the Willamowski-Rossler network) under spatial confinement and crowding conditions. Our results show that the presence of inert crowders has a non-nontrivial effect on the dynamics of the network and, consequently, that effective modeling efforts aiming at a general understanding of the behavior of biochemical networks in vivo should be stochastic in nature and based on an explicit representation of both spatial confinement and macromolecular crowding.
physics.bio-ph
physics
Chemical reaction networks under spatial confinement and crowding conditions Brownian dynamics simulations of an idealized chemical reaction network under spatial confinement and crowding conditions Benjamin B. Bales1 and Giovanni Bellesia2 1)Department of Computer Science, University of California Santa Barbara 93106 Santa Barbara, CA 2)Los Alamos National Laboratory, 87545 Los Alamos, NMa) (Dated: 1 July 2021) 234 PACS numbers: 82.40.Qt,05.40.Jc,83.10.Rs,87.10.Mn,87.18.Vf Keywords: chemical networks, reaction-diffusion systems, spatial confinement, crowd- ing, Brownian dynamics 5 1 0 2 v o N 2 1 ] h p - o i b . s c i s y h p [ 2 v 4 7 4 5 0 . 0 1 5 1 : v i X r a a)Electronic mail: [email protected] 1 Chemical reaction networks under spatial confinement and crowding conditions I. INTRODUCTION Biochemical networks in vivo are typically open to the exchange of energy and matter with the surrounding environment1 -- 4. They often contain autocatalytic steps5 -- 9 and their dynamics tends to be strongly influenced by thermal and intrinsic noise10,11, macromolecular crowding and spatial confinement12 -- 16. In this study we present a simple computational model of a generic biochemical network in vivo and we investigate how its dynamics is affected by spatial confinement and particle crowding.12 -- 16. Our model is based on the Willamowski-Rossler (WR) chemical network17. The WR network (see Figure 1(a)) is a non- linear chemical system based on zeroth, first and second order chemical reactions. It contains three autocatalytic steps involving species A,B and C and it is thermodynamically open1 -- 4. Its rate equations display a rich and complicated dynamics comprising fixed point, limit cycle and chaotic attractors. The WR network has been previously studied via deterministic and non-spatial stochastic simulation methods18 -- 23 but never as a stochastic reaction -- diffusion system where crowding and spatial confinement are explicitly taken into account. In detail, we investigate the effects of spatial confinement and crowding on a minimal version of the WR network (MWR) (see Figure 1(b) and Ref. 23) using hard-sphere24,25 Brownian dynamics simulations integrating chemical reactivity26,27. We fix the population numbers for species E1, E2, E3, P1 and P2 (consequently the rates k1, k3 and k5 become pseudo-first order) so that the MWR network is thermodynamically open. The following chemical reactions describe the MWR system used in our simulations23 (see also Figure 1(b)). k1−−(cid:42)(cid:41)−− k−1 ¯E1 + A 2 A A + B k2−→ 2 B A + C k4−→ P 2 ¯E2 + B k3−→ P 1 k5−−(cid:42)(cid:41)−− ¯E3 + C 2 C. k−5 (1a) (1b) (1c) (1d) (1e) The main assumption in the MWR system17,23 is that three of the backward reaction rate constants shown in Figure 1(a), namely k−2, k−3 and k−4, are much smaller than their forward counterparts and, hence, can be neglected (See Figure 1(b)). The MWR system is composed by two main subsystems: a Lotka-Volterra oscillator28 -- 30 involving species A 2 Chemical reaction networks under spatial confinement and crowding conditions and B and a chemical switch 18 that couples the Lotka-Volterra component to species C. Similarly to the 'full' WR network, the MWR rate equations derived from the set of chemi- cal reactions (1a) − (1e) display a diverse dynamical behavior comprising fixed point, limit cycle and chaotic attractors17,23. We quantify the effects of confinement and crowding on the population dynamics, flux of information and spatial organization within the MWR net- work. Our approach and analysis can be naturally extended to more complicated chemical networks and can be potentially relevant to a number of open problems in biochemistry such as the synthesis of primitive cellular units (protocells) and the definition of their role in the chemical origin of life, the characterization of vesicle-mediated drug delivery processes and, more generally, the study of biochemical networks in vivo4,31 -- 33. We make the case for a more widespread development and use of spatial stochastic simulation methods for bio- chemical networks in vivo that explicitly take into account confinement and macromolecular crowding12,34 -- 36. II. METHODS All three autocatalytic species A, B and C are spatially confined within a spherical container, E1 and E3 catalyze the synthesis of A, and C, respectively, whereas E2 catalyzes the degradation of C. P1 and P2 are the products of reactions (1d) and (1c), respectively, and they get instantaneously eliminated from the reaction pool, i.e., their constant population number is zero. The constant population numbers of E1, E2 and E3 and the instantaneous elimination of P1 and P2 lead to a biochemical network composed by A, B and C which is spatially enclosed and thermodynamically open, i.e., it exchanges matter and energy with the surrounding environment by means of three sources (E1, E2 and E3) and two sinks (P1 and P2). The constant values of E1, E2, E3 are incorporated into the pseudo-first order rates k1, k3, k5, respectively (see Figure 1). The different chemical species in the MWR system are modeled as reactive, Brownian hard spheres confined in a spherical container. The details of the Brownian integrator used in our simulations can be found in Refs. 26 and 27. The radius of the hard spheres for species A, B, and C is 0.01 µm and the diffusion coefficient is D = 0.01 µm2s−1. In all our simulations the time step is fixed at ∆t = 0.01 s. To study the effects of crowding and confinement we run two separate sets of reactive 3 Chemical reaction networks under spatial confinement and crowding conditions Brownian dynamics simulations. In the first set we consider six different spherical containers with radius varying between 0.4 and 0.65 µm. The containers are implemented as 'hard- wall' spherical boundary conditions. For each of the six spherical containers we run a total of 30 independent simulations, each of total time ttot = 1000 s. Three sets of values for the reaction rate constants (kset1, kset2, kset3 ) are used for each one of the six different spherical containers. They correspond to three distinct dynamical behaviors in the deter- ministic implementation of the MWR model: fixed point, limit cycle and chaotic dynamics, respectively. The first set (kset1, fixed point attractor) is k1 = 30.0, k−1 = 0.25, k2 = 1.0, k3 = 10.0, k4 = 0.4, k5 = 16.5, k−5 = 0.5. To generate the second set (kset2 - limit cycle attractor) we simply consider the first set of parameters and change the value of k4 to 0.6. In the third set (kset3 - chaotic attractor) we set k4 = 0.6, k5 = 18.5 and k−5 = 0.4. In other words, kset2 is generated from kset1 by increasing the degradation of A and C (increasing the coupling between the Lotka-Volterra component and the switch) while kset3 is obtained from kset1 by increasing both the A − C coupling and decreasing the ratio k−5/k5. We run 10 independent simulations for each of the three parameter sets. The starting point for each simulation is generated randomly placing A = B = C = 100 hard spheres within the proper spherical container. In the second set of simulations we take into account the presence of a variable number of 'chemically inert' crowders modeled as hard spheres of radius r = 0.01 µm and with diffusion coefficient D = 0.01 µm2s−1. The starting point for each simulation in the second set is generated randomly placing A = B = C = 100 hard spheres and a variable number of inert crowders in a spherical container with radius R = 0.4 µm. We run independent simulations for five different crowder population numbers: varying between 2 × 103 and 8 × 103. For each of the five crowder population numbers we run a total of 30 independent simulations (10 for each of the three parameters sets), each of total time ttot = 1000 s. III. RESULTS AND DISCUSSION A. Population dynamics We focus our analysis on the stationary37 portion of our Brownian simulations. In Figure 2 we show a set of representative time windows for the population numbers of species A, B 4 Chemical reaction networks under spatial confinement and crowding conditions and C related to simulations with variable container volume (left panel), and to simulations with constant container volume and varying number of inert crowders (right panel). All data refer to parameterization kset3 (see Section II). Time series population data generated under kset1 and kset2 are not shown as they display analogous temporal patterns. Figure 2 qualitatively shows that (1) both average population and fluctuations increase with increas- ing container volume for all species and (2) the presence of an increasing number of inert crowders affects the average population of species A, B and C. It also appears to lower both the fluctuations in the population dynamics and the temporal interdependence between the different species. A quantitative assessment of the mean and fluctuations dependence from both the container volume and the crowders number is given in Figure 3. In the left panel we show that in the limited range of container volumes considered in our simulations, the mean population increases linearly with increasing container volume. The fluctuations cal- culated as the standard deviation from the mean also have a tendency to increase although the actual functional dependency is not immediately clear. The effects of the presence of inert crowders are shown in the right panel of Figure 3. All species show a decrease in their average population for increasing crowders numbers which can be intuitively related to the diminished availability of free volume within the spherical container. An additional observation on the data in Figure 3 relates to the dependence of the average population from the parameterization set. First, the population dynamics of species A, B and C does not change significantly when the parameterization set changes from kset1 to kset2. Second, the transition from parameterization kset1 and kset2 to kset3 has opposite effects on species B and C. Third, species A does not show any quantifiable dependence from the parameter set (under both volume and crowders' number varying conditions). It is easy to connect the increase in the slope of the C(volume) linear fit to the increase in C's net synthesis going from kset1 and kset2 to kset3. The decrease in the linear fit's slope for species B is less clear since species B is not directly affected by the changes in the parameterization set and species A, which is directly coupled to B, is insensitive to those changes. The insensitiv- ity to parameter changes in species A can be qualitatively explained considering that A is the connection point in the MWR network between the Lotka-Volterra component and the switch component18 and therefore benefits from the 'modulation' given by the interaction with both species B and C. 5 Chemical reaction networks under spatial confinement and crowding conditions B. Information flux and statistical complexity A possible explanation of the peculiar behavior of species A and its relation with A's 'double coupling' within the MWR network comes from the analysis of the information flux quantified by the transfer entropy defined as a particular case of the conditional mutual information38 -- 40: (2) where I(X, Y Z) is the mutual information between X and Y given Z 41 and τ is the time delay. We employ transfer entropy to estimate both the amount and the direction of the max I(Y (t), X(t + τ )X(t)), τ information flux in the MWR network. The value of the delay parameter τ considered in our calculations corresponds to the maximum of the transfer entropy for a given trajectory pair. A number of interesting conclusions can be inferred from the analysis of the transfer entropy data in Figures 4 and 5. Considering first the chemical network as a whole it is worth noting that the varying container volume does not significantly affect the information flux between the different species in the network. For systems with variable number of inert crowders there is a small but noticeable systematic increase in the transfer entropy with differences between the less and the most crowded systems of the order of 0.3 − 0.4 bits. A further look at the behavior of the single species shows the pivotal role of species A as a common influencer of the dynamics of species B and C. Both Figure 4 and 5 show that the amount of information transferred from species A is systematically larger than the information transferred to species A in both volume-varying and crowding number-varying systems. This asymmetry in the information flux (common to all three parameterization sets kset1, kset2 and kset3 - data not shown) can be linked to an increased ability of A to 'absorb' external perturbations and therefore to its lower sensitivity to parameter change (see previous Section). The main difference in terms of information transfer between systems with and without inert crowders (see Figures 4 and 5, respectively) is in the characteristic time delay τ (see Equation 2) at which the information transfer is maximal. While the characteristic delay τ is not affected by changes in the container volume (right panel in Figure 4), the presence of an increasing number of inert crowders in a constant volume container decreases the speed at which the information is transferred within the network (right panel in Figure 5). In order to improve the clarity and conciseness of our manuscript, from now on we focus 6 Chemical reaction networks under spatial confinement and crowding conditions only on simulations performed under parameterization set kset3 as this set of parameters seems to have an additional layer of complexity with respect to kset1 and kset2 (data not shown) and carries all the significant information about our system. Information theory functionals can be also used to estimate the degree of complexity in the time evolution of the chemical network and its dependence from the container volume and from the presence of crowders. The complexity estimation quantity that we choose is an intensive statistical complexity measure which is the product of the normalized spectral entropy S(Pr) and the intensive Jensen-Shannon divergence Q(Pr, Pe)42,43 defined respectively as: S(Pr) = −S0 with Pr(cid:48) log2 Pr(cid:48) Nf(cid:88) r(cid:48) r(cid:80)Nf f 2 r(cid:48) f 2 r(cid:48) , Pr = (3) (4) where fr are the frequencies in the Fourier spectrum and Nf = 4000 is the number of frequencies considered, and Q(Pr, Pe) = Q0 (cid:104) S (cid:16) Pr + Pe 2 (cid:17) − 1 2 S(Pe) − 1 2 (cid:105) S(Pr) , (5) where S(Pe) = log2 Nf = S−1 0 and Q0 is the normalization factor for Q and Pe = 1/Nf . The statistical complexity S Q is zero for both Pr = {1, 0, 0,··· , 0} and Pr = Pe = 1/Nf , i.e., for spectral entropy S = 0 and S = log2 Nf (fully ordered and fully stochastic systems)43. The results for the statistical complexity S Q are shown in Figure 6. The top panel shows that container volume variability does not significantly affect the average statistical complexity for species A, B and C (both Sand Q do not vary significantly - Deltaleq0.02). Conversely, for systems with constant volume and variable crowders number the statistical complexity decreases with increasing number of crowders. In detail, the decrease is almost exclusively due to a decrease in the normalized spectral entropy from 0.62 to 0.55, 0.61 to 0.54 and 0.58 to 0.50, for species A, B and C, respectively. The intensive Jensen-Shannon divergence remains constant at around 0.39− 0.40. As a general conclusion from our information theoretic analysis, we can state that the presence of a growing number of inert crowders drives the chemical network toward a lower degree of complexity which 7 Chemical reaction networks under spatial confinement and crowding conditions is possibly driven by a more efficient information transfer Figure 5) between the reactive chemical species. C. Spatial statistics The spatial organization of the chemical species in the network and its coupling with their population sizes are investigated employing a deterministic implementation of the DBSCAN clustering algorithm44, where 'boundary' particles are discarded as noise and with parameterization  = 0.06µm and nc = 4.  is the cutoff distance defining particle pairs belonging to the same cluster and therefore 'connected' to each other, and nc is the minimum number of 'connections' that defines a 'core' particle44. The presence of inert crowders strongly influences the spatial organization of the chemical species A, B and C in the MWR network. In Figure 7 we show the average number of clusters (top) and average maximum cluster size (bottom) as a function of the number of inert clusters. On the one hand, the average number of clusters shows a weak tendency to increase for all three chemical species. On the other hand, the average maximum cluster size decreases with denser crowding conditions. Among the three reactive species the maximum cluster size in species C displays both the largest values and the largest decrease rate. Figure 7 basically shows that the presence of an increasing number of crowders opposes the natural tendency of the reactive particles in our system to accumulate in well-defined regions of the available space. An interesting feature of the maximum cluster size temporal evolution is shown in Figure 8. For small numbers of crowders the maximum cluster size for species A tightly mirrors the time evolution of the population of species A (species B and C show very similar behavior - data not shown). The 'correlation' between population dynamics and maximum cluster dynamics weakens with increasing crowders number. Indeed, table I shows that the mutual information41 between population and maximum cluster dynamics decreases with increasing crowders numbers. IV. CONCLUSIONS In this study we investigate the dynamical behavior of a simple chemical network under spatial confinement and crowding. We observe that the presence of inert crowders affects in a 8 Chemical reaction networks under spatial confinement and crowding conditions non-trivial way the population dynamics of the reactive species in the network. The detailed analysis of the population dynamics of the MWR network under different confinement and crowding conditions presented in Section III represents, from a more general perspective, an example of the level of detail, not accessible to deterministic and stochastic well-mixed models, that can be resolved when spatial confinement and crowding are explicitly taken into account. In conclusion, we try to make the case for the use of spatial stochastic simulations as an elective method to complement experiments and to improve our understanding of complex systems where dynamics is both spatially confined and compartmentalized. ACKNOWLEDGMENTS The authors would like to thank Linda Petzold, Zachary Frazier, Frank Alber, Marco J. Morelli and Steve Plimpton for useful discussions and for the help with the testing of our Brownian simulator. This work has been supported by NIH grant 1R01EB014877-01. REFERENCES 1H. Qian, J Phys Chem B 110, 15063 (2006). 2H. Qian, Annu Rev Phys Chem 58, 113 (2007). 3F. Ritort, Nonequilibrium fluctuations in small systems: from physics to biology (John Wiley and Sons, Inc., 2008), vol. 137, chap. 2, pp. 31 -- 123. 4P. Stano and P. L. Luisi, Curr Opin Biotechnol 24, 633 (2013). 5A. D. McNaught and A. Wilkinson, IUPAC. Compendium of Chemical Terminology (Blackwell Scientific Publications, Oxford, 1997), 2nd ed. 6D. H. Lee, K. Severin, and M. R. Ghadiri, Current Opinion in Chemical Biology 1, 491 (1997). 7Z. Dadon, N. Wagner, and G. Ashkenasy, Angew Chem Int Ed Engl 47, 6128 (2008). 8N. Virgo and T. Ikegami, in ECAL - General Track (2013). 9W. Hordijk, J. Hein, and M. Steel, Entropy 12, 1733 (2010). 10H. A. Johnson, Q Rev Biol 62, 141 (1987). 11D. T. Gillespie, The Journal of Physical Chemistry 81, 2340 (1977). 12A. P. Minton, Biopolymers 20, 2093 (1981). 9 Chemical reaction networks under spatial confinement and crowding conditions 13A. P. Minton, Methods in Enzymology 295, 127 (1998). 14A. P. Minton, Journal of Biological Chemistry 276, 10577 (2001). 15K. Richter, M. Nessling, and P. Lichter, J Cell Sci 120, 1673 (2007). 16K. Richter, M. Nessling, and P. Lichter, Biochim Biophys Acta 1783, 2100 (2008). 17K. D. Willamowski and O. Rossler, Zeitschrift fur Naturforschung 35a, 317 (1980). 18B. D. Aguda and B. L. Clarke, The Journal of Chemical Physics 89, 7428 (1988). 19J. Guemez and M. A. Matias, Physical Review E 48, R2351 (1993). 20P. Geysermans and G. Nicolis, The Journal of Chemical Physics 99, 8964 (1993). 21A. Goryachev and R. Kapral, Physical Review Letters 78, 1619 (1996). 22F. Chavez and R. Kapral, Physical Review E 65, 056203 (2002). 23J. W. Stucki and R. Urbanczik, Zeitschrift fur Naturforschung 60a, 599 (2005). 24T. Ando and J. Skolnick, Proceedings of the National Academy of Sciences 107, 18457 (2010). 25T. Ando and S. Jeffrey, in Proceedings of the International Conference of the Quantum Bio-Informatics IV (Georgia Institute of Technology, World Scientific Publishing, 2011), vol. 28, pp. 413 -- 426. 26M. J. Morelli and P. R. ten Wolde, J Chem Phys 129, 054112 (2008). 27Z. Frazier and F. Alber, J Comput Biol 19, 606 (2012). 28J. A. Lotka, The Journal of Chemical Physics 14, 271 (1910). 29J. A. Lotka, Journal of the American Chemical Society 42, 1595 (1920). 30V. Volterra, Nature 118, 558 (1926). 31S. Rasmussen, M. A. Bedau, L. Chen, D. Deamer, D. C. Krakauer, N. H. Packard, and P. F. Stadler, eds., Protocells, Bridging Nonliving and Living Matter (MIT Press (Cambridge Mass.), 2008). 32P. L. Luisi, F. Ferri, and P. Stano, Naturwissenschaften 93, 1 (2006). 33E. Szathmary, Nature 433, 469 (2005). 34M. Beck, M. Topf, Z. Frazier, H. Tjong, M. Xu, S. Zhang, and F. Alber, Journal of Structural Biology 173, 483 (2011). 35D. T. Gillespie, A. Hellander, and L. Petzold, The Journal of Chemical Physics 138, 170901 (2013). 36O. S. Andersen, J Gen Physiol 125, 3 (2005). 10 Chemical reaction networks under spatial confinement and crowding conditions 37N. G. Van Kampen, Stochastic Processes in Physics and Chemistry (Elsevier, 2007), 3rd ed. 38T. Schreiber, Physical Review Letters 85, 461 (2000). 39K. Hlavackova-Schindler, M. Palus, M. Vejmelka, and J. Bhattacharya, Physics Reports 441, 1 (2007). 40M. Vejmelka and M. Palus, Phys Rev E Stat Nonlin Soft Matter Phys 77, 026214 (2008). 41T. M. Cover and J. A. Thomas, Elements of information theory (Wiley-Interscience, Hobo- ken, N.J., 2006), 2nd ed. 42G. E. Powell and I. C. Percival, Journal of Physics A: Math. Gen. 12, 2053 (1979). 43O. A. Rosso, H. A. Larrondo, M. M. T, A. Plastino, and M. A. Fuentes, Physical Review Letters 99, 154102 (2007). 44M. Ester, H.-P. Kriegel, J. Sander, and X. Xiaowei, in Proceedings of the Second In- ternational Conference on Knowledge Discovery and Data Mining (KDD-96), edited by E. Simoudis, J. Han, and U. M. Fayyad (1996). 11 Chemical reaction networks under spatial confinement and crowding conditions TABLE I. Mutual information between population size and largest cluster size for species A, B and C and for different numbers of inert crowders. Inert Crowders A B C 0 2000 4000 6000 8000 0.969 0.811 1.574 0.893 0.643 1.359 0.847 0.619 0.806 0.611 0.583 0.735 0.568 0.471 0.442 12 Chemical reaction networks under spatial confinement and crowding conditions FIG. 1. Schematic view of the Willamowski-Rossler chemical network. (Left) the full version. (Right) the minimal version analyzed in this study, obtained by setting k−2 = k−3 = k−4 = 0. In red we highlight the Lotka-Volterra component of the network. The rate equations for both versions exhibit three different dynamical regimes: fixed point, limit cycle and chaotic17,23. 13 Chemical reaction networks under spatial confinement and crowding conditions FIG. 2. Population time series (partial time window) obtained each from a single trajectory of our brownian dynamics simulations of the MWR network with kset3 parameterization. Left panel: population time series for systems with varying volume (radii varying from 0.4 to 0.65 µm). Right panel: population time series for systems with varying number of crowders and with constant container volume (radius is 0.4 µm). 14 Chemical reaction networks under spatial confinement and crowding conditions FIG. 3. Average population values for species A (top), B (middle) and C (bottom) with param- eterizations kset1, kset2 and kset3. Left panel: the average populations are plotted against the volume of the spherical container, no inert crowders are present. The average population of the reactive species grows linearly with the volume of the spherical container . Right panel: the aver- age populations are plotted against the number of inert crowders. A slight decrease in the average population is observed for larger crowders numbers. 15 Chemical reaction networks under spatial confinement and crowding conditions FIG. 4. Left panel: transfer entropy as a function of the container volume. Right panel: character- istic time delay correspondent to the maximum value of the transfer entropy. Data refer to kset3 parameterization. 16 Chemical reaction networks under spatial confinement and crowding conditions FIG. 5. Left panel: transfer entropy as a function of the number of inert crowders. Right panel: characteristic time delay correspondent to the maximum value of the transfer entropy. Data refer to kset3 parameterization. 17 Chemical reaction networks under spatial confinement and crowding conditions FIG. 6. Statistical complexity measure as a function of the container volume (top) and number of inert crowders (bottom). Data refer to species A, B, C under kset3 parameterization. 18 Chemical reaction networks under spatial confinement and crowding conditions FIG. 7. Average number of clusters (top) and average maximum cluster size (bottom) as a function of the number of inert clusters. Data refer to species A, B, C under kset3 parameterization. The average number of clusters shows a weak tendency to increase for all three chemical species. The average maximum cluster size decreases with denser crowding conditions. The max. cluster size in species C displays the largest decrease rate. 19 Chemical reaction networks under spatial confinement and crowding conditions FIG. 8. Time evolution for the population size (blue) and the maximum cluster size (yellow). The temporal pattern in the maximum cluster size accurately mirrors the population size. Data refer to species A under kset3 parameterization. The mutual information between the two sets of temporal data decreases with increasing number of crowders (see Table I). 20
1509.01029
1
1509
2015-09-03T10:51:42
Immune response to functionalized mesoporous silica nanoparticles for targeted drug delivery
[ "physics.bio-ph", "cond-mat.soft", "q-bio.CB" ]
Multifunctional mesoporous silica nanoparticles (MSN) have attracted substantial attention with regard to their high potential for targeted drug delivery. For future clinical applications it is crucial to address safety concerns and understand the potential immunotoxicity of these nanoparticles. In this study, we assess the biocompatibility and functionality of multifunctional MSN in freshly isolated, primary murine immune cells. We show that the functionalized silica nanoparticles are rapidly and efficiently taken up into the endosomal compartment by specialized antigen-presenting cells such as dendritic cells. The silica nanoparticles showed a favorable toxicity profile and did not affect the viability of primary immune cells from the spleen in relevant concentrations. Cargo-free MSN induced only very low immune responses in primary cells as determined by surface expression of activation markers and release of pro-inflammatory cytokines such as Interleukin-6, -12 and -1\beta. In contrast, when surface-functionalized MSN with a pH-responsive polymer capping were loaded with an immune-activating drug, the synthetic Toll-like receptor 7 agonist R848, a strong immune response was provoked. We thus demonstrate that MSN represent an efficient drug delivery vehicle to primary immune cells that is both non-toxic and non-inflammagenic, which is a prerequisite for the use of these particles in biomedical applications.
physics.bio-ph
physics
Primary immune response to MSN response to silica functionalized mesoporous Immune nanoparticles for targeted drug delivery Simon Heidegger1,2, Stefan Niedermayer3, Alexandra Schmidt3, Dorothée Gössl3, Christian Argyo3, Stefan Endres1, Thomas Bein3,*, Carole Bourquin1,4,* 1 Center for Integrated Protein Science Munich (CIPSM), Division of Clinical Pharmacology, Medizinische Klinik und Poliklinik IV, Ludwig-Maximilians-Universität München, 80336 Munich, Germany 2 III. Medizinische Klinik, Klinikum Rechts der Isar, Technische Universität München, 81675 Munich, Germany 3 Department of Chemistry and Center for NanoScience (CeNS), University of Munich (LMU), 81377 Munich, Germany 4 Chair of Pharmacology, Department of Medicine, Faculty of Science, University of Fribourg, 1700 Fribourg, Switzerland * These authors contributed equally. Corresponding Authors: Carole Bourquin, Chair of Pharmacology, Department of Medicine, Faculty of Science, University of Fribourg, 1700 Fribourg, Switzerland ([email protected]); Thomas Bein, Department of Chemistry and Center for NanoScience (CeNS), University of Munich (LMU), 81377 Munich, Germany ([email protected]) 1 ABSTRACT Primary immune response to MSN Multifunctional mesoporous silica nanoparticles (MSN) have attracted substantial attention with regard to their high potential for targeted drug delivery. For future clinical applications it is crucial to address safety concerns and understand the potential immunotoxicity of these nanoparticles. In this study, we assess the biocompatibility and functionality of multifunctional MSN in freshly isolated, primary murine immune cells. We show that the functionalized silica nanoparticles are rapidly and efficiently taken up into the endosomal compartment by specialized antigen-presenting cells such as dendritic cells. The silica nanoparticles showed a favorable toxicity profile and did not affect the viability of primary immune cells from the spleen in relevant concentrations. Cargo-free MSN induced only very low immune responses in primary cells as determined by surface expression of activation markers and release of pro-inflammatory cytokines such as Interleukin-6, -12 and -1β. In contrast, when surface- functionalized MSN with a pH-responsive polymer capping were loaded with an immune- activating drug, the synthetic Toll-like receptor 7 agonist R848, a strong immune response was provoked. We thus demonstrate that MSN represent an efficient drug delivery vehicle to primary immune cells that is both non-toxic and non-inflammagenic, which is a prerequisite for the use of these particles in biomedical applications. 2 INTRODUCTION Primary immune response to MSN In recent years the development of nanoparticles for various biomedical applications has been the focus of intense research.1 Rationally designed nanoparticles are engineered for targeted delivery of various drugs or vaccines.2 With spatio-temporally controlled release of their therapeutic cargo, such nanoparticles have the potential to increase drug efficacy while minimizing undesired off-target effects. However, due to their nanoscale size, chemical composition and surface reactivity, nanoparticles can be potentially detected by and interact with the host immune response.3 While in certain applications (e.g. vaccine delivery) an immunostimulatory function may be desirable,4 uncontrolled systemic immune activation will limit their therapeutic use. Thus, the profound understanding of a nanomaterial’s interaction with the immune system and its possible stimulatory and suppressive actions are a critical prerequisite for any clinical application. Specialized antigen-presenting cells (APC) of the innate immune system, such as dendritic cells (DC), constantly sample their surroundings, taking up cell debris and foreign materials. These cells are equipped with a variety of pattern-recognition receptors that allow for the detection of invading pathogens or signs of cell stress and damage. Ligation of these receptors results in maturation of APCs associated with expression of costimulatory molecules (such as CD80 and CD86), the release of proinflammatory cytokines and eventually the initiation of a subsequent adaptive immune response.5 DCs have been shown to efficiently engulf various kinds of nanoparticles both in vitro and in vivo,6,7 but the consequences with regard to subsequent immune responses remain poorly understood. Previous work of some of us has been focused on the development of a portfolio of core- shell colloidal mesoporous silica nanoparticles.8, 9 Due to their different molecular functionalization in the core and the shell, the MSN can be equipped with internal functionality for controlled host-guest interactions, a system for accurate cargo release upon external stimuli, as well as targeting ligands towards the required type of cell. A broad variety of different triggered-release capping systems were presented in the past years, based on external stimuli such as light10, 11, heat,12 or magnetic fields,13 or triggered by intracellular events including a change in pH,14, 15 redox reactions,16 or the presence of enzymes.17 For example, we recently reported a multifunctional system based on a pH-responsive polymer (poly(2-vinylpyridine), MSN-PVP) that allows for the facile delivery of membrane-permeable cargos into cancer cells.18 With the possibility to integrate almost any functionality of interest, as well as the efficient synthesis, this system holds promise for wide-ranging biological and medical applications, especially in cancer therapy. 3 Primary immune response to MSN In this study, we investigate the immune-modulatory properties of MSN in primary murine splenocytes. We show that MSN themselves provoke only a marginal immune response but when loaded with a synthetic immune activator can function as an efficient delivery tool for potent immune activation. METHODS tetrahydrofuran Materials. All reagents were purchased from commercial suppliers. Tetraethyl orthosilicate (TEOS, Sigma-Aldrich, >98%), cetyltrimethylammonium chloride (CTAC, Fluka, 25wt% in H2O), triethanolamine (TEA, Sigma-Aldrich, 98%), (3-aminopropyl)-triethoxysilane (APTES, Fluka, 95%), N-(3-dimethylaminopropyl)-N-ethylcarbodiimide hydrochloride (EDC, Sigma- Aldrich, 97%), ammonium nitrate (Sigma-Aldrich, 99%), conc. hydrochloric acid (Sigma- Aldrich, >95%, 37 wt%), α-amino-ω-carboxy terminated poly(2-vinylpyridine) (NH2-PVP- COOH, Polymer Source, Mn = 10.000, PDI = 1.08), Boc anhydride (Aldrich, 95%), N- hydroxysulfosuccinimide sodium salt (sulfo-NHS, Sigma-Aldrich, 98%), ethanol (EtOH, Sigma-Aldrich, >99.5%), (THF, anhydrous, Sigma-Aldrich, ≥99.9%), trifluoroacetic acid (TFA, Acros Organics, 99%), magnesium sulfate (MgSO4, anhydrous, Sigma-Aldrich, ≥99.5%), triethylamine (Sigma-Aldrich, ≥99%), fluorescein isothiocyanate (FITC, Fluka, >90%), saline-sodium citrate buffer concentrate (SSC-buffer (20x), Sigma- Aldrich), citric acid buffer solution (citric acid/HCl/NaCl buffer, pH 2, Sigma-Aldrich), TLR7 agonist R848 (Invivogen, Toulouse). All chemicals were used as received without further purification. Doubly distilled water from a Millipore system (Milli-Q Academic A10) was used for all synthesis and purification steps. Preparation of shell-functionalized colloidal mesoporous silica nanoparticles. Polymer- capped mesoporous silica nanoparticles (MSN-PVP) were synthesized as described previously.8, 18 In brief, colloidal MSN with an amino-functionality (MSN-NH2) were prepared by a delayed co-condensation approach with cetyltrimethylammonium chloride (CTAC) as structure directing agent and tetraethylorthosilicate (TEOS) as primary silica source.8 In order to introduce the outer shell functionalization, 3-aminotriethoxysilane (APTES) was added during the condensation process. The synthesis was followed by a facile extraction of the organic template from the mesopores. Then, 50 mg of the template-extracted MSN-NH2 were subsequently reacted with a boc-protected bi-functional polymer (COOH-PVP-NHboc) dissolved in THF in an EDC-assisted amidation reaction. The reaction was followed by the deprotection of the polymer with trifluoroacetic acid, yielding the sample MSN-PVP.18 4 Primary immune response to MSN Fluorescein labeling of the MSN. Fluorescein isothiocyanate (FITC)-labeling of MSN-NH2 was performed following a described procedure.19 In brief, 50 mg of MSN-NH2 dispersed in ethanol (25 mL) were added to an ethanolic solution of FITC (25 mL, containing 7.4 mg FITC, 0.19 mmol). The suspension was stirred at ambient temperature in the dark for 24 h. The resulting sample MSN-FITC was collected by centrifugation (19.000 rpm, 43.146 RCF, 20 min) and washed three times with ethanol (25 mL each) by subsequent centrifugation and redispersion. After the final centrifugation step the particles were resuspended in absolute ethanol. For cell experiments, 1 mg of the labeled MSN were centrifuged and resuspended in 1 mL phosphate-buffered saline (PBS). MSN loading with the synthetic TLR7 agonist R848. 1 mg MSN-PVP were dispersed in 240 µL of sterile water. To open the pores and enable the uptake of the respective drug molecule, 50 µL of citrate buffer (pH 2) was added, followed by the addition of 10 µL R848 of the stock solution (1 mg/mL), yielding an overall drug concentration of 10 µg/300 µL in the solution. The particles were stirred overnight, centrifuged, and re-dispersed in 700 µL SSC buffer (pH 7) to enable the closure mechanism of the pH-responsible polymer. The resulting particles were washed extensively with SSC buffer (pH 7) and finally redispersed in 1 mL SSC. For incubation with primary cells, MSN were centrifuged and redissolved in complete RPMI medium (see below). Characterization of MSN. Centrifugation was performed using a Sorvall Evolution RC equipped with a SS-34 rotor or an Eppendorf centrifuge 5418 for small volumes. All samples were investigated with an FEI Titan 80-300 transmission electron microscope operating at 300 kV with a high-angle annular dark field detector. A droplet of the diluted MSN solution in absolute ethanol was dried on a carbon-coated copper grid. Nitrogen sorption measurements were performed on a Quantachrome Instruments NOVA 4000e. All samples (15 mg each) were heated to 60 °C for 12 h in vacuum (10 mTorr) to outgas the samples before nitrogen sorption was measured at 77 K. Pore size and pore volume were calculated by a NLDFT equilibrium model of N2 on silica, based on the desorption branch of the isotherms. In order to remove the contribution of the interparticle textural porosity, pore volumes were calculated only up to a pore size of 8 nm. A BET model was applied in the range of 0.05 – 0.20 p/p0 to evaluate the specific surface area of the samples. Dynamic light scattering (DLS) measurements were performed on a Malvern Zetasizer-Nano instrument equipped with a 4 mW He-Ne laser (633 nm) and an avalanche photodiode. The hydrodynamic radius of the particles was determined by dynamic light scattering in ethanolic or aqueous suspension. For this purpose, 100 µL of an ethanolic suspension of MSN particles (ca. 10 mg/mL) was diluted with 3 mL of ethanol or water prior to the measurement. Zeta potential measurements of the 5 Primary immune response to MSN samples were performed on a Malvern Zetasizer-Nano instrument equipped with a 4 mW He- Ne laser (633 nm) and an avalanche photodiode. Zeta potential measurements were performed using the add-on Zetasizer titration system (MPT-2) based on diluted NaOH and HCl as titrants. For this purpose, 1 mg of the particles was diluted in 10 mL bi-distilled water. Thermogravimetric analysis was performed on a Netzsch STA 440 C TG/DSC with a heating rate of 10 K / min in a stream of synthetic air of about 25 mL/min. The mass was normalized to 100% at 150 °C for all samples to take into account solvent desorption. Mice. Female C57Bl/6 mice were purchased from Harlan-Winkelmann. Mice were at 6-8 weeks of age at the time of the experiment. Animal studies were approved by the local regulatory agency (Regierung von Oberbayern, Munich, Germany). Media, reagents and cell culture. Single cell suspensions from spleens were obtained by filtering through a 100 µm cell strainer (BD Biosciences, Heidelberg, Germany). Erythrocytes were lysed with ammonium chloride buffer (BD Biosciences). Cells were cultured in complete RPMI (Roswell Park Memorial Institute) 1640 medium (10% fetal calf serum (FCS), 2 mM L-glutamine, 100 µg/mL streptomycin and 1 IU/mL penicillin) at 37 °C in 10% CO2. In some conditions, CpG 1826 (unmethylated CpG sequence-containing oligonucleotides, a TLR9 agonist, 3 µg/mL, from Invivogen, Toulouse, France) was added to the culture. For maximal IL-1β release, DCs were primed with lipopolysaccharide (LPS, a TLR4 agonist, 20 ng/ml, from Invivogen) overnight and ATP (5 mM, from Sigma-Aldrich) was added to the culture 2 hours prior to the analysis. For exposure to nanoparticles, splenocytes were seeded in complete RMPI at a density of 1.0 x 106 / mL in 96-well tissue culture plates. Silica nanoparticles were added in complete RPMI medium at the indicated concentration. After 18 - 24 hours, cells and culture supernatant were analyzed. Quantification of cytokines. Cell supernatants were analyzed for cytokine secretion by ELISA (R&D Systems or eBioscience) according to the manufacturers’ protocol. Flow cytometry and apoptosis assay. Cell suspensions were stained in PBS with 1% FCS. Fluorochrome-coupled antibodies against the surface antigens B220, CD3, CD4, CD8, CD11b, CD11c, CD69, CD80, F4/80 and appropriate isotype controls were purchased from BioLegend. Data were acquired on a FACSCanto II (BD Biosciences) and analyzed using FlowJo software (Tree Star, Ashland, OR). The Annexin V-FITC Apoptosis Detection KIT 1 (BD Biosciences) was used for detection of apoptotic cells. Following surface staining and 2 washing steps, single-cell suspensions were resuspended in the provided buffer, incubated with Annexin V - FITC and propidium iodide (PI) and subsequently analyzed by flow 6 Primary immune response to MSN cytometry. Terminal deoxynucleotidyltransferase dUTP nick end labeling (TUNEL staining) was done with the APO-BRDU™ Kit (BD Pharmingen) according to the manufacturer’s protocol. Confocal Microscopy. Splenocytes were incubated for 4 hours with fluorescein-tagged NPs, washed and re-suspended in culture medium. 75 nM LysoTracker(cid:31) (Invitrogen) and 3 µg/mL Hoechst dye (Invitrogen) were used for lysosomal and nuclear staining. Stained cells were visualized using a confocal laser scanning microscope (TCS SP5II, Leica). Statistics. All data are presented as mean ± S.E.M. Statistical significance of single experimental findings was assessed with the independent two-tailed Student’s t-test. For multiple statistical comparison of a data set the one-way ANOVA test with Bonferroni post-test was used. Significance was set at p-values p < 0.05, p < 0.01 and p < 0.001 and was then indicated with an asterisk (*, ** and ***). All statistical calculations were performed using Graphpad Prism (GraphPad Software). 7 RESULTS MSN particle characteristics. Primary immune response to MSN As reported previously,18 the template-free MSN-NH2 show a wormlike pore structure with an average pore size of 3.8 nm and feature a large surface area (1097 m2/g) which is typical for MSN (Table 1 and Fig. 1). Figure 1: Structure of mesoporous silica nanoparticles. (A) Transmission electron micrograph of a template-extracted MSN-NH2, exhibiting a worm-like pore structure. (B) Schematic illustration of the pH-responsive nanocarrier system (MSN-PVP) employed in this work at a pH value of 5 (open state). The inorganic-organic hybrid material consists of a mesoporous silica core (grey) and a covalently attached pH-responsive polymer (poly(2-vinylpyridine), blue). The covalent modification of MSN-NH2 with a boc-protected poly(2-vinylpyridine) and the following de-protection was monitored with several characterization methods; the accessible pore size as determined by nitrogen sorption measurements is barely affected by the surface modification with the pH-responsive polymer, whereas the decrease of the specific surface area is relatively large. This can be explained by the addition of non-porous polymer to the outer surface and the blocking of some pores by frozen polymer on the surface layer of the MSN. Dynamic light scattering (DLS) measurements in aqueous media revealed an average particle size of 160 nm for MSN-NH2 and 550 nm for MSN-PVP. The polymer-modified sample shows some aggregation behaviour in water due to the hydrophobicity at pH 7, indicated by the apparent size increase to 550 nm. However, transmission electron microscopy (TEM) revealed that the polymer-functionalized sample MSN-PVP still features a narrow particle size distribution, which makes them excellent candidates as drug delivery vehicles. 8 Table 1. Key features of pH-responsive MSN-PVP. Primary immune response to MSN Sample Particle sizea [nm] BET surface area [m²/g] Pore sizeb [nm] Relative mass lossc [%] MSN-NH2 MSN-PVP-NH2 160 550 1097 617 3.8 3.4 15 62 a) Particle size refers to the peak value derived from dynamic light scattering (DLS). b) Non-linear density functional theory (NLDFT) pore size refers to the peak value of the pore size distribution. c) Relative mass loss obtained by thermogravimetric analysis (TGA). All curves were normalized to 150 °C.18 Efficient uptake of mesoporous silica nanoparticles by specialized antigen-presenting cells To test whether MSN-NH2 can serve as a delivery tool in primary immune cells, freshly isolated mouse splenocytes that harbor a variety of different immune cells were cultured in the presence of fluorescently-labeled MSN (MSN-FITC) overnight. The uptake of labeled MSN-FITC by different cell types was analyzed by flow cytometry. Cells of the innate immune system, that includes macrophages and dendritic cells, showed high uptake of MSN-FITC, as measured by fluorescence signal-positive cells (Fig. 2A). As such, dendritic cells, which are highly specialized in antigen uptake, processing and presentation, were more efficient in uptake than macrophages. In contrast, T and B cells, which are the effector cells of the adaptive immune system, showed only trace fluorescence signal positivity. The intracellular uptake of MSN-FITC was clearly concentration-dependent (Fig. 2B). Fluorescent microscopy showed a speckled distribution pattern of fluorescent signals within dendritic cells, suggesting uptake of labeled MSN into distinct intracellular compartments but not into the cytosol or nucleus (Fig. 2C). Indeed, counter-staining with a fluorescent marker for lyso- /endosomes showed co-localization with the fluorescein-labeled MSN. In summary, MSN show rapid and efficient uptake into the endosomal compartment of specialized antigen- presenting cells such as dendritic cells but not into adaptive immune cells. 9 Primary immune response to MSN Figure 2: Mesoporous silica nanoparticles are efficiently taken up by specialized antigen- presenting cells. (A, B) Freshly isolated, total splenocytes were incubated for 18 h with different concentrations of fluorescein-labeled MSN (MSN-FITC). The uptake of fluorescence signals by different cell populations was determined by flow cytometry. (A) Representative histograms are gated on the indicated cell subset. The numbers give the percentage of fluorescein highly positive-stained cells. (B) Diagrams show the mean percentage of fluorescein-positive cells of triplicate samples ± s.e.m. An asterisk indicates comparison with unstimulated cells. (C) Complete splenocytes were incubated for 3 h with 0.1 µg/mL MSN-FITC. Cell endosomes were stained with LysoTracker™ and intracellular MSN fluorescence microscopy. All results are representative of at least two independent experiments. No stim., no stimulation. localization was determined by MSN-NH2 are non-toxic and do not affect the viability of splenocytes, unless used in very high concentrations We next sought to determine the cytotoxicity of MSN for primary cells. For this purpose, freshly isolated splenocytes were cultured in the presence of increasing concentrations of MSN-NH2 and 18 hours later, we performed Annexin V / propidium iodide (PI) analysis by flow cytometry. Annexin V binds to phosphatidylserins, which in apoptotic cells are translocated from the inner to the outer leaflet of the plasma membrane, and are thus exposed to the external cellular environment.20 PI is a small molecule that intercalates into double-stranded DNA and becomes fluorescent upon intercalation. PI can only reach nuclear DNA when the cell’s integrity is severely compromised during late apoptosis and cell death. The gating strategy for early and late apoptotic cells is depicted in the representative dot blot (Fig. 3A). MSN-NH2 showed a favorable (low) toxicity profile in primary cells and induced 10 Primary immune response to MSN marked apoptosis only when used in very high concentrations of 200 µg mL-1. The common cytotoxic chemotherapeutic drug oxaliplatin was used as a positive control. To confirm these data, we also performed terminal deoxynucleotidyltransferase dUTP nick end labeling (TUNEL). During the late phase of apoptosis endonucleases degrade the higher order chromatin structure into small DNA pieces. With the TUNEL assay these DNA fragments can be identified through addition of bromolated deoxyuridine triphosphates (Br-dUTP) to the 3'-hydroxyl (OH) termini of double- and single-stranded DNA by the endogenous enzyme terminal deoxynucleotidyl transferase (TdT) and subsequent staining with an FITC-labeled anti-BrdU antibody. The TUNEL analysis confirmed that MSN-NH2 are non-toxic to primary murine splenocytes over a wide concentration range (Fig. 3B). incubated Figure 3: MSN-NH2 are non-toxic and do not affect the viability of splenocytes, unless used in very high concentrations. Complete splenocytes were for 18 h with different concentrations of MSN-NH2. (A) Cell integrity and viability of splenocytes as determined by propidium iodide (PI) exclusion and Annexin V staining was analyzed by flow cytometry. The dot blot shows the gating strategy of viable (PI- Annexin V-), early (PI-, Annexin V+) and late apoptotic (PI+, Annexin V+) cells. (B) DNA fragmentation was assessed by TUNEL assay. The histogram shows the gating strategy for TUNEL+ cells. All data give the mean percentage of apoptotic cells of triplicate samples ± s.e.m. The mean base-line level of apoptosis in the untreated control group was set as zero %. An asterisk indicates comparison to untreated cells. Results are representative of at least two independent experiments. 11 Mesoporous silica nanoparticles induce only very low immunological responses in primary myeloid immune cells Primary immune response to MSN As a next step, we focused on the immunological response of mammalian primary immune cells to cargo-free MSN. For the implementation of any nanoparticle constructs as molecular delivery system, it is essential to fully understand their immune-modulatory potential in order to tightly control the initiation of immune responses according to the therapeutic goal. Thus, freshly isolated splenocytes were cultured in the presence of MSN-NH2. After 24 hours, surface expression of the B7 family member CD80, a co-stimulatory molecule and activation marker, was examined on different antigen-presenting cells by flow cytometry with specific fluorochrome-coupled antibodies. Additionally, the concentration of the secreted pro- inflammatory cytokines IL-6 and IL-12p70 in the culture supernatant of stimulated cells was quantified by enzyme-linked immunosorbent assay (ELISA). Synthetic unmethylated CpG sequence-containing oligonucleotides (CpG-ODN) that resemble bacterial DNA were used as a positive control. CpG-ODN bind to the endosomal Toll-like receptor (TLR) 9, thereby initiating a full-blown immune response.21 In comparison to CpG-ODN, MSN-NH2 induced only low levels of CD80 expression on monocytes and dendritic cells, indicating that the cargo-free nanoparticles barely activate immune responses (Fig. 4A). The secretion profile of pro-inflammatory cytokines confirmed these findings, as cells cultured with MSN-NH2 released only low amounts of IL-6 and IL-12p70 (Fig. 4B). Crystalline silica (found in nature as sand or quartz) have been shown to activate a cytosolic multi-protein complex called the NALP3 inflammasome resulting in the release of bioactive IL-1β, a very potent pro- inflammatory cytokine.22 In contrast, MSN-NH2 only induced trace levels of IL-1β (Fig. 4C). In summary, these data demonstrate that MSN-NH2 only mildly activate primary murine APCs. 12 Primary immune response to MSN Figure 4: Mesoporous silica nanoparticles induce only very low immunological responses in primary myeloid immune cells. Complete splenocytes were incubated for 18 h with different concentrations of MSN-NH2. (A) The surface expression of the co-stimulatory molecule CD80 on myeloid antigen-presenting cells was analyzed by flow cytometry. Data give the mean fluorescence intensity (MFI) of the indicated marker on triplicate samples ± s.e.m. The amounts of the pro- inflammatory cytokines (B) IL-6, IL-12p70 and (C) IL-1β in the cell culture supernatant were determined by ELISA. Data give the mean values of triplicate samples ± s.e.m. An asterisk indicates comparison to unstimulated cells. All results are representative of at least two independent experiments. Bystander lymphoid cells are not stimulated by mesoporous silica nanoparticles T and B lymphocytes are the effector cells of the adaptive immune system. We have shown that these cell types do not efficiently take up MSN-FITC and are thus unlikely to directly recognize these nanoparticles (Fig. 3A,B). However, T and B cells may react to low levels of pro-inflammatory cytokines released by antigen-presenting cells in response to MSN. In order to investigate the immunostimulatory effect of MSN on such bystander lymphocytes, complete splenocytes (containing both antigen-presenting cells and lymphocytes) were cultured in the presence of MSN-NH2 and expression of the transmembrane C-type lectin 13 Primary immune response to MSN CD69, an early activation marker on B and T cells, was analyzed by flow cytometry. We found that neither B nor T cells showed upregulation of the activation marker CD69 (Fig. 5), indicating that the low-level cytokine release by antigen-presenting cells in response to MSN-NH2 is not sufficient for activation of bystander lymphocytes. In summary, our findings show that MSN-NH2 are rapidly taken up into specialized antigen-presenting cells but are non-toxic and only weakly immunostimulatory to primary murine immune cells. Figure 5: Mesoporous silica nanoparticles do not result in activation of bystander lymphoid cells. Complete splenocytes were incubated for 18 h with different concentrations of MSN-NH2. Expression of the activation marker CD69 on effector (A) B lymphocytes and (B) T lymphocytes was analyzed by flow cytometry. The representative histogram is gated on B220+ B cells and illustrates CD69 expression (black dotted line, unstimulated cells; red line, 1 µg/mL MSN; blue line, CpG-DNA). Data in the graphs give the mean values of triplicate samples ± s.e.m. An asterisk indicates comparison to unstimulated cells. All results are representative of at least two independent experiments. 14 Functionalized mesoporous silica nanoparticles are an efficient delivery tool for the synthetic immunostimulatory TLR7 ligand R848 Primary immune response to MSN To test whether MSN can principally function as delivery tool in primary immune cells, MSN coated with a pH-responsive polymer (MSN-PVP) were loaded with R848. A defined amount of the drug was adsorbed at low pH values into the mesopores of MSN-PVP, followed by the subsequent closure of the polymer coat on the mesoporous nanoparticles at pH 7. Following internalization and shuttling into the endosome, the local acidic environment allows for re- opening of the mesopores and release of the cargo. The low molecular weight, synthetic imidazoquinoline compound R848 (also called resiquimod) immune responses upon uptake and ligation to endosomal TLR 7.23 Indeed, we found that both unbound (molecules suspended in liquid) as well as MSN-PVP-encapsulated R848 induced activation of dendritic cells with potent upregulation of CD80 and release of pro-inflammatory IL-6 (Fig. 6). These data show that MSN-PVP can be used as a drug delivery tool in primary immune cells. A therapeutic approach to use the targeted release of MSN cargo in the endosome of immune cells in order to target endosomal receptors such as TLR7 with stimuli- responsive capping mechanisms in cancer immunotherapy will be the subject of future studies. induces potent Figure 6: Mesoporous silica nanoparticles are an efficient delivery tool for the synthetic immunostimulatory TLR7 ligand R848. Complete splenocytes were incubated for 18 h either with the small molecule immunostimulant R848, cargo-free, or R848-loaded MSN-PVP, respectively. (A) The surface expression of the co-stimulatory molecule CD80 on dendritic cells was analyzed by flow cytometry. (B) The level of Interleukin-6 in the cell culture supernatant was determined by ELISA. Data give the mean values of triplicate samples ± s.e.m. An asterisk indicates comparison to unstimulated cells. All results are representative of two independent experiments. 15 DISCUSSION Primary immune response to MSN Despite intense research in the field of nanomedicine, fundamental knowledge about the interaction of nanomaterials with the cellular components of the host immune system remains scarce. We demonstrate in this work that MSN-NH2 particles are rapidly and efficiently taken up by specialized antigen-presenting cells (APCs) such as dendritic cells and macrophages and are delivered into endo-/lysosomes. Efficient uptake of MSN via endocytosis into non-immune cells, in particular different types of tumor cell lines, has been observed previously.24, 25, 26 For example, for HeLa cervical adenocarcinoma cells, Mou and co-workers reported a size-dependent endosomal uptake of MSN, favouring a size range of 50 - 120 nm.27 In contrast to tumor cells, APCs are specialized to scavenge their environment by taking up and sampling cell debris and foreign material, and are also equipped with a variety of germ-line encoded immune receptors to identify invading microorganisms. Many of these receptors, especially several members of the family of Toll- like receptors (TLR), are localized in the endosome and upon ligation of pathogenic material lead to innate immune stimulation. However, our results demonstrate that the uptake of cargo-free MSN-NH2 into APCs did not result in immune activation, as dendritic cells and macrophages showed only mild upregulation of the co-stimulatory molecule CD80 and released only low levels of pro-inflammatory cytokines. Similarly, Lee et al. showed that incubation of a macrophage cell line or peritoneal macrophages with MSN resulted in only trace release of cytokines and that short-term in vivo application of MSN in mice did not lead to contact hypersensitivity.28 However, treatment of mice with MSN over a time course of several weeks resulted in histological changes in liver and spleen.29 We note that such findings are expected to be very dependent on the size, surface functionalization and zeta potential of the particles and thus cannot be generalized for different types of MSN. Generally, the impact of repetitive treatments with MSN must be carefully evaluated before long-term clinical applications are conducted. Crystalline silica have been shown to potently activate the NLRP3 inflammasome, a cytosolic multiprotein complex, triggering the release of the bioactive form of the potent pro- inflammatory cytokine IL-1β.30 Similar to crystalline silica, non-functionalized, amorphous silica nanoparticles can activate the inflammasome, leading to significant IL-1β secretion.31 In contrast, here we demonstrate that the molecularly functionalized MSN-NH2 particles do not induce relevant levels of IL-1β release from primary murine splenocytes. Initial reports suggested that phagocytosis of crystalline silica with a median particle size of 5 µm results in presumably osmotic swelling and damage of the lysosome, leading to activation of the NALP3 inflammasome, which is triggered by lysosomal rupture and content release, not the 16 Primary immune response to MSN thus potentiating antigen-specific T-cell crystal structure itself.30 While others linked the formation of reactive oxygen species during an oxidative stress situation to NLRP3 activation,32 the exact molecular mechanism and prerequisite for inflammasome activation remains to be determined. We propose that due to their size and molecularly functionalized surface structure, spherical MSN-NH2 in comparison to silica crystals do not induce lysosomal damage and subsequent inflammasome-mediated IL-1β release. Generally, besides the physicochemical properties of particle size and surface chemistry, the biological activity of MSN has also been attributed to shape features such as aspect ratio and morphology.33 MSN have attracted much interest for their potential as drug delivery vehicles to control various cell functions by the stimuli-responsive delivery of bioactive cargos.34, 35, 36 An ideal drug delivery vehicle based on MSN may be composed of a multifunctional silica core able to specifically control the interaction with diverse active cargo components. The cargo molecules (e.g. pharmaceutically active drugs) are adsorbed in the mesopores of the nanoparticles, yielding an effective shielding from external degradation in biological fluids.37, 38 Such multi-functional MSN have been successfully evaluated as antigen carriers and adjuvants for vaccine delivery.39 Thereby, MSN have shown intrinsic adjuvant activity under responses.40,41 certain conditions, Interestingly, MSN have been described to enhance MHC class I-restricted presentation of antigens by human dendritic cells.42 This process called cross-presentation is a vital prerequisite for the induction of adaptive T-cell immunity against exogenous antigens such as tumor proteins. We found that the functionalized MSN-NH2 without protein or adjuvant cargo do not interact with or activate T- and B-lymphocytes in vitro. These findings underline the important role of dendritic cells at the interface of innate and adaptive immunity. Importantly, the lack of unspecific lymphocyte priming by the MSN carrier system is a promising requisite for future in vivo applications in order to use the high specificity of molecular immunostimulants either on the surface of the MSN or delivered as cargo from its pore system. Generally, the application of immunostimulatory adjuvants or vaccines via MSN harbors the risk of undesired systemic inflammatory responses upon the uptake of cargo-loaded MSN and subsequent cargo release. These dangerous adverse events can possibly be circumvented by context-dependent, spatiotemporally controlled cargo release. We and other groups pursue a promising approach that takes advantage of internal triggers such as an intracellular change in pH.14, 15 The efficient pH-responsive closing and opening mechanism of a reversible polymer cap system has been previously demonstrated by time-based fluorescence release experiments; fluorescent dyes were used in these studies.18, 43 In this immune 17 Primary immune response to MSN work, the synthetic TLR7 agonist R848 (resiquimod) was used as active cargo. R848-loaded MSN-PVP particles induced strong activation of dendritic cells with potent release of pro- inflammatory cytokines. As a defined ligand for endosomal TLR7/8, R848 was presumably released after pH-dependent reopening of the mesopores in the endosome. Whether such spatially controlled release can augment the efficacy and regulation of the subsequent immune response is the focus of ongoing research. Similarly, temporally defined release of the MSN-PVP cargo also appears attractive, as the kinetics of receptor sensitivity strongly influence the outcome of R848-based cancer immunotherapy.44 In summary, we demonstrate in this study that MSN-NH2 nanoparticles are non-toxic to primary murine leukocytes and provoke only trace immune activation. In addition, surface functionalized MSN-PVP can serve as a pH-triggered drug delivery tool for the synthetic TLR7/8 ligand R848 to induce potent immune activation in responder cells. The controlled release of their immunomodulatory cargo by otherwise non-immunogenic MSN is a promising tool in future therapies in order to achieve localized immune activation (e.g. in the tumor microenvironment) while preventing undesired, systemic adverse effects. 18 ACKNOWLEDGEMENT Primary immune response to MSN This study was supported by the Swiss National Science Foundation (projects 138284 and 310030-156372 to C.B. and National Centre of Competence in Research Bio-Inspired Materials), the Swiss Foundation for Cancer Research (grant KFS-2910-02-2012 to C.B.), the German Research Foundation Graduiertenkolleg 1202 (to C.B., S.E. and S.H.) and Else- Kröner Fresenius Stiftung (to S. H.). S.N., A.S., D.G., C.A. and T.B. thank the German Research Foundation (DFG, SFB 749 and SFB 1032), the Center for NanoScience (CeNS) and the Nano Initiative Munich (NIM) for financial support. S.N. received a Kekulé grant from the Verband der Chemischen Industrie. AUTHORSHIP CONTRIBUTION S.H., A.S., T.B. and C.B. designed the research, analyzed and interpreted the results and prepared the manuscript. S.N. and D.G. designed and synthesized the mesoporous silica nanoparticles. S.H. performed experiments with primary cells. S.E. and C.A. gave methodological support and conceptual advice. T.B. and C.B. guided the study. The authors declare no financial conflicts of interest. 19 REFERENCES Primary immune response to MSN 1. 2. 3. 4. 5. 6. 7. 8. 9. 11. 12. 13. Yohan D, Chithrani BD. Applications of nanoparticles in nanomedicine. Journal of biomedical nanotechnology 2014, 10(9): 2371-2392. Shah MA, He N, Li Z, Ali Z, Zhang L. Nanoparticles for DNA vaccine delivery. Journal of biomedical nanotechnology 2014, 10(9): 2332-2349. Smith MJ, Brown JM, Zamboni WC, Walker NJ. From immunotoxicity to nanotherapy: the effects of nanomaterials on the immune system. Toxicological sciences : an official journal of the Society of Toxicology 2014, 138(2): 249-255. Prashant CK, Kumar M, Dinda AK. Nanoparticle based tailoring of adjuvant function: the role in vaccine development. Journal of biomedical nanotechnology 2014, 10(9): 2317-2331. Iwasaki A, Medzhitov R. Control of adaptive immunity by the innate immune system. Nature immunology 2015, 16(4): 343-353. Goya GF, Marcos-Campos I, Fernandez-Pacheco R, Saez B, Godino J, Asin L, et al. Dendritic cell uptake of iron-based magnetic nanoparticles. Cell biology international 2008, 32(8): 1001-1005. Manolova V, Flace A, Bauer M, Schwarz K, Saudan P, Bachmann MF. Nanoparticles target distinct dendritic cell populations according to their size. European journal of immunology 2008, 38(5): 1404-1413. Cauda V, Schlossbauer A, Kecht J, Zurner A, Bein T. Multiple core-shell functionalized colloidal mesoporous silica nanoparticles. Journal of the American Chemical Society 2009, 131(32): 11361-11370. Kobler J, Moller K, Bein T. Colloidal suspensions of functionalized mesoporous silica nanoparticles. ACS nano 2008, 2(4): 791-799. 10. Mal NK, Fujiwara M, Tanaka Y, Taguchi T, Matsukata M. Photo-Switched Storage and Release of Guest Molecules in the Pore Void of Coumarin-Modified MCM-41. Chem Mater 2003, 15(17): 3385-3394. Schlossbauer A, Sauer AM, Cauda V, Schmidt A, Engelke H, Rothbauer U, et al. Cascaded photoinduced drug delivery to cells from multifunctional core-shell mesoporous silica. Adv Healthcare Mater 2012, 1(Copyright (C) 2012 American Chemical Society (ACS). All Rights Reserved.): 316-320. Chen KJ, Chaung EY, Wey SP, Lin KJ, Cheng F, Lin CC, et al. Hyperthermia- mediated local drug delivery by a bubble-generating liposomal system for tumor- specific chemotherapy. ACS nano 2014, 8(5): 5105-5115. Thomas CR, Ferris DP, Lee JH, Choi E, Cho MH, Kim ES, et al. Noninvasive remote- controlled release of drug molecules in vitro using magnetic actuation of mechanized nanoparticles. J Am Chem Soc 2010, 132(31): 10623-10625. 20 Primary immune response to MSN Lee CH, Cheng SH, Huang IP, Souris JS, Yang CS, Mou CY, et al. Intracellular pH- responsive mesoporous silica nanoparticles for the controlled release of anticancer chemotherapeutics. Angew Chem Int Ed 2010, 49(44): 8214-8219. Schlossbauer A, Dohmen C, Schaffert D, Wagner E, Bein T. pH-responsive release of acetal-linked melittin from SBA-15 mesoporous silica. Angew Chem Int Ed 2011, 50(30): 6828-6830. Lai C-Y, Trewyn BG, Jeftinija DM, Jeftinija K, Xu S, Jeftinija S, et al. A mesoporous silica nanosphere-based carrier system with chemically removable CdS nanoparticle caps for stimuli-responsive controlled release of neurotransmitters and drug molecules. J Am Chem Soc 2003, 125(Copyright (C) 2012 American Chemical Society (ACS). All Rights Reserved.): 4451-4459. Coll C, Mondragon L, Martinez-Manez R, Sancenon F, Marcos MD, Soto J, et al. Enzyme-mediated controlled release systems by anchoring peptide sequences on mesoporous silica supports. Angew Chem Int Ed 2011, 50(9): 2138-2140. Niedermayer S, Weiss V, Herrmann A, Schmidt A, Datz S, Muller K, et al. Multifunctional polymer-capped mesoporous silica nanoparticles for pH-responsive targeted drug delivery. Nanoscale 2015, 7(17): 7953-7964. Salmio H, Brühwiler D. Distribution of Amino Groups on a Mesoporous Silica Surface after Submonolayer Deposition of Aminopropylsilanes from an Anhydrous Liquid Phase. The Journal of Physical Chemistry C 2007, 111(2): 923-929. Koopman G, Reutelingsperger CP, Kuijten GA, Keehnen RM, Pals ST, van Oers MH. Annexin V for flow cytometric detection of phosphatidylserine expression on B cells undergoing apoptosis. Blood 1994, 84(5): 1415-1420. Hemmi H, Takeuchi O, Kawai T, Kaisho T, Sato S, Sanjo H, et al. A Toll-like receptor recognizes bacterial DNA. Nature 2000, 408(6813): 740-745. Dostert C, Petrilli V, Van Bruggen R, Steele C, Mossman BT, Tschopp J. Innate immune activation through Nalp3 inflammasome sensing of asbestos and silica. Science (New York, NY) 2008, 320(5876): 674-677. Hemmi H, Kaisho T, Takeuchi O, Sato S, Sanjo H, Hoshino K, et al. Small anti-viral compounds activate immune cells via the TLR7 MyD88-dependent signaling pathway. Nature immunology 2002, 3(2): 196-200. Tao Z, Toms BB, Goodisman J, Asefa T. Mesoporosity and Functional Group Dependent Endocytosis and Cytotoxicity of Silica Nanomaterials. Chem Res Toxicol 2009, 22(11): 1869-1880. Zhu J, Tang J, Zhao L, Zhou X, Wang Y, Yu C. Ultrasmall, Well-Dispersed, Hollow Siliceous Spheres with Enhanced Endocytosis Properties. Small 2010, 6(2): 276-282. Sauer AM, Schlossbauer A, Ruthardt N, Cauda V, Bein T, Bräuchle C. Role of endosomal escape for disulfide-based drug delivery from colloidal mesoporous silica evaluated by live-cell imaging. Nano Lett 2010, 10(9): 3684-3691. Lu F, Wu S-H, Hung Y, Mou C-Y. Size Effect on Cell Uptake in Well-Suspended, Uniform Mesoporous Silica Nanoparticles. Small 2009, 5(12): 1408-1413. 21 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. Primary immune response to MSN Lee S, Yun H-S, Kim S-H. The comparative effects of mesoporous silica nanoparticles and colloidal silica on inflammation and apoptosis. Biomaterials 2011, 32(35): 9434-9443. Lee S, Kim M-S, Lee D, Kwon TK, Khang D, Yun H-S, et al. The comparative immunotoxicity of mesoporous silica nanoparticles and colloidal silica nanoparticles in mice. International Journal of Nanomedicine 2013, 8: 147-158. Hornung V, Bauernfeind F, Halle A, Samstad EO, Kono H, Rock KL, et al. Silica crystals and aluminum salts activate the NALP3 inflammasome through phagosomal destabilization. Nature immunology 2008, 9(8): 847-856. 31. Winter M, Beer H-D, Hornung V, Krämer U, Schins RPF, Förster I. Activation of the inflammasome by amorphous silica and TiO2 nanoparticles in murine dendritic cells. Nanotoxicology 2011, 5(3): 326-340. Zhou R, Tardivel A, Thorens B, Choi I, Tschopp J. Thioredoxin-interacting protein links oxidative stress to inflammasome activation. Nature immunology 2010, 11(2): 136-140. Hao N, Li L, Tang F. Shape-mediated biological effects of mesoporous silica nanoparticles. Journal of biomedical nanotechnology 2014, 10(10): 2508-2538. Vivero-Escoto JL, Slowing II, Trewyn BG, Lin VSY. Mesoporous Silica Nanoparticles for Intracellular Controlled Drug Delivery. Small 2010, 6(18): 1952-1967. Rosenholm JM, Sahlgren C, Linden M. Towards multifunctional, targeted drug delivery systems using mesoporous silica nanoparticles - opportunities & challenges. Nanoscale 2010, 2(10): 1870-1883. Argyo C, Weiss V, Bräuchle C, Bein T. Multifunctional Mesoporous Silica Nanoparticles as a Universal Platform for Drug Delivery. Chem Mater 2013, 26(1): 435-451. Cauda V, Argyo C, Bein T. Impact of different PEGylation patterns on the long-term bio-stability of colloidal mesoporous silica nanoparticles. J Mater Chem 2010, 20(39): 8693-8699. 28. 29. 30. 32. 33. 34. 35. 36. 37. 38. 40. 42. Lin YS, Abadeer N, Haynes CL. Stability of small mesoporous silica nanoparticles in biological media. Chem Commun 2011, 47(1): 532-534. 39. Mody KT, Popat A, Mahony D, Cavallaro AS, Yu C, Mitter N. Mesoporous silica nanoparticles as antigen carriers and adjuvants for vaccine delivery. Nanoscale 2013, 5(12): 5167-5179. Kupferschmidt N, Qazi KR, Kemi C, Vallhov H, Garcia-Bennett AE, Gabrielsson S, et al. Mesoporous silica particles potentiate antigen-specific T-cell responses. Nanomedicine 2014, 9(12): 1835-1846. 41. Mahony D, Cavallaro AS, Stahr F, Mahony TJ, Qiao SZ, Mitter N. Mesoporous silica nanoparticles act as a self-adjuvant for ovalbumin model antigen in mice. Small 2013, 9(18): 3138-3146. Jimenez-Perianez A, Abos Gracia B, Lopez Relano J, Diez-Rivero CM, Reche PA, Martinez-Naves E, et al. Mesoporous silicon microparticles enhance MHC class I 22 Primary immune response to MSN cross-antigen presentation by human dendritic cells. Clinical & developmental immunology 2013, 2013: 362163. Schlossbauer A, Kecht J, Bein T. Biotin-avidin as a protease-responsive cap system for controlled guest release from colloidal mesoporous silica. Angewandte Chemie 2009, 48(17): 3092-3095. Bourquin C, Hotz C, Noerenberg D, Voelkl A, Heidegger S, Roetzer LC, et al. Systemic cancer therapy with a small molecule agonist of toll-like receptor 7 can be improved by circumventing TLR tolerance. Cancer research 2011, 71(15): 5123- 5133. 43. 44. 23
1702.05325
1
1702
2017-02-17T13:01:23
Fractional Cable Model for Signal Conduction in Spiny Neuronal Dendrites
[ "physics.bio-ph", "q-bio.NC" ]
The cable model is widely used in several fields of science to describe the propagation of signals. A relevant medical and biological example is the anomalous subdiffusion in spiny neuronal dendrites observed in several studies of the last decade. Anomalous subdiffusion can be modelled in several ways introducing some fractional component into the classical cable model. The Chauchy problem associated to these kind of models has been investigated by many authors, but up to our knowledge an explicit solution for the signalling problem has not yet been published. Here we propose how this solution can be derived applying the generalized convolution theorem (known as Efros theorem) for Laplace transforms. The fractional cable model considered in this paper is defined by replacing the first order time derivative with a fractional derivative of order $\alpha\in(0,1)$ of Caputo type. The signalling problem is solved for any input function applied to the accessible end of a semi-infinite cable, which satisfies the requirements of the Efros theorem. The solutions corresponding to the simple cases of impulsive and step inputs are explicitly calculated in integral form containing Wright functions. Thanks to the variability of the parameter $\alpha$, the corresponding solutions are expected to adapt to the qualitative behaviour of the membrane potential observed in experiments better than in the standard case $\alpha=1$.
physics.bio-ph
physics
Fractional Cable Model for Signal Conduction in Spiny Neuronal Dendrites Silvia Vitali email [email protected] Francesco Mainardi email [email protected] June 22, 2021 Abstract The cable model is widely used in several fields of science to describe the propagation of signals. A relevant medical and biological example is the anomalous subdiffusion in spiny neuronal dendrites observed in sev- eral studies of the last decade. Anomalous subdiffusion can be modelled in several ways introducing some fractional component into the classical cable model. The Chauchy problem associated to these kind of models has been investigated by many authors, but up to our knowledge an explicit solution for the signalling problem has not yet been published. Here we propose how this solution can be derived applying the generalized convo- lution theorem (known as Efros theorem) for Laplace transforms. The fractional cable model considered in this paper is defined by re- placing the first order time derivative with a fractional derivative of order α ∈ (0, 1) of Caputo type. The signalling problem is solved for any input function applied to the accessible end of a semi-infinite cable, which sat- isfies the requirements of the Efros theorem. The solutions corresponding to the simple cases of impulsive and step inputs are explicitly calculated in integral form containing Wright functions. Thanks to the variability of the parameter α, the corresponding solutions are expected to adapt to the qualitative behaviour of the membrane potential observed in experiments better than in the standard case α = 1. Keywords: Fractional cable equation, Sub-diffusion, Wright functions PACS: 87.19.L-, 05.40-a, 87.16.A- Introduction The one dimensional cable equation has been used from a long time to describe the spatial and the temporal dependence of trans-membrane potential Vm(x, t) along the axial x direction of a cylindrical nerve cell segment. It can be derived directly from the Nernst-Planck equation for electro-diffusive motion of ions. The resulting differential equation for the trans-membrane potential takes the form of a standard diffusion equation with a shift: λ2 ∂2Vm(x, t) ∂x2 − γ ∂Vm(x, t) ∂t 1 − Vm(x, t) = 0 , (1) where λ and γ are space and time constants related to the membrane resistance and capacitance per unit length, see e.g. [10]. For simplicity in the rest of this work, following [10], we will use the dimensionless scaled variables X = x/λ and T = t/γ, so that we consider the equation ∂2Vm(X, T ) ∂X 2 − ∂Vm(X, T ) ∂T − Vm(X, T ) = 0 . (2) In signalling problems the cable is considered of semi-infinite length (0 ≤ X < ∞, initially quiescent for T < 0 and excited for T ≥ 0 at the accessible end (X = 0) with a given input in membrane potential Vm(0, T ) = g(t). Fundamental problems are the cases of an impulsive input g(t) = δ(t) and of a unit step input g(t) = θ(t) where δ(t) and θ(t) denote the Dirac and the Heaviside functions, respectively. The solutions corresponding to these inputs read in our notation and X 2 e−( X 4T +T) , Gs(X, T ) = √4πT 3 Hs(X, T ) =Z T 0 Gs(X, T ′) dT ′ . (3) (4) We refer to Gs to as the fundamental solution or the Green function for the signalling problem of the (standard) cable equation (2), whereas to Hs to as the step response. As known, the Green function is used in the time convolution integral to represent the solution corresponding to any given input g(T ) as follows Vm(X, T ) =Z T 0 g(T − T ′)Gs(X, T ′) dT ′ . (5) The spatial variance associated to this model is known to evolve linearly in time, while it has been observed that the spatial variance of an inert tracer in spiny Purkenje cell dendrites evolves as a sub-linear power law of time, as spines may trap and release diffusing molecules, and the diffusion with smaller values of the power exponent is associated to higher spine density [20]. To model anomalous sub-diffusion we substitute the first-order time deriva- tive in Eq.(2) with a fractional time derivative of Caputo type [3], [17] of order α ∈ (0, 1): ∂2Vm(X, T ) ∂αVm(X, T ) ∂X 2 − ∂T α − Vm(X, T ) = 0 . (6) This kind of model is a simple extension to fractional behaviour of the Neuronal Cable Model from a mathematical point of view and it turns to be in some way equivalent to the equation developed in a relevant study [5], which has been de- rived from a modified Nernst-Planck equation, with diffusion constant replaced by fractional derivatives of Riemann-Liouville type. Other studies consider sim- ilar approaches [5],[7],[8],[9],[14]. We will see that beside the apparent simplicity our approach allows to reproduce at least qualitatively the main characteristics observed in experiments [15], [6],[2],[18]. Further generalizations of this model introducing a second fractional time derivative to the shift term could be anal- ysed in future, to refine the biological relevance of the model. 2 Solution of the signalling problem Applying the Laplace transform to Eq. (6) with the boundary conditions re- quired by the signalling problem, that is Vm(X, 0+) = 0, Vm(0, T ) = g(T ), we have: which is a second order equation in the variable X with solution: = 0, ∂X 2 ∂2fVm(X, s) (sα + 1)fVm(X, s) − fVm(X, s) =eg(s)e−√(sα+1)·X . (7) (8) The inversion of the Laplace transform now requires special effort with respect to the case where the term Vm(X, T ) is not present in Eq. (6), that is for ∂2V ∗m(X, T ) ∂X 2 − ∂αV ∗m(X, T ) ∂T α = 0 . (9) Indeed for Eq. (9), known as the time-fractional diffusion equation, the solutions of the corresponding Cauchy and signalling problems have been found in the 1990's by Mainardi in terms of 2 auxiliary Wright functions (of the second type) [11, 12]. Specifically for the signalling problem the general solution provided by Mainardi in integral convolution form reads V ∗m(X, T ) =Z T 0 g(T−T ′)G∗α,s(X, T ′) dT ′ , G∗α,s(X, T ) = 1 T W −α/2,0(cid:16)−X/T α/2(cid:17) , (10) where G∗α,s(X, T ) denotes the Green function of the signalling problem of the fractional time diffusion equation (9) and W −α/2,0(·) is a particular case of the transcendental function known as Wright function Wλ,µ(z) := zn n! Γ[λn + µ] ∞Xn=0 , λ > −1, µ ≥ 0 . (11) This function, entire in the complex plane, is extensively discussed in the Ap- pendix F of Mainardi's book [13] where the interested reader can find the fol- lowing relevant Laplace transform pairs, rigorously derived by Stankovi´c[19]: tµ−1 W −ν,µ (x/tν ) ÷ s−µ exp (−xsν) , 0 ≤ ν < 1 , µ > 0 . (12) Here we have adopted an obvious notation to denote the juxtaposition of a locally integrable function of time t with its Laplace transform in s with x a positive parameter. It is worth to recall the distinction of the Wright functions in first type (λ ≥ 0) and second type (−1 < λ ≤ 0) and, among the latter ones, the relevance of the two auxiliary functions introduced in [11]: Fν (z) = W −ν,0(−z) , Mν(z) = W −ν,1−ν (−z) , 0 < ν < 1 , (13) inter-related as Fν (z) = νzMν(z). Indeed the relevance of both the Wright func- tions has been outlined by several authors in diffusion and stochastic processes. Particular attention is due to the M -Wright function (also referred to as the Mainardi function in [17]) that, since for ν = 1/2 reduces to exp (−z2/4)/√π, 3 is considered a suitable generalization of the Gaussian density, see [16] and ref- erences therein. Then the Green function for the signalling problem of the time fractional diffusion equation (9) can be written in the original form provided in [11] as (14) G∗α,s(X, T ) = 1 T Fα/2(cid:16)X/T α/2(cid:17) = α 2 X T α/2+1 Mα/2(cid:16)X/T α/2(cid:17) , where the superscript ∗ is added to distinguish the time fractional diffusion equation from our fractional cable equation, both depending on the order α ∈ (0, 1). Because of the shift constant in the square root of the Laplace transform in Eq.(8), the inversion is no longer straightforward with the Wright functions as it is in the time fractional diffusion equation (9). Consequently, we have overcome this difficulty recurring to the application of the Efros theorem that generalizes the well known convolution theorem for Laplace transforms. For sake of convenience let us hereafter recall this theorem, usually not so well-known in the literature. The Efros theorem [4] states that if we can write a Laplace transform ef (s) where the function eF (s) has a known inverse Laplace transform F (T ), the in- verse Laplace transform is: (15) as: For the solution (8) of our signalling problem we thus have: 0 F (τ )G(τ, T )dτ ef (s) = φ(s) · eF (ψ(s)), f (T ) =Z ∞ G(τ ; T ) ÷ eG(τ, s) = φ(s)e−τ ψ(s) φ(s) =eg(s), ψ(s) = sα , eF (s)X = e−X√s+1 . where: and of Laplace transforms, we obtain: Then, having eG(τ, s) =eg(s) e−τ sα G(τ, T ) =Z T 0 where W (16) (17) (18) (19) (20) (21) , thanks to the standard convolution theorem g(T − T ′) T ′ −α,0(−τ /T ′α)dT ′ W −α,0 is the F-Wright function, and √4πT 3 F (T )X = X 2 4T +T ) e−( X is the solution (3) of the standard cable equation (2). Then, the general solution of our signalling problem can be written in terms of known functions: Vm(X, T ) 0 =Z ∞ =Z T 0 X e−( X √4πτ 3 g(T − T ′)(cid:20)Z ∞ 0 2 4τ +τ )"Z T 0 g(T − T′) T′ W e−( X 2 4τ +τ ) 1 T ′ X √4πτ 3 4 −α,0(−τ /T′α)dT′# dτ −α,0(−τ /T ′α) dτ(cid:21) dT ′ . (22) W Example 1: g(T ) = δ(T ) Substituting g(T ) = δ(T ) in the general solution (22) we obtain the Green function for the fractional model (6): Vm(X, T ) := Gα,s(X, T ) =Z ∞ =Z ∞ 0 Gs(X, τ ) 0 Gs(X, τ ) 1 T 1 T −α,0(−τ /T α)dτ W Fα( τ T α )dτ (23) This solution is plotted versus X for T = 0.1 and T = 100 and versus T for X = 1 in Fig.1. 2.5 2 0 V 1.5 / ) X ( 1 0.5 0 0 m V α=1. T=0.1 1 x 10−4 α=0.25 T=100 0 V / ) X ( m V α=0.5 α=0.75 α=1. 0.5 0 0 5 X 10 α=0.75 α=0.5 α=0.25 102 α=0.25 100 0 V α=0.5 / ) T ( V m 10−2 α=0.75 α=1. 1 2 X 3 4 5 10−4 10−4 10−2 T 100 102 Fig.1: Plots of Vm(X, T ) versus X in T = 0.1 and T = 100. in the inset (left) and plots of Vm(X, T ) versus T in X = 1. (right), for g(T ) = δ(T ). Example 2: g(T ) = θ(T ) When g(T ) = θ(T ) we obtain the step response of our fractional cable equation : Vm(X, T ) := Hα,s(X, T ) =Z ∞ 0 X √4πτ 3 e−( X 2 4τ +τ )"Z T 0 −α,0(−τ /T′α)dT′# dτ . (24) W 1 T′ After some manipulations including the change of variable z = τ /T ′α and inte- grating by parts after using the recurrence relation of Wright functions: dWλ,µ(z) dz = Wλ,λ+µ(z) (25) and the relation between the auxiliary functions: Fν (z) = νz Mν(z) we may rewrite the step-response solution as: Vm(X, T ) := Hα,s(X, T ) =Z ∞ 0 Hs(X, τ ) · 1 T α Mα( τ T α )dτ (26) This solution is plotted versus X for T = 1 and versus T for X = 1 in Fig.2. 5 0 V / ) X ( m V 100 10−1 10−2 10−3 10−4 100 α α=1. α=0.75 α=0.5 α=0.25 2 4 X 6 8 10 0 V / ) T ( m V 10−1 α=0.25 α=0.5 α=0.75 10−2 10−2 α=1. 10−1 100 T 101 102 Fig.2: Plots of Vm(X, T ) versus X in T = 1. (left) and plots of Vm(X, T ) versus T in X = 1. (right), for g(T ) = θ(T ). The plotted solutions were computed without any particular difficulty except the careful truncation of the series defining the auxiliary F and M functions in our routines. Biological interpretation Fractional cable models are used to describe subthreshold potentials, or passive potentials, associated to dendritic processes in neurons. The travelling potential is summed up at the soma and the cell produces an action potential when a threshold is exceeded. It has been observed in [20] that diffusion results more anomalous, i.e. the fractional exponent α decreases, with increasing spine density. Decreasing spine density is characteristic of aging [6],[2], pathologies as neurological disorders [15] and Down's sindrome [18]. It has been suggested that increasing spine density should serve to compensate time delay of postsynaptic potentials along dendrites and to reduce their long time temporal attenuation [5]. Looking at our plotted solutions for the fractional cable equation short and long space and time behaviour can be distinguished about the evolution of the sub-diffusion process. When an impulsive potential is applied at the accessible end it can be noted from Fig.1 that peak high decreases more rapidly with decreasing α at early times, viceversa is less suppressed at longer times, and the cross over time increases with decreasing α. Looking at the potential versus time it can also be noted that potential functions associated to lower α last for longer time at appreciable intensity and arrive faster at early times with respect to the normal diffusion case. By the way, when a constant potential is applied at the accessible end we note from Fig.2 that the exponential suppression of the potential along the dendrite is reduced for high X values with respect to normal diffusion. Instead for small X the potential results just slightly more suppressed in the sub-diffusion process. 6 Conclusions The presented fractional cable model satisfies the main biological features of the dendritic cell signalling problem. With respect to models solved as Cauchy prob- lem, our approach could include specific time dependent boundary conditions, which will allow to reconstruct with accuracy the expected signal at the soma if the model will result capable to predict real data behaviour. Furthermore the solutions can be computed directly, i.e. calculating the integral associated, as well as by Laplace transform inversion [1] without any remarkable issue. Aknowledgements The work of F. M. has been carried out in the framework of the activities of the National Group of Mathematical Physics (INdAM-GNFM). The authors are indebted to the Interdepartmental Center "Luigi Galvani" for integrated studies of Bioinformatics, Biophysics and Biocomplexity of the University of Bologna for partial support. References [1] J. Abate and W. Ward. A unified framework for numerically inverting laplace transforms. INFORMS J. on Computing, 18(4):408 -- 421, January 2006. [2] H. Duan. Age-related dendritic and spine changes in corticocortically pro- jecting neurons in Macaque monkeys. Cerebral Cortex, 13, 09 2003. [3] R. Gorenflo and F. Mainardi. Fractional Calculus: Integral and Differential Equations of Fractional Order, pages 223 -- 276. CISM Courses and Lecture Notes, Vol. 378. Springer-Verlag, Wien, 1997. [4] U. Graf. Applied Laplace Transforms and z-Transforms for Scientists and Engineers. Springer, 2004. [5] B.I. Henry, T.A.M. Langlands, and S.L. Wearne. Fractional cable models for spiny neuronal dendrities. Phys. Rev. Lett, 100:128103/1 -- 3, 2008. [6] Bob Jacobs, Lori Driscoll, and Matthew Schall. Life-span dendritic and spine changes in areas 10 and 18 of human cortex: A quantitative golgi study. The Journal of Comparative Neurology, 386, 10 1997. [7] T.A.M. Langlands, B I. Henry, and S.L. Wearne. Fractional cable equa- infinite domain tion models for anomalous electrodiffusion in nerve cells: solutions. Journal of Mathematical Biology, 59, 2009. [8] T.A.M. Langlands, B.I. Henry, and S.L. Wearne. Fractional cable equa- tion models for anomalous electrodiffusion in nerve cells: Finite domain solutions. SIAM Journal on Applied Mathematics, 71, 2011. [9] Fawang Liu, Qianqian Yang, and Ian Turner. Two new implicit numerical methods for the fractional cable equation. Journal of Computational and Nonlinear Dynamics, 6, 2011. 7 [10] R.L Magin. Fractional Calculus in Bioengineering. Begell House Publishers, 2006. [11] F. Mainardi. Fractional relaxation-oscillation and fractional diffusion-wave phenomena. Chaos, Solitons and Fractals, 7:1461 -- 1477, 1996. [12] F. Mainardi. Fractional Calculus: Some Basic problems in Continuum and Statistical Mechanics, pages 291 -- 348. CISM Courses and Lecture Notes, Vol. 378. Springer-Verlag, Wien, 1997. [13] F. Mainardi. Fractional Calculus and Waves in Linear Viscoelasticity. Im- perial College Press, London, 1st edition, 2010. [14] K. Moaddy, A. G. Radwan, K. N. Salama, S. Momani, and I. Hashim. The fractional-order modeling and synchronization of electrically coupled neuron systems. Comput. Math. Appl., 64(10):3329 -- 3339, November 2012. [15] Esther A. Nimchinsky, Bernardo L. Sabatini, and Karel Svoboda. Sructure and function of dendritic spines. Annual Review of Physiology, 64, 03 2002. [16] G. Pagnini. The M-Wright function as a generalization of the Gaussian density for fractional diffusion processes. Fract. Calc. Appl. Anal, 16, 2013. [17] I. Podlubny. Fractional Differential Equations. Mathematics in Science and Engineering 198. Academic Press, San Diego, 1st edition, 1999. [18] M. Suetsugu and P. Mehraein. Spine distribution along the apical dendrites of the pyramidal neurons in Down's syndrome. Acta Neuropathologica, 50, 1980. [19] B. Stankovi`c. function of E.M. Wright. l'InstitutMath`ematique, Beograd, Nouvelle S`er., 10:113 -- 124. On the Publ. de [20] F. Santamaria, S. Wils, E. De Schutter, and George J. Augustine. Anoma- lous diffusion in purkinje cell dendrites caused by spines. Neuron, 52, 2006. 8
1911.02529
1
1911
2019-11-03T04:41:43
Quantum Biology: Can we explain olfaction using quantum phenomenon?
[ "physics.bio-ph", "quant-ph" ]
The sense of smell is an important part of living organisms. It assists with the interaction of the living organism with its environment. The mechanism by which smell is detected and identified is not fully known. Earlier, shape theory was proposed as an explanation for odor perception mechanism. This theory posits that the shape of the odorant must fit in its complementary olfactory receptor for an odor to be identified -- just like key fits in a lock. This theory turns out to have some limitations, thus leading to the proposition of new theory called the vibrational theory of olfaction. In vibrational theory, the nose is regarded as a spectroscope that detects vibration of odorants. However, this happens to be currently controversial and actively debated. In this paper, I will give a review of the history of olfaction, the challenges, and then explain some new experiments (both on human and fruit flies) that support or refute the vibrational theory of olfaction.
physics.bio-ph
physics
Can we explain olfaction using quantum phenomenon? Quantum Biology: Chukwuemeka Asogwa∗ Department of Physics and Astronomy, University of Waterloo, Waterloo, Ontario, N2L 3G1, Canada (Dated: November 7, 2019) The sense of smell is an important part of living organisms. It assists with the interaction of the living organism with its environment. The mechanism by which smell is detected and identified is not fully known. Earlier, shape theory was proposed as an explanation for odor perception mechanism. This theory posits that the shape of the odorant must fit in its complementary olfactory receptor for an odor to be identified -- just like key fits in a lock. This theory turns out to have some limitations, thus leading to the proposition of new theory called the vibrational theory of olfaction. In vibrational theory, the nose is regarded as a spectroscope that detects vibration of odorants. However, this happens to be currently controversial and actively debated. In this paper, I will give a review of the history of olfaction, the challenges, and then explain some new experiments (both on human and fruit flies) that support or refute the vibrational theory of olfaction. I. INTRODUCTION A. Sensing of odor by living organisms To understand how life began was one of the earlier challenges of scientists and philosophers. As back as the nineteenth century, a good number of scientists accepted life as a kind of magic matter which has no chemical nor physical explanation. This ideology lasted until an assumption that life has a chemical recipe. Life was then known to be a combination of two or more chemicals. This led to the suggestion of 'producing life' by mixing up chemicals in the lab. This idea was a blind alley as life has not been produced using inanimate objects till date. How scientists understand the beginning of life was rev- olutionized in 1944 by Erwin Schrodinger in his published lectures; titled What is life? [1]. This famous publication gave rise to the study of molecular biology. Schrodinger was of the opinion that there is a quantum mechanical process involved in the stable transmission of genetic in- formation from one generation gene to another genera- tion and that this occurs unaware of the gene encoding roles. This was backed up by other founders of quantum mechanics, Neil Bohr, Werner Heisenberg and Eugene [2]. As science develops, quantum mechanics explains the atomic and molecular structure component of a mat- ter and the states it can appear to be. Life for a physics at this time was considered weird, therefore it could only be explained by a weird phenomenon like quantum me- chanics [3]. The study of molecular biology was actively growing, arguing that the shapes of molecules and their chemical affinities compost the functionality of the cell. These properties are phenomena that quantum mechan- ics can explain. There are several proposals that the fundamental level of all things is quantum mechanical. To understand life ∗ [email protected] on a higher level, some quantum concepts such as en- tanglement, superposition states, spooky action at a dis- tance and quantum tunneling need to be both experimen- tally and theoretically proven to be connected with the processes involved in the definition of life. How energy is captured through photosynthesis and the olfaction sci- ence are suggested to be connected with some of these quantum concepts [4, 5]. Among these two processes, we consider olfaction in detail. Olfaction is the ability to detect and distinguish dif- ferent odors. These odors come as airborne molecules. Different living things adapt peculiar method in detect- ing and distinguishing odors. Mammals and insects, in general, can use the intensity, whether it is offensive or fragrant to remark an odor and able to identify it next time. Humans find it difficult to effectively identify odor with little or no recognizable intensity, fragrance or offen- siveness. To detect where the smell is emerging from or even notice the smell is also difficult. Some animals like dogs, anemonefish, reef fish and sharks have good ability to detect the smallest strength of odor beyond human's capability. Example: • Fish have a good sense of smell that helps them to locate their home even when ocean current drift them away from their home. • Shark, in the same way, has the ability to detect the smell of a drop of blood as far as a kilometer away. • Unlike human nostrils that are used mainly for breathing, sharks' nostrils are located on the un- derside of the snout and used mainly for smelling. As they swim in the ocean, water flows through their nostrils and down to the sensory cells. They have high sensitive sensory cells. The side of the nostrils the smell hits first tells the shark where the smell is coming from, with this, they hunt for the wounded prey. In 2007, Gabriele Gerlach [6] designed an experiment to verify the theory behind larval reef fish identifying their home. She called her experimental setup a 'two- channel olfactory choice flumes test'. She collected two flumes of seawater; one from reef they were hatched and the other from a far away reef. She then placed the larval reef fish downstream of the two flumes. She observed the preference of the larval reef fish between the two flumes of seawater. It came out to be that the larval reef fish preferred the seawater which they were hatched. The ability of the larval reef fish to distinguish between the two seawater was ascribed to their smelling ability. Similarly, Daniella Dixson [7] confirmed that an anemonefish was able to distinguish between the collected water from their habitat reef and the water collected from an offshore. Michael Arvedlund [8, 9] designed a simi- lar experiment to Gabriele Gerlach's, to understand the ability of anemonefish in identifying their species. It was confirmed that anemonefish was able to distinguish their host species from others uncolonized by them. These ex- periments, together with earlier experiments and theories confirmed that fish follow a scent trail in identifying their home and species. Atmospheric molecules undergo vibrational interac- tion, turbulence and migration from one point to the other just like molecules in water. Molecules in air di- lute and disperse faster than the molecules in water [3]. Terrestrial creatures possess their own feature of olfac- tion. Example: • Rabbits have 100 million scent cells which help them in identifying other rabbits and animals. This sense of smell is inherent from birth, guiding the bunnies to find their mothers' teat even when their eyes are closed. • Bears have been identified to possess the best sense of smell. It can smell 7 times better than the blood- hound, and 2,100 times better than humans [3]. This helps them to identify food as far as 20 miles away. In addition to finding food, they can detect mates, avoid danger coming from competing fellow bears and track the whereabouts of their cubs when lost while journeying. • Elephants, in the same way, can smell water, es- pecially the underground water 12 miles away and still remember where they found it for future use. • Snakes have a different method of identifying a smell. They taste the air with their tongues gath- ering the particles using their damp surface, then transport them to Jacobson organ in the mouth where they will be identified as food or danger. • Dogs, especially the bloodhound can identify and track an individual by following the smell of the butyric. Butyric acid is different for every animal and human, even for identical twins. • Just like in animals discussed above, smell helps human in sniffing out bad odors and their location, food using its aroma, and in communication. 2 FIG. 1. A structure of the human olfactory system. As odor is perceived, odorant moves through the nose which acts as a channel, to meet the olfactory epithelium situated at the top of the nasal cavity. From the epithelium, it journeys to the cilia which extend from the main body of the neuron in the mucus layer. The odorant meets the mucus and the inside of the cell. Figure from [10]. This indicates that the nostrils have the capability of identifying and distinguishing several millions of odors. II. ANATOMY OF THE HUMAN NOSTRILS The nose is one of the chemosensory systems which is an intrinsic feature every human possesses. Chemosen- sory systems (of which taste and the vomeronasal sys- tems are inclusive) work hand-in-hand to identify a smell. They all pick up information in the form of molecules. Molecules can enter the olfactory system either through the nostrils normally or the back of the throat, which happens mainly when eating food. The nostrils act as a chamber that allows air pass into the nasal cavity. The nasal cavity then assists in filtering and warming the air with the mucus lining in its inner part. The odorants are then identified inside the nasal cavity as they hit the olfactory epithelium which is situated on the roof of our nasal cavity (FIG. 1). A scent signal is then sent to the brain through the olfactory nerve and olfactory bulb and an interpretation is made. The olfactory epithelium har- bors millions of olfactory receptors which bind up with a particular odorant. This helps in identifying several varieties of odors. Odorants are considered to have a specific shape that binds to the olfactory receptors. The receptors are ac- tivated by different molecules of the odor. There are always variations in the strength of binding of molecules to its receptor, which can affect the ability of the brain to fully interpret a smell. Our ability to detect various smells lies in the complexity of the interaction between the receptors and the odorants. This, in the actual sense, shows that the final smell we perceive is an integration of 3 FIG. 2. The red circles represent 'combinatorial code'. Large circles represent large responses and small circles, small re- sponses. The addition of specific individual 'codes' leads to the odorant detection. Figure from [11]. various odorants interacting with various receptors and generating encoded information [11]. In human olfactory, there are about 390 types of func- tional receptors which somehow 'tuned' in response to the different molecular stimuli given by the odorants in order to accommodate potentially thousands of odorants [12]. This is assumed to occur in a combinatorial process, where all the different receptors have a 'code' that deter- mines the characteristics of a smell [13]. One odorant activates more than one receptor which the combination of their responses give the smell of the odorant (FIG. 2). An odorant existing as a mixture of many odorants actives very many receptors. The process that we are most interested in is the func- tionality of capturing of the odorant by the olfactory neu- ron. Does this process involve quantum phenomena? III. THEORIES OF OLFACTION A. The 'lock and key' model An intriguing question one would ask is; "How does each receptor recognize its own set of odorant and binds with them?". In 1963 Amoore [14] first proposed that the response to scent works by the mechanism of a 'lock and key'. The lock and key model explains that the shapes of receptor molecules and the odorants are in complemen- tarity with each other, implying that the odorants fix FIG. 3. Top: The structure of benzaldehyde and hydrogen cyanide. They both have the same smell (bitter almond) but have different structures [19]. Bottom: Models of space-filling ferrocene and nickelocene. They have almost identical shapes and very different smells. Ferrocene and nickelocene have a spicy and oily/chemical smell respectively. Figure from [4]. into an olfactory receptor which has its shape; just like a key fits into its padlock. This idea was generated from the molecular mechanism of the behavior of enzyme [15]. Another illustration of this theory is the shape fitting games that toddlers play with excitement. This involves fitting a cut-out shape into its complementary opening in a wooden or plastic board. The odorants can be imag- ined as the shapes trying to fit in the opening on the board, which is the olfactory molecule. Since odorants come in different shapes, a reason- able assumption is made that molecules having the same shape should smell alike, and odorants having different shapes should have obvious distinct smells. The study of the structures of different molecules has shown that this assumption does not work [4]. Notwith- standing this powerful explanation of binding of odor structurally, it has been shown that there exist molecules with different shapes yet smell alike, and compounds with similar structures yet smell differently [16 -- 18]. Odorants like ferrocene and nickelocene have similar structures but different odors. Nickelocene has a cycloalkene odor, while ferrocene smells camphoraceous. Also, hydrogen cyanide and benzaldehyde have different structures but the same odor (bitter almond). FIG. 3 shows the space-filling model of the odorants. Other examples are illustrated in FIG. 4. Compounds (a)− (d) have different structures but smell the same (musk), while (e) and (f ) have a simi- lar shape and still possess distinct smells. Compound (e) is identified to be odorless while (f ) smells like a urine. These pose a problem in 'key and lock' theory. 4 FIG. 4. Odorants (a) to (d) have different structures, but the same odor (musk), while odorants (e) and (f ) have the same structure but different odors. Figure from [3]. FIG. 5. Both limonene and its mirror-image dipentene have different smell while they are observed to have the same spec- trum. Figure from [3]. B. Vibrational theory In 1938, Malcolm Dyson proposed that the nose de- tects the vibrational frequency of an odorant rather than its shape [5]. This revolutionized the olfactory science from the traditionally accepted 'key and lock' to the vi- brational theory of the molecule. Dyson observed that chemicals that have the same smell are usually made up of compounds having the same chemical groups (eg., C=O). These chemical groups define to large extent the properties these molecules possess; smell being part of the properties. Furthermore, there exist some other chemi- cal groups (such as thiol (SH) group) that determine the smell of an odorant irrespective of their shape. Dyson no- ticed this in some chemical groups with sulphydryl (SH), having a hydrogen atom attached to a sulfur atom. This smells like rotten-egg. He measured the vibrational fre- quencies of the compounds following the Raman principle of light scattering [20]. Light bounces off an atom when hit on surface elas- tically just like a ball bounces off a hard surface when hit against a hard surface. This occurs without any loss of energy. This means the energy is conserved. Chan- drasekhara Venkata Raman in 1930 earned a Noble prize for his work in the field of light scattering. He no- ticed that the wavelength of some of a deflected light changes as it transverses a transparent material. This phenomenon was named after him as Raman scattering [20]. This, in principle, means that light can also scatter inelastically, thereby losing some energy to the molecular bonds they hit. As the light hits the molecule, vibration occurs and the observed scattered light surfaces with less energy. This decrease in energy causes a decrease in the frequency of the photon. The amount of the lost en- ergy gives the Raman spectrum [5]. This spectrum has a specific feature called the 'signature', attributed to the particular chemical bonds. This underpins Dyson's follow-up theory that the nose could actually be seen as the 'spectroscopy' that detects the vibrational signature frequencies of different chemical bonds. He noticed that there is a strong correlation of some frequencies in the Raman spectra with a particular odor [for more details see 5]. All compounds having a terminal sulfur-hydrogen bond were identified to have a Raman frequency within the same range. Applying Dyson's theory to the explanation of olfac- tion was a difficult one. One is the methodology by which our nose acts like spectroscopy to collect the smell in form of a scattered light. Another is the involvement of light in the process. The theory became futile when it was observed that chiral molecules having the same chemical structure and identical Raman spectra could be easily distinguished by our nose. This shows that molecule can have different smell even when they have the same chem- ical structure and identical Raman spectra. A general example is the limonene, regarded to be right-handed molecules and its left-handed (mirror-image) molecule as dipentene (FIG. 5). They both have the same molecu- lar bond, thus the same Raman spectrum but different odors. Also, carvone (one of the chemical components of seeds) and caraway (belonging to the family of flowering plant called Apiaceae) have the same Raman spectrum but different odors. C. 'Weak shape' or odotope model The reason why molecules with different shapes are identified by olfactory receptors as the same odor led to the proposal of the 'weak shape' or odotope theory pro- posed in 1994 by Kensaku Mori and Gordon Shepherd 5 FIG. 6. Vanillin and isovanillin have the same chemical parts but can have different odors. Figure from [3]. [21]. This theory argues that there must be molecular shape recognition somewhat in odor detection. This pro- poses that the binding of the molecule to the olfactory receptor is based on the shape of the substructure (ie., the component chemical group) rather than the entire shape of the molecule. This theory fails for molecules with the same component of a chemical group but arranged in a different pattern. An example is vanillin and isovanillin (FIG. 6) having the same chemical parts but arranged differently, thus having different odors. The vibrational theory is also not yet able to solve this problem. D. Turin's theory of inelastic electron tunneling spectroscopy (IETS) Based on Dyson theory of vibrational theory, Turin proposed that the principle which the olfactory receptors use to detect vibrations of chemical bonds could be re- lated to the concept of quantum tunneling of electrons [4]. Quantum tunneling involves the seeping through of electrons or photons as they encounter a barrier which they do not have the energy to classically surmount. This concept of quantum tunneling of an electron was earlier developed by Henri Becquerel in 1896 while researching on radioactivity. This principle involves the quantization of the energy levels within a molecule with an arbitrary energy gap E = ω0, where ω is the resonant frequency of a particular vibrational mode and  is the reduced Planck constant. FIG. 7 shows an illustration of the energy process involve in inelastic tunneling of an electron. There are two metal plates closely placed to each other and separated by a miniature gap (the tunneling layer). Metal 1 is negatively charged by accommodating electrons, while metal 2 is positively charged (electron acceptor). The two Fermi levels of the metals are separated by eV , where V is the applied voltage. The Fermi energy in metal 1 E(1) is F FIG. 7. A demonstration of electron tunneling through a metal junction. The electron can tunnel through the tunnel- ing layer (insulating barrier) either (a) elastically (ordinary tunneling) or (b) inelastically. The electron only tunnels in- elastically if there is a molecule present in the tunneling layer with a vibrational mode ω0, which the tunneling electron puts into excitation. Energy loss occurs for the electron to tunnel to the other metal if only eV ≥ eV0 = ω0, where V0 is the threshold voltage. Figure from [22]. F − F , and can be related by E(1) higher than in metal 2 E(2) E(2) F = eV . Once there is an application of V in the junction, electrons present in the metal 1 can classically transit horizontally to the empty state in metal 2 without losing or gaining any energy. This process of tunneling across the barrier without loss of energy is called elastic tunneling. In the actual sense, electrons can only tunnel from the metal plate to the other if there is an empty spot available at the acceptor side of the metal plate with the same energy level. The electron will happen to lose its energy if it has to enter a slot with different energy level. This process then results in an inelastic tunneling. When a molecule is present in the tunneling layer between the two metals, as the electron tunnels across the molecule, its energy is absorbed by the molecule, and an excitation of a phonon in the molecule occurs. This creates bias voltages that couple to give the vibrational spectrum of the molecule between the two metals. The current-voltage characteristic (I/V ) of the elas- tic tunneling gives a linear relationship. For inelastic tunneling, the process of the tunneling of the electron inelastically results to an increase in conductance at the voltage eV0 (cid:39) ω0 of which a second derivative of the current-voltage characteristics (d2I/dV 2) shows clearly a defined peak in the spectra which appear like Dirac delta distribution. This tells the energies of the vibra- tional transitions of the molecule in the tunneling layer. A probe of the energy difference between the electron at the donor and the acceptor plates gives some insight on the molecular bonds of the molecule between the two metal plates. Following this explanation, Turin proposed that the ol- factory system works with the same principle, with the olfactory receptor serving as the IETS plates, and the electron that is at the acceptor site results in the produc- tion of the G-protein molecular torpedo, which triggers the olfactory neuron to send a signal to the brain for in- terpretation. The explanation of the biological IETS is given below. 1. Biological IETS The production of electron or hole by a source that enables the flow of charge requires some energy input. It is assumed that since human cells have some volt- age of order 0.5V, then the source of the voltage is not considered a problem [23]. In the biological system, for Turin's inelastic electron tunneling to occur, we require a source of the electron, removal mechanism from the elec- tron source, the right energy levels, and a possible donor and acceptor. It is accepted that the electron transfer in biology is aided by series of oxidation and reduction reactions within several biomolecules [24]. The trans- fer uses, to a very large degree, metalloproteins [25], so the theory of IETS can be applied in the biological sys- tem. Turin [4] suggested that the source of the electron could be the nicotinamide adenine dinucleotide phos- phates (NADPH). NADPH binds to the electron donor at one side of the gap through several amino-acid mo- tifs which build up the tertiary structure of a protein [26], and at the other side of the gap, an acceptor hav- ing zinc-allegedly coming from the anosmic attributed to zinc deficiency in the diet, replenished through supple- mentary and dietaries [27]. The zinc anchors the olfac- tory G-protein as shown in Turin's proposal in FIG. 8. The G-protein is release as a result of the oxidation of disulphide bridge as the electron tunnels to the acceptor. This suggests the pathway for olfactory signal transduc- tion in which odors are bind to specific receptors and activate specific G-proteins, and then transduce an in- tracellular signal causing the activation of second mes- senger systems for possible interpretation by the brain [13, 24]. The vibrational frequency of odorant could be measured via regarding the receptor protein to act as a spectrometer designed to detect a particular quantized vibration that relates to the difference in energy. Turin posits that there are about 10 receptors which get tuned during this process of electron transfer, and the signal produced is a combination of overlapping messages from different cells involved in the process. Olfactory cells have been found [28, 29] (using the whole-cell patch Clamp technique [30]) to have a resting potential in the range of (−50 to −65mV) ± 12mV and with average mem- brane capacitance of 3.9 ± 1pF. A biological IETS does not involve a scan over a range of frequencies like the conventional IETS does, rather it involves a build-up of spectrum piecewise by a series of receptors tuned to a range of different frequencies [4]. The frequency range is limited only by the emf from the NAPDH at the re- 6 FIG. 8. Turin's model of IETS. An electron from NADPH is accepted at the receptor protein. The electron will not be able to tunnel when the receptor binding site is empty (as shown on top) because there is no empty level available with the right energy. This means when no odor is present to normalize the energy difference at the receptor and acceptor, no tunneling happens. On the presence of an odorant (represented by the dipole) at the binding site (as shown at the bottom), the electron tunnels losing energy during the tunneling through excitation of the odorant's vibrational mode matching the energy gap between the donor and acceptor. Electrons flow through the protein and reduce the disulphide bridge via a zinc ion thereby a release of the G-protein is triggered. Figure from [4]. ceptor site. Most observed molecular bonds in odorants have been calculated [4] to have vibrational frequencies ranging from 0 to 4, 000cm−1 and also a corresponding energy range. Obtaining the actual measurement of the energy transfer associated with every receptor would help in the classification and grouping of odorants according to the observed energy. This implies that several re- ceptors would be involved in order to cover the vibra- tional spectrum and each can be tuned to a particular range of the vibrational spectrum. The acceptor and the donor energy levels are prone to thermal broadening of the range 2kT(∼ 400cm−1) implying that the spec- trometer of the nose can have poor resolution [18]. Thus about 10 receptor can cover receptor types within the range of 0 − 400cm−1. This arrangement is synonymous to that found in other spectral senses like vision and hear- ing where the broadly tuned receptor groups lie within the complete spectrum. However, electron transfer does not happen at all if the donor is not filled with the specific energy level and the acceptor having empty energy levels just like in ordinary IETS. 7 FIG. 9. Top: The structure of ambergris odorants having similar odors but different structures. The similar odors were identified to have obvious similar CHYPRE spectra despite having different structures. Figure from [4]. 2. IETS spectra calculation If electron tunneling is the mechanism of odor detec- tion, then it implies that there should be a correlation between the tunneling spectra of the molecules mea- sured by the biological detector and their odors. The comparison of spectra from different odorants is based on 1) the frequency of a given vibrational mode, 2) its intensity recorded by the sensor and 3) the resolution of the detector system. The theory and calculation for the measurement of IETS spectra of different compounds in metal-insulator-metal junctions have been developed. The application of the theory to the biological system which involves odorant is difficult because the odorants are prone to evaporating during the vacuum deposition processes involved in the manufacturing of the tunnel- ing junctions. Turin [4] developed an algorithm called CHYPRE (CHaracter PREdiction) which calculates the biological IETS spectra. Apparently, the difference be- tween the calculated value from IR and the CHYPRE is that the region of fingerprint below 1500cm−1 that is gen- erally used as a guild to determining molecular structure is empathized more by CHYPRE than by IR. CHYPRE algorithm was then used to examine the postulate that smell of compounds can be predicted from their vibra- tional spectra notwithstanding the characteristic of their structures. As an example, a comparison of the spectra of the chemically related ambergris odorants having dif- ferent structures and similar odor was done. The amber- gris odorant used were cedramber, karanal, jeger's ketal and timberol (FIG. 9). The observed convolved vibra- tional spectra of the ambergris odorants were seen to be very similar, in consistency with their similar odor. Con- versely, molecules having very similar structures and dif- ferent odors were observed to have different spectra. As an example, the spectra of three undecanones that differ FIG. 10. Top: The structure of three undecanones differing in the position of the carbonyl group. As a result of the shift in the position of the carbonyl from 4 to 6, the peaks of the intensities (indicated by arrow) are affected. The spectra of 2- and 6-undecanone are clearly different while 4-undecanone has an intermediate spectrum between the two, except within the region of 5 − 600cm−1 wavenumber where it has a close match with that of 2-undecanone. Figure from [4]. in the position of the carbonyl group were calculated. 2- Undecanone has the odor of ruewort, 6-undecanone has a fruity smell, while 4-undecanone has a smell that is intermediate between the two undecanones (FIG. 10). Notwithstanding that the structures of the odorants are closely similar, 2- and 6-undecanone were identified to possess obvious different spectra. These two examples suggest that the olfactory system uses vibrational transduction mechanism in discriminat- ing molecules that are closely related. E. Swipe-card model Swipe-card model is a hybrid of the theory of 'lock and key' and the vibrational theory. This model was first proposed by Marshall Stoneham and published as Brookes et al. [31]. Stoneham and colleagues argued that the shape of the olfactory receptor and the bond vibrations of the odorant give rise to what we identify as a smell [31]. The binding part of the olfactory receptor is said to function just like the swipe-card machine. A swipe-card has a magnetic strip with encoded infor- mation on it. It can only work in a swipe-card machine if the swipe-card has the same shape, thickness, placed in the right position and has the information already in the system. For chiral molecules, the odorant with right or left chiral fits into the right or left olfactory receptor and can only be identified if the olfactory receptor is able to recognize it. This, in general, means that the odor- ant first fits into its complementary receptor shape, then triggers the vibration-induced electron tunneling. The right-handed receptor detects the right-handed molecule and gives its smell which differs from corresponding left- handed molecule detected by the left-handed receptor. The point still unclear is how the shape of the olfac- tory receptor looks, how both the acceptor and donor molecules are positioned, and the binding of the right- handed and left-handed molecules to either same or dif- ferent receptor. IV. ELECTRON TUNNELING RATE Electron transfer in biological system involves oxida- tion (reduction) of certain species (X) in the cell fluid [24], though unclear on the particular biological origin [31]. The time interval τX involved in the diffusion of the electron through an aqueous medium can be estimated using the standard method for computing diffusion of ma- terial through a solution from the diffusion equation and the Stokes-Einstein relation for the diffusion coefficient [32]. This gives an estimate of diffusion time as τX = 3η 2nX kβT , (1) where η is the viscosity of water (0.89× 10−3kgm −1s−1), nX is the concentration of X, kβ is the Boltzmann con- stant and T is the temperature (in Kelvin). The nature of X or the receptor does not contribute to the result of the estimate. It has been estimated that nX will lie within 1µM to 100µM, thus a substitution of typical bi- ological system in the equation results to a value of ηX within the range 0.01 to 1ms [31]. The crossing process of the electron through the odorant to the odorant receptor can be described using Marcus theory [33], and the rate in which the electron tunnels can be approximated using Fermi's Golden Rule:  (cid:68) 2π (cid:69)2ρ, ψf Hψi 1 τi−f = (2) where ψi represents the eigenfunction of the initial eigen- state, ψf is the eigenfunction of the final state and the density of the final state or the Franck-Condon (FC) fac- tor is ρ. H is the Hamiltonian tunneling matrix. Eyring-Polanyi equation [34] is applied to describe the effect of temperature on the rate of chemical reaction. The transitional rate constant of an electron from donor to acceptor will be given as [35] (cid:19) (cid:18) − ∆G‡ kβT k = κB , (3) where the electron collision frequency B depends on the phase of the reacting molecule. For bimolecular reac- tion, B becomes the liquid phase collision frequency, 8 FIG. 11. A configuration diagram showing the total potential energy as a function of reaction coordinate Q. The activation energy barrier E‡, which separates the reactant and the prod- uct states (respectively donor D, and acceptor A) is modeled as an adiabatic reaction with κ = 1. The energy difference between D and A at coordinates XD and XA is represented as ∆E. Xc represents the crossing point for the transiting particle at the asterisk. Figure from [35]. B ∼ 1011M−1s−1, but for monomolecular reaction (in- tra molecular reaction), B becomes a vibrational fre- quency 1/(β), B ∼ 1013s−1 [36]. ∆G‡ is the Gibbs energy of activation, kβ is the Boltzmann's constant and T is the absolute temperature. κ is associated to the electron-transfer matrix element and gives the probabilis- tic value of the electron-transfer in the reaction. It deter- mines whether the reaction is adiabatic or non-adiabatic. For κ = 1, the adiabatic reaction dominates, while for κ < 1, the non-adiabatic regime occurs. At the clas- sical regime, κ → HDA2. Using the Gibbs energy, ∆G◦ = ∆H − T ∆S◦, equation 3 can be written as (cid:18) ∆S‡ (cid:19) kβ (cid:19) (cid:18) − ∆H‡ kβT k = κB exp exp , (4) where ∆S‡ is the entropy of activation and ∆H‡ is the enthalpy of activation. Comparison of equation 4 with the Arrheminus equation given as [37] k = κ∞ exp(cid:0)−∆E‡/kβT(cid:1) (cid:1). (5) where ∆E‡ represents the energy of activation, implies that ∆E‡ corresponds to ∆H‡ and k∞ corresponds to κB exp(cid:0)∆S‡/kβ The configuration coordinate diagram (FIG. 11) shows the terms. Considering a parabolic geometry, E‡ can be expressed as [35] E‡ = (λ + ∆E)2 4λ , (6) where λ is the reorganization energy (also called relax- ation energy) coupled to the electron transfer and ∆E is the energy difference between the donor and the accep- tor. λ is a combination of the reorganization energies due to the inner shell atoms λi and the surrounding solvent molecules λo. Combining the parameters from the inner shell vibrational modes, λi can be calculated as kijQ2 j , (for normal modes). (7) (cid:88) j λi = 1 2 kij is the Hook's law force constant and Qj represents all the harmonic oscillators arising from the displacement of the equilibrium position of the vibrational coordinates due to the electron transfer. The reorganization energy from the surrounding sol- vent molecules λo is calculated by assuming the solvent as a continuous polar medium. This gives λo as [36] (cid:19)(cid:18) 1 (cid:19) (cid:18) 1 λo = (∆e) 4π0 + 1 2r2 − 1 r12 2r1 − 1 Ds Dop , (8) where ∆e is the electron transfered from D to A, r1 and r2 are the radii of the two reactants, r12 is the radius of the reactant in contact, Dop is the square of the refractive index of reaction [36], Ds is the static dielectric constant and 0 is the vacuum permitivity. λ can in general be approximated as [35] λ = (µω2Q2) 2 , (9) where µ represents the reduced mass, Q is the normal mode, and the angular frequency of the harmonic oscil- lation ω = 2πν. The probability of an electron to penetrate barrier dur- ing tunneling is small for inelastic channel and greater for elastic channel. From quantum mechanics, the width of the barrier contributes highly to the prediction of the probability of penetration. The probability for a particle to penetrate a square barrier is given as [38] √ P ∝ e(−2/) 2m∆E‡r. (10) The probability varies as an exponential decay as the barrier width r increases and the decay constant varies as the square root of the product of the barrier height E‡ and the particle's mass m. To estimate the electron tunneling rate in a biologi- cal system, a generalized formula of non-adiabatic semi- classical Marcus theory describing the rate of olfaction can be obtained as [19] (cid:18) − (En − λ)2 (cid:19) exp , (11) 1 = 2π  t2 σn(cid:112)4πλkβT 4λkβT τD,0→A,n where En = D − A − nω0, D and A are the en- ergy states of the donor and acceptor, and ω0 is the vibrational mode of the odorant. The factor n rep- resents the number of phonon excitation on the odor- ant. n = 1 means one phonon excitation and n = 0 means zero phonon excitation. σn is poisson expres- sion for the dependence on Huang-Rhys factor S, σn = exp(−S)Sn/n!, λ is the environment reorganization en- q Sqωq, where Sq is the Huang-Rhys fac- tor for all oscillations in the environment. t contains an ergy. λ = (cid:80) 9 electronic coupling matrix element that determines the strength between D and A. Substituting values for a typical biological system, Brookes et al. [31] found the times characterizing elas- tic (τT0) and the inelastic (τT1 ) electron tunneling from donor to acceptor to be different. Their result gives τT0 ∼ 87ns and τT1 ∼ 1.3ns, thus implies τT1 (cid:28) τT0. This indicates that the inelastic electron tunneling could be the mechanism involved in the detection of odorants by the human nose. The result of the calculation also supports that the process of the detection of odorants happens within the order of milliseconds. The total time estimate includes all the time from different stages in the process of detection of the odorant. The time from the different stages include: • The time it takes the electron source to diffuse to the donor site (receptor protein). • The time it takes the electron (moving from the source) to be accepted at the donor site. • The time it takes the electron to either tunnel elas- tically or inelastically across the tunneling barrier. • The time it takes the electron to finally move from the acceptor site. V. A MODEL OF CHIRAL RECOGNITION IN OLFACTION One of the shortcomings of the vibrational-based model is the detection of enantiomers chiral odorants as having the same spectra and different smells. It is dif- ficult to use a simple model of vibrational olfaction to give an explanation to this [5]. Turin's explanation to this states that as a result of the different geometry of the enantiomers, the olfactory receptor seems not to be detecting some of the vibrational modes. For the case of carvone, the carbonyl in one of the two enantiomers is detected lesser intensely than the other as a result of wrong orientation of the molecule. The C=O group in (S)-carvone is silent. To solve this difference in smell, he suggested that 'adding back' some carbonyl stretch fre- quency back to the (R)-carvone would shift odor charac- ter from mint to caraway. This can be achieved through the mixture of (R)-carvone with a molecule (eg., 2- pentanone) having C=O stretch as its dominated vibra- tional spectrum. In fact a mixture of 2-pentanone and (R)-carvone results to a change of its mint smell to car- away smell. Arash Tirandaz et al. [39] in their paper presented a quantum model of olfaction for chiral recognition. This model first checks the physical viability of odorant- mediated inelastic electron tunneling in olfactory science. Unlike the vibrational model of olfaction which is rep- resented by a simple harmonic oscillator, the molecules are assumed to be undergoing a contorsion vibration as they oscillate between double-well potential energy sur- face. The contorsional mode gives the representation of the chiral recognition. To incorporate all the biological effects happening in the environment as one perceives smell, a set of harmonic oscillators stands as a represen- tation of the biological environment. The model takes into account three main components: 1) the chiral odor- ant with Hamiltonian Hod, 2) the electron that tunnels through the odorant to the receptor with Hamiltonian He and 3) the surrounding environment with Hamiltonian HEv, all sum up to contribute to the total Hamiltonian of the system, given by H0 = Hod + He + HEv. (12) Due to the chiral and the fundamental parity-violating interactions, asymmetry was introduced to account for the interactions of the molecules in the environment and the odorant. The odorant which is considered as an asymmetric double-well potential has left- and right- handed states L(cid:105) and R(cid:105) of the minima potential. As the electron tunnels through the barrier, the handed states are inter-converted. The Hamiltonian of the odor- ant includes the effect of the handed states. This is rep- resented as [39] Hod = − ωz 2 σz − ωx 2 σx, (13) where σi is the i-component of Pauli operator, ωz and ωx are the asymmetry and the tunneling frequencies re- spectively. The tunneling electron from the donor state D(cid:105) to the acceptor state A(cid:105) has energy D and A for donor and acceptor states respectively. These to- gether gives a Hamiltonian of the electron as He = A A(cid:105)(cid:104)A + D D(cid:105)(cid:104)D. The Hamiltonian of the biologi- bi where ωi is cal environment is given as HEv =(cid:80) ωi † b i i the i-frequency in the environment, and the creation and † annihilation operators are b i and bi respectively. In a broader view, the total Hamiltonian of the interaction is a combination of the three individual components, which are: 1) interaction between the donor and acceptor of the receptor with tunneling strength ∆, 2) the interac- tion between donor and the odorant, and acceptor and the odorant, with γD and γA as the coupling strengths respectively, and lastly, 3) interaction between the donor (acceptor) and the environment's i-th harmonic oscilla- tor having coupling strength γiD (γiA). The interaction Hamiltonian is given as [39] Hint = ∆ (A(cid:105)(cid:104)D + D(cid:105)(cid:104)A) (cid:88) + (γD D(cid:105)(cid:104)D + γA A(cid:105)(cid:104)A) σx + † (γiD D(cid:105)(cid:104)D + γiA A(cid:105)(cid:104)A) (b i + bi). (14) i The combination of the total Hamiltonian of the sys- tem gives the evolution used in calculating the electron tunneling rates. Tirandaz et al. [39] proposed measur- able parameters -- temperature and pressure which can 10 FIG. 12. The transitions in chiral odorant. There is no tran- sition for elastic electron tunneling. For the inelastic elec- tron tunneling, there is a vibrational transition from left- and right-handed states to the first excited energy state as the electron tunnels inelastically through the respective left- and right-handed enantiomers. Figure from [39]. be used to distinguish between elastic and inelastic tun- neling through the potential. FIG. 12 shows the transi- tion of the chiral odorant for an inelastic electron tun- neling. The result from Tirandaz et al. shows that there are thresholds for which the olfactory system would be able to recognize an odor. In their calculation, they ob- tained different rates for elastic and inelastic tunneling. The calculated ratio of the inelastic electron tunneling rate for the left-handed enantiomers to the right-handed counterparts was found to increase with the ratio of the tunneling frequency to the asymmetry frequency. Arash and co-workers [15, 39] proposed that there is an energy difference which occurs as a result of the interactions of chiral between the donor and the acceptor. The energy difference between two enantiomers of a chiral odorant is the main factor for recognizing chirals. VI. EXPERIMENTAL RESULTS ON VIBRATIONAL THEORY OF OLFACTION There are various experiments conducted both on mammals and insects to confirm the vibrational theory of olfaction. One of the methods of the experiments is the isotope exchange; a replacement of hydrogen (mass 1.007) with deuterium (mass 2.014). Deuteration does not change molecular shape, atom size or bond length or stiffness, rather the vibrational modes of the odorant as a result of doubling the hydrogen mass. This results to different smells of the isotopomers. Impurities can arise in the process of preparing the deuterium of the odorant. The presence of an impurity in an odorant can affect the perception odor. To avoid a bias interpretation of smell, adequate precautions are taken while preparing the synthesis of the odorant and the final molecule collected through efficient method (eg., Gas chromatography) that would only give a pure mono-molecule odorant. The olfactory system of mammals and insects have cer- tain common features as well as some unrelated features. For example, in mammals, the olfactory receptors which are proteins indicated in the cell membrane of olfactory receptor neurons (ORNs) are located in the nasal cavity, whereas in insects, they are located within sensilla pores on the antennas [40]. The odorant receptors in insects are genetically different from those possessed by vertebrates [41]. Also, the number of glomeruli (a spherical structure located in the olfactory bulb of the brain where synapses happen) are different in insects and vertebrates. Fruit- fly has about 62 glomeruli, 165 in the honeybee, 1800 in mice and about 1100 to 1200 in human [40]. An intuitive question that always arises is whether both the insects and the mammals use the same mecha- nism to identify smell despite having some differences in their olfactory system. One would not finally conclude that both use the same mechanism or not without the full knowledge of some suspected phenomena (still unclear) that could determine odor detection. A knowledge of the structures of olfactory receptors of both insect and mam- mal at the atomic level would also help in the conclusion. Nevertheless, there is a common fundamental mechanism followed in distinguishing different odors, so it wouldn't be surprising when the final answer turns out to be the same mechanism. In fact, a test on anosmic Drosophila (explained herein) has strongly suggested that flies use olfaction just like mammals to detect an odor. The idea that olfaction is related to the vibrational frequency of odorant is still regarded as speculative since no research is yet to give the structures of the odorant receptors, the binding sites or the processes involved in the activation and binding of odorant to the odorant re- ceptors [42, 43]. Both humans and insects have been confirmed to have the ability to distinguish between different isotopomers. Eric et al. [42] in their research tested the response of human musk-recognizing receptor OR5AN1, and also the mouse (methylthio) methanethiol-recognizing recep- tor, MOR244-3 to 1) deuterated, 2) nondeuterated and 3) 13C isotopomers. In doing so, they considered the effects of impurities and isotope effects in the interpre- tation of the odor perceived since some experimental re- sults in olfaction have been criticized of having impuri- ties. Example, an experiment by Haffenden et al. [44] using benzaldehyde-d6 and benzaldehyde gave that both isotopomers have statistically significant different odors supporting the vibrational theory. This result was crit- icized of not accounting for the perireceptor events (ie., the enzyme-mediated biochemical conversion of odorants in the nasal mucus before reaching the olfactory recep- tor) and not having double-blind controls to get rid of bias in their duo-trio test [45]. Indeed, there are several conflicting results on whether a human can distinguish [18, 44, 46] between deuterated odorants (benzaldehyde and acetophenone) or not [45, 47]. It appears that musk isotopomers are accepted to be easily distinguished by human [47]. Also, studies show that Drosophila melanogaster (fruit flies) can distinguish between isotopomers of acetophenone [48, 49]. Likewise, training Apis mellifera (the honey bee) makes them able to distinguish pairs of isotopomer [50]. Though concerns have been made on the response of the Drosophila to be behavioral and not related to olfactory receptors, and 11 since signaling of the olfactory receptors in Drosophila is different to the human's, then the result of the test on Drosophila should not be finally attributed to the ability of the human to differentiate isotopomers [48, 51]. To this end, instead of relying on the behavioral tests, Block et al. [42] deemed it to be essential to test the vibrational theory of olfaction at a molecular level us- ing receptor-based assays. They found that human musk receptor, OR5AN1 was actually able to distinguish be- tween muscone and muscone-d30 suggesting that the vi- bration theory does not apply to it neither does it apply to the mouse thiol receptor MOR244-3 or other olfactory receptors tested. Their theoretical analysis of vibrational theory also suggests that it is unrealistic in the biological content. However, it would be interesting to look into details of one experiment each supporting or refuting the vibra- tional theory of olfaction. It is worth noting that the choice of the experiments we will be focusing on was based on my behavioral response (interest) on the meth- ods used in the experiments. We, therefore, discuss the experiments below. A. Experimental support of the vibrational theory of olfaction To test whether animals can distinguish an odor or not, a behavior action in responding to a given odor is exam- ined. Franco et al. [48] examined this in their experiment using a T-maze shape olfactomers and fruit flies placed between the arms of the T-maze having odorants. The preferential response of the flies was observed by count- ing the number of flies that moved to each arm of the T-maze in preference to the odorant placed in there. Acetophenone (ACP, C8H8O) and its deuterium atoms three, five and eight (d3, d5, d8) were used in the first stage of the experiment. It is expected that the flies should respond differently to ACP and deuterated ACP due to their different smells and vibrational frequencies. The odorants are > 99% impurity free. ACP was di- luted in nonvolatile odorless isopropyl myristate (IPM, C17H34O2) and flies placed at the T-maze arm to choose between ACP and IPM. The flies (> 15% excess flies in the preferred arm) exhibited a vivid natural preference for ACP. When d3-, d5- and d8-ACP were tested, a con- trary outcome was obtained. It was found that the flies start to show preference towards the IPM. A clear pref- erence about 15% excess flies in the preferred arm IPM was seen when d8-ACP was used. They also examined whether flies could discriminate between ACP and its deuterated counterpart d8-ACP and whether the amount of concentration of the odorants can affect the behavioral preference action. Their result showed a preference of ACP against equal concentration (1:1) of d8-ACP. But reducing the amount of concentration of d8-ACP to 50% gave no significant preferential outcome of the flies. The flies approximately equally filled the arms of the T-maze. In a similar way, they tested whether the flies could discriminate between isotopomers 1-octanol (OCT, C8H18O) and its deuterated counterpart d17-OCT (C8D17OH). The flies showed preference to the OCT with > 30% excess flies in the arm. This discrimination against d17-OCT was removed by reducing the concen- tration of d17-OCT by 75% (1:0.25). They observed that the arms were filled with approximate equal flies, signify- ing that the discrimination is based on odor perception. To eliminate the doubt that flies might be using other features rather than olfactory sense to discriminate against a particular odorant, Franco and colleagues used anosmic mutants in their test. This should lead to an elimination of the differential response to the deuterated odorants. The anosmic Drosophila Or83b1 and Or83b2 mutants showed no preferential discrimination against d8-ACP and also against d17-OCT. The mutants were distributed equally at the arms of the T-maze as would expect. This implies that the flies use olfaction alone to discriminate between a normal odorant and its deuter- ated counterpart [48]. Additionally, there should be a salient feature that results in the spontaneous discrimi- nation against deuterated odorant from its normal coun- terpart. Flies can be trained to recognize and avoid a particular odorant by the use of an electric foot-shock punishing stimulus [52]. In all the tests on 3 pairs of the odorants used by Franco and colleagues, flies continuously avoided shock-associated odor. This strengthens the idea that flies can distinguish between isotopomers. To examine the possibility that the spontaneous pref- erences seen on the flies are not as a result of impuri- ties which the odorant might be contained even though the odorants are of high impurity-free, the flies were conditioned to generally avoid either deuterated or nor- mal odorant from different unrelated odorants. The flies avoided all the unrelated deuterated odorants when trained to avoid a particular deuterated odorant. The same behavior is seen when the flies are conditioned to avoid particular natural hydrogenated odorants, they avoided all the unrelated normal odorants. This suggests that a salient feature such as the molecular vibrations of the odorant could be the feature the flies sense to dis- tinguish isotopomers. If so, the flies use the modes most affected by deuteration (e.g., the C-H stretch) to general- ize and distinguish deuterium from the normal odorants. The C-H stretch occurs approximately at 3,000 cm−1 re- gion but reduces to 2,200 cm−1 region for deuterated odorants, C-D. Indeed, if the molecular vibrational frequency is the feature that determines the identification of odor, then flies should not be able to distinguish between odorants with the same known vibrational frequency and the same odor. Citronellyl nitrile (NIT, C10H17N) and citronel- lal (ALD, C10H18O) were used in the test since they possess similar odor characteristics to the human nose. Both have a lemongrass-like smell. Just like odor per- ception of human, flies showed no preference between the 12 FIG. 13. A result of a T-maze experiment to show fruit flies use vibrational frequency in odor discrimination. FIG.(A) shows the computed vibrational spectra of OCT and d17- OCT. As a result of deuteration, a reduction of the vibration stretch from 3,000cm−1 (OCT) to 2,150cm−1 (d17-OCT) is obtained. FIG.(B) shows the computed vibration spectra for citronellal (ALD) and citronellyl nitrile (NIT). Both have al- most the same vibrational frequency. ALD has two noticeable stretch frequencies; for C=O stretch at 1750cm−1 and for C- H stretch at 2,765cm−1, while NIT has a different stretch at 2,150cm−1. FIGs. (C) and (B) show the odor discriminations by the flies when conditioned to avoid a particular odorant. The lightning symbol represents the electric foot shock used to condition the flies. Figure from [48]. two odorants (FIG. 13C). The IR spectra of both odor- ants in the fingerprint region are almost the same (FIG. 13B). They only differ in vibrations stretchs. Aldehyde has C=O stretch around 1,740cm−1 and aldehydic C- H stretch around 2,765cm−1. NIT and d17-OCT share the same vibrational stretch at 2,150cm−1. However, on conditioning of the flies to avoid d17-OCT resulted in a discrimination against NIT. Likewise, conditioning the flies to avoid NIT (FIG. 13D) showed a discrimination against d17-OCT. This suggests that flies detect the vi- bration of odorant's functional group and use it to dis- criminate among odorants. Some other researchers [eg., 49, 50] used similar method and similar odorant samples to show that just like mammals, insects can discriminate between isotopomers. B. Experiment refuting the vibrational theory of olfaction Inasmuch as so many authors accepted the vibrational theory as a plausible theory in olfaction, some still argue that it is not a viable theory for it fails to account for some differences in the smell of enantiomers. To prove the implausibility of vibration theory, Rajeev et al. [43] 13 smell of oranges) [55]. Limonene mainly occurs as the (R)-enantiomers with a strong smell of a fresh citrus or- ange, while (S)-limonene isomer has a hash, turpentine- like lemony smell [56]. It is noteworthy to point out that some other odorants can be used for this experiment. The choice for these odorants was influenced by their commercial availability, non-harmful nature and relatively affordable. Using a proper wafting technique, the students determined and recorded whether the odorants have the same smell or not. A compilation of the result is shown in table 15. Comparing the percentage of the students reporting whether the enantiomers smell the same or different, it indicates that the acetophenone and acetophenon- d3 have an identical odor, while the enantiomeric car- vaones and limonenes have different odors for the first two years the tests were performed. In the third year, there was an equal percentage report of acetophenone and acetophenone-d3 having a different smell and the same smell. But when acetophenone and acetophenone- d8 were tested, the students reported they smell the same. The inconsistency of the result from first two years and the third year was attributed to be possibly due to the perireceptor events [57] in the nasal mucus as reported by Brock et al. [42]. The biochemical interac- tions between the odorants and the nasal mucus (in the vertebrates) or the sensillar lymph (in insects) can affect the final odor of the odorants. The presence of impurities in one of the commercial samples of acetophenone-d3 was reported as possibly being the reason for different percep- tions recorded by the students during the first two years and the third year. In conclusion, their perception result does not support the vibrational theory of olfaction as reported in litera- ture because in the case of the isotopomers having the same odor, this is not in agreement with the vibrational theory stating that molecules having different vibration frequencies would smell differently. Also, if the shapes of the two enantiomers are not considered, then they should have the same vibration, thus the same smell, but that was not the case observed in the experiment. VII. CONCLUSION AND SUMMARY In this paper, we have studied the theories of olfac- tion. The studied theories include the 'lock and key' model, the odotope model, vibrational theory by Dyson, Swipe-card model by Stoneham, Turin theory of inelastic electron tunneling spectroscopy (IETS) and chiral recog- nition models. Among all these, Turin's model has gained more popularity. There are some experimental evidence that support the idea of detection of smell to be a func- tion of the molecular frequency, while some experiments have showcased that vibrational theory of olfaction is im- plausible and as such, should not be regarded as an ex- planation of odor detection and disquisition. It is still an unsolved question whether recognition of FIG. 14. The structures of the odorants used in the experi- ment. Figure from [43]. reported two different tests that confirmed the theory of vibration cannot be solely used to explain olfaction. They designed two different tests for groups of students for three years. These different groups had to test 1) the vibrational theory of olfaction using isotopomeric odor- ants and 2) the enantiomeric odorants. A sample of acetophenone and aceptophenone-d3 were used to test whether deuterated odorants and their non- deuterated counterparts have different odors. In the sec- ond test, a sample of (R)- and (S)-carvone, and (R)- and (S)-limonene (FIG. 14) were used to test whether enan- tiomeric odorants have the same odor since they ought to have the same bond vibrations. These tests can only validate the vibrational theory of olfaction: • if the end of the experiment confirms that the deuterated compounds have a different odor com- pared to the nondeuterated compounds as pre- dicted in literature. • if in the conclusion of the experiment, the enan- tiomers were confirmed to have the same odor since they have identical vibrational spectra. Firstly, the students predicted and compared the stretching frequency of a C-D bond stretch and C-H bond by using the simple harmonic approximation to bond vi- brations and applied Hooke's law. C-H bond was found to have a higher bond stretching frequency than the C- D bond confirming that deuterated compounds and its counterparts should have different vibrational frequen- cies. In the other hand, the measured IR spectra of the enantiomers were found to be identical while they have different odors. This does not necessarily generally con- clude that all enantiomeric odorant have different odors even though the majority do [17]. (R)-isomers of car- vone are the most abundant compound in the essen- tial oil from different species of mint thus, has a strong spearmint-leaves odor, while its mirror image (S)-carvone smells like caraway seeds [53, 54]. The biosynthesis of carvone is by oxidation of limonene (which has a strong 14 FIG. 15. A compilation of the result from the students on determining whether the odorants have the same smell or not [43]. smell is either solely by shape theory or the molecular vi- brations of the odorant or a combination of the two [47]. However, since olfactory receptors belong to the family of class A G protein-coupled receptors, and since proteins are chiral, then shape theory provides an insufficient ex- planation to why most pair of enantiomers has the same odor. The vibrational theory has faced a lot of antagonism over its inability to reconcile the fact that some enan- tiomers have different smell even when they have the same vibrational spectrum. An experiment by Franco and colleagues has sug- gested that fruit flies use the molecular vibration of an odorant for its preferential selection and discrimina- tion. In contrast, the result of the experiment by Rajeev and colleagues refutes the vibrational theory of olfaction since they obtained the same smell for isotopomers used in their test, while the vibrational theory states that molecules with different vibrational frequencies should have different smells. Also, enantiomers with the same vi- bration frequency should smell the same because if shape is neglected, bond is the only thing that affects vibration. But in their own case, they observed different smells for their enantiomers. Although the theories of olfaction are faced with some challenges and some unanswered questions (like how the shape of the olfactory receptor looks), I believe olfaction has a connection with a quantum mechanical principle, thus a focus on a biological mechanism obtained from a principle in quantum mechanics should result to a per- manent solution to the unanswered questions in the field of quantum biology. VIII. REFERENCES [1] E. Schrodinger, What is life? (University Press: Cam- [12] K. Harini and R. Sowdhamini, PloS one 10, e0131077 bridge, 1943). (2015). [2] D. Abbott, P. C. Davies, and A. K. Pati, Quantum as- pects of life (World Scientific, 2008). [3] J. McFadden and J. Al-Khalili, Life on the edge: the com- ing of age of quantum biology (Broadway Books, 2016). [4] L. Turin, Chemical senses 21, 773 (1996). [5] G. Malcolm Dyson, Journal of Chemical Technology and Biotechnology 57, 647 (1938). [6] G. Gerlach, J. Atema, M. J. Kingsford, K. P. Black, and V. Miller-Sims, Proceedings of the national academy of sciences 104, 858 (2007). [7] D. L. Dixson, G. P. Jones, P. L. Munday, S. Planes, M. S. Pratchett, M. Srinivasan, C. Syms, and S. R. Thorrold, Proceedings of the Royal Society of London B: Biological Sciences 275, 2831 (2008). [8] M. Arvedlund and K. Kavanagh, in Ecological connec- tivity among tropical coastal ecosystems (Springer, 2009) pp. 135 -- 184. [9] M. Arvedlund, Journal of the Marine Biological Associ- ation of the United Kingdom 89, 863 (2009). [13] L. Buck and R. Axel, Cell 65, 175 (1991). [14] J. E. Amoore, Nature 199, 912 (1963). [15] A. Tirandaz, F. T. Ghahramani, and V. Salari, arXiv preprint arXiv:1701.01050 (2017). [16] J. C. Brookes, A. P. Horsfield, and A. M. Stoneham, Journal of The Royal Society Interface 6, 75 (2009). [17] R. Bentley, Chemical reviews 106, 4099 (2006). [18] L. Turin and F. Yoshii, Handbook of olfaction and gus- tation , 275 (2003). [19] J. C. Brookes, Contemporary Physics 52, 385 (2011). [20] C. V. Raman, (1930). [21] K. Mori and G. M. Shepherd, in Seminars in cell biology, Vol. 5 (Elsevier, 1994) pp. 65 -- 74. [22] J. Lambe and R. Jaklevic, Physical Review 165, 821 (1968). [23] M. Mohseni, Y. Omar, G. S. Engel, and M. B. Plenio, Quantum effects in biology (Cambridge University Press, 2014). [24] N. E. Rawson and G. Gomez, Microscopy research and [10] F. from Internet, (Accessed on January 8, 2018) technique 58, 142 (2002). https://sites.google.com/a/edmail.edcc.edu/savanna- test/literature-research/pheromones-and-neurocircuits. [11] B. Malnic, J. Hirono, T. Sato, and L. B. Buck, Cell 96, 713 (1999). [25] R. Williams (ACS Publications, 1990). [26] H. Lodish, A. Berk, S. L. Zipursky, P. Matsudaira, D. Baltimore, J. Darnell, et al., Molecular cell biology, Vol. 3 (Scientific American Books New York, 1995). 15 [27] J. C. Brookes, A microscopic model of signal transduc- tion mechanisms: olfaction, Ph.D. thesis, UCL (Univer- sity College London) (2009). [28] D. Restrepo, Y. Okada, J. H. Teeter, L. D. Lowry, B. Cowart, and J. Brand, Biophysical journal 64, 1961 (1993). N. Ragoussis, E. M. Skoulakis, and L. Turin, PloS one 8, e55780 (2013). [48] M. I. Franco, L. Turin, A. Mershin, and E. M. Skoulakis, Proceedings of the National Academy of Sciences 108, E350 (2011). [49] E. R. Bittner, A. Madalan, A. Czader, and G. Roman, [29] N. Thurauf, M. Gjuric, G. Kobal, and H. Hatt, European The Journal of chemical physics 137, 22A551 (2012). Journal of Neuroscience 8, 2080 (1996). [30] O. P. Hamill, A. Marty, E. Neher, B. Sakmann, and F. Sigworth, Pflugers Archiv European journal of physi- ology 391, 85 (1981). [31] J. C. Brookes, F. Hartoutsiou, A. Horsfield, and A. Stoneham, Physical review letters 98, 038101 (2007). [32] P. Atkins and J. De Paula, Atkin's physical chemistry , 25 (2002). [50] W. Gronenberg, A. Raikhelkar, E. Abshire, J. Stevens, E. Epstein, K. Loyola, M. Rauscher, and S. Buchmann, Proceedings of the Royal Society of London B: Biological Sciences 281, 20133089 (2014). [51] C. S. Sell, Chemistry and the Sense of Smell (John Wiley & Sons, 2014). [52] T. Tully and W. G. Quinn, Journal of Comparative Phys- iology A 157, 263 (1985). [33] R. A. Marcus, Annual Review of Physical Chemistry 15, [53] S. L. Murov and M. Pickering, J. Chem. Educ 50, 74 155 (1964). (1973). [34] H. Eyring, Chemical Reviews 17, 65 (1935). [35] J. C. Brookes, in Proc. R. Soc. A, Vol. 473 (The Royal Society, 2017) p. 20160822. [54] T. J. Leitereg, D. G. Guadagni, J. Harris, T. R. Mon, and R. Teranishi, Journal of Agricultural and Food Chem- istry 19, 785 (1971). [36] D. DeVault, Quantum-mechanical tunnelling in biological [55] A.-T. Karlberg, K. Magnusson, and U. Nilsson, Contact systems (Cambridge University Press, 1984). Dermatitis 26, 332 (1992). [37] D. Devault, Quarterly reviews of biophysics 13, 387 (1980). [38] H. B. Gray and J. R. Winkler, Proceedings of the Na- tional Academy of Sciences of the United States of Amer- ica 102, 3534 (2005). [39] A. Tirandaz, F. T. Ghahramani, and A. Shafiee, Physical Review E 92, 032724 (2015). [40] A. Horsfield, A. Haase, Physics: X 2, 937 (2017). and L. Turin, Advances in [41] U. B. Kaupp, Nature Reviews Neuroscience 11, 188 (2010). [42] E. Block, S. Jang, H. Matsunami, S. Sekharan, B. De- thier, M. Z. Ertem, S. Gundala, Y. Pan, S. Li, Z. Li, et al., Proceedings of the National Academy of Sciences 112, E2766 (2015). [56] L. Friedman and J. G. Miller, Science 172, 1044 (1971). [57] P. Pelosi, Developmental Neurobiology 30, 3 (1996). [58] Q. Fu, Y. Luo, J. Yang, and J. Hou, Physical Chemistry Chemical Physics (Incorporating Faraday Transactions) 12, 12012 (2010). [59] A. J. Leggett, S. Chakravarty, A. Dorsey, M. P. Fisher, A. Garg, and W. Zwerger, Reviews of Modern Physics 59, 1 (1987). [60] L. Turin, Chem Technol Flavors Fragrances , 261 (2009). [61] E. Block, S. Jang, H. Matsunami, V. S. Batista, and H. Zhuang, Proceedings of the National Academy of Sci- ences 112, E3155 (2015). [62] L. Turin, S. Gane, D. Georganakis, K. Maniati, and E. M. Skoulakis, Proceedings of the National Academy of Sciences 112, E3154 (2015). [43] R. S. Muthyala, D. Butani, M. Nelson, and K. Tran, [63] L. B. Vosshall, Proceedings of the National Academy of Journal of Chemical Education 94, 1352 (2017). Sciences 112, 6525 (2015). [44] L. Haffenden, V. Yaylayan, and J. Fortin, Food chem- [64] K. Maniati, K.-J. Haralambous, L. Turin, and E. M. istry 73, 67 (2001). Skoulakis, eNeuro 4, ENEURO (2017). [45] A. Keller and L. B. Vosshall, Nature neuroscience 7, 337 [65] R. H. Wright, Journal of Theoretical Biology 64, (2004). 473IN1475 (1977). [46] L. Turin, Chemistry and industry , 866 (1997). [47] S. Gane, D. Georganakis, K. Maniati, M. Vamvakias,
1811.00909
1
1811
2018-11-02T14:56:48
Hydrodynamics of bacteriophage migration along bacterial flagella
[ "physics.bio-ph", "cond-mat.soft", "physics.flu-dyn" ]
Bacteriophage viruses, one of the most abundant entities in our planet, lack the ability to move independently. Instead, they crowd fluid environments in anticipation of a random encounter with a bacterium. Once they land on the cell body of their victim, they are able to eject their genetic material inside the host cell. Many phage species, however, first attach to the flagellar filaments of bacteria. Being immotile, these so-called flagellotropic phages still manage to reach the cell body for infection, and the process by which they move up the flagellar filament has intrigued the scientific community for decades. In 1973, Berg and Anderson (Nature, 245, 380) proposed the nut-and-bolt mechanism in which, similarly to a rotated nut that is able to move along a bolt, the phage wraps itself around a flagellar filament possessing helical grooves (due to the helical rows of flagellin molecules) and exploits the rotation of the flagellar filament in order to passively travel along it. One of the main evidence for this mechanism is the fact that mutants of bacterial species such as Escherichia coli and Salmonella typhimurium that possess straight flagellar filaments with a preserved helical groove structure can still be infected by their relative phages. Using two distinct approaches to address the short-range interactions between phages and flagellar filaments, we provide here a first-principle theoretical model for the nut-and-bolt mechanism applicable to mutants possessing straight flagellar filaments. Our model is fully analytical, is able to predict the speed of translocation of a bacteriophage along a flagellar filament as a function of the geometry of both phage and bacterium, the rotation rate of the flagellar filament, and the handedness of the helical grooves, and is consistent with past experimental observations.
physics.bio-ph
physics
Hydrodynamics of bacteriophage migration along bacterial flagella Panayiota Katsambaa and Eric Lauga† Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Cambridge CB3 0WA, United Kingdom (Dated: November 5, 2018) Bacteriophage viruses, one of the most abundant entities in our planet, lack the ability to move independently. Instead, they crowd fluid environments in anticipation of a random encounter with a bacterium. Once they 'land' on the cell body of their victim, they are able to eject their genetic material inside the host cell. Many phage species, however, first attach to the flagellar filaments of bacteria. Being immotile, these so-called flagellotropic phages still manage to reach the cell body for infection, and the process by which they move up the flagellar filament has intrigued the scientific community for decades. In 1973, Berg and Anderson (Nature, 245, 380-382) proposed the nut-and-bolt mechanism in which, similarly to a rotated nut that is able to move along a bolt, the phage wraps itself around a flagellar filament possessing helical grooves (due to the helical rows of flagellin molecules) and exploits the rotation of the flagellar filament in order to passively travel along it. One of the main evidence for this mechanism is the fact that mutants of bacterial species such as Escherichia coli and Salmonella typhimurium that possess straight flagellar filaments with a preserved helical groove structure can still be infected by their relative phages. Using two distinct approaches to address the short-range interactions between phages and flagellar filaments, we provide here a first-principle theoretical model for the nut-and-bolt mechanism applicable to mutants possessing straight flagellar filaments. Our model is fully analytical, is able to predict the speed of translocation of a bacteriophage along a flagellar filament as a function of the geometry of both phage and bacterium, the rotation rate of the flagellar filament, and the handedness of the helical grooves, and is consistent with past experimental observations. 8 1 0 2 v o N 2 ] h p - o i b . s c i s y h p [ 1 v 9 0 9 0 0 . 1 1 8 1 : v i X r a a Current address: School of Mathematics, University of Birmingham, Edgbaston, Birmingham, B15 2TT, UK † [email protected] 2 FIG. 1. Bacteriophages: (A) A typical morphology of a bacteriophage, such as the Enterobacteria T4 phage (Adenosine, Wikimedia Commons); (B) Electron micrograph of bacteriophages attached to a bacterial cell, (Dr. G. Beards, Wikimedia Commons); (C-E) Phages can come in various shapes: (C) Myoviridae; (D) Podoviridae; (E) Siphoviridae [8]. Panels C-E: Reprinted by permission from Suttle CA, "Viruses in the sea", Nature, 437 (356), 356-361, Copyright 2005 Springer Nature. I. INTRODUCTION As big as a fraction of a micrometre, bacteriophages (in short phages), are 'bacteria-eating' viruses (illustrated in Fig. 1) that infect bacteria and replicate within them [1]. With their number estimated to be of over 1031 on the planet, phages are more abundant than every other organism on Earth combined [2 -- 8]. Phages have been used extensively in genetic studies [1, 4, 9], and their future use in medicine is potentially of even greater impact. The global rise in antibiotic resistance, as reported by the increasing number of multidrug-resistant bacterial infections [10], poses one of the greatest threats to human health of our times, and phages could offer the key to resolution. Indeed, phages have been killing bacteria for way longer than humanity has been fighting against bacterial infections, with as many as 1029 infections of bacterial cells by oceanic phages taking place every day [11, 12]. Phage therapy is an alternative to antibiotics that has been used for almost a century and offers promising solutions to tackle antibiotic-resistant bacterial infections [13]. Furthermore, the unceasing phage-bacteria war taking place in enormous numbers offers the scientific community great opportunities to learn. For example, the ability of phages to update their infection mechanisms in response to bacterial resistance could offer us valuable insight into updating antibiotics treatment against multi drug-resistant pathogenic bacteria [14]. In addition, the high selectivity of the attachment of a phage to the receptors on the bacterial cell surface and the species it infects could help identify possible target points of particular pathogenic bacteria for drugs to attack [15]. In general, extensive studies of bacteriophage infection strategies could not only reveal vulnerable points of bacteria, but may help uncover remarkable biophysical phenomena taking place at these small scales. Infection mechanisms can vary across the spectrum of phage species [15, 16]. Lacking the ability to move indepen- dently, phages simply crowd fluid environments and rely on a random encounter with a bacterium in order to land on its surface and accomplish infection using remarkable nanometre size machinery. Typically, the receptor-binding proteins located on the long tail fibres recognise and bind to the receptors of the host cell via a two-stage process called phage adsorption [15]. The first stage is reversible, and is followed by irreversible attachment onto the cell surface. Subsequently, the genetic material is ejected from their capsid-shaped head, through their tail, which is a hollow tube, into the bacterium [17, 18]. While all phages need to find themselves on the surface of the cell body for infection to take place, there is a class of phages, called flagellotropic phages, that first attach to the flagellar filaments of bacteria. Examples include the χ-phage infecting Escherichia coli (E. coli) and Salmonella typhimurium (Salmonella), the phage PBS1 infecting Bacillus subtilis (B. subtilis) and the recently discovered phage vB VpaS OWB (for short OWB) infecting Vibrio parahaemolyticus (V. parahaemolyticus) [19], illustrated in Fig. 2. Given the fact that phages are themselves incapable of moving independently and that the distance they would have to traverse along the flagellar filament is large compared to their size, they must find an active means of progressing along the flagellar filament. In Ref. [22], electron microscopy images of the flagellotropic χ-phage, shown in Figs. 2C and D, were provided to show that the mechanism by which χ-phage infects E. coli consists of travelling along the outside of the flagellar filament until it reaches the base of the flagellar filament where it ejects its DNA. A possible mechanism driving the translocation of χ-phage along the flagellar filament was first proposed in Berg and Anderson's seminal paper as the 'nut-and-bolt' mechanism [23]. Their paper is best known for establishing that bacteria swim by rotating their flagellar filaments. One of the supporting arguments was the proposed mechanism HeadCollarTailLong Tail FibresBase PlateDNAProtein2D3D 3 FIG. 2. Flagellotropic phages: (A) Attachment of phage OWB to V. parahaemolyticus [19]. Red arrows indicate phage particles. (B) Phage PBS1 adsorbed to the flagellar filament of a B. Subtilis bacterium with its tail fibres wrapped around the flagellar filament in a helical shape with a pitch of 35 nm [20]. The phage hexagonal head capsid measures 120 nm from edge to edge [21]. Reprinted (amended) by permission from American Society for Microbiology from Raimondo LM, Lundh NP, Martinez RJ, "Primary Adsorption Site of Phage PBS1: the Flagellum of Bacillus", J. Virol., 1968, 2 (3), 256-264, Copyright 1968, American Society for Microbiology. (C) χ-phage of E. coli [22]. The head measures 65 to 67.5 nm between the parallel sides of the hexagon [22]; (D) χ-phage at different times between attachment on the flagellar filament of E. coli and reaching the base of the filament [22]. Arrows point to the bases of the flagella. Panels C-D: Reprinted (amended) by permission from American Society for Microbiology from Schade SZ, Adler J, Ris H, "How Bacteriophage χ Attacks Motile Bacteria", J. Virol., 1967, 1 (3), 599-609, Copyright 1967, American Society for Microbiology. where the phage plays the role of the nut and the bolt is the flagellar filament, with the grooves between the helical rows of flagellin molecules making up the flagellar filament serving as the threads [23] (Figs. 3A and B). A phage would then wrap around the flagellar filament and the rotation of the latter would result in the translocation of the phage along it. A mutant of Salmonella that has straight flagellar filaments, but possesses the same helical screw-like surface due to the arrangement of the flagellin molecules [24] is non-motile due to the lack of chiral shape yet fully sensitive to χ-phage [26], i.e. the phages manage to get transported to the base of the flagellar filament. This is consistent with the nut-and-bolt mechanism and was used as evidence that the flagellar filament is rotating [23]. More evidence in support of the nut-and-bolt mechanism were provided 26 years after its inception in a work studying strains of Salmonella mutants with straight flagellar filaments whose motors alternate from rotating clockwise (CW) and counter-clockwise (CCW) [27]. The directionality of rotation is crucial to the mechanism as CCW rotation will only pull the phage toward the cell body if the phage slides along a right-handed groove. In order to test the directionality, the authors used a chemotaxis signalling protein that interacts with the flagellar motor, decreasing the CCW bias. They found that strains with a large CCW bias are sensitive to χ-phage infection, whereas those with small CCW bias are resistant, in agreement with the proposed nut-and-bolt mechanism. Details of the packing of the flagellin molecules that give rise to the grooves can be found in Ref. [25] and examples are shown in Figs. 3A and B. It is important to note that the packing of flagellin molecules produces two overlapping sets of helical grooves, a long-pitch and a short-pitch set of grooves which are of opposite chirality [24]. Once in contact with a rotating flagellar filament, it is anticipated that the phage fibres will wrap along the short-pitch grooves. Indeed, the findings of Ref. [27] show that the directionality of phage translocation correlates with the chirality of the short-pitch grooves. The flagellar filaments of bacteria can take one of the twelve distinct polymorphic shapes as illustrated in Figs. 3B and C. The authors in Ref. [27] examined flagellar filaments with different polymorphic forms, since the different arrangements of the flagellin subunits give rise to grooves with different pitch and chirality [25], as shown in Fig. 3B. The L-type straight flagellar filament (f0) that was used by Ref. [27] has both left-handed long-pitch grooves and right- handed short-pitch grooves [24]. Given that the short-pitch grooves are relevant to the wrapping of the fibres, the findings of Ref. [27] that bacteria with their flagellar filament in the f0 polymorphic state (i.e. with right-handed short- pitch grooves) and with a large CCW bias are sensitive to χ-phage infection, are in agreement with the nut-and-bolt 4 FIG. 3. Bacterial flagellar filaments and polymorphism: (A) Structure of straight flagellar filament from a mutant of Salmonella typhimurium [24]. Reprinted with permission from O'Brien EJ, Bennett PM, "Structure of straight flagella from a mutant Salmonella", J. Mol. Biol., 70 (1), 145-152, Copyright 1972 Elsevier. The L-type straight flagellar filament (left) has two types of helical grooves, left-handed long-pitch and right-handed short-pitch grooves, whereas the R-type straight filament (right) has right-handed long-pitch and left-handed short-pitch grooves. Examples of short-pitch grooves are marked by thin blue lines and indicated by arrows. (B) Schematic of some polymorphic states of the flagellar filament, from left to right: L-type straight, normal (left-handed shape), curly (right-handed) and R-type straight. The top panel shows the shape of the filaments while the bottom panel displays the arrangements of flagellin subunits. Examples of short-pitch grooves are marked by thin blue lines and indicated by arrows [25]. Reprinted with permission from Namba K, Vonderviszt F, "Molecular architecture of bacterial flagellum", Q. Rev. Biophys., 1997, 30 (1), 1-65, Copyright 1997 Cambridge University Press. mechanism. The same study also argued that the translocation time of the phage to the cell body is less than the flagellar filament reversal interval, a necessary condition for successful infection by the virus of wild-type bacteria whose motors alternate between CCW and CW rotation. Their estimated translocation speeds on the order of microns per seconds give a translocation time which is less than the CCW time interval of about a 1 s [27]. Relevant to the nut-and-bolt mechanism are also the findings of an alternative mechanism for adsorption of the flagellotropic phages φCbK and φCb13 that interact with the flagellar filament of Caulobacter crescentus using a filament located on the head of the phage [28], instead of the tail or tail fibres that other flagellotropic phages use, such as χ and P BS1. This study also reports on a higher likelihood of infection with a CCW rotational bias that is consistent with the nut-and-bolt mechanism. Notably, phages can also attach to curli fibres, which are bacterial filaments employed in biofilms; however due to the lack of helical grooves and rotational motion of these filaments, phages are unable to move along them [29]. In this paper, we theoretically examine the nut-and-bolt mechanism from a quantitative point of view and perform a detailed mathematical analysis of the physical mechanics at play. We focus on the virus translocation along straight flagellar filaments in mutants such as the mutant of Salmonella used in Ref. [23]. A flagellotropic phage can wrap around a given flagellar filament using its tail fibres (fibres for short), its tail, or in some cases a filament emanating from the top of its head [28] and the models we develop can address all these relevant morphologies. A schematic diagram of the typical geometry we consider is shown in Fig. 4. A phage floating in a fluid whose fibres suddenly collide with a flagellar filament rotating at high frequency will undergo a short, transient period of wrapping, during which the length of the fibres that are wrapped around the filament is increasing. In this paper we study the translocation of the phage once it has reached a steady, post-wrapping state, and assume that it is moving rigidly with no longer any change in the relative virus-filament configuration. In order to provide first-principle theoretical modelling of the nut-and-bolt mechanism, we build in the paper a hierarchy of models. In §II, we start with a model of drag-induced translocation along smooth flagellar filaments that ignores the microscopic mechanics of the grooves yet implicitly captures their effect by coupling the helical shape of the fibres with anisotropy in motion in the local tangent plane of the flagellar filament. Having acquired insight into the key characteristics of the mechanism, we proceed by building a refined, more detailed model of the guided 5 FIG. 4. Schematic model of the translocation of a flagellotropic phage along the straight flagellar filament of a mutant bacterium. The phage, illustrated in dark green, is wrapped around the straight flagellar filament (light blue cylinder) using its fibres, with its tail and head protruding in the bulk fluid. translocation of phages along grooved flagellar filaments by incorporating the microscopic mechanics of the grooves in §III. This is done by including a restoring force that acts to keep the fibres in the centre of the grooves, thereby guiding their motion, as well as a resistive force acting against the sliding motion. In both models, we proceed by considering the geometry, and the forces and torques acting on the different parts of the phage. We use the resistive-force theory of viscous hydrodynamics in order to model the tail and tail fibres which are both slender [30, 31]. The portion of the phage wrapping around the flagellar filament is typically the fibres. They experience a hydrodynamic drag from the motion in the proximity of the rotating flagellar filament along which they slide in the smooth flagellum model, or a combination of a guiding and resistive forces in the grooved flagellum model. Parts sticking out in the bulk away from the flagellar filament experience a hydrodynamic drag due to their motion in an otherwise stagnant fluid. We build in our paper a general mathematical formulation relevant to a broad phage morphology. In our typical geometry of phages wrapping around flagellar filaments using their fibres, two limits arise for long-tailed and short- tailed phages. Long-tailed phages have their tail and head sticking out in the bulk, away from the flagellar filament, whereas for short-tailed phages only the head is exposed to the bulk fluid. The hydrodynamic torque actuating the translocation is provided by the parts sticking out in the bulk. We compare the results from the two models addressing the two geometrical limits and find these to be consistent with each other and with the predictions and experimental observations of Refs. [23, 27]. In particular, we predict quantitatively the speed of phage translocation along the flagellar filament they are attached to, and its critical dependence on the interplay between the chirality of the wrapping and the direction of rotation of the filament, as well as the geometrical parameters. Most importantly we show that our models capture the correct directionality of translocation, i.e. that CCW rotation will only pull the phage toward the cell body if the phage slides along a right- handed groove, and predict speeds of translocation on the order of µms−1, which are crucial for successful infection in the case of bacteria with alternating CCW and CW rotations. II. DRAG-INDUCED TRANSLOCATION ALONG SMOOTH FLAGELLAR FILAMENTS II.1. Geometry As our first model, we consider the flagellar filament as a straight, smooth rod aligned with the z-axis and of radius Rf l. The phage has a capsid head of size 2ah, a tail of length Lt and fibres that wrap around the flagellar filament. We implicitly capture the effect of the grooves (i) by imposing that the fibres that emanate from the bottom of the tail of the phage are wrapped around the flagellar filament in a helical shape and (ii) via the anisotropy in the drag arising from the relative motion between the fibres and the rotating flagellar filament. The helical shape of the fibres has helix angle α, as shown in Fig. 5. With the assumption that the gap between the fibres and the flagellar filament is negligible compared to the radius Rf l of the flagellar filament, the centreline rf ib(s) of the fibres, parametrised by 6 FIG. 5. Mathematical model of drag-induced translocation along a smooth flagellar filament. The fibres experience an anisotropic drag due to their motion in the proximity of the rotating flagellar filament. The phage, shown in green, has a capsid head of size 2ah, a tail of length Lt, fibres of helical shape with helix angle α and cross-sectional radius rf ib, and is translocating along a straight flagellar filament, shown in light blue, of radius Rf l. The flagellar filament is rotating at a rate ωf l. The phage is translocating at speed U and rotating with rate ωp about the flagellar filament. The inset shows a short segment of the phage fibre at a distance d from the local tangent plane of the flagellar filament. the contour length position s, is described mathematically as (cid:18) (cid:18) (cid:19) (cid:18) rf ib(s) = Rf l cos , hRf l sin s Rf l/ sin α s Rf l/ sin α (cid:19) (cid:19) , s cos α , −LL f ib < s < LR f ib, (1) f ib + LR f ib (left side) and LR f ib, where we allow for fibres extending to both sides of the base of the tail f ib (right side). The helix wrapping is right-handed or left-handed according to with total contour length Lf ib = LL to have lengths LL whether the chirality index h takes the value +1 or −1 respectively. Assuming the phage to move rigidly and working in the laboratory frame, every point r on the phage moves with velocity U ez + ωpez ∧ r. The flagellar filament is assumed to rotate at rate ωf l along its axis, and thus its velocity is given by ωf lez ∧ r in a fluid that is otherwise stationary, where the value of ωf l is known. The purpose of our calculation is to compute the two unknown quantities, U and ωp, in terms of ωf l by enforcing the overall force and torque balance on the phage along the z-axis. II.2. Forces and moments In order to calculate the forces and torques acting on the tail and fibres we use the resistive-force theory of viscous hydrodynamics (RFT in short) [30, 31]. This theoretical framework predicts the viscous tractions due to the motion of a slender filament in a viscous fluid by integrating fundamental solutions of the Stokes equations of hydrodynamics [32] along the centreline of the filament. In an infinite fluid, the instantaneous hydrodynamic force per unit length exerted on a filament due to its motion (cid:20) (cid:18) ∂r (cid:12)(cid:12)(cid:12)(cid:12)s (cid:19) ∂r (cid:21) (cid:12)(cid:12)(cid:12)(cid:12)s (cid:18) ∂r ∂s (cid:12)(cid:12)(cid:12)(cid:12)s − ζ(cid:107) (cid:19) ∂r (cid:12)(cid:12)(cid:12)(cid:12)s , f (s) = −ζ⊥ v(s) − (cid:12)(cid:12)s and v(s) are the local unit tangent and velocity of the filament relative to the fluid at contour-length where ∂r ∂s position s respectively, and ζ(cid:107), ζ⊥ are the drag coefficients for motion parallel and perpendicular to the local tan- gent [30, 33]. For a slender rod of length L and radius r in an infinite fluid, we have .v(s) .v(s) (2) ∂s ∂s ∂s in an otherwise stagnant viscous fluid is given by where µ is the dynamic viscosity of the fluid. ζ⊥ ≈ 4πµ ln (L/r) , ρ ≡ ζ(cid:107)/ζ⊥ ≈ 1/2, (3) The fact that the perpendicular drag coefficient is twice the parallel one captures the fact that it is twice as hard to pull a rod through a viscous fluid in a direction perpendicular to its length than lengthwise. This drag anisotropy is at the heart of the propulsion physics for microorganisms such as bacteria and spermatozoa [31]. 7 As a result, the total hydrodynamic force and the z-component of the torque on the phage tail due to its motion in the fluid are given by Ftail = − ez · Mtail = −ez · Lt(cid:90) (cid:2)ζ(cid:107),tttailttail + ζ⊥,t(1 − ttailttail)(cid:3) · urel Lt(cid:90) rt(s) ∧(cid:8)(cid:2)ζ(cid:107),tttailttail + ζ⊥,t(1 − ttailttail)(cid:3) · urel tail(s) ds, 0 tail(s)(cid:9) ds, (4) (5) (7) (8) (9) (10) (13) (14) 0 where rt(s) and ttail(s) are the position and tangent vectors of the fibre centreline at contour length position s respectively. The symbols ζ⊥,t, ζ(cid:107),t are the drag coefficients for motion perpendicular and parallel to the local tangent, with ζ(cid:107),t ≡ ρtζ⊥,t and the velocity of the tail relative to the fluid is tail(s) = ωp (ez ∧ rt) + U ez. urel (6) For the fibres, we use the version of RFT modified to capture the motion of slender rods near a surface. The flagellar filament is rotating at rate ωf l, thus the velocity of the fibres relative to the flagellar filament is given by f ib(s) = Ωrel (ez ∧ rf ib) + U ez, urel where the relative angular velocity is given by Ωrel = ωp − ωf l. The expressions for the fibres are similar, and we have L(R) f ib(cid:90) f ib(cid:90) L(R) −L(L f ib) (cid:2)ζ(cid:107),f ibtf ibtf ib + ζ⊥,f ib(1 − tf ibtf ib)(cid:3) · urel rf ib(s) ∧(cid:8)(cid:2)ζ(cid:107),f ibtf ibtf ib + ζ⊥,f ib(1 − tf ibtf ib)(cid:3) · urel f ib(s) ds, f ib(s)(cid:9) ds. Ff ib = − ez · Mf ib = −ez · −L(L) f ib The difference between the expressions in Eqs. 4,5 and Eqs. 9,10 is that in the latter we use the appropriate resistance coefficients, ζ⊥,f ib and ζ(cid:107),f ib, for motion at a small, constant distance d from a nearby surface, in directions perpendicular and parallel to the local tangent of the fibre respectively, given by ζ⊥,f ib ≈ 4πµ (11) with ζ(cid:107),f ib = ρf ibζ⊥,f ib and again ρf ib ≈ 1/2 (see Ref. [34] and references therein). These results are valid in the limit in which the distance d between the fibre and the surface of the flagellar filaments is much smaller than the radius of the flagellar filament (d (cid:28) Rf l), such that the surface of the smooth flagellar filament is locally planar. Importantly, the very drag anisotropy that allows the rotation of helical flagellar filaments to propel bacteria in the bulk will also enable the rotation of helical fibres around a smooth filament to lead to translocation along the axis of the filament. If the tail is straight with total length Lt and the head is spherical with radius ah, the position of the centre of the ln (2d/rf ib) , head is given by rh = rb + Ltttail + ahth, (12) where rb = (Rf l, 0, 0) is the base of the tail from which the fibres emanate. The drag force and torque due to the motion of the head in the otherwise stagnant fluid are given by Fhead = −6πµahurel ez · Mhead = ez ·(cid:2)−6πµah head, (cid:0)rh ∧ urel head (cid:1)(cid:3) − 8πµa3 hωp, with urel h given by h = ωp (ez ∧ rh) + U ez. urel 8 (15) Taking th = ttail, the centre of the head will be located at position rh = rb + (Lt + ah) ttail. Evaluating the integrals in Eqs. 4,5,9 and 10 and the expressions of Eqs. 13,14 with this geometry we obtain the forces and torques exerted on the different parts of the phage (projected along the z-axis), (cid:2)(sin2 α + ρf ib cos2 α)U − (1 − ρf ib) sin α cos αhΩrelRf l (cid:3) , (cid:2)hRf lΩrel(cos2 α + ρf ib sin2 α) − (1 − ρf ib) sin α cos αU(cid:3) , U(cid:2)1 − (1 − ρt)t2 (cid:20) (cid:3) − ωpRf l(1 − ρt)tytz (cid:21) (cid:27) , z ez · Ff ib = −ζ⊥,f ibLf ib ez · Mf ib = −hζ⊥,f ibRf lLf ib ez · Ftail = −ζ⊥,tLt (cid:26) ez · Mtail = −ζ⊥,t ωp LtR2 (cid:26) 1 3 L3 t Rf ltx + f l + L2 − (1 − ρt)ty(U tz + ωpRf lty)Rf lLt x + t2 y) t (t2 (cid:27) , ez · Fhead = −6πµahU, ez · Mhead = −6πµahωp (cid:2)R2 f l + 2Rf l(Lt + ah)tx + (Lt + ah)2(t2 x + t2 y)(cid:3) − 8πµa3 hωp, where we use the notation ttail = (tx, ty, tz) for the components of the tangent of the tail. II.3. Phage translocation: General formulation The overall force and torque balance on the phage along the z-axis is written as (16) (17) (18) (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) which leads to 0 = − ζ⊥,f ibLf ib − ζ⊥,tLt 0 = − hζ⊥,f ibRf lLf ib 0 = ez · [Ff ib + Ftail + Fhead], 0 = ez · [Mf ib + Mtail + Mhead], (cid:26) (cid:27) (cid:2)(sin2 α + ρf ib cos2 α)U − (1 − ρf ib) sin α cos αhΩrelRf l (cid:3) U(cid:2)1 − (1 − ρt)t2 (cid:2)hRf lΩrel(cos2 α + ρf ib sin2 α) − (1 − ρf ib) sin α cos αU(cid:3) (cid:20) (cid:2)R2 (cid:3) − ωpRf l(1 − ρt)tytz (cid:21) y)(cid:3) − 8πµa3 f l + 2Rf l(Lt + ah)tx + (Lt + ah)2(t2 − 6πµahU, t Rf ltx + f l + L2 x + t2 y) x + t2 LtR2 hωp. t (t2 L3 1 3 z (cid:26) − ζ⊥,t ωp − 6πµahωp − (1 − ρt)ty(U tz + ωpRf lty)Rf lLt (cid:27) Writing this system in a formal matrix form for U and ωp in terms of ωf l gives (cid:19) (cid:19) (cid:18)Z (cid:19)(cid:18) U (cid:3) + ζ⊥,f ibLf ib(sin2 α + ρf ib cos2 α) + 6πµah, ωf l, ωp H = B D (cid:18)A B (cid:2)1 − (1 − ρt)t2 (cid:20) A = ζ⊥,tLt B = − [ζ⊥,tLt(1 − ρt)tytz + hζ⊥,f ibLf ib(1 − ρf ib) sin α cos α] Rf l, z D = ζ⊥,tLt f l + LtRf ltx + (1 − t2 z) − (1 − ρt)R2 f lt2 y L2 t 3 f l(cos2 α + ρf ib sin2 α) f l + 2Rf l(Lt + ah)tx + (Lt + ah)2(1 − t2 (cid:21) z)(cid:3) R2 (cid:2)R2 + ζ⊥,f ibLf ibR2 + 6πµah + 8πµa3 h, Z = −hζ⊥,f ibLf ibRf l(1 − ρf ib) sin α cos α, H = ζ⊥,f ibLf ibR2 f l(cos2 α + ρf ib sin2 α). where we have Inverting Eq. 25 gives the translation linear and rotational speeds as (cid:18) U (cid:19) (cid:18) D −B −B A (cid:19)(cid:18)Z 1 (cid:19) 9 (31) The full expressions for (DZ − BH),(AD − BC) and (−CZ + AH) for a general phage geometry are given in the Supplementary Material (see [35]). AD − B2 ωf l. ωp H = II.4. Two limits: long vs short-tailed phages From the experimental images in Fig. 2 we can distinguish two geometries of wrapping according to how far the tail and head are sticking out in the bulk fluid and away from the flagellar filament. We thus proceed by considering the two limiting geometries of long- and short-tailed phages. II.4.1. Long-tailed phages Examples of long-tailed morphology include the χ-phage of E. coli and the PBS1 phage of B. subtilis shown in Figs. 2B and D. We use below the χ-phage as a typical long-tailed phage, whose detailed dimensions are reported in Ref. [22]. The hexagonal head measures 650 − 675 A between parallel sides (that is 2ah ≈ 650 − 675 A). The tail is a flexible rod that is 2, 200 A long and 140 A wide, and the tail fibres are 2, 000− 2, 200 A long and 20− 25 A wide. The flagellar filaments are 5 − 10 µm long and have a diameter 20 nm, hence Rf l ≈ 100 A [36]. For this specific phage, we thus have ah ≈ 30 nm, Lt ≈ 220 nm, Lf ib ≈ 200 nm, rf ib ≈ 1 nm and Rf l ≈ 10 nm. From this we see that we can safely assume that Rf l, ah (cid:28) Lt, Lf ib. Notice however that Lf ib ≈ Lt and that Rf l and ah are of the same order of magnitude. Our variables are thus divided into the short lengthscales of ah, Rf l and the long lengthscales of Lf ib, Lt. With these approximations we obtain the translocation speed as Ulong ≈ − hωf lRf l(1 − ρf ib) sin α cos α Glong, Glong = ζ⊥,f ibLf ib (32) (33) (cid:2) 1 3 ζ⊥,tLt(1 − t2 ζ⊥,tLt [1 − (1 − ρt)t2 (cid:20) z)(cid:3)(cid:3) z) +(cid:2)ζ⊥,tRf ltx + 6πµah(1 − t2 (cid:21) , z] + ζ⊥,f ibLf ib(sin2 α + ρf ib cos2 α) (cid:21) 1 3 ζ⊥,tLt(1 − t2 z) (cid:20) ≈ ζ⊥,f ibLf ib (34) ζ⊥,tLt [1 − (1 − ρt)t2 z] + ζ⊥,f ibLf ib(sin2 α + ρf ib cos2 α) with a relative error of O (ah/Lt, ah/Lf ib, Rf l/Lt, Rf l/Lf ib). Details of the approximation are given in the Supple- mentary Material (see [35]). II.4.2. Short-tailed phages Phages with very short tails that use their fibres to wrap around flagellar filaments are equivalent geometrically to phages that use their entire tail for wrapping since in both cases there is a filamentous part of the phage wrapped around the flagellar filament and the head is sticking out in the bulk close to the surface of the filament. For example, the phage OWB that infects V. parahaemolyticus studied in Ref. [19] and shown in Fig. 2A, uses its tail for wrapping. In order to avoid any confusion, we will carry out the calculations of this section using the geometry of short-tailed phages, and assume that (i) the tail is negligible and (ii) the fibres are wrapping around the flagellar filament. In this case we obtain a translocation speed of Ushort = −hRf lωf l(1 − ρf ib) sin α cos α Gshort, Gshort = (cid:2)(Rf l + ahtx)2 + a2 (cid:2)ζ⊥,f ibLf ib(sin2 α + ρf ib cos2 α) + 6πµah f l + 6πµahζ⊥,f ibLf ibR2 ζ⊥,f ibLf ib6πµah ρf ibζ 2⊥,f ibL2 +6πµah f ibR2 (cid:20) ht2 3 a2 y + 4 (cid:3)(cid:2)(Rf l + ahtx)2 + a2 f l(cos2 α + ρf ib sin2 α) h (cid:3) (35) (36) (cid:3)(cid:21) , ht2 y + 4 3 a2 h with details of the calculation given in the Supplementary Material (see [35]). Note that the results for phages which use their tail to wrap around the flagellar filament can be readily obtained by replacing Lf ib with Lt and all relevant quantities in the above result. 10 II.4.3. Interpretation and discussion of the asymptotic results We now interpret and compare the results we obtained in Eqs. 32 and 35. As we now see, our formulae give the cor- rect directionality and speed of translocation in agreement with the qualitative predictions and the experimental data of Ref. [27], as well as the requirements for translocation, thereby providing insights to the translocation mechanism. Firstly, and most importantly, both results for the translocation speeds in Eqs. 32 and 35 have the common factor −hRf lωf l(1 − ρf ib) sin α cos α which is multiplying the positive dimensionless expressions Glong and Gshort respectively. The factor −hωf l gives a directionality for U in agreement with the qualitative prediction of Ref. [27] that CCW rotation will only pull the phage toward the cell body if the phage slides along a right-handed groove. Indeed, our model captures this feature: for right-handed helical wrapping (h = +1) and CCW rotation of the flagellar filament when viewing the flagellar filament towards the cell body (ωf l < 0), the phage moves towards the cell body (i.e. U > 0). Secondly, the factor (1 − ρf ib) reveals that translocation requires anisotropy in the friction between the fibres and the surface of the flagellar filament (i.e. ρf ib (cid:54)= 1). We interpret the requirement for anisotropy as an indication of the important role of the grooves in guiding the motion of the fibres. The assumption of a helical wrapping of the fibres coupled with this anisotropy simulates the guiding effect of the grooves in this first model by resisting motion perpendicular to the local tangent of the grooves and promoting motion parallel to it. Thirdly, the presence of the factor sin α cos α shows the requirement of a proper helix i.e. there is no translocation in the limiting cases of a straight (α = 0) or circular wrapping (α = π/2). Fourthly, translocation requires a non-vanishing value of Lf ib in the numerator of both Eqs. 32 and 35. This is because the fibres are providing the 'grip' by wrapping around the flagellar filament. Fifthly, the terms involving the tail and head appear in both the numerator and denominator of Eq. 32. Similarly, terms involving the head appear in both the numerator and denominator of Eq. 35. These show that the parts of the phage that are sticking out in the bulk are contributing to both the torque actuating the motion of the phage relative to the flagellar filament and the drag. In the case of short phages, only the head is providing the torque, hence the terms in the numerator of Eq. 35. Finally, focusing on the χ-phage, the lengths of the tail and the fibres are similar and the logarithmic dependence of the resistance coefficients allow us to estimate the fraction in Eq. 32 to be of O(1). The grooves have a pitch of approximately 50 nm [27] and the radius of the flagellar filament is approximately 10 nm, giving rise to helix angle α ≈ 51◦. With ωf l ≈ 100 Hz, we have that U = O(µms−1). Importantly, this means that it is possible for χ-phage to translocate along a flagellar filament of a few µm long within the timescale of a second, in agreement with the CCW time interval for bacteria with alternating CCW and CW rotation, thereby enabling the phage to reach the cell body and infect the bacterium. II.5. Dependence of translocation speed on geometrical parameters We now illustrate the dependence of the translocation speed on the geometrical parameters of the phage, namely the lengths Lt and Lf ib. The asymptotic formulae we obtained above and discussed in §II.4 will help verify the asymptotic behaviour of U for large values of Lt and for vanishing tail length, as well as explain the trends for the translocation speed with increasing Lt and Lf ib. To fix ideas, we consider the specific case of the χ phage and hence the following set of parameter values (as in §II.4.1 and Ref. [22]): ah ≈ 30 nm, Lt ≈ 220 nm, rtail = 7 nm, Lf ib ≈ 200 nm, rf ib ≈ 1 nm, Rf l ≈ 10 nm and √ µ = 10−3 Pa.s. For the helix angle we take α = 51◦ (see §II.4.3). We also take (tx, ty, tz) = (1, 1, 1)/ 3. Note that for simplicity we do not include the slow variation of the resistive coefficients with Lt, but instead keep their constant non dimensional values, ζ⊥,t = 4π/ ln(220/7), and ζ⊥,f ib = 4π/ ln(4) based on ζ⊥,t ≈ 4πµ/ ln (Lt/rtail) with Lt/rtail = 220/7 and ζ⊥,f ib ≈ 4πµ/ ln (2d/rf ib) with d = 2rf ib. We then use Eq. 31 to plot U versus Lt and Lf ib in Fig. 6. For simplicity we have non-dimensionalised lengths by Rf l, time by ω−1 f l and viscosity by µ, and denote dimensionless quantities using a hat. In Fig. 6A we observe that the phage translocation speed is a decreasing function of Lt. The value of U for vanishing tail length is well captured by the theoretical approximation for short-tailed phages of Eq. 36 (red star). For large values of Lt, it approaches the theoretical approximation for long-tailed phages (black dash-dotted line, inset). For the long-tailed approximation we used the expressions in Eq. A1-A2 in the Supplementary Material (see [35]) keeping terms up to and including L3, for L either Lt or Lf ib. We note that the smaller the phage head size, ah, the better the convergence between Eq. 31 and the long-tailed approximation (which assumes ah, Rf l (cid:28) Lt, Lf ib). In Fig. 6A we used the dimensions for the χ phage, that correspond to ah/Lf ib = 3/22, and the long-tailed approximation also requires ah (cid:29) Lt. The long-tailed approximation from Eq. 34 captures the decreasing behaviour of U with Lt and 11 FIG. 6. Smooth flagellum model: Dependence of the translocation speed, U , on the geometrical parameters of the phage. Left: U vs. the length of the phage tail, Lt, based on Eq. 31 (blue solid line). The speed is a decreasing function of Lt. The value of U for vanishing tail length is well captured by the theoretical approximation for short-tailed phages (red star). For large values of Lt it approaches the theoretical approximation for long-tailed phages (black dash-dotted line, inset). Right: U vs. the length of the fibres, Lf ib, following Eq. 31. The speed is an increasing function of Lf ib. The long-tailed approximation of Eq. 34 also captures this increasing behaviour. can be used to explain this result. Specifically, the term involving Lt in the denominator of Eq. 34 shows that this decay arises from the resistive part of the hydrodynamic force on the tail (the first term in Eq. 18) which increases as Lt increases. In other words, when Lt increases the drag increases and therefore the speed decreases. In Fig. 6B we next show the speed U as a function of the length of the fibres, Lf ib. Clearly the speed is an increasing function of Lf ib and the long-tailed approximation of Eq. 34 captures this increasing behaviour. We observe that terms involving Lf ib appear in both the numerator and denominator of Eq. 34. Dividing top and bottom by Lf ib, we see that as Lf ib increases, U increases (at some point U will asymptote to a constant value, but the length at which this happens appears to be too large to be relevant biologically). The increase of U with Lf ib stems from the propulsive forces per unit length on the fibres that integrate to a larger propulsive torque as Lf ib increases. Of course there is also the resistive drag on the fibres but that starts to become more important only at larger values of Lf ib. III. GUIDED TRANSLOCATION ALONG GROOVED FLAGELLAR FILAMENTS III.1. Geometry As a more refined physical model, we now include in this section the mechanics arising from the microscopic details of the grooved surface of the flagellar filament due to the packing of the flagellin molecules and modify the previous calculation in order to account for the motion of the phage fibres sliding along the helical grooves. If the phage slides with speed V along the grooves of helix angle α in the frame of the straight flagellar filament, as illustrated in Fig. 7, then the translocation velocity and rotation rate measured in the laboratory frame, U and ωp, become U = V cos α, hV sin α ωp = Rf l + ωf l. (37) (38) 12 FIG. 7. Guided translocation of phage along a grooved flagellar filament. The phage, shown in green, has a capsid head of size 2ah, a tail of length Lt and fibres of cross-sectional radius rf ib. The flagellar filament (light blue) has helical grooves of helix angle α and is rotating at a rate ωf l. The phage slides along the grooves with speed V in the frame of the flagellar filament. As shown in the inset, the force acting on the fibre sliding along the grooves consists of two parts: (i) a drag resisting the sliding motion of magnitude µV in the −tf ib direction and (ii) a restoring force acting to keep the fibre in the centre of the groove of magnitude kδ in the hbf ib direction, where δ is the local offset of the centre of the fibre cross-section from the centre of the groove and bf ib is the local binormal to the fibre centreline that lies in the local tangent plane of the surface of the flagellar filament and is perpendicular to the tangent of the fibre centreline. With this substitution we obtain the forces and torques acting on the tail and head as (39) (40) (41) (42) (cid:26) (cid:26) U(cid:2)1 − (1 − ρt)t2 (cid:20) z (cid:3) − ωpRf l(1 − ρt)tytz (cid:21) (cid:27) , 1 3 L3 t (t2 t Rf ltx + f l + L2 − (1 − ρt)ty(U tz + ωpRf lty)Rf lLt x + t2 y) (cid:27) , ez · Ftail = −ζ⊥,tLt ez · Mtail = −ζ⊥,t ωp LtR2 ez · Fhead = −6πµahU, ez · Mhead = −6πµahωp (cid:2)R2 f l + 2Rf l(Lt + ah)tx + (Lt + ah)2(t2 x + t2 y)(cid:3) − 8πµa3 hωp. III.2. Forces and moments The details of the interactions between the phage fibres and the grooves are expected to be complicated as they depend on the parts of the flagellin molecules that make up the groove surface and interact with the proteins that the fibres consist of. These interactions could originate from a number of short range intermolecular forces, for example electrostatic repulsion or Van der Waals forces. We model here the resultant of the interaction forces acting on the fibre sliding along the grooves as consisting of two parts, a drag and a restoring force, as shown in the inset of Fig. 7. Firstly, the fibre is subject to a viscous drag, −µV tf ib per unit length where µ is a hydrodynamic resistance coefficient against the sliding motion (with dimensions of a viscosity). A simple approximation for that coefficient is to assume that there is a fully-developed shear flow resisting the sliding between the fibres and the surface of the grooves. Assuming the cross-section of the latter to be a circular arc, so that a fraction fcov of the circumference of a cross-section of the fibres lies inside the groove, we obtain approximately µ ≈ 2πrf ibfcovµ hgap , (43) where hgap is the size of the gap between the grooves and the fibres and rf ib is the radius of the fibres. Secondly, there should be a restoring force acting to keep the fibre in the centre of the groove arising from the physical interactions between the fibre and the groove. A simple modelling approach consists of viewing each side of the groove as repelling the fibre, with the resultant of these forces providing a restoring force hkδbf ib(s) per unit length, arising from a potential well 1 2 kδ2 where δ is the distance from the centre of the well, and bf ib is the local binormal vector to the fibre centreline, (cid:20) (cid:18) (cid:19) (cid:18) (cid:19) (cid:21) bf ib = h cos α sin s Rf l/ sin α ,− cos α cos s Rf l/ sin α , h sin α , (44) 13 that lies in the local tangent plane of the surface of the flagellar filament and is perpendicular to the tangent vector tf ib of the fibre centreline. Assuming δ to be uniform along the length of the fibres, the expressions for the force and torque on the fibres given as the integrals, (cid:90) L(R) (cid:90) L(R) −L(L) f ib f ib f ib −L(L) f ib Ff ib = Mf ib = −µV tf ib(s) − hkδbf ib(s) ds, rf ib(s) ∧ [−µV tf ib(s) − hkδbf ib(s)] ds, when projected along the z direction become ez · Ff ib = [−µV cos α − kδ sin α] Lf ib, ez · Mf ib = − [µV hRf l sin α − hkδRf l cos α] Lf ib. III.3. Phage translocation: General formulation The total force and torque balance on the phage along the z-axis, 0 = ez · [Ff ib + Ftail + Fhead], 0 = ez · [Mf ib + Mtail + Mhead], give the system to be solved in order to find the two unknown quantities, V and kδ, in terms of ωf l, (cid:3) − (cid:18) hV sin α Rf l (cid:19) + ωf l Rf l(1 − ρt)tytz (cid:27) LtR2 f l + L2 t Rf ltx + x + t2 y) 1 3 t (t2 L3 (cid:18) hV sin α Rf l (cid:21) (cid:19) V cos αtz + + ωf l Rf lty Rf lLt f l + 2Rf l(Lt + ah)tx + (Lt + ah)2(t2 x + t2 (cid:21) (cid:27) y)(cid:3) 0 = − [µV hRf l sin α − hkδRf l cos α] Lf ib 0 = [−µV cos α − kδ sin α] Lf ib − ζ⊥,tLt − 6πµahV cos α, (cid:26) z Rf l + ωf l V cos α(cid:2)1 − (1 − ρt)t2 (cid:19)(cid:20) (cid:26)(cid:18) hV sin α (cid:20) (cid:19)(cid:2)R2 (cid:18) hV sin α (cid:19) (cid:18) hV sin α (cid:18)A B − (1 − ρt)ty + ωf l + ωf l Rf l Rf l h . − ζ⊥,t − 6πµah − 8πµa3 Using matrix notation, these two equations take the form (cid:19)(cid:18) V (cid:19) (cid:18)Z (cid:19) H = ωf l, C D kδ (45) (46) (47) (48) (49) (50) (51) (52) where A = (cid:0)cos α(cid:2)1 − (1 − ρt)t2 z ζ⊥,tLt (cid:1) + µLf ib cos α + 6πµah cos α (cid:27) , (cid:3) − h sin α(1 − ρt)tytz (cid:20) C = h sin α f l + L2 hµLf ibRf l sin α + ζ⊥,t − (1 − ρt)tyζ⊥,t [cos αtz + h sin αty] Rf lLt LtR2 Rf l t Rf ltx + L3 t (t2 1 3 f l + 2Rf l(Lt + ah)tx + (Lt + ah)2(t2 x + t2 (cid:26) (cid:26) B = Lf ib sin α, + 6πµah h sin α Rf l D = − hRf lLf ib cos α, Z = ζ⊥,tLtRf l(1 − ρt)tytz, H = − (cid:26) (cid:20) f l + L2 LtR2 ζ⊥,t (cid:2)R2 (cid:2)R2 (cid:27) , h sin α Rf l x + t2 y) (cid:21) y)(cid:3) + 8πµa3 (cid:21) y)(cid:3) + 8πµa3 (cid:27) h h . 14 (53) (54) (55) (56) (57) (58) (59) (60) t Rf ltx + L3 t (t2 x + t2 y) − (1 − ρt)t2 yR2 f lLt 1 3 + 6πµah f l + 2Rf l(Lt + ah)tx + (Lt + ah)2(t2 x + t2 The details of the calculation are given in the Supplementary Material (see [35]). Inverting Eq. 52, V and kδ are obtained as V = ωf l, DZ − BH AD − BC −CZ + AH AD − BC (61) where the full expressions for (DZ − BH),(AD − BC) and (−CZ + AH) for a general phage geometry are given in the Supplementary Material (see [35]). Finally, from Eq. 60, the translocation velocity along the z-axis is calculated as U = V cos α. kδ = ωf l, We now proceed by considering the two limiting geometries of long-tailed and short-tailed phages similarly to §II.4. III.4. Two limits: long vs short-tailed phages Under the approximations relevant for long-tailed phages such as χ-phage described in §II.4, i.e. Rf l, ah (cid:28) Lt, Lf ib, III.4.1. Long-tailed phages the translocation velocity along the z-axis gets simplified to Ulong = −hRf lωf l sin α cos αGlong, Glong = 3 ζ⊥,tLt + 6πµah (cid:3) (1 − t2 (cid:2) 1 L2 t (cid:2) 1 (cid:3) (1 − t2 z) f lLf ib with the details of the approximation given in the Supplementary Material (see [35]). 3 ζ⊥,tLt + 6πµah z) sin2 α + µR2 L2 t · (62) (63) In the case of short-tail phages, we assume that the tail is negligible and that the fibres are wrapping around the III.4.2. Short-tail phages flagellar filament. The translocation velocity simplifies then to Ushort = −hRf lωf l sin α cos αGshort, Gshort = 6πµah µLf ibR2 f l + 6πµah R2 3 − t2 f l + 2Rf lahtx + a2 z) f l + 2Rf lahtx sin2 α + a2 h( 7 R2 (cid:105) h sin2 α(cid:0) 7 3 − t2 z (cid:104) (cid:104) (cid:1)(cid:105)· (64) (65) with all calculation details in the Supplementary Material (see [35]). 15 FIG. 8. Grooved flagellum model: Dependence of the translocation speed on the geometrical parameters of the phage. Left: U as a function of the length of the phage tail, Lt, based on Eq. 60 (blue solid line). The speed is an increasing function of Lt. The value of U at vanishing tail length is well captured by the theoretical approximation for short-tailed phages (red star). For large values of Lt, it approaches the theoretical approximation for long-tailed phages (black dash-dotted line). Right: U vs. the length of the fibres, Lf ib, based on Eq. 60. The long-tailed approximation of Eq. 63 captures the decreasing behaviour of U with Lf ib. III.4.3. Interpretation and discussion of the results Similarly to §II.4.3, we interpret and compare the results in Eqs. 62 and 64. Here again, the crucial factor −hRf lωf l sin α cos α appears in both equations multiplying a positive, non-dimensional expression, and we obtain the correct directionality and speed of translocation in agreement with Ref. [27]. The prefactor −hωf l gives the cor- rect directionality for U , i.e. for right-handed helical wrapping (h = +1) and CCW rotation of the flagellar filament (ωf l < 0), the phage moves towards the cell body (U > 0), in agreement with Ref. [27]. Again, the translocation speed of O(µm/s) allows translocation during the CCW time interval for bacteria that alternate between CCW and CW sense, in agreement with Ref. [27]. The factor sin α cos α shows that a proper helix is needed for translocation. The presence of the term µR2 f lLf ib in the denominator implies that the sliding drag from the fibre decreases the translocation speed, and longer fibres give a decreased speed. Further, and similarly to §II.4.3, the terms inside the square brackets in both the numerator and denominator of Eqs. 63 and 65 show that the parts of the phage that are sticking out in the bulk (for the long phages these are the tail and the head, for the short phages it is only the head), are contributing to both the torque that is actuating the motion of the phage relative to the flagellar filament and to the drag. III.5. Dependence of translocation speed on geometrical parameters We now illustrate the dependence of the translocation speed on the geometrical parameters of the phage, namely Lt and Lf ib, according to our model of translocation along grooved flagellar filaments. We use the same approach, parameter values and non-dimensionalisation as in §II.5 and as there we denote dimensionless quantities using a hat. For µ ≡ µ/µ ≈ 2πrf ibfcov/hgap, we take rf ib/hgap = 2, i.e. assume the gap between the grooves and the fibres is half the fibre cross-sectional radius, and fcov = 1/2, so that the cross-section of the grooves is a semi-circle. These parameter values give µ = 2π. We then use Eq. 60 to plot U (calculated as U = V cos α) as a function of both Lt 16 FIG. 9. Grooved flagellum model: Dependence of the phage translocation speed, U , on the 'effective' viscosity, µ (non-dimensionalised by the fluid viscosity µ). As µ increases, the resistive part of the force from the motion of the fibres in the grooves increases, which slows down the phase and leads to a decrease of U . and Lf ib in Fig. 8. In Fig. 8A we observe that the translocation speed is an increasing function of Lt. The value of U at vanishing tail length is well captured by the theoretical approximation for short-tailed phages of Eq. 65 (red star in figure). For large values of Lt, the result approaches the theoretical approximation for long-tailed phages (black dash-dotted line). The long-tailed approximation from Eq. 63, through the terms with Lt in both numerator and denominator, is able to capture the increasing behaviour of U with Lt. Physically, this trend is caused by the propulsive terms in ez · Mtail in Eq. 40 (proportional to L3 t ) that increase as Lt increases. Next in Fig. 8B we show the speed U as a function of the length of the fibres, Lf ib, as predicted by Eq. 60, and observe a decreasing trend. The long-tailed approximation of Eq. 63 is able to capture this behaviour. The presence of the term involving Lf ib in the denominator of Eq. 63 leads to a decrease of U with Lf ib, and is physically due to an increase of the viscous drag on the fibres as Lf ib increases. Finally, as shown in Fig. 9, we obtain that the translocation speed is a decreasing function of the 'effective' viscosity in the grooves, µ, due to the resistive part of the force from the motion of the fibres in the grooves, as expected. IV. CONCLUSION In this work, we carried out a first-principle theoretical study of the nut-and-bolt mechanism of phage translocation along the straight flagellar filaments of bacteria. The main theoretical predictions from our two models, Eqs. 32, 35, 62 and 64, give the phage translocation speed, U , in terms of the phage and groove geometries and the rotation rate of the flagellar filament, in the two relevant limits of long- and short-tailed phages. These mathematical results capture the basic qualitative experimental observations and predictions of Refs. [23, 27] for the speed and directionality of translocation which are both crucial for successful infection. The common prefactor in the formulae for the translocation speed along the filament, U ∼ −hωf lRf l sin α cos α, appears in the expressions from both models. This provides the expected directionality in agreement with Refs. [23, 27]: sliding of the fibres along flagellar filaments with right-handed helical grooves (h = +1), combined with CCW rotation of flagellar filament (ωf l < 0), will give rise to phage translocation towards the cell body (U > 0) for infection to follow, whereas CW rotation (ωf l > 0) would translocate the phage away from the cell body (U < 0), towards the free end of the flagellar filament, and thus away from the cell body. Quantitatively, the speeds predicted by our model are estimated to a few micrometres per second, U = O(µms−1). This is important for phages infecting bacteria which alternate between CCW and CW senses of rotation. The phage needs translocation speeds of this magnitude in order to move along a flagellar filament of a few µm long within a timescale of 1 s, which is approximately the CCW time interval [27]. Furthermore, the two limits of long- and short-tailed phages clarify that the physical requirements for translocation along the flagellar filament are the grip from the part of the phage that is wrapped around the flagellar filament, in a helical shape (indeed as dictated by the shape of the grooves), combined with torque provided by the parts of the phage sticking out in the bulk, away from the flagellar filament. The plots for the translocation speed as a function of the phage tail length approach the asymptotic approximations for long- and short- tailed phages at large and vanishing tail lengths respectively in both models. 17 The important point where the two models deviate from each other is their opposite predictions for the translocation speed as a function of the phage tail length and the phage fibre length. We conjecture that the second model with its explicit inclusion of the grooves should be closer to the real-life situation. According to it, U increases when Lt increases because the propulsive terms in the axial torque on the tail increase with Lt. In contrast, the decreasing behaviour of U with Lf ib is caused by the resistive part of the force exerted by the motion of the fibres in the grooves that increases with Lf ib, and thus slows down the motion. Having modelled phage translocation along straight flagellar filaments of mutant bacteria, the next natural step will be to model phage translocation along the naturally helical flagellar filaments of wild type bacteria. In this case, the geometry is more complicated as it involves motion along helical grooves on top of a flagellar filament whose centreline is also a helix. This is a more complicated system geometrically, as the helical fibres are sliding along a helical flagellar filament with a spatially-varying local tangent and therefore we expect that numerical computations would be required in order to tackle it. Notably, there will be an additional hydrodynamic drag on the phage due to the rotation and translation of the flagellar filament. The hydrodynamic drag from the translation will oppose or enhance the translocation of the phage towards the cell body depending on the chirality of the flagellar filament. This opens up the possibility of a competition between the nut-and-bolt translocation effect and the possibly opposing drag due to translation, which will vary with the helical angle of the flagellar filament. Different regimes are expected to arise as the helical angle of the flagellar filament is increased from 0, hinting to a rich, nonlinear behaviour to be investigated with the aid of numerical simulations. In this work, we focus on the translocation of the phage once it has reached a steady, post-wrapping state, and thus assuming that it is moving rigidly. Future studies could address the transient period of wrapping, where the length of the fibres wrapped around the filament is increasing and the 'grip' is possibly becoming tighter. Some phages (in the Siphoviridae family) have long, flexible tails, thereby requiring the addition of the elasticity of the tail and fibres into the model. We hope that the modelling developed in this paper will motivate not only further theoretical studies along those lines but also more experimental work clarifying the processes involved in the wrapping and motion of the fibre in the grooves. This work was funded by the EPSRC (PK) and by the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation programme (grant agreement 682754 to EL). ACKNOWLEDGEMENTS [1] G. P. C. Salmond and P. C. Fineran, "A century of the phage: past, present and future," Nat. Rev. Microbiol. 13, 777 -- 786 (2015). [2] O. Bergh, K. Y. Borsheim, G. Bratbak, and M. Heldal, "High abundance of viruses found in aquatic environments." Nature 340, 467 -- 468 (1989). [3] K. E. Wommack and R. R. Colwell, "Virioplankton: viruses in aquatic ecosystems." Microbiol. Mol. Biol. Rev. 64, 69 -- 114 (2000). [4] H. Brussow and R. W. Hendrix, "Phage genomics: small is beautiful." Cell 108, 13 -- 16 (2002). [5] S. W. Wilhelm, W. H. Jeffrey, C. A. Suttle, and D. L. Mitchell, "Estimation of biologically damaging uv levels in marine surface waters with DNA and viral dosimeters." Photochem. Photobiol. 76, 268 -- 273 (2002). [6] R. W. Hendrix, "Bacteriophage genomics." Curr. Opin. Microbiol. 6, 506 -- 511 (2003). [7] E. Hambly and C. A. Suttle, "The viriosphere, diversity, and genetic exchange within phage communities." Curr. Opin. Microbiol. 8, 444 -- 450 (2005). [8] C. A. Suttle, "Viruses in the sea." Nature 437, 356 -- 361 (2005). [9] A. D. Hershey and M. Chase, "Independent functions of viral protein and nucleic acid in growth of bacteriophage." J. Gen. Physiol. 36, 39 -- 56 (1952). [10] J. M. A. Blair, M. A. Webber, A. J. Baylay, D. O. Ogbolu, and L. J. V. Piddock, "Molecular mechanisms of antibiotic resistance," Nat. Rev. Microbiol. 13, 42 -- 51 (2015). [11] C. A. Suttle, "Marine viruses -- major players in the global ecosystem." Nat. Rev. Microbiol. 5, 801 -- 812 (2007). [12] C. P. Brussaard, S. W. Wilhelm, F. Thingstad, M. G. Weinbauer, G. Bratbak, M. Heldal, S. A. Kimmance, M. Middelboe, K. Nagasaki, J. H. Paul, D. C. Schroeder, C. A. Suttle, D. Vaque, and K. E. Wommack, "Global-scale processes with a nanoscale drive: the role of marine viruses." ISME J. 2, 575 -- 578 (2008). [13] W. C. Summers, "Bacteriophage therapy," Annu. Rev. Microbiol. 55, 437 -- 451 (2001). [14] S. J. Labrie, J. E. Samson, and S. Moineau, "Bacteriophage resistance mechanisms," Nat. Rev. Microbiol. 8, 317 -- 327 (2010). 18 [15] D. V. Rakhuba, E. I. Kolomiets, and E. S. Deyand G. I. Novik, "Bacteriophage receptors, mechanisms of phage adsorption and penetration into host cell." Pol. J. Microbiol. 59, 145 -- 155 (2010). [16] E. Kutter and A. Sulakvelidze, Bacteriophages: Biology and Applications (CRC Press, Boca Raton, Florida, 2004). [17] P. Grayson and I. J. Molineux, "Is phage DNA 'injected' into cells - biologists and physicists can agree," Curr. Opin. Microbiol. 10, 401 -- 409 (2007). [18] I. J. Molineux and D. Panja, "Popping the cork: mechanisms of phage genome ejection," Nat. Rev. Microbiol. 11, 194 -- 204 (2013). [19] H. Zhang, L. Li, Z. Zhao, D. Peng, and X. Zhou, "Polar flagella rotation in vibrio parahaemolyticus confers resistance to bacteriophage infection," Sci. Rep. 6, 26147 (2016). [20] L. M. Raimondo, N. P. Lundh, and R. J. Martinez, "Primary adsorption site of phage pbs1: the flagellum of bacillus subtilis," J. Virol. 2, 256 -- 264 (1968). [21] Eiserling FA., "The structure of bacillus subtilis bacteriophage pbs 1." J Ultrastruct Res. 17, 342 -- 347 (1967). [22] S. Z. Schade, J. Adler, and H. Ris, "How bacteriophage χ attacks motile bacteria," J. Virol. 1, 599 -- 609 (1967). [23] H. C. Berg and R. A. Anderson, "Bacteria swim by rotating their flagellar filaments," Nature 245, 380 -- 382 (1973). [24] E. J. O'Brien and P. M. Bennett, "Structure of straight flagella from a mutant salmonella," J. Mol. Biol. 70, 145 -- 152 (1972). [25] K. Namba and F. Vonderviszt, "Molecular architecture of bacterial flagellum," Q. Rev. Biophys. 30, 1 -- 65 (1997). [26] T. Iino and M. Mitani, "A mutant of salmonella possessing straight flagella," J. Gen. Microbiol. 49, 81 -- 88 (1967). [27] A. D. T. Samuel, T. P. Pitta, W. S. Ryu, P. N. Danese, and H. C. Berg, "Flagellar determinants of bacterial sensitivity to χ-phage," Proc. Natl. Acad. Sci. U. S. A. 96, 9863 -- 9866 (1999). [28] R. C. Guerrero-Ferreira, P. H. Viollier, B. Elyt, J. S. Poindexter, M. Georgieva, G. J. Jensen, and E. R. Wright, "Alternative mechanism for bacteriophage adsorption to the motile bacterium caulobacter crescentus," Proc. Natl. Acad. Sci. U. S. A. 108, 9963 -- 9968 (2011). [29] Vidakovic L, Singh PK, Hartmann R, Nadell CD, and Drescher K., "Dynamic biofilm architecture confers individual and collective mechanisms of viral protection," Nat. Microbiol. 3, 26 -- 31 (2018). [30] J. Gray and G. J. Hancock, "The propulsion of sea-urchin spermatozoa," J. Exp. Biol. 32, 802 -- 814 (1955). [31] E. Lauga and T. R. Powers, "The hydrodynamics of swimming microorganisms," Rep. Prog. Phys. 72, 096601 (2009). [32] L. G. Leal, Advanced Transport Phenomena: Fluid Mechanics and Convective Transport Processes (Cambridge University Press, Cambridge, UK, 2007). [33] G. J. Hancock, "The self-propulsion of microscopic organisms through liquids," Proc. R. Soc. Lond. A 217, 96 -- 121 (1953). [34] C. Brennen and H. Winet, "Fluid mechanics of propulsion by cilia and flagella," Annu. Rev. Fluid Mech. 9, 339 -- 398 (1977). [35] "See supplemental material at [url will be inserted by publisher] for [give brief description of material]." . [36] S. Chattopadhyay, R. Moldovan, C. Yeung, and X. L. Wu, "Swimming efficiency of bacterium escherichia coli," Proc. Natl. Acad. Sci. U. S. A. 103, 13712 -- 13717 (2006).
1606.06850
2
1606
2016-11-24T05:59:57
Online games: a novel approach to explore how partial information influences human random searches
[ "physics.bio-ph", "cond-mat.stat-mech", "physics.soc-ph", "q-bio.PE" ]
Many natural processes rely on optimizing the success ratio of a search process. We use an experimental setup consisting of a simple online game in which players have to find a target hidden on a board, to investigate the how the rounds are influenced by the detection of cues. We focus on the search duration and the statistics of the trajectories traced on the board. The experimental data are explained by a family of random-walk-based models and probabilistic analytical approximations. If no initial information is given to the players, the search is optimized for cues that cover an intermediate spatial scale. In addition, initial information about the extension of the cues results, in general, in faster searches. Finally, strategies used by informed players turn into non-stationary processes in which the length of each displacement evolves to show a well-defined characteristic scale that is not found in non-informed searches.
physics.bio-ph
physics
Online games: a novel approach to explore how partial information influences human random searches Ricardo Mart´ınez-Garc´ıa,1, ∗ Justin M. Calabrese,2, 3 and Crist´obal L´opez4 1Department of Ecology and Evolutionary Biology, Princeton University, Princeton, NJ 08544, USA 2Smithsonian Conservation Biology Institute, National Zoological Park, Front Royal, VA 22630, USA 3Department of Biology, University of Maryland, College Park, MD 20742, USA 4IFISC, Instituto de F´ısica Interdisciplinar y Sistemas Complejos (CSIC-UIB), E-07122 Palma de Mallorca, Spain. (Dated: September 6, 2018) Many natural processes rely on optimizing the success ratio of a search process. We use an experimental setup consisting of a simple online game in which players have to find a target hidden on a board, to investigate the how the rounds are influenced by the detection of cues. We focus on the search duration and the statistics of the trajectories traced on the board. The experimental data are explained by a family of random-walk-based models and probabilistic analytical approximations. If no initial information is given to the players, the search is optimized for cues that cover an intermediate spatial scale. In addition, initial information about the extension of the cues results, in general, in faster searches. Finally, strategies used by informed players turn into non-stationary processes in which the length of each displacement evolves to show a well-defined characteristic scale that is not found in non-informed searches. Introduction The problem of searching for targets whose location is unknown arises in many fields and at different scales [1 -- 3]. Numerous examples appear in the natural sciences including in ecology [4 -- 8], biochemistry [9 -- 11] and chemistry [12]. In addition, many human activities involve situations where a target has to be found. Some instances are the location of a lost object, rescue operations, or fugitive prosecutions [13]. More recently, the development of eye-tracking technology has allowed the study of visual searches on screens [14 -- 16]. In order to understand the social, biological and physical mechanisms behind these processes, it is essential to have empirical evidence of the performance of different strategies and how they are affected by environmental cues, regardless of whether they are employed by humans, animals or bacteria [17]. Such data are also required to verify the mathematical models that have been proposed [18 -- 25], and to develop improved protocols. Situations in which a target has to be located appear in a large variety of scenarios, which allows the design of multiple strategies to find a successful solution. Such strategies can be classified in many different ways, according to one or more of their properties [1]. For instance, stochastic or systematic processes are distinguished depending on the type of search rule [2] and the amount of directional information available determines the existence of bias towards preferred regions [26, 27]. Finally, differences may also be attributable to the movement pattern, such as cruising versus ambush [28] and to the frequency of the reorientation events, such as intensive (frequent) versus extensive (infrequent) [29, 30]. The effectiveness of a particular choice within each category is determined by the properties and the state of the searcher, the target, and the environment where the task has to be accomplished. For instance, searchers with memory that navigate relatively predictable environments do not employ purely random strategies but combine a stochastic component with knowledge acquired through previous experience. There is therefore a learning process that plays an important role in the emergence of new rules [31, 32]. In other scenarios, individuals who live in groups may incorporate information gathered by conspecifics with their own in order to improve foraging efficiency. It has been recently showed that intermediate combinations between both types of cues result in more efficient searches regardless of the nature of the mobility pattern [33] and the spatial distribution of the targets [34, 35]. However, the precise optimal balance between social and individual information is determined by each specific setup. In all of these scenarios, interactions with the environment provide the searcher with information that may alter the effectiveness of a given strategy over the course of the search. Therefore, in the most general case, search strategies must be interpreted as dynamical processes consisting of several components rather than fixed procedures. For instance, many predators respond to the detection of cues indicating the proximity of prey by increasing their turning angles and reducing their speed in order to scan the local environment more carefully [36, 37], which leads ∗Electronic address: [email protected] 2 to concentration of the search activity in areas of high prey density [38]. This behavior has been reported in several species of insects [39], seabirds [40, 41] and also in human searchers looking for hidden resources in open environments [42]. Other phenomena that trigger sudden changes in individual movement behavior are changes between habitats [43] and changes in the amount and quality of information gathered by the searcher [7]. In this work we propose the use of computer games as a new experimental approach with which it is possible to address these and related questions in humans. This is particularly intriguing since, due to their cognitive abilities, individuals might show a large diversity of complex responses to the same stimulus. Despite substantial efforts aimed at understanding the theoretical concepts behind many search processes, a reliable and unifying empirical framework in which these ideas may be tested is still lacking. The family of games presented here is a good candidate to fill this gap, as they can be accessed online by a large number of players. This results in the generation of large and clean datasets. In addition, the rules and setup of the game can be experimentally manipulated so that different mechanisms or strategies can be rigorously tested. Firsy, we address several questions related to search efficiency and investigate how the strategies change due to the amount and the quality of information acquired by the player at different stages of the game. In a second step, the main features of these patterns are extracted from the data and used to develop a family of random walk models that can be applied to predict human search behavior in other configurations of the game. The variety of experiments shown in this work reinforces the flexibility of our approach and aims to open a new route for the study of search problems. In the following section, after presenting the characteristics of the game, we show the empirical results obtained from two different setups. In the first case, players have no information about the configuration of the board, whereas in the second study they are provided with partial information about it. Then, we formulate a family of models that capture the main mechanisms behind the experimental results and derive analytical approximations to show the robustness of the results. Finally, all the previous steps are combined to develop a comprehensive framework in which it is possible to predict the optimal configuration of the landscape that yields faster searches. The paper finishes with a discussion of the results and opportunities for new lines of research. Results Experimental setup We consider a simple game in which a single target has to be found. It slightly resembles the classic minesweeper, although the objective is to find a unique target (mine) instead of avoiding a collection of them. The interface consists of N × N squares that can be explored by the player through successive clicks with the mouse. There are three classes of cells depending on their color after being clicked (unclicked cells are always blue): (i) black cells are typically far from the target, (ii) yellow cells indicate that the target may be one of the neighboring cells and (iii) the single red cell is the target. The target is randomly located within a patch of yellow cells. Therefore, it provides partial information about the configuration of the board. Two different geometries for this set of yellow cells are explored here. First, in the next two sections they form a Ny × Ny neighborhood square region (Fig. 1a). Second, in the last section of the Results they will outline a random patch whose size will be measured in terms of the number of yellow cells. Further details about the implementation of these random neighborhoods will be provided in that section. The discovery of a yellow cell indicates that the player is in the neighborhood of the target and thus reduces the area that needs to be scanned. For simplicity we fixed N = 20 in all the experiments and then manipulated Ny. To generate the dataset players access anonymously the game online and are asked to find the target using as few clicks (jumps on the board) as possible. Since players are not identified separately, we cannot identify the number of rounds played by each user. The rounds are all independent (different configurations of the board) and each one is represented by the trajectory traced by the player on the board. Finally, the experimental setup also includes a timer. In order to study the limit in which searches are more stochastic, players are requested to find the target as quickly as possible. This constraint also mimics many real situations both for humans and non-humans in which time is a limitation for the search. Some instances are human rescue operations or animal foraging while avoiding predators. In the following sections we investigate i) how the duration of the search, represented by the number of mouse clicks, changes with the size of the target's neighborhood (also called yellow region); and ii) the statistical properties of the searching patterns as defined by the distance between clicks di (jump length) and the turn angles θi. By definition, we consider turns to the left to be between 0◦ and 180◦ and turns to the right to be between 180◦ and 360◦ (see Fig. 1b for a definition of both quantities). We consider two classes of experiments: a) blind searches, where the player is given no a priori knowledge of the size of the neighborhood, and b) searches with initial information, where the value of Ny is given to the player at the beginning of the round. The objective of performing both classes of experiments is twofold: on the one hand to investigate whether players adapt their searching strategies when they have better information about the landscape and, on the other hand, to examine how search efficiency changes when 3 the reliability of the information provided by the yellow cells increases. Experiments with blind searchers and square neighborhoods For this first series of experiments neither the exact size, the position of the yellow region, nor a range of possible dimensions was given to the searchers. Before starting the round, each player only knew that a target (red square) was hidden in the board and it might be randomly placed inside a square vicinity of yellow cells of unknown size. The uncertainty in the size of the neighborhood reduces the reliability of the information acquired by the player when a yellow cell is clicked and favors the efficiency of random strategies [1]. Our dataset consists of 500 rounds with Ny ranging from Ny = 1, which means that the target does not have a neighborhood, to Ny = 13. A distribution of the number of rounds for each value of Ny is shown in the Supplementary Table I. We first measure the mean number of clicks needed to find the target as a function of the lateral length of its yellow neighborhood (black squares in Fig. 2a). Due to the design of the experiments, there is a tradeoff between finding the yellow region and finding the target inside it. Larger neighborhoods are easier to locate but make the final detection of the target inside them harder. Smaller neighborhoods, however, need on average more steps to be found but make the target within them easier to locate (Fig. 2b). According to our results, this tradeoff is balanced at intermediate sizes of the neighborhood, N opt y = 5. This resembles the foraging dynamics of animals that exchange information about food location with their conspecifics, so that both spreading information over distances that are either too large or too short may slow down the search [35]. Following this analogy, we refer to the the size of the yellow area that minimizes the number of clicks needed to find the target as the optimal interaction range. The standard deviation of the number of jumps is also minimal at the optimal range, which means a narrowing in the distribution of clicks used to detect the target and therefore a reduction in the stochasticity of the search. In the limit of zero information (i.e. no yellow cells or N y = 1, or the whole board is yellow, N y = N ), the probability of finding the target on the first click is given by the inverse of the number of available cells, 1/N 2. In any subsequent movement, m, this probability is given by Pm = N 2 − (m − 1) N 2 × 1 N 2 − (m − 1) , (1) where the first term yields the probability of not having found the target in the previous m − 1 clicks and the second term yields the probability of hitting the target once m− 1 squares have been visited. Equation (1) reduces to 1/N 2 regardless of the value of m. Therefore, the probability of detection in the limit Ny = 1 (and Ny = N ) follows a uniform distribution of mean N 2/2 and standard deviation N 2/√12, which is in good agreement with data (black squares in Fig. 2). Next, we analyze all the trajectories traced by the players in every round. To facilitate this, the experimental setup saves the sequence of clicks in each round, from which we calculate the length of each displacement and the angle of each turn. We identify extensive and intensive searching modes that depend on whether the player has detected a yellow cell or not respectively (Fig. 3a). In both situations the jump lengths can be fitted using exponential distributions, with the intensive phase showing a lower mean value 1/λint = 2.04 and 1/λext = 3.70 (1/λint, 1/λext are the mean length of the displacements in the intensive and the extensive phase respectively). Therefore, the typical size of the jumps is reduced once the player finds the yellow area as the detection of the cue (represented by a yellow cell) triggers an area-restricted search [36, 37]. Although the player does not know how big the neighborhood is and therefore how reliable the information is, the trajectories recorded after the discovery of the yellow region still show shorter distances between turns, suggesting that players switch to an intensive search mode once they find the yellow region [1]. It is important to remark that, although alternation between extensive (motion phase) an intensive modes (scanning phase) is also characteristic of intermittent searches, the player is not performing an intermittent search as it has been defined in the literature before [2]. The differences lie in two points. First, in our study the switch between reorientation modes is triggered by the external cue instead of taking place at random. Second, the detection of the target may take place in both phases instead of being limited to the extensive one. Regarding to the type of motion, we study, however, a spatially intermittent search since the player performs a saltatory trajectory in which the target can be found only if the searcher lands on it. This differs from the case of a cruise forager who looks for targets while moving and that would constitute a completely different study. Regarding the turn angles, both the extensive (before the first encounter with a yellow cell) and the intensive phases (after detecting the first yellow cell) show correlations between subsequent turn angles (Fig. 3b,c, respectively). This could indicate that the strategies are not completely random but contain some systematic features. In fact, a frequent strategy consists of tracing a series of short jumps in the same direction. To reduce searching times players show a tendency to scan a direction doing several consecutive clicks. This behavior is also seen in the distributions of jump lengths, since they show a large deviation from the exponential for one-cell length jumps, which are overrepresented 4 in the dataset (Figure 3a). The higher frequency of turning angles closer to zero is linked to the higher presence of jumps of length one. The explanation for this persistence in the direction of movement shown in Fig. 3b,c is probably a combination between the attempt of some players to design purely systematic strategies and the intrinsic tendency of humans to keep visually scanning in the same direction [16]. As an exception, movements done immediately after a yellow-to-black transition show a strong tendency to reverse the direction, as this sequence in the colors of the cells indicates that the player is moving away from the target (Fig. 3d). Experiments with initial information and square neighborhoods. The case of Ny = 5 as compared to the blind case In this second series of experiments the players know the size of the yellow region, which is fixed at the optimal interaction range Ny = 5. This increases the quality of the information obtained when one of the yellow cells is found as the player can limit the search area. The position of this area, as well as the location of the target inside it, is random, changes from round to round and is unknown to the player. Data from 230 rounds were collected. As a general result, a priori information accelerates the search and reduces its stochasticity. Blind searchers need on average 31.30 clicks to find the target when Ny = 5 (subset of 65 rounds from the 500 trajectories analyzed for blind experiments, see Supplementary Table I ), while informed players use 25.5 clicks. The two-tailed P value on the difference of these mean values obtained using an unpaired t-test, 3× 10−4, is highly statistically significant. The standard deviation also decreases, indicating a narrowing in the distribution of the number of displacements and therefore in the randomness of the process: σb = 14.10 for blind searchers and σi = 10.50 for the informed ones. To find out what stage of the search is more strongly affected by the initial information, we analyze the number of clicks done in each phase of the search. We repeat this factorization for the blind and the informed cases and compare both of them (Fig. 4). From Fig. 4b and 4c, we observe that all of the reduction in the number of jumps accumulates in the intensive phase, while the extensive stage remains unaltered by the initial information. More interestingly, if the number of displacements that take place between two yellow cells is subtracted from the total number of jumps of the intensive phase (Fig. 4d), we observe that this quantity remains almost unchanged. There is, however, an important reduction in the number of displacements that correspond to the rest of the combinations of cells (black to yellow, yellow to black and black to black jumps; Fig. 4e). In fact, the percentage of yellow-to-yellow movements that take place during the intensive phase increases from 54% to a 75% in the informed searches. This result indicates that having information about the size of the yellow zone allows a faster detection of its limits and therefore reduces the number of movements spent to find the target. Regarding to the statistical analysis of the trajectories, initial information about Ny also yields some differences in the distributions of the lengths of the jumps and the turning angles. Informed players adapt their displacements during the extensive phase, concentrating the length of their movements around the size of the yellow neighborhood, Ny = 5 (Fig. 5a). If we analyze the whole set of informed rounds, we observe a strong dominance of movements of length Ny = 5 (green squares in Figure 5a). This is due to the presence of approximately 50 rounds in the dataset where players performed optimally designed systematic strategies that consist of moving in jumps of fixed length Ny during the extensive phase. We will come back to this in the description of a random walk based model for this process. For the purposes of this section we will remove these systematic rounds and focus on the subset of stochastic strategies formed by the other 180 rounds. The distribution of the length of the displacements is still dominated by jumps that cover a distance of the order of Ny (red circles in Fig. 5a). For the subset of blind searchers with Ny = 5, however, this distribution does not show a well defined typical scale and instead, players explore several scales as they look for the yellow region (Fig. 5b). For the intensive phase, informed searches also show a higher abundance of one-cell displacements than the distribution of the blind searches (inset of Fig. 5a and 5b respectively). This result is independent of whether or not the systematic deterministic strategies are included within the analyzed dataset and is due to the fact that knowing the neighborhood size reduces exploration during this phase. Finally, giving the size of the yellow neighborhood to the players in advance also has an effect on the distribution of turns made by the searcher immediately after a yellow-to-black displacement. This distribution is shown in Fig. 5c for informed strategies and in Fig. 5d for blind searches. Although in both cases the movement shows a strong bias backwards, informed searches result in distributions with a stronger peak around θ = 180◦. This is due to the fact that players do not have to find out the size of the neighborhood of the target and consistent with the factorization of the number of clicks shown in Fig. 4 We conclude this section with an analysis of the trajectories during the extensive phase, in order to find the mechanism by which a characteristic length scale appears in the jump length distribution. We find the existence a feedback between the searcher and the environment that makes the extensive phase non-stationary (the mean value of the distribution changes with time). This feedback allows a progressive narrowing of the jump length distribution 5 around Ny as the extensive phase evolves and the searcher gathers and accumulates information from the landscape. Since the player has perfect memory about his trajectory (visited cells remain open), trajectories that start with large displacements tend to create landscapes that are fragmented in patches of length Ny in which long movements are inefficient. To show the existence of this feedback we split the data of the extensive phase in four subdivisions: (i) from jump 1 to 5, (ii) 6 to 10, (iii) 11 to 15, and (iv) 16 to the end. The distributions for each of these pieces are shown in Fig. 6a, 6b, 6c and 6d respectively, and they can be fitted by a family of gamma distributions (dashed lines in each panel) of decreasing mean, mode and variance (See Table I for numerical values of these parameters and details of the distributions). Then the total distribution of Fig. 5a can be approximated by a gamma function defined in terms of the parameters of the distributions of the pieces (dashed line in Fig. 5a). This approach shows an excellent agreement with a direct fitting of the whole extensive phase (full line in Fig. 5a). At this point, we have shown the existence of an optimal size for the neighborhood of the target, as well as an improvement in the search efficiency when the value of Ny is revealed at the beginning of the round. In addition, these informed strategies evolve through information gathering during the extensive phase towards a dominant jumping distance equal to the lateral length of the neighborhood of the target, Ny. In the following sections we develop a theoretical framework and a family of models based on random walks to study the basic principles behind these results and how they can be transfered to more general scenarios, with irregular shapes for the information region. Model for blind searchers: numerical simulations and analytical approximation We develop a minimalistic searching model based on random walks to explain previous experimental results on the basis of simple dynamical rules. The model has the three main ingredients obtained from the data analysis: (i) two modes of movement defined by the mean length of the displacements: 1/λint and 1/λext; (ii) in the absence of any information (no yellow cell clicked) the direction is completely random (uniform distribution in the turning angles); and (iii) when cues are obtained (a yellow cell has been detected), the searcher has a bias towards unvisited cells surrounding a yellow one. The choice of a uniform distribution for the turning angles is a consequence of using a purely exponential distribution for the length of the displacements (see Fig. 3a). The high persistence shown by the experimental turning angle distribution, which can be approximated by a uniform distribution except for that peak at θ = 0 (see Fig. 3b,c), comes from the high presence of jumps of length one. Disregarding the high frequency of unity-length movements also implies disregarding the higher abundance of turning angles close to zero and therefore using a uniform distribution for θ. The third assumption aims to capture the influence of the information provided to the searcher when a yellow cell is found, as well as the strong tendency to go back to yellow cells exhibited by the distribution of turning angles in Fig. 3d. The results of the simulations (green curve in Fig. 2) show an excellent agreement with the experimental data (black curve) both in the mean average number of jumps and in its standard deviation. Simulations reproduce at least the two first moments of the number of clicks distribution. Except in the limits Ny = 1 (no yellow cells) and Ny = N (yellow cells occupy the whole board), it is hard to obtain exact analytical expressions for the average total number of jumps needed to find the target. However, it is possible to obtain the distribution for the length of the extensive phase Pi(Ny) = pi(Ny) i−1 Y j=1 (1 − pj(Ny)) , (2) where Pi(Ny) is the probability of having an extensive phase of i jumps when the neighborhood of the target has a lateral length Ny and pi(Ny) = Ny N −i+1 is the probability of finding a yellow cell in the i − th mouse click. In words, the probability of having an extensive phase with i jumps is given by the probability of not finding a yellow cell in all the previous movements multiplied by the probability of finding one in the i − th movement. Given Eq. (2), the mean length of the extensive phase is Mext = N 2 y +1 −N 2 X i=1 iPi(Ny). (3) For the length of the intensive phase however we can only give and upper and a lower limit, assuming that after the detection of the first yellow cell all the movements are to neighboring cells. Therefore, the target is found on average after N 2 y /2 jumps in the intensive phase when the neighborhood of the target is large and after (Ny + 2)2/2 movements when the neighborhood is small. These two limits account for the decreasing probability of visiting cells outside the neighborhood when increasing its size. For small values of Ny it is very likely to reach the border of the neighborhood before detecting the target and thus to return to the black region. Combining these two results for the intensive phase with the length of the extensive phase obtained in Eq. 3, we obtain two theoretical approximations to the total number of clicks 6 M up = N 2 y +1 −N 2 X iPi + (Ny + 2)2 2 , M low = N 2 i=1 −N 2 X y +1 i=1 iPi + y N 2 2 . (4) (5) The combination of these two expressions gives an approximated range for the length of the search (magenta region in Fig. 2) that shows an excellent agreement with empirical data and numerical simulations of the model. Model for searches with initial information. The design of optimal strategies. Knowing the size of the yellow region at the beginning of the game changes the nature of the search as the information gathered by the player with each movement may be used to design the next displacement. This reinforces the non-Markovian nature of the informed search process as the player uses all the previous steps to discard cells that have not been visited yet and results in self-adaptive strategies that evolve towards displacements of length Ny. Also, as the value of Ny is known, the number of exits from the neighborhood of the target diminishes (Fig. 4). In a first approach to model this effect, we modify the model used for blind searches using the new experimental distribution of the length of the displacements in both the extensive and the intensive modes (Fig. 5a). Therefore, instead of using the exponential distributions of Fig. 3 we sample the histograms of Fig. 5a (red circles) and its inset, that are obtained from experimental searches with initial information. This approach overestimates both the length of the extensive and the intensive phases, which results in a clearly higher average number of movements; 33.50 jumps, σ = 19.00 for the model and 25.50 jumps, σ = 10.50 in the data (MB green bars and DI gray bars in Fig. 7 respectively). This is due to the fact that the model does not integrate the information about the size of the target to a priori discard some of the cells during the intensive and the extensive phase. In a first approach to remove this discrepancy, we hypothesize that the most important differences arise in the modeling of the intensive phase. During this stage, given a certain number of yellow cells and some of their neighboring black squares, our experimental results suggest that human players are able to discriminate the real border of the neighborhood of the target and thus reduce the number of erroneous displacements. The model that we developed for blind searches lacks this ingredient, which increases the duration of the intensive phase since more black cells are open. To correct this, we modify the model and include the effect that previous movements, together with knowing the size of the neighborhood of the target, have on the intensive phase (See Methods for a detailed description). In this new approach, once the first yellow cell has been detected and based on all the previous movements, only those cells that can possibly be part of a 5× 5 yellow square have a non-zero probability of being visited by the searcher. This mechanism reduces the number of times that black cells are visited once a yellow cell has been found as the model is able to discriminate all the possible borders of the neighborhood of the target. With this new ingredient the efficiency of the model increases (MI blue bar in Fig. 7a) and the number of jumps in the intensive phase shows excellent agreement with the experimental data (DI gray and MI blue bars in Fig. 7c). However, despite this substantial improvement as compared to the blind model, significant differences still remain between empirical data and numerical results. The source of this disagreement arises from the extensive phase (DI gray and MI blue bars in Fig. 7b). To correct this, we next modify the extensive phase of the model. During the extensive phase, players are able to discriminate regions where the target cannot be placed as a 5 × 5 square would not fit. To incorporate this in the model, we first compute the probability of jumping to each of the non-visited cells of the board according to the histogram in Fig. 5a. Then, for each cell we obtain all the possible squares of lateral length 5 to which it could belong and set the probability of jumping to that cell to zero if all these squares contain at least one open black cell (See Methods for more details). With this mechanism the extensive phase becomes more efficient and the agreement of the model with the experimental data is excellent. More importantly, this comes from a precise fitting of both the intensive and the extensive phase individually (DI gray and MII black bars in Fig. 7a,b,c). Optimal strategy.- However, both actual player strategies and random walk models are much less efficient than entirely systematic protocols. Knowing a typical size of the target in advance allows the design of optimized strategies 7 that minimize the number of incorrect steps. Particularly important is to shorten the extensive phase, as within the neighborhood of the target all the cells are equivalent and it is equally likely to find the target in any position. In fact, during the experimental rounds with initial information, one of the players developed one of these searching methods by repeatedly playing with the same size of Ny = 5. This strategy optimizes the extensive phase and only allows for two yellow-to-black transitions during the intensive phase (Fig. 8a). Given a value for Ny, the search rule is given the following steps: 1. Divide the board in theoretical squares of size Ny × Ny (see Fig. 8a) 2. Click in the upper right corner of each subdivision. Start with those squares whose upper right cell has more neighbors and continue with those in the borders. This reduces the length of the extensive phase on average as corners that are farther from the border are more likely to contain a yellow cell. 3. Once a yellow cell is found, visit consecutive squares in a given direction (horizontal in Fig. 8a for Ny = 5) until finding a black position. Then, if the number of yellow cells in the row is lower than Ny, complete it. 4. Repeat the same operation in the other direction starting from one already known yellow cell. 5. Once the neighborhood of the target has been delimited, move inside it until finding the target. In the particular case of Ny = 5, the average number of movements before target detection following this strategy is 19.03 (104 realizations) and it is always lower than 42. In addition the extensive phase has a duration of 5.90 clicks on average, which is about 50% lower than the experimental result. This improvement is much higher than the one observed for the intensive phase, which can be optimized by players once they are provided with initial informaition about the landscape (see gray and red bars in Fig. 7b, c for a comparison). In real human scenarios, this result suggests that efforts put into optimizing the extensive phase may pay off more than equivalent efforts to optimize the intensive phase. Applying this optimal strategy to many sizes of the yellow region (Fig. 8b) we observe that the tradeoff between finding the neighborhood of the target (yellow diamonds in Fig. 8b) and finding the target inside it (blue circles in Fig. 8b) balances at intermediate values of Ny. Following theoretical results for blind experiments, analytical expressions can be obtained for the mean number of movements during both phases and therefore for the optimal interaction range. The mean number of clicks during the intensive phase is N 2 y /2 as the target can be in any cell with the same probability (green dashed line in Fig. 8b) (we only consider the lower bound obtained for blind experiments since this optimal protocol minimizes the number of erroneous movements). To obtain the mean number of movements in the extensive stage, we assume that the upper right corner of each subdivision of the board (Fig. 8a) is equally likely to have a yellow cell. Therefore, the number of steps is given by N 2/2N 2 y . This is not completely true, as cells close to the border have a lower probability of being yellow, but it is a good approximation (black dashed line in Fig. 8b y = √N = 4.47 fitting yellow diamonds). At the optimal interaction range both functions intersect, which gives N opt for our experimental setup with N = 20. This result is in excellent agreement with the value obtained from the experiments (Fig. 8c) and suggests, together with the theoretical approximation, that the optimal interaction range is independent of the searching strategy. This result suggests the possibility of using this theoretical framework to predict the optimal size of the neighborhood of the target in more general scenarios. Anticipating the optimal range of interaction for random neighborhoods. In this section we allow the target to adopt different sizes and random shapes across rounds. In order to facilitate the formulation of theoretical predictions, the neighborhood is built starting from a triangle of varying base by (see Methods for a detailed description and Fig. 9) where the target is embedded. Then, the region is randomized by turning 30% of the cells black. In this way, we implement random neighborhoods that vary in form and size from round to round but with an underlying fixed pattern. Before starting the game, players know that the neighborhood has now a varying form and size (Fig. 9), but they are given no information about the way it is constructed. The optimal interaction range can be evaluated from an independent estimation of the number of movements needed in the extensive and the intensive phases. The length of the extensive phase is obtained following the same steps used for square neighborhoods; the probability of finding a yellow cell in the i − th movement is given by Eq. (2) from where the mean length of the extensive phase is obtained using Eq.3. This quantity is shown by the magenta circles in Fig. 10b. To approximate the number of movements used in the intensive phase, which will give us the optimal interaction range we used the underlying triangle shape of the neighborhood of the target. This calculation provides lower and upper bounds for the average duration of the intensive phase. The lower bound is obtained assuming that all the cells from the original target have the same probability of being visited but all the cells that do not belong to 8 it will never be clicked. The total number of cells that form this original triangle is (by/2 + 0.5)2 and since all the cells can be visited with the same probability, the lower limit for the length of the intensive phase is given by (by/2+0.5)2 . The upper limit is obtained assuming that the first cells that do not belong to the triangle in each direction also have a non-zero probability of being visited. This results in an upper bound for the length of the intensive phase given by [(by+4)/2+0.5]2 . Both limits are shown by the magenta circles in Fig. 10c. Finally, the total number of movements, i.e., the sum of the extensive and the intensive phase, is shown by the magenta circles in Fig. 10a, with an estimated optimal neighborhood size in between 18 and 25 yellow cells. It is important to note the difference between the optimal interaction range for random and square neighborhoods, which shows the non triviality of predicting optimal interaction ranges for different geometries. 2 2 We tested these a priori predictions with a series of experiments using an experimental setup with neighborhoods that consist of 5, 16, 33, 55 and 69 cells (plus the target red cell, see Supplementary Table II for a distribution of the number of rounds with each size). 301 rounds were analyzed and the observed mean number of clicks is shown in Fig. 10a. We also split each round into the extensive and intensive phases and the results are shown in panels b and c of Fig. 10. The good agreement between the predicted values and the results obtained with the experiments shows the robustness of the theoretical approach developed in simpler scenarios. Discussion We have developed a novel approach to study human search problems by building a simple game that can be accessed online. This approach facilitates the collection of large and clean experimental datasets. By combining data analysis with probabilistic calculations and numerical simulations of existing and new models, it is possible to obtain a deeper understanding of how humans approach simple search tasks and how their strategies differ from optimal patterns. A comprehensive analysis of the trajectories on the board of the game (length of the displacements and turning angles) shows that players follow strategies consisting of two modes. The detection of cues about the location of the target triggers an area-restricted search mainly characterized by shorter movements on average [36, 37]. In the context of existing studies, these processes are usually modeled by composite random walks that consist of an extensive phase and an intensive one. In the particular instance of animal foraging, the latter is triggered by encountering a food item and is characterized by shorter steps and larger turning angles (relative to the extensive mode) [1, 44 -- 46]. Our findings show that the duration of the search is minimal when the cues extend over intermediate spatial scales as compared to the system size. The tradeoff between locating a cue and finding the target among the cues is balanced, which results in faster searches. Although this result seems to be robust against changes in the total system size, considering larger landscapes could offer a richer phenomenology in the analysis of the trajectories on the board as well as in the features of both phases. In the simplest scenario studied here, in which no information is given about the size of the neighborhood of the target, developing a systematic searching rule as opposed to following a stochastic trajectory does not provide a significant advantage. A systematic scan of the environment usually provides higher efficiencies by minimizing the probability of revisiting a certain region. In this setup, however, cells remain open once they are visited, providing players with a perfect memory about the history of their movements. As a consequence, neither random nor systematic players click more than once on a cell, regions are not revisited, and both protocols offer equivalent results. This scenario however changes when some information about the nature of the target is provided to the players. In that case an optimal systematic strategy can be constructed based on this information. Interestingly, our data show that one of these optimal strategies was developed by a particular player who repeatedly played several rounds in the same landscape. This result opens the door to explore a broad range of questions at the interface between landscape variability, the searcher's memory, and learning abilities, which has recently become an important topic in movement ecology [32]. Most animals do not follow completely random strategies, but combine this stochastic component with spatial memory and learning [31, 47, 48]. To investigate the importance of cognitive skills such as learning or memory in the development of optimal strategies, our approach could easily be extended to allow landscapes where the position of the target exhibits a certain degree of persistence across rounds of the game. In addition, in order to compare how more complex decision-making processes come into play, it would be particularly interesting to compare the results presented here with the outcome of a new round of experiments in which players are not requested to find the target in the shortest possible time. In fact, we have shown that, when they have some knowledge about the landscape (size of the neighborhood of the target), players use the additional information obtained in each movement step to increase search efficiency. In this scenario, the effect of the information gathered during the whole process has to be included in theoretical models to reproduce experimental results. Introducing a more realistic finite memory by allowing clicked cells to revert back to 9 the unclicked state after some time arises as a future line of research. More importantly, however, the excellent agreement between our experimental data and simple theoretical models suggest that this online-game based methodology could be applicable to address more complex scenarios. Energy budget related questions can be addressed by introducing a metabolic cost that penalizes longer movements and evolutionary aspects of search problems may be addressed by allowing pairs of players to compete and selecting those using more efficient strategies. This would mimic environments where different individuals compete for limited resources and could shed some light on the driving forces behind the evolution of optimal searching. The effect of cooperative interactions among players on search efficiency could also be addressed. Many species forage in groups as opposed to individually. The methodology that has been presented here would facilitate, given a certain landscape, exploration of the level of confidence that players place on movements performed by previous participants. Before every movement of the new player, the choice of previous searchers at that same moment can be shown to the new player to investigate whether and how much the current player trusts on previous participants. In addition, if the neighborhood of the target is changed, or multiple targets are included, it would be possible to explore the relationships between use of social information versus personal experience for tasks of increasing difficulty. Finally, in this study we have focused on the case of saltatory searches, in which the target can be detected only if the searcher lands on it. A next step should consider the more general scenario of cruising searches, in which the target can be detected at any point of the displacements [1]. Such setup would provide a higher flexibility in constructing more complex landscapes with different gradients of information that could allow the study of taxis-driven searches. In summary, and in view of the large and exciting range of possibilities for future exploration, we expect that this general framework will complement purely theoretical efforts to unveil the fundamental mechanisms that drive a wide variety of search scenarios. Methods Ethics statement. The anonymity of all the participants was maintained during the whole experimental protocol. Participants accessed the game remotely through internet and non of their personal data was stored. No ethical concerns are involved other than preserving the anonymity of participants. Informed consent was obtained from all subjects. The procedure was checked and approved by the Committee of Ethics in Research of the University of the Balearic Islands, since the game was hosted in the web domain of one of its research institutes, the Institute for Cross- Disciplinary Physics and Complex Systems (IFISC) The experiments were subsequently carried out in accordance with the approved guidelines. Fitting of the partial distributions of displacement lengths to gamma distributions in informed searches We showed that, for informed searches, the length of the displacements when players are given a priori information about the landscape follow a series of gamma distributions whose probability density function is given by f (x; α, β) = β−αe−x/βx−1+α Γ(α) , (6) where α and β are real positive parameters. For known values of α and β, the mean value of the distribution can be obtained as αβ, the variance as αβ2 and the mode (the value that appears most often in the distribution) as β(α− 1). All the parameters shown in Table I were obtained using the maximum likelihood estimation. Results shown in the last row of Table I correspond to a distribution that is a mixture of all four component distributions. Given the mean value and variance of these distributions, we can assume that they all have the same weight in the composition since all the subsets of the trajectory have the same length. The mixed distribution can be obtained as: µmix = σ2 mix = 1 4 1 4 4 X i=1 4 X i=1 µi (cid:0)µ2 i + σi(cid:1) − µ2 mix (7) (8) 10 Implementation of the random walk model for blind searches We have developed a minimalistic model based on composite random walks to understand the basic features of the search strategies used by the players. We initialize the model from a random configuration of the board in which the target is placed in a random position inside a smaller square of lateral length Ny. To mimic the experimental setup, we fix the size of the board so it has 20 cells on each side and explore Ny varying between 1 and 20. The searcher is placed in a random position of the board and the dynamics starts. The algorithm consists of the following steps: 1. Obtain the probability of jumping from the current position, i to the rest of the cells in the board j. This is given by the experimental jump length distributions, so Pij = exp(−λγrij)/λγ, where γ ≡ {in, ext} and rij is the distance between two cells. The two values of λ are obtained from the experimental data and define the extensive and the intensive phase: 1/λint = 2.05 and 1/λext = 3.70. 2. As in the game the player has perfect memory of previous moves, so the probability of jumping to already visited cells is set to zero. 3. If any of the visited cells belongs to the neighborhood of the target (yellow cell), then we multiply the probability of jumping to each of its unvisited neighbors by a bias factor η = 103 whose effect is to keep the searcher around the cues and avoid unrealistic escapes from them. The existence of such a bias is suggested by the distribution of turn angles shown in Fig. 3d that shows a high probability of returning to the yellow region when it is left. Our results are, however, independent of the numerical value of this bias provided that it is strong enough to trap the searcher close to the yellow cells. 4. Renormalize all the jumping probabilities so N 2 P j=1 Pij = 1. 5. Sort a uniform random number u between 0 and 1 and move to a cell k when k P j=1 Pij ≥ u. These steps are repeated until the target is found, then the number of movements is saved and the system restarted for a new realization. Implementation of the random walk model for searches with initial information To introduce the effect of having initial information about the configuration of the landscape (size of the yellow region) we modify the random-walk model presented for blind searches. Simulations are set as in the first model, starting from a 20× 20 cells board where the target is randomly placed inside a square region of lateral length Ny = 5. The position of this region is also random in the board and changes across realizations. The searcher is placed at an initial random position and the dynamics starts. The algorithm has two well differentiated parts for the intensive and the extensive phase: • Extensive phase: 1. Obtain the distance from the current position of the searcher, i, to every other cell in the board, j, and assign a jumping probability, Pij, by taking a random sample from the histogram in Fig. 5a. 2. As in the game the player has perfect memory of previous moves, so the probability of jumping to already visited cells is set to zero. 3. For every cell j in the board obtain all the possible 5 × 5 squares to which it can belong. If all of them have any open black cell, then set the probability of jumping to j to zero. This step is skipped in the intermediate model where only the intensive phase is improved. 4. Renormalize all the jumping probabilities so they sum one. 5. Sort a uniform random number u between 0 and 1 and move to a cell k when k P j=1 Pij ≥ u. • Intensive phase, after the first yellow cell is hit: 1. Obtain all the possible neighborhoods of the target to which the first detected yellow cell can belong. 2. Count the number of open cells of both classes (black and yellow) in each of those possible neighborhoods of the target. 3. Pick those 5 × 5 squares that include all the open yellow cells and none of the black ones. 4. Set the probability of jumping to all other of the rest of the cells of the board to zero. 5. From the histogram in the inset of Fig. 5a, obtain the probability Pij of jumping to the cells that belong to the chosen 5 × 5 squares. 6. Renormalize all the jumping probabilities so they sum one. 11 7. Sort a uniform random number u between 0 and 1 and move to a cell k when k P j=1 Pij ≥ u. [1] M´endez, V., Campos, D. & Bartumeus, F. Random search strategies. In Stochastic Foundations in Movement Ecology, Springer Series in Synergetics, 177 -- 205 (Springer Berlin Heidelberg, 2014). [2] B´enichou, O., Loverdo, C., Moreau, M. & Voituriez, R. Intermittent search strategies. Reviews of Modern Physics 83, 81 -- 129 (2011). [3] Kagan, E. & Ben-Gal, I. Search and Foraging: Individual Motion and Swarm Dynamics (CRC Press, 2015). [4] Viswanathan, G., da Luz, M. G. E., Raposo, E. P. & Stanley, H. E. The physics of foraging: an introduction to random searches and biological encounters (Cambridge University Press, 2011), 1 edn. [5] Viswanathan, G. M. et al. Optimizing the success of random searches. Nature 401, 911 -- 4 (1999). [6] M´endez, V., Campos, D. & Bartumeus, F. Biological searches and random animal motility. In Stochastic Foundations in Movement Ecology, Springer Series in Synergetics, 267 -- 288 (Springer Berlin Heidelberg, 2014). [7] Bartumeus, F., da Luz, M. G. E., Viswanathan, G. M. & Catalan, J. Animal search strategies: a quantitative random-walk analysis. Ecology 86, 3078 -- 3087 (2005). [8] Edwards, A. M. et al. Revisiting l´evy flight search patterns of wandering albatrosses, bumblebees and deer. Nature 449, 1044 -- 1048 (2007). [9] Gorman, J. & Greene, E. C. Visualizing one-dimensional diffusion of proteins along dna. Nature structural & molecular biology 15, 768 -- 774 (2008). [10] Kantsler, V., Dunkel, J., Blayney, M. & Goldstein, R. E. Rheotaxis facilitates upstream navigation of mammalian sperm cells. eLife 3 (2014). [11] Bonnet, I. et al. Sliding and jumping of single ecorv restriction enzymes on non-cognate dna. Nucleic acids research 36, 4118 -- 4127 (2008). [12] Hanggi, P., Talkner, P. & Borkovec, M. Reaction-rate theory: fifty years after kramers. Rev. Mod. Phys. 62, 251 -- 341 (1990). [13] Frost, J. & Stone, L. D. Review of search theory: advances and applications to search and rescue decision support. Tech. Rep., DTIC Document (2001). [14] Najemnik, J. & Geisler, W. S. Optimal eye movement strategies in visual search. Nature 434, 387 -- 391 (2005). [15] Credidio, H. F., Teixeira, E. N., Reis, S. D., Moreira, A. A. & Andrade Jr, J. S. Statistical patterns of visual search for hidden objects. Scientific reports 2 (2012). [16] Amor, T. A., Reis, S. D., Campos, D., Herrmann, H. J. & Andrade Jr, J. S. Persistence in eye movement during visual search. Scientific reports 6 (2016). [17] Levin, S. A. The problem of pattern and scale in ecology: the robert h. macarthur award lecture. Ecology 73, 1943 -- 1967 (1992). [18] Bartumeus, F., Catalan, J., Fulco, U., Lyra, M. & Viswanathan, G. Optimizing the encounter rate in biological interactions: L´evy versus brownian strategies. Physical Review Letters 88, 097901 (2002). [19] Hein, A. M. & McKinley, S. A. Sensing and decision-making in random search. Proceedings of the National Academy of Sciences of the United States of America 109, 12070 -- 12074 (2012). [20] B´enichou, O., Coppey, M., Moreau, M., Suet, P.-H. & Voituriez, R. Optimal search strategies for hidden targets. Phys. Rev. Lett. 94, 198101 (2005). [21] Chupeau, M., B´enichou, O. & Voituriez, R. Cover times of random searches. Nature Physics advance on (2015). [22] Vergassola, M., Villermaux, E. & Shraiman, B. I. 'Infotaxis' as a strategy for searching without gradients. Nature 445, 406 -- 9 (2007). [23] Campos, D., Abad, E., M´endez, V., Yuste, S. & Lindenberg, K. Optimal search strategies of space-time coupled random walkers with finite lifetimes. Physical Review E 91, 052115 (2015). [24] Abe, M. S. & Shimada, M. L´evy walks suboptimal under predation risk. PLoS Comput Biol 11, e1004601 (2015). [25] Roberts, W. M. et al. A stochastic neuronal model predicts random search behaviors at multiple spatial scales in C. elegans. eLife;10.7554/eLife.12572 (2016). [26] Patlak, C. S. Random walk with persistence and external bias. The Bulletin of mathematical biophysics 15, 311 -- 338 (1953). 12 [27] Codling, E. A., Plank, M. J. & Benhamou, S. Random walk models in biology. Journal of the Royal Society Interface 5, 813 -- 834 (2008). [28] Obrien, W. J., Browman, H. I. & Evans, B. I. Search strategies of foraging animals. American Scientist 78, 152 -- 160 (1990). [29] Jonsen, I. D., Flemming, J. M. & Myers, R. A. Robust state-space modeling of animal movement data. Ecology 86, 2874 -- 2880 (2005). Prototypical state-space model paper. Animals analyzed one at a time. 2 phase CRW model. Error esimated separately and plugged these estimates in. Done in Winbugs, used seals as example datasets. [30] McClintock, B. T. et al. A general discrete-time modeling framework for animal movement using multistate random walks. Ecological Monographs 82, 335 -- 349 (2012). [31] Merkle, J., Fortin, D. & Morales, J. A memory-based foraging tactic reveals an adaptive mechanism for restricted space use. Ecology letters 17, 924 -- 931 (2014). [32] Fagan, W. F. et al. Spatial memory and animal movement. Ecology letters 16, 1316 -- 1329 (2013). [33] Mart´ınez-Garc´ıa, R., Calabrese, J. M. & L´opez, C. Optimal search in interacting populations: Gaussian jumps versus L´evy flights. Physical Review E 89, 032718 (2014). [34] Bhattacharya, K. & Vicsek, T. Collective foraging in heterogeneous landscapes. Journal of The Royal Society Interface 11, 20140674 (2014). [35] Mart´ınez-Garc´ıa, R., Calabrese, J. M., Mueller, T., Olson, K. A. & L´opez, C. Optimizing the Search for Resources by Sharing Information: Mongolian Gazelles as a Case Study. Physical Review Letters 110, 248106 (2013). [36] Hassell, M. P. The dynamics of arthropod predator-prey systems (Princeton University Press, 1978). [37] Curio, E. The ethology of predation, vol. 7 (Springer Science & Business Media, 2012). [38] Kareiva, P. & Odell, G. Swarms of predators exhibit "preytaxis" if individual predators use area-restricted search. American Naturalist 233 -- 270 (1987). [39] Kareiva, P. Patchiness, dispersal, and species interactions: consequences for communities of herbivorous insects. Commu- nity ecology 192 -- 206 (1986). [40] Weimerskirch, H., Pinaud, D., Pawlowski, F. & Bost, C.-A. Does prey capture induce area-restricted search? a fine-scale study using gps in a marine predator, the wandering albatross. The American Naturalist 170, 734 -- 743 (2007). [41] Fauchald, P. & Tveraa, T. Using first-passage time in the analysis of area-restricted search and habitat selection. Ecology 84, 282 -- 288 (2003). [42] Hills, T. T., Kalff, C. & Wiener, J. M. Adaptive l´evy processes and area-restricted search in human foraging. PLoS One 8, e60488 (2013). [43] Ovaskainen, O. Habitat-specific movement parameters estimated using mark-recapture data and a diffusion model. Ecology 85, 242 -- 257 (2004). [44] Auger-Mth, M., Derocher, A. E., Plank, M. J., Codling, E. A. & Lewis, M. A. Differentiating the lvy walk from a composite correlated random walk. Methods in Ecology and Evolution 6, 1179 -- 1189 (2015). [45] Benhamou, S. Efficiency of area-concentrated searching behaviour in a continuous patchy environment. Journal of Theo- retical Biology 159, 67 -- 81 (1992). [46] Morales, J., Haydon, D., Frair, J., Holsinger, K. & Fryxell, J. Extracting more out of relocation data: building movement models as mixtures of random walks. Ecology 85, 2436 -- 2445 (2004). [47] Boyer, D. & Solis-Salas, C. Random walks with preferential relocations to places visited in the past and their application to biology. Physical review letters 112, 240601 (2014). [48] Polansky, L., Kilian, W. & Wittemyer, G. Elucidating the significance of spatial memory on movement decisions by african savannah elephants using state -- space models. Proceedings of the Royal Society of London B: Biological Sciences 282, 20143042 (2015). Acknowledgements We acknowledge Ant`onia Tugores, Rub´en Tolosa and Iharob al Asimi Espina for advice in the development of the experimental setup. We are also grateful to George W. Constable for useful discussions and to Frederic Bartumeus for useful discussions and a critical reading of the manuscript. This work is funded by the Gordon and Betty Moore Foun- dation through Grant GBMF2550.06 to RMG, Universitat de les Illes Balears through a 2015 Young Visiting Scholar grant to RMG, the US National Science Foundation through grant ABI 1458748 to JMC and Ministerio de Econom´ıa y Competitividad and Fondo Europeo de Desarrollo Regional through project CTM2015-66407-P (MINECO/FEDER) to CL. Author contributions statement R.M-G conceived the study, implemented the experimental setup, and did the numerical simulations. All the authors designed the experiments, analyzed and discussed the results and contributed to the writing and revision of the manuscript. The authors declare no competing financial interests. Additional information a y b 3 20 15 1 y 10 2 θ 2 d1 x 5 1 1 5 10 x 15 20 13 12 8 4 0 C l i c k n u m b e r FIG. 1: Experimental setup. a) Single realization as shown in the game interface. Blue cells have not been visited, black and yellow cells represent the two types of cues and the red square is the target. Yellow crosses mark those squares that belong to the neighborhood of the target and have not been visited yet. They are used here to indicate the layout of the board but they are not shown to the player. b) Reconstruction of the round in A from the saved data. Small circles correspond to black cells, bigger circles to the yellow ones and the biggest circle is the target. Circles are labeled with blue numbers, di is the distance jumped starting from node i and θi is the turn angle relative to the direction at node i. 300 200 100 s k c i l c f o r e b m u N 0 0 a Simulations Data Theory b Total Intensive Extensive 200 150 100 50 3 6 9 12 15 18 21 0 0 2 4 6 8 10 12 14 Ny Ny FIG. 2: Number of movements for the blind searches as a function of the lateral length of the yellow neighborhood Ny. a) Data-model-theory comparison of the total search length. Ny = 1 means that there are no yellow cells around the target. Black squares are averages taken from experimental data, light green squares are obtained from numerical simulations (averages over 104 independent realizations) and the magenta region is the theoretical approximation. Dashed lines are interpolations and the error bars represent the standard deviation of the data. b) Decomposition of the total number of clicks between the intensive and the extensive phase. Dashed lines are interpolations and the error bars represent the standard error. When the bar is not shown the error is lower than the size of the point. 14 a 0 10 -1 10 -2 10 -3 10 -4 10 0 λext= -0.27 λint = -0.49 Extensive Intensive b 150° 180° 210° 120° 90° 60° 30° 0° 330° 5 10 15 20 Jump length 240° 300° 270° c 120° 90° 60° d 120° 90° 60° 150° 30° 150° 180° 210° 0° 180° 330° 210° 30° 0° 330° 240° 300° 270° 240° 300° 270° FIG. 3: Statistical analysis of the trajectories on the board. a) (Linear-log plot) Jump length distribution during the extensive (blue squares) and intensive (green circles) phase. Magenta lines are exponential fits with mean value given by 1/λ. b, c) Turn angle distributions during the extensive and the intensive phase respectively. d) Turn angle distribution for movements performed immediately after a yellow-to-black transition. a c s p m u j f o r e b m u N s p m u j f o r e b m u N 40 30 20 10 0 40 30 20 10 0 Blind Informed Blind Informed b d e s p m u j f o r e b m u N s p m u j f o r e b m u N 40 30 20 10 0 15 10 5 0 15 10 5 0 Blind Informed Blind Informed Blind Informed FIG. 4: Factorization of the number of clicks comparing blind and informed searches with Ny = 5. a) Total number of clicks, b) extensive phase, c) intensive phase, d) yellow-to-yellow jumps of the intensive phase, e) black-to-yellow and yellow-to-black transitions and black-to-black movements during the intensive phase. 0.2 0.1 a ■ 1 0.5 ● ● ■ ● ■ ● ● ■ 0 0 ● ■ ● 0.3 0.2 0.1 5 10 ■ ● ■ 0 ■ 0 ■ ■ ● ■ ● ■ ● ● ● ● ■ ■ ■ ■ ● ■ ● ● ■ 5 10 15 Jump length ■ ■ ● ● ● 20 0 0 15 b 1 0.5 0 0 5 10 5 10 15 20 Jump length c 120° 90° 60° d 120° 90° 60° 150° 180° 210° 30° 150° 0° 180° 330° 210° 30° 0° 330° 240° 270° 300° 240° 270° 300° FIG. 5: Comparison of the jump length and turning angle distributions for informed and blind searchers with Ny = 5 . a) Jump length distribution for the extensive phase. Green squares correspond to the whole set of rounds and red circles to the subset of random strategies. The dashed and full lines show two analytical approximations. Inset: distribution for the intensive phase. b) Equivalent to a) but for the subset of 65 blind rounds with Ny = 5. c) Turning angle distribution for movements in the intensive phase made immediately after a yellow-to-black jump. d) Same as c) but for the subset of blind searches with Ny = 5. 0.2 a 0.1 ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● c ● ● ● ● ● ● ● ● ● ● ● ● ● ● 0 0.2 0.1 ● 0 0 ● ● ● ● ● ● ● ● ● ● b d ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● 10 ● ● ● ● ● ● ● ● ● ● ● 20 30 10 20 30 0 Jump length FIG. 6: Non-stationary jump length distributions for the extensive phase of the informed searches . The extensive phase is divided in four parts and the distribution of each subset is shown: steps 1-5 (a), 6-10 (b), 11-15 (c) and 16-end (d). Red circles show experimental data and black dashed lines the theoretical fitting. Parameter estimates for each fit are shown in Table I. 16 s k c i l c f o r e b m u N 40 30 20 10 0 a D I M B M I M I I O S 40 30 20 10 0 b D I M B M I M I I O S 40 30 20 10 0 c D I M B M I M I I O S FIG. 7: Comparison between informed searches experimental data and the models. a) Number of clicks before target detection (extensive + intensive). b) Number of clicks during the extensive phase. c) Number of clicks during the intensive phase. The magenta dashed line shows the value obtained from the data for informed searchers. Labels of the x-axis: DI data informed, MB model blind, MI model informed, MII model informed 2 and OS optimal strategy. 30 25 20 15 10 5 0 C l i c k n u m b e r FIG. 8: Analysis of the optimal systematic strategy. a) Typical realization. The color of the squares indicates the temporal sequence of the jumps and its size the location outside (smaller squares) or inside the neighborhood (intermediate squares). The biggest square represents the target. b) Typical length of search as a function of the size of the neighborhood (red squares). This quantity is divided between the extensive (yellow diamonds) and the intensive (blue circles) phases. Analytical approximations are shown by dashed lines. c) Comparison between the mean number of jumps needed using an optimized systematic search rule (red squares) and the blind experimental data (black squares). Error bars represent the standard deviation, lines are interpolations. FIG. 9: Construction of random information neighborhoods starting from triangles of different size. Black cells highlighted in green belonged to the original triangle and have been removed in the randomization process. They are used here to indicate the original layout of the board but they are not shown to the player. 75 a Theoretical prediction Experimental data 17 s k c i l c f o r e b m u N s k c i l c f o r e b m u N 60 45 30 0 60 40 20 0 10 20 30 40 50 60 70 80 Number of yellow cells b c 80 60 40 20 0 0 20 40 60 80 0 20 40 60 80 Number of yellow cells Number of yellow cells FIG. 10: Prediction of the optimal size of random neighborhoods . a) Total number of movements, b) extensive phase, c) intensive stage. Black squares correspond to experimental data and magenta circles to theoretical predictions. Dashed lines are interpolations and the error bars represent the standard error, when not shown they are smaller than the size of the square. Part Mean Variance Mode α 1-5 6-10 11-15 16 -- Total Mix 8.48 6.66 6.09 5.20 7.07 6.61 22.56 12.86 9.36 7.88 16.10 16.02 5.83 4.72 4.56 3.69 4.80 4.19 3.19 3.45 3.97 3.44 3.11 2.73 β 2.66 1.93 1.54 1.51 2.28 2.42 TABLE I: Parameters obtained fitting the jump length distributions to gamma distributions. The extensive phase of the informed searches is divided in four pieces and the partial distributions fitted to gamma distributions. Changes in the mean value show the non-stationarity of the process .
1108.0538
1
1108
2011-08-02T10:45:42
Mechanochemical modeling of dynamic microtubule growth involving sheet-to-tube transition
[ "physics.bio-ph", "physics.comp-ph", "q-bio.SC" ]
Microtubule dynamics is largely influenced by nucleotide hydrolysis and the resultant tubulin configuration changes. The GTP cap model has been proposed to interpret the stabilizing mechanism of microtubule growth from the view of hydrolysis effects. Besides, the microtubule growth involves the closure of a curved sheet at its growing end. The curvature conversion also helps to stabilize the successive growth, and the curved sheet is referred to as the conformational cap. However, there still lacks theoretical investigation on the mechanical-chemical coupling growth process of microtubules. In this paper, we study the growth mechanisms of microtubules by using a coarse-grained molecular method. Firstly, the closure process involving a sheet-to-tube transition is simulated. The results verify the stabilizing effect of the sheet structure, and the minimum conformational cap length that can stabilize the growth is demonstrated to be two dimers. Then, we show that the conformational cap can function independently of the GTP cap, signifying the pivotal role of mechanical factors. Furthermore, based on our theoretical results, we describe a Tetris-like growth style of microtubules: the stochastic tubulin assembly is regulated by energy and harmonized with the seam zipping such that the sheet keeps a practically constant length during growth.
physics.bio-ph
physics
Mechanochemical modeling of dynamic microtubule growth involving sheet-to-tube transition Xiang-Ying Ji and Xi-Qiao Feng* Institute of Biomechanics and Medical Engineering, Department of Engineering Mechanics, Tsinghua University, Beijing 100084, China keywords: microtubule growth, stabilizing mechanism, GTP cap, conformational cap, coarse-grained mechanochemical model, tubulin conformation, sheet structure ABSTRACT Microtubule dynamics is largely influenced by nucleotide hydrolysis and the resultant tubulin configuration changes. The GTP cap model has been proposed to interpret the stabilizing mechanism of microtubule growth from the view of hydrolysis effects. Besides, the microtubule growth involves the closure of a curved sheet at its growing end. The curvature conversion also helps to stabilize the successive growth, and the curved sheet is referred to as the conformational cap. However, there still lacks theoretical investigation on the mechanical-chemical coupling growth process of microtubules. In this paper, we study the growth mechanisms of microtubules by using a coarse-grained molecular method. Firstly, the closure process involving a sheet-to-tube transition is simulated. The results verify the stabilizing effect of the sheet structure, and the minimum conformational cap length that can stabilize the growth is demonstrated to be two dimers. Then, we show that the conformational cap can function independently of the GTP cap, signifying the pivotal role of mechanical factors. Furthermore, based on our theoretical results, we describe a Tetris-like growth style of microtubules: the stochastic tubulin assembly is regulated by energy and harmonized with the seam zipping such that the sheet keeps a practically constant length during growth. INTRODUCTION As the most rigid cytoskeletal element in cells, microtubules have a well-marked lattice structure, consisting of regularly arranged α- and β-tubulin heterodimers (e.g., (1,2)). The dimers bind head-to-tail along their longitudinal direction into polar protofilaments, which, in turn, associate laterally in a staggered manner, rendering an elegant tubular structure. In spite of their mechanical firmness and lattice regularity, the dynamic and evolutionary attributes are intrinsic and essential for microtubules to fulfill their various significant functions in cell divisions and other intracellular biological processes (e.g., (2-4)). *Correspondence: [email protected]. 1 Microtubules suffer stochastic transitions between growing and shrinking. Such dynamic processes are highly coupled with the hydrolysis of nucleotides bonded on the assembled tubulins (5,6). Dimers in different nucleotide states assume different curvatures (7), and two distinct protofilament configurations, curved or straight, are resulted in, depending on whether GDP or GTP is bonded (8,9). The intrinsic bending characteristic of GDP-tubulins is incompatible with the canonical microtubule lattice (10). To elucidate the physical mechanisms by which a microtubule composed mainly of bending GDP-tubulins can still keep stable growth, the GTP cap model has been proposed (11,12). It hypothesizes that only when the rate of tubulin assembly exceeds the rate of GTP hydrolysis, can some newly-added GTP-tubulin layers maintain as a cap at the growing end of the microtubule. The GTP cap can sustain the uniform lattice and the microtubule growth, and its disappearance will cause the depolymerization. Despite the logical elegance of this cap model, yet there is a shortage of sufficient evidence for the existence of the GTP cap, and disputes exist about its size (13,14) and the inside GTP distributions (15,16). Experimental observations also reveal that a microtubule end can assume far more colorful conformations than mere straight or curved (17). Typically, an open and outward-curved sheet is imaged at the growing end by cryoelectron microscopy (cryo-EM) (18,19). Such a sheet structure has been considered an interphase during microtubule growth (20-22). The growth is achieved by the sheet closure which involves a distinct transition of curvature, from longitudinal to lateral. The feasibility of this growth pathway has been experimentally validated by Nogales and coworkers (10). They showed that at low temperature, tubulins binding GMPCPP, a non-hydrolysable GTP analogue, can form ribbons, in which the protofilaments have a radial bend of about 5 ° between two adjacent dimers. As temperature rises, the ribbons directly convert into tubes. It is thus suggested that the GMPCPP ribbons structurally correspond to the curved sheets at the growing microtubule end (10). The sheet-to-tube growth mode can interpret, quite successfully, the formation of the seam in a microtubule. The seam, apparently being a linear lattice defect, may virtually offer important binding sites for associated proteins to help zip the microtubule (23). Prominently, the above-described growth mode, which involves a conversion of end conformations, itself provides a stabilizing mechanism for microtubule growth (9,24). Due to this mechanism, which is referred to as the conformational cap model, the sheet is more stable than the zipped microtubule body (25). The sheet closure happen stochastically (9), and its complete closure into a blunt-ended tube would induce microtubule shrinkage (26). Despite this well-conceived essentiality of the sheet structure in microtubule growth, the detailed mechanisms remain vague. Though the conformational cap and the GTP cap models offer two different stabilizing mechanisms, a microtubule does not face an either-or choice. Actually, the conformational cap does not contradict the GTP cap but provides further guarantee for a stable microtubule growth. Also, the conformational change contributes bonus energy accumulated in microtubule lattice in addition to the energy from GTP hydrolysis (27,28), i.e., the elastic energy caused by the curvature transition (25). However, the detailed relationship between these two caps and the relation between closure and hydrolysis remain unclear. Structure analysis reveals that only GTP-tubulins can form lateral contacts compatible with the 2 microtubule lattice, and the hydrolysis is not necessary for sheet closure and microtubule growth (10). However, this conclusion only suggests that hydrolysis could happen after closure, but does not dictate a direct link between closure and hydrolysis. Further investigation is desired to determine whether closure triggers hydrolysis (18,29). On the other hand, if a sheet totally composed of GDP-tubulins can stabilize the growth phase (25,30), the essential role of pure mechanical factors in cell physiology is signified since the conformational cap works purely mechanically. These previous experimental findings clearly demonstrate the intrinsic and strong mechanics-biochemistry coupling, which is prevalent and vital for the dynamic behavior of microtubules. A deeper understanding of the appealing microtubule characters calls for well-defined theoretical models. Much effort has already been directed towards the modeling of microtubule dynamics. The switch between growing and shrinking and the corresponding length variation have been pithily described by differential equations (e.g., (13,31,32)). A detailed probe of the evolutions of conformation and energy inevitably needs finer three-dimensional simulations. Some examples have been given by VanBuren et al. (33) and Molodtsov et al. (34), who successfully accounted for the strain energy changes induced by the association and disassociation of individual dimers. As yet, however, there is still a lack of theoretical investigation on the dynamics of the sheet-to-tube growth mode. The spatial energy distribution and variation, as well as the complete growth process, under mechanical–chemical coupling regulations have rarely been addressed. Recently, we established a coarse-grained model for studying the macroscopic behavior of microtubules (35). Our simulation on the intricate sheet-ended microtubule conformation and the radial indentation technique demonstrated the efficacy of the model. The sheet structure is shown to be energetically stable. In this paper, we will employ this model to investigate the microtubule growth process which involves a sheet-to-tube transition and is mechanical-chemical coupled. We comprehensively calculate the potential energy by taking into account the intrinsic curvatures of both GTP- and GDP-tubulins. Events as subunit association and sheet closure are treated as stochastic processes regulated by the coupled changes of chemical association energy and mechanical potential energy. The influence of GTP hydrolysis is also taken into account. This study gives insight into the conformational cap hypothesis and the stabilizing mechanisms in the microtubule growth process. METHODS Coarse-grained model of microtubule In our previous work, we have developed a model to simulate the dynamic behavior of microtubules by considering their structural complexity and mechanical-chemical coupling features. Fig. 1 shows our model, which take accounts of seven types of monomer interactions according to the actual biophysical condition. Each interaction potential is assumed to be a quadratic function regarding the corresponding degree of freedom of deformation. The definition of interactions and interaction constants are summarized in Table 3 1 and five of them are illustrated in the enlarged view in Fig. 1. This model was used to reveal the detailed conformation of a sheet-ended microtubule in which protofilaments are both bending and twisting, and to simulate the indentation test for measuring the mechanical properties of a microtubule (35). The results have demonstrated the efficacy of this model in representing a dynamic process involving structural and energy evolution. In the present paper, the model established in Ref. (35) will be further employed to study the growth process and stabilizing mechanism of microtubules. To more exactly reflect the complete growth process of a microtubule, a sequence of events involving the addition of a new subunit, the closure of sheet, and the GTP hydrolysis will be incorporated in the simulations. We further develop the previous established algorithm (35) to achieve this goal. For a newly happening event, an iterative computation of monomer positions following the previous algorithm is continued until the total potential energy becomes stable. Then, a new event is allowed to occur, and the corresponding changes of conformation and energy are calculated in the same way. Thus a continuous dynamic process can be simulated. Free energy of association For polymerizing polymers, the free energy of association can be divided into two additive parts (36,37). One is beneficial for association, including the free energy associated latG with the interfaces or bonds between subunits, called denote the bond energies for longitudinal and lateral associations, respectively. The other part sG . It is the free energy required to immobilize a is unfavorable for association, denoted by "bond energy ". Let longG and subunit in the polymer, to which the most important contribution is the entropy lost due to association. Noncovalent bonds as electrostatic and hydrophobic interactions are a bit loose rather than rigid. Therefore, the subunits are not totally fixed but have some freedom to rotate and vibrate. Evidently, the loss of entropy in this case is less than that when the translational longG = –19 and rotational degrees of freedom are completely lost (38,39). Here, we take latG = –4 kBT/dimer, and kBT/dimer, results of chemical reaction kinetics (40) and are verified by computer simulations (41). sG = 11 kBT/dimer. These values match the theoretical Fig. 2 shows the free association energies for the assembly of a tubulin dimer at different sites. In the first case, a dimer filling into the gap between two long protofilaments (Fig. 2 a) will form two lateral bonds and a longitudinal bond with the pre-assembled subunits and, therefore, the corresponding association energy is calculated by =+ GGG . For + G s2 longlat the second case, a dimer associating at the side of a long protofilament (Fig. 2 b) gets a lateral bond and a longitudinal bond. Correspondingly, the association energy equals 4 =+ GGG + G longlat s . In the third case, only a longitudinal bond can be formed when a dimer assembles at the crest of the microtubule (Fig. 2 = GG c) and the association energy is . The competition and balance of the association energy and the potential + G long s energy will dictate the assembly process. Integrated thermodynamic description of microtubule growth A growing microtubule can experience three possible events, namely, the assembly of tubulins at the tip of each protofilament, the closure of the seam, and the hydrolysis for GTP-tubulins after polymerization. The probabilities for the occurrence of these dynamic events are regulated by the corresponding changes of energy in the microtubule. The assembly of tubulins is dimer-based and can be described as a thermodynamic process. For a tubulin dimer to assemble into the microtubule, all the 13 protofilament tips are potential polymerization sites, corresponding to 13 possible equilibrium energy states for the growing microtubule. All the energies at these 13 states are calculated and compared with nED the initial energy, and the differences are denoted as includes the variations of the interaction energy U and the free energy of association G regarding the n th protofilament, that is, + EUU nnini ), respectively. n = - , (1) nED ( 113 G n D=- where iniU and nU are the total potential energies before and after the assembly of a dimer, respectively. The probability for the occurrence of assembly at each tip, np , is assumed to be proportional with the absolute value of nED . To date, no direct experimental observation has been reported on the closure process of microtubules. Considering the lateral interaction is between monomers, we characterize the closure as monomer-based, i.e., the seam is zipped by the consecutive linking of monomer pairs. The closure process is simulated as follows. When a pair of monomers at the opposite edges of the sheet is about to bond, the relevant lateral interactions come into play and the whole conformation of the microtubule evolves to a new equilibrium. Ambiguity also exists for the GTP hydrolysis process, particularly for the relationship between sheet closure and GTP hydrolysis, and the distribution of GTP-tubulins in the microtubule, adding difficulties to the modeling. In this paper, unless otherwise specified, we assume that the closed part has been totally hydrolyzed whereas the sheet has not. The hydrolysis for the whole helical turn at the sheet root is treated as a synchronous event with the closure of the monomer pair. This relation is irrespective of the causality of the two processes. 5 RESULTS The sheet-to-tube transition process contributes stored energy to the microtubule lattice It is widely accepted that the hydrolysis of GTP bonded on the added dimers helps the accumulation of energy constrained in the microtubule lattice. In addition, the conformational cap model proposes that the curvature transition during the sheet-to-tube transition would also contribute to the lattice energy (25). We simulate the closure process of a microtubule and analyze the variation of the potential energy by the presented model. The sheet structure composed of GTP-tubulins measures ten monomers in length at the start. Once a monomer pair closes, the hydrolysis and the associated conformation change of tubulins on the same helical turn happen simultaneously. We take this assumption as the standard model hypothesis and will further investigate different geometrical and chemical conditions of microtubules in the following subsections. Fig. 3 shows the evolution of the total potential energy and the seven components of interaction energy during a successive closure process of three monomers. The continuous sheet-to-tube transition process and the energy changes are given in Movie S1 in the Supporting Material. It is demonstrated that the closure is steadily propelled. Both the energy barrier and the energy difference between two subsequent equilibrium states are identical during the whole closure process. An activation energy of about 104 kBT is needed for a single monomer pair to close. When a monomer pair has been zipped and the whole microtubule has evolved into stable, the accumulated energy amounts to about 2400 kBT, which is about 10 time higher than the energy from hydrolysis (29,42). A mere conformational cap plays a stabilizing role in microtubule growth Now we turn to consider the stabilizing effect of the sheet structure. Experimental works have already suggested that the blunt end is a metastable intermediate between shrinking and growing. In other words, a blunt end is not as stable as a sheet. Thus, a sheet-ended microtubule is in a stable growth state, and only when the whole sheet has closed into a tube, can the microtubule shrinkage happen (26). In the last subsection, it has been shown that the zipping of a monomer pair requires an activation energy much higher than the energy difference between the two equilibrium states. Therefore, an open sheet structure does help prevent the microtubule from shrinking by providing an energy barrier for the sheet-to-tube transition. The stabilizing effect of a conformational cap is thus supported by our simulations. The relationship between the GTP cap and the conformational cap has long been speculated but remains unclear. Here, we test whether the conformational change caused by GTP hydrolysis will resist the sheet-to-tube process and thus inhibit the microtubule growth. 6 Two microtubule models in different nucleotide states of the sheet are tested and compared with the standard structure described in the last subsection. In the first model, we assume that the microtubule, including the extending sheet, is composed merely of GDP-tubulins. In the second model, the sheet is half-hydrolyzed and, in other words, the upper part of the sheet is composed of GTP-tubulins whereas the lower part is composed of GDP-tubulins. The two models have the same configuration. In the standard model and the half-hydrolyzed sheet model, both of which have GTP tubulins at the sheets, the GTP hydrolysis process of the lowest GTP-helical turn is accompanied by its closure. Fig. 4 evolutions during the closure process in the three microtubule models. It is found that the sheet-to-tube transition can continue even when the sheet has been totally hydrolyzed. For the three sheets with different GTP distributions, both the values of the activation energies and equilibrium energy differences are similar. a compares the energy Nogales and the co-workers have stated that the bending of GDP-tubulins is incompatible with the formation of canonical lateral interaction in microtubules (10). Here, our modeling is based on the premise that the lateral interaction has been pre-defined except that along the seam. Though insufficient to completely testify the relation between the longitudinal bending conformation and the lateral interaction, our results demonstrate that the curvature transition and the sheet closure can be achieved by highly curved GDP-tubulins. This indicates that the conformational cap could be uncoupled with the GTP cap. Despite the likelihood that the substructure change in hydrolyzed tubulins would resist the formation of lateral interactions, at least from the view point of energy mechanism, hydrolysis is allowed to the tubulins on the sheet before closure and the microtubule growth will not be interfered by the nucleotide state. In addition, our simulations evidence the stabilizing role of a mere conformational mechanism in microtubule growth, and highlight the significance of mechanical factors in the microtubule behaviors. The intrinsic curvature of GTP-tubulins scarcely affects the function of conformational cap Though straighter than GDP-tubulins (43), GTP-tubulins are not perfectly straight as widely assumed (44). A slight bend of about 5 ° was exhibited at the intra-dimer interface of each GTP-tubulin (45). Using the defined model, we now explore the influence of this intrinsic curvature of GTP-tubulins on the growth process of microtubules. The energy variations during the closure of GTP sheets with an intrinsic curvature of 5 and 15 ° are compared in Fig. 4 difference either in the activation energy or the equilibrium energy stepping. This means that neither the conformation nor the energy evolution during the sheet-to-tube transition is sensitive to the longitudinal curvature. The intrinsic curvature should be dictated mainly by some structural factors and seems to be not critical to the growth process and stabilizing mechanisms of microtubules. This result supports the recent investigation about the intrinsic bending of microtubule protofilaments by Grafmüller and Voth (46), who, through large-scale b. It is seen that these three cases do not have distinct ° (standard), 0° 7 atomistic simulations, concluded that no observable difference exists between the mesoscopic properties of the intra-dimer and inter-dimer. The distinct curvature difference between polymerizing and depolymerizing protofilaments may majorly due to their lattice constraint. A conformational cap should include at least two tubulin dimer layers The cap length has long been a controversial issue (16,47,48). In this subsection, we test the dependence of the energy barrier and equilibrium energy stepping during the zipping of a monomer pair on the sheet length from ten monomers to one. The results are shown in Fig. 5. Clearly, the energy barrier and the energy stepping for the sheets containing four to ten monomers in each protofilament along the length direction are almost identical, but they are much higher than those in shorter sheets. If the sheet is shorter than the length of two dimers, the activation energy to close the sheet drops distinctly and thus the whole open sheet would experience a swift and unstable transition into a tube. In this case, the sheet structure will lose its stabilizing effect and then depolymerization will occur. Recent experimental and theoretical works about the size of GTP cap also conclude that an effective GTP cap should include at least two dimer layers (15,34). This minimum size of the GTP cap accords to that of the conformational cap estimated by our model. Such a consonance hints the potential direct relevance between closure and hydrolysis. As estimated by some researchers (29), hydrolysis can be catalyzed by the closure due to the energy accumulated. Microtubule growth plays Tetris We have demonstrated that the open sheet structure at the growing end can stabilize the microtubule growth and have testified the sheet-to-tube growth style. However, if the closure speed is faster than the tubulin assembly rate at the tip, the sheet structure will vanish and depolymerization may be triggered at the blunt microtubule end. Therefore, a stable growth should rest on harmonized closure and polymerization, keeping the sheet length practically constant, of at least two dimer layers. In this case, the seam 's zipping will experience a self-similar propagation, which conjures up the reverse process of a stable crack advancement. Further, we extrapolate that the assembly and closure is regulated as Tetris. Tubulin assembly can happen at each protofilament tip. In view of the fact that the change of potential energy due to the assembly of a single tubulin is tiny and intrinsically fluctuates, the assembly site is majorly determined by the free energy of association. That is, the position A in Fig. 2 a is the most likely site to accommodate a tubulin dimer, whereas an assembly at position C in Fig. 2 c has the smallest possibility. This association process shares the idea of Kossel-Stranski model in the kinetic theory of crystal growth. When a top layer of the sheet has been stochastically filled in by the coming tubulins, the closure of a dimer-length will 8 happen subsequently. Fig. 6 shows some snapshots of the "fill in–close up" process, and the dynamic process is shown in Movie S2. The sheet length is nearly constant so that the growth is steadily advanced; conversely, if the sheet is quickly closed into tube, the game will be soon over. DISCUSSIONS Physiological indications The actual sheet length The above calculations put forward the hypothesis that the sheet keeps a nearly constant length of more than two dimers during the sheet-to-tube growth process. However, the actual sheet length is yet unknown. We speculate that the sheet length may correlate with the nucleation process, and the length determined in the nucleation is sustained in the subsequent growth stage. It has been supposed that the microtubule nucleation template is likely to be some sheet structure composed of the laterally associated short protofilaments. The sheet closes to form an embryo of the tube (19,21). Thus the microtubule nucleation and assembly share the same basic mechanism (25). For a longitudinally curved sheet to transit its curvature and close into tube, it needs to overcome an energy barrier, which should be higher if more tubulins were zipped. The available energy that can be employed to overcome the barrier may determine the nucleated conformation and, in turn, the sheet length. It is inspiring that the sheet-to-tube nucleation pathway can be simulated under the same framework as presented in this paper. Our further study will focus on simulating the nucleation process to probe the critical nucleus, the structural templates, and the influence of nucleation on the subsequent growth. Alternate likelihood is that the sheet length is relevant with the hunt of the microtubule ends for regulators of assembly such as microtubule-associated proteins (MAPs) and coming tubulins. As can be seen from the form-finding results of sheet-ended microtubules (35), the sheets with different lengths have different shapes: a shorter sheet has a rounded top edge and its lateral alignment is more compact. The sheet end offers special recognition sites for tubulins to assemble. The inter-protofilament interfaces may accommodate MAPs (45,49). Specifically, some plus-end-tracking proteins (+TIPs) bind the microtubule end with a higher affinity than the wall (50), and one of the proposed mechanisms is their recognition of a unique structural feature of the growing end (51,52). These effects should be of direct relevance with the space and structure of the edge line, the lateral gap between protofilaments, and, therefore the sheet length. Indicated roles of some +TIPs such as EB1/Mal3p Our simulations clearly reveal that there exists a steady energy barrier during the closure of a sheet, and the sheet-to-tube process needs to be activated. It has been supposed by experiments that a kind of microtubule +TIPs, EB1 in vertebrates and its homolog Mal3p in 9 schizosaccharomyces pombe, can bind the seam. In the presence of this protein, microtubule growth is promoted (23). We guess that the binding of EB1/Mal3p can lower the activation energy required for a tubulin pair to bond and catalyze the sheet-to-tube process. Besides the function of helping the seam to zip, +TIPs are speculated to have some other roles in promoting the microtubule growth. Firstly, +TIPs may facilitate the assembly of longer tubulin oligomers in addition to individual tubulin dimers by associating the dimers along its length in solution and adding the whole oligomer into the growing microtubule (53). Secondly, because of this template effect of +TIPs for oligomer assembly, the dimers are pre-straightened before adding into the microtubule, thus the whole microtubule structure is more firm and the closure is facilitated. These influences of assembly units and configurations on the global microtubule growth can also be conveniently investigated by this model. Mechanical influences of GTP hydrolysis GTP hydrolysis is directly bound up with the microtubule dynamics. Tubulin dimers bind two GTP molecules. GTP at the N-site of α-tubulin is non-exchangeable, whereas GTP at the E-site of β-tubulin will hydrolyze into GDP after assembly (54). It is commonly agreed that the main body of a microtubule is made of GDP-tubulins, although the GTP distribution at the sheet structure at the growing end is yet unclear. However, some experiments suggest that the microtubule lattice also contains scattered GTP-tubulin remnants, meaning that the hydrolysis is sometimes incomplete during polymerization (55,56). Currently, the moment and condition for GTP to hydrolyze, as well as the influence of GTP hydrolysis on the properties and behavior of microtubules, is little known. In this paper, microtubule models with sheets in different nucleotide states have been compared. We have also tested the models in which the tubulins in the closed lattice are not in the same nucleotide state and, namely conformational state. No distinction is found with respect to the sheet-to-tube curvature conversion process, implying that the conformational change of tubulins resulted from GTP hydrolysis hardly interferes with the mechanic requirement of the global conformation evolution of a growing microtubule. The conformational change may mainly influence the depolymerization process. For example, it weakens the lateral interaction and facilitates the ram's horn-like peeling of protofilaments during shrinking, and the GTP remnants in the lattice could help rescue the microtubule from shortening (53,54). Model discussion Influences of the interaction definitions In our model, three kinds of interactions that are not experimentally based, and their values are assumed as k k= dihedralbend latlat /50 , k k= dihedralbend longlong /50 , and k k= diaglong (35). Our previous work has demonstrated that the variations of the two dihedral angles have little influence on the total energy (35). The diagonal interaction majorly acts to restrain the 10 fluctuations of tubulin positions and make the calculation converge faster. Here, we examine the influence of these values on the energy evolution during the closure process and validate our assumptions. We vary each constant value independently by keeping the other two at their originally and assumed values. Fig. 7 shows the results. No surprisingly, the 10-fold changes of latk dihedral longk dihedral do not make notable differences on the energy and conformation evolutions. With changing k diagonal , the energy barriers alternate. A smaller k diagonal -value results in a larger barrier, due to the increased flexibility of the model and the resultant larger tubulin displacements during the calculation process, but the stable state is soon found. Moreover, the equilibrium energy stepping remains the same and the energy barrier value is stable for each closure, so all conclusions in the paper can be validated. Limitations and further directions It is desired if a time factor is incorporated in the model. This requires the knowledge about the rate of assembly, closure and hydrolysis, by which a spatiotemporal evolution of conformation and energy could be elucidated. Besides, the bond rupture, namely, the depolymerization is not considered in the presented model. We hope that a systematic modeling of the integrated dynamic process of tubulin assembly, sheet closure, and protofilament peeling can be accomplished. CONCLUSION In this study, we have simulated the chemical-mechanical coupled microtubule growth process which involves a sheet-to-tube transition. The conformation and energy evolution is exhibited and the stabilizing mechanism of the conformational cap is analyzed. We demonstrate that an effective conformation cap should comprise at least two dimer layers, and the cap length is maintained during a stable growth process by the harmonized tubulin assembly and sheet closure. SUPPORTING MATERIAL Movie S1. Movie S2. This work was supported by the National Natural Science Foundation of China (Grant No. 10732050 to X.Q.F). 11 References 1. Hawkins, T., M. Mirigian, M. S. Yasar, and J. L. Ross. 2010. Mechanics of microtubules. J. Biomech. 43:23–30. 2. Howard, J. and A. A. Hyman. 2003. Dynamics and mechanics of the microtubule plus end. Nature. 422:753–758. 3. Desai, A. and T. J. Mitchison. 1997. Microtubule polymerization dynamics. Annu. Rev Cell Dev. Biol. 13:83–117. 4. Heald, R. and E. Nogales. 2002. Microtubule dynamics. J. Cell Sci. 115:3–4. 5. Carlier, M. F. 1989. Role of nucleotide hydrolysis in the dynamics of actin-filaments and microtubules. Int. Rev. Cytol. 115:139–170. 6. Erickson, H. P. and D. Stoffler. 1996. Protofilaments and rings, two conformations of the tubulin family conserved from bacterial FtsZ to alpha/beta and gamma tubulin. J. Cell Biol. 135:5–8. 7. Krebs, A., K. N. Goldie, and A. Hoenger. 2005. Structural rearrangements in tubulin following microtubule formation. EMBO Rep. 6:227–232. 8. Mandelkow, E. M., E. Mandelkow, and R. A. Milligan. 1991. Microtubule dynamics and microtubule caps - a time-resolved cryoelectron microscopy study. J. Cell Biol. 114:977–991. 9. Hyman, A. A. and E. Karsenti. 1996. Morphogenetic properties of microtubules and mitotic spindle assembly. Cell. 84:401–410. 10. Wang, H. W. and E. Nogales. 2005. Nucleotide-dependent bending flexibility of tubulin regulates microtubule assembly. Nature. 435:911–915. 11. Mitchison, T. and M. Kirschner. 1984. Dynamic instability of microtubule growth. Nature. 312:237–242. 12. Erickson, H. P. and E. T. O'Brien. 1992. Microtubule dynamic instability and GTP hydrolysis. Annu. Rev. Biophys. Biomol. Struct. 21:145–166. 13. Martin, S. R., M. J. Schilstra, and P. M. Bayley. 1993. Dynamic instability of microtubules - monte-carlo simulation and application to different types of microtubule lattice. Biophys. J. 65:578–596. 12 14. Drechsel, D. N. and M. W. Kirschner. 1994. The minimum GTP cap required to stabilize microtubules. Curr. Biol. 4:1053–1061. 15. Schek, H. T., M. K. Gardner, J. Cheng, D. J. Odde, and A. J. Hunt. 2007. Microtubule assembly dynamics at the nanoscale. Curr. Biol. 17:1445–1455. 16. Gardner, M. K., A. J. Hunt, H. V. Goodson, and D. J. Odde. 2008. Microtubule assembly dynamics: new insights at the nanoscale. Curr. Opin. Cell Biol. 20:64–70. 17. Zovko, S., J. P. Abrahams, A. J. Koster, N. Galjart, and A. M. Mommaas. 2008. Microtubule plus-end conformations and dynamics in the periphery of interphase mouse fibroblasts. Mol. Biol. Cell. 19:3138–3146. 18. Chr étien, D., S. D. Fuller, and E. Karsenti. 1995. Structure of growing microtubule ends – 2-dimensional sheets close into tubes at variable rates. 129:1311–1328. J. Cell Biol. 19. Mozziconacci, J., L. Sandblad, M. Wachsmuth, D. Brunner, and E. Karsenti. 2008. Tubulin Dimers Oligomerize before Their Incorporation into Microtubules. PLoS One. 3:e3821. 20. Pampaloni, F. and E.-L. Florin. 2008. Microtubule architecture: inspiration for novel carbon nanotube-based biomimetic materials. Trends Biotechnol. 26:302–310. 21. Amos, L. A. 1995. The microtubule lattice – 20 years on. Trends Cell Biol. 5:48–51. 22. Vitre, B., F. M. Coquelle, C. Heichette, C. Garnier, D. Chr étien, and I. Arnal. 2008. EB1 regulates microtubule dynamics and tubulin sheet closure in vitro. Nat. Cell Biol. 10:415–421. 23. Sandblad, L., K. E. Busch, P. Tittmann, H. Gross, D. Brunner, and A. Hoenger. 2006. The Schizosaccharomyces pombe EB1 homolog Mal3p binds and stabilizes the microtubule lattice seam. Cell. 127:1415–1424. 24. Tran, P. T., R. A. Walker, and E. D. Salmon. 1997. A metastable intermediate state of microtubule dynamic instability that differs significantly between plus and minus ends. J. Cell Biol. 138:105–117. 25. Chr étien, D., I. Jan ósi, J. C. Taveau, and H. Flyvbjerg. 1999. Microtubule's conformational cap. Cell Struct. Funct. 24:299–303. 26. Arnal, I., E. Karsenti, and A. A. Hyman. 2000. Structural transitions at 13 microtubule ends correlate with their dynamic properties in Xenopus egg extracts. Biol. 149:767–774. J. Cell 27. Caplow, M., R. L. Ruhlen, and J. Shanks. 1994. The free-energy for hydrolysis of a microtubule-bound nucleotide triphosphate is near zero - all of the free-energy for hydrolysis is stored in the microtubule lattice. J. Cell Biol. 127:779–788. 28. Elie-Caille, C., F. Severin, J. Helenius, J. Howard, D. J. Muller, and A. A. Hyman. 2007. Straight GDP-Tubulin Protofilaments Form in the Presence of Taxol. Biol. 17:1765–1770. 29. Mickey, B. and J. Howard. 1995. Rigidity of microtubules is increased by stabilizing agents. J. Cell Biol. 130:909–917. 30. Mahadevan, L. and T. J. Mitchison. 2005. Cell biology - Powerful curves. Curr. Nature. 435:895–897. 31. Bayley, P. M., M. J. Schilstra, and S. R. Martin. 1989. A simple formulation of microtubule dynamics - quantitative implications of the dynamic instability of microtubule populations in vivo and in vitro. J. Cell Sci. 93:241–254. 32. Chen, Y. D. and T. L. Hill. 1987. Theoretical-studies on oscillations in microtubule polymerization. Proc. Natl. Acad. Sci. U. S. A. 84:8419–8423. 33. VanBuren, V., L. Cassimeris, and D. J. Odde. 2005. Mechanochemical model of microtubule structure and self-assembly kinetics. Biophys. J. 89:2911–2926. 34. Molodtsov, M., E. Ermakova, E. Shnol, E. Grishchuk, J. McIntosh, and F. Ataullakhanov. 2005. A molecular-mechanical model of the microtubule. 88:3167–3179. Biophys. J. 35. Ji, X. Y. and X. Q. Feng. 2011. A coarse-grained mechanochemical model for simulating the dynamic behavior of microtubules. Submitted. 35. Erickson, H. P. and D. Pantaloni. 1981. The role of subunit entropy in cooperative assembly – nucleation of microtubules and other two-dimensional polymers. Biophys. J. 34:293–309. 36. Chothia, C. and J. Janin. 1975. Principles of protein-protein recognition. Nature. 256:705–708. 37. Steinberg, I. Z. and H. A. Scheraga. 1963. Entropy Changes Accompanying Association Reactions of Proteins. J. Biol. Chem. 238:172–181. 14 38. Page, M. I. and W. P. Jencks. 1971. Entropic Contributions to Rate Accelerations in Enzymic and Intramolecular Reactions and the Chelate Effect. Proc. Natl. Acad. Sci. U. S. A. 68:1678–1683. 39. Erickson, H. and D. Pantaloni. 1981. The role of subunit entropy in cooperative Biophys. J. assembly. Nucleation of microtubules and other two-dimensional polymers. 34:293–309. 40. VanBuren, V., D. J. Odde, and L. Cassimeris. 2002. Estimates of lateral and Proc. Natl. Acad. Sci. U. S. A. longitudinal bond energies within the microtubule lattice. 99:6035–6040. 41. Efremov, A., E. L. Grishchuk, J. R. McIntosh, and F. I. Ataullakhanov. 2007. In search of an optimal ring to couple microtubule depolymerization to processive chromosome motions. Proc. Natl. Acad. Sci. U. S. A. 104:19017–19022. 42. Felgner, H., R. Frank, J. Biernat, E. Mandelkow, E. Mandelkow, B. Ludin, A. Matus, and M. Schliwa. 1997. Domains of neuronal microtubule-associated proteins and flexural rigidity of microtubules. J. Cell Biol. 138:1067–1075. 43. Jan ósi, I. M., D. Chr étien, and H. Flyvbjerg. 2002. Structural microtubule cap: Stability, catastrophe, rescue, and third state. Biophys. J. 83:1317–1330. 44. Nogales, E. and H. W. Wang. 2006. Structural mechanisms underlying nucleotide-dependent self-assembly of tubulin and its relatives. 16:221–229. Curr. Opin. Struct. Biol. 45. Grafm üller, A. and G. A. Voth. 2011. Intrinsic Bending of Microtubule Protofilaments. Structure 19:409-417. 46. Bayley, P., M. Schilstra, and S. Martin. 1989. A lateral cap model of microtubule dynamic instability. FEBS Lett. 259:181–184. 47. Caplow, M. and J. Shanks. 1996. Evidence that a single monolayer tubulin-GTP cap is both necessary and sufficient to stabilize microtubules. Mol. Biol. Cell. 7:663–675. 48. Wu, Z. H., H. W. Wang, W. H. Mu, Z. C. Ouyang, E. Nogales, and J. H. Xing. 2009. Simulations of Tubulin Sheet Polymers as Possible Structural Intermediates in Microtubule Assembly. PLoS One. 4:e7291. 49. Schuyler, S. C. and D. Pellman. 2001. Microtubule "plus-end-tracking 15 proteins": The end is just the beginning. Cell. 105:421–424. 50. Carvalho, P., J. S. Tirnauer, and D. Pellman. 2003. Surfing on microtubule ends. Trends Cell Biol. 13:229–237. 51. Diamantopoulos, G. S., F. Perez, H. V. Goodson, G. Batelier, R. Melki, T. E. Kreis, and J. E. Rickard. 1999. Dynamic Localization of CLIP-170 to Microtubule Plus Ends Is Coupled to Microtubule Assembly. J. Cell Biol. 144:99–112. 52. Kerssemakers, J. W. J., E. L. Munteanu, L. Laan, T. L. Noetzel, M. E. Janson, and M. Dogterom. 2006. Assembly dynamics of microtubules at molecular resolution. Nature. 442:709–712. 53. Galjart, N. 2010. Plus-End-Tracking Proteins and Their Interactions at Microtubule Ends. Curr. Biol. 20:R528–R537. 54. Dimitrov, A., M. Quesnoit, S. Moutel, I. Cantaloube, C. Pous, and F. Perez. 2008. Detection of GTP-Tubulin Conformation in Vivo Reveals a Role for GTP Remnants in Microtubule Rescues. Science. 322:1353–1356. 55. Cassimeris, L. 2009. Microtubule Assembly: Lattice GTP to the Rescue. Curr. Biol. 19:R174–R176. 16 Interaction potential Interaction constant Value longk latk diagk longk bend latk bend longk dihedral latk dihedral 3.0 nN/nm 14.0 nN/nm 3.0 nN/nm 2.0 nN(cid:215) nm 8.5 nN(cid:215) nm 0.04 nN(cid:215) nm 0.17 nN(cid:215) nm Interaction Longitudinal tension or compression Lateral tension or compression Diagonal tension or compression 1 2 3 4 5 6 7 Uk d= longlonglong Uk d= latlatlat Uk d= diagdiagdiag 2 2 2 / 2 / 2 / 2 / 2 Longitudinal bending U = k bendbend longlonglong 2 q Lateral bending U = bendbend latlatlat k 2 q / 2 Longitudinal dihedral bending U = k dihedraldihedral longlonglong 2 y / 2 Lateral dihedral bending U k dihedraldihedral latlatlat = 2 y / 2 Table 1 Interaction definitions in the model 17 · · · · · · · Figure Legends FIGURE 1 Model of a sheet-ended microtubule. The enlarged view shows five of the defined interactions, including: (1) longitudinal tension or compression interaction, (2) lateral tension or compression interaction, (3) diagonal tension or compression interaction, (4) longitudinal bending interaction, and (5) lateral bending interaction. For the detailed mathematical depictions and the definitions of the longitudinal and lateral dihedral bending interactions, please refer to Ref. (35). FIGURE 2 Sites for a coming tubulin dimer to assemble. ( a) the dimer inserting into a gap, (b) the dimer associating a single-sided neighbor, and (c) the dimer falling upon the crest. FIGURE 3 Potential energy evolution during the sheet-to-tube transition process. A continuous zipping of the seam counting three pairs of monomers is characterized. The consecutive closure happens when the microtubule is in an equilibrium conformation. The upper panel shows the evolution of the total potential energy, from which the energy barrier and energy difference between two equilibrium states are clearly detected. The lower panel exhibits the evolution of the seven energy components. Respecting the differences of orders of magnitude, a semilogarithmic coordinate is adopted. FIGURE 4 Comparison of the energy barriers and energy differences during the sheet-to-tube transition under different sheet structures: ( a) the sheets are in three different nucleotide states, (b) the intrinsic curvatures of GTP-tubulins are of three different values. In both panels, the red lines represent the result for the standard model shown in Fig. 3. For clarity, the three sets of data have been offset horizontally, but not vertically. FIGURE 5 Energy barrier and energy stepping for a monomer pair closure of sheets of different lengths, varying from 1 to 10 monomers in the longitudinal direction. The ten microtubule models are composed of protofilaments of the same length. The total potential energies are at different levels since the length of closed parts of the ten models are different. FIGURE 6 Coupled assembly and closure during microtubule growth. ( a) Snapshots of the sheet structure. When a helical turn at the end has been fully filled by tubulins, the seam will be zipped a same length and the sheet will be closed up. Two subsequent closures are shown. (b) Schematic drawing of the sheet structure evolution exhibited in ( a). Three stages colored black, blue and red in chronological order are involved. The short dashes at the bottom represent the root of the sheet, and the arrays of 13 short lines at the top represent the tubulin distributions at the tips of the 13 protofilaments. The first "fill in –close up " process is characterized in blue, and the second in red; the corresponding conformations are highlighted with the same colors in (a). FIGURE 7 Influences of interaction constants on the energy barrier and energy difference 18 between two equilibrium states during closure. (a–c) Influences of the interaction constants of lateral dihedral, longitudinal dihedral, and diagonal tension or compression, respectively. 19 1 2 20 3 21 4 5 22 6 7 23
1003.4481
1
1003
2010-03-23T17:57:33
Topological Solitons and Folded Proteins
[ "physics.bio-ph", "hep-th", "physics.chem-ph" ]
We propose that protein loops can be interpreted as topological domain-wall solitons. They interpolate between ground states that are the secondary structures like alpha-helices and beta-strands. Entire proteins can then be folded simply by assembling the solitons together, one after another. We present a simple theoretical model that realizes our proposal and apply it to a number of biologically active proteins including 1VII, 2RB8, 3EBX (Protein Data Bank codes). In all the examples that we have considered we are able to construct solitons that reproduce secondary structural motifs such as alpha-helix-loop-alpha-helix and beta-sheet-loop-beta-sheet with an overall root-mean-square-distance accuracy of around 0.7 Angstrom or less for the central alpha-carbons, i.e. within the limits of current experimental accuracy.
physics.bio-ph
physics
Topological Solitons and Folded Proteins M.N. Chernodub,1, 2, ∗ Shuangwei Hu,1, 3, † and Antti J. Niemi3, 1, ‡ 1Laboratoire de Math´ematiques et Physique Th´eorique, Universit´e Fran¸cois-Rabelais Tours, F´ed´eration Denis Poisson - CNRS, Parc de Grandmont, 37200 Tours, France 2Department of Mathematical Physics and Astronomy, Krijgslaan 281, 59, Gent, B-9000, Belgium 3Department of Physics and Astronomy, Uppsala University, P.O. Box 803, S-75108, Uppsala, Sweden (Dated: October 28, 2018) We propose that protein loops can be interpreted as topological domain-wall solitons. They interpolate between ground states that are the secondary structures like α-helices and β-strands. Entire proteins can then be folded simply by assembling the solitons together, one after another. We present a simple theoretical model that realizes our proposal and apply it to a number of biologically active proteins including 1VII, 2RB8, 3EBX (Protein Data Bank codes). In all the examples that we have considered we are able to construct solitons that reproduce secondary structural motifs such as α-helix -- loop -- α-helix and β-sheet -- loop -- β-sheet with an overall root-mean-square-distance accuracy of around 0.7 Angstrom or less for the central α-carbons, i.e. within the limits of current experimental accuracy. PACS numbers: 87.15.A-,87.15.Cc,87.14.hm Solitons are ubiquitous and widely studied objects that can be materialized in a variety of practical and theo- retical scenarios [1], [2]. For example solitons can be deployed for data transmission in transoceanic cables, for conducting electricity in organic polymers [1], and they may also transport chemical energy in proteins [3]. Solitons explain the Meissner effect in superconductiv- ity and dislocations in liquid crystals [1]. They also model hadronic particles, cosmic strings and magnetic monopoles in high energy physics [1] and so on. The first soliton to be identified is the Wave of Translation that was observed by John Scott Russell in the Union Canal of Scotland. This wave can be accurately described by an exact soliton solution of the Korteweg-de Vries (KdV) equation [1]. At least in principle it can also be con- structed in an atomary level simulation where one ac- counts for each and every water molecule in the Canal, together with all of their mutual interactions. However, in such a Gedanken simulation it would probably be- come a real challenge to unravel the collective excitations that combine into the Wave of Translation without any guidance from the known soliton solution of the KdV equation since solitons can not be constructed simply by adding up small perturbations around some ground state: A (topological) soliton emerges when non-linear interac- tions combine elementary constituents into a localized collective excitation that is stable against small pertur- bations and cannot decay, unwrap or disentangle [1], [2]. In this Letter we propose that (topological) solitons can also explain and describe the folding of proteins into their native state [4], [6]. We characterize a folded pro- tein by the Cartesian coordinates ri of its N central α- carbons, with i = 1, ..., N . For many biologically active proteins these coordinates can be downloaded from Pro- tein Data Bank (PDB) [7]. Alternatively, the protein can be described in terms of its bond and torsion angles that can be computed from the PDB data. For this we intro- duce the tangent vector ti and the binormal vector bi ti = ri+1−ri bi = ti−1×ti ti−1−ti ri+1−ri & (1) Together with the normal vector ni = bi × ti we then have three vectors that are subject to the discrete Frenet [8] equation  = exp{−κi · T 2} · exp{−τi · T 3} ni+1 bi+1 ti+1  (2) ni bi ti Here T 2 and T 3 are two of the standard generators of three dimensional rotations, explicitely in terms of the permutation tensor we have (T i)jk = ijk. From (1), (2) we can compute the bond angles κi and the torsion angles τi using PDB data for ri. Alternatively, if we know these angles we can compute the coordinates ri. The common convention is to select the range of these angles so that κi is positive. In the continuum limit where (2) becomes the standard Frenet equation for a continuous curve, κi → κ(x) then corresponds to local curvature. As an example we consider the 35 residue villin head- piece protein with PDB code 1VII that has been widely investigated, both theoretically and experimentally [4]. For example in the state of the art simulation [5] suc- ceeded in producing its fold for a short time within an accuracy of ∼ 2 − 3 A. From the PDB data we compute the values of bond angles κi and torsion angles τi and the result is displayed in Figure 1(a), where we use the (standard) convention that the discrete Frenet curvature κ is positive. In 1VII there are three α-helices that are separated by two loops. When we use the PDB (NMR) convention for indexing the residues the first, longer, loop is located at sites 49-54 and the second, shorter, between 59-62. 2 FIG. 2: The potential energy on (κ, τ ) plane that corresponds qualitatively to the data in Figure 1(b), the soliton between sites 49-54 corresponds to the red dashed trajectory and the soliton between sites 59-62 to the blue solid trajectory. defined by the energy functional (3) (a) (b) FIG. 1: (a) The bond and torsion angles of 1VII, computed with the (standard) convention that the discrete Frenet cur- vature κ is positive. (b) The Z2 gauge transformed bond and torsion angles. We shall now show that Figure 1(a) describes two soli- ton configurations, albeit in an encrypted form. In order to decrypt the data in Figure 1(a) so that these solitons become unveiled we observe that the equation (2) has the following local Z2 gauge symmetry: At every site we can send (cid:26)κi → κi · cos(∆i+1) τi → τi + ∆i − ∆i+1 Z2 : and when we choose at each site ∆i = 0 or ∆i = π where ∆i = π is the nontrivial element of the Z2 gauge group, the Cartesian coordinates ri computed from the discrete Frenet equation remain intact. If we judiciously implement this Z2 gauge transformation in the data dis- played in Figure 1(a) we arrive at the apparently quite different Figure 1(b). Unlike in Figure 1(a), the profile of κi in Figure 1(b) clearly displays the hallmark profile of a topological soliton-(anti)soliton pair in a double-well potential: The two solitons are located around the sites with indices 49-54 and 59-62 which are the locations of the two loops in 1VII. These solitons interpolate between the two "ground state" values κi ≈ ±π/2 that pinpoint the locations of the α-helices in 1VII. Moreover, the two downswings in the value of τi from the value τi ≈ 1 that mark the locations of the α-helices, coincide with the lo- cations of the two solitons. The ensuing combined profile of κi and τi is qualitatively consistent with a double- well potential structure in the (κ, τ ) plane that has the form displayed in Figure 2: When we move from left to right in Figure 1(b), we follow a trajectory in the (κ, τ ) plane that starts by fluctuating around the potential en- ergy minimum at (κ, τ ) ≈ (−π/2, 1) in Figure 2, corre- sponding to the first α-helix. The trajectory then moves through the first loop a.k.a. soliton (the red dashed line) to the second potential energy minimum i.e. α-helix at (κ, τ ) ≈ (+π/2, 1) in Figure 2, and finally back through the second loop a.k.a. soliton (the blue solid line) to the first potential energy minimum at (κ, τ ) = (−π/2, 1). We now present a simple theoretical model [9], [10] that reproduces the (κ, τ ) profile in Figure 1(b) as a combi- nation of two soliton solutions, with a very high atomary level accuracy for the central α-carbons. The model is (κi+1 − κi)2 + c · (κ2 N(cid:88) i=1 N−1(cid:88) N(cid:88) i=1 E = + (cid:8)b κ2 i − m2)2 (cid:9) i τi i τ 2 i + d τi + e τ 2 i + q κ2 (4) i=1 Here N is the number of central α-carbons and (c, m, b, d, e, q) are parameters. The first sum describes nearest neighbor interactions along the protein. The sec- ond sum describes a local self-interaction of the bond angles. The third sum describes local interactions be- tween bond and torsion angles, its first term has an ori- gin in a Higgs effect which is due to the potential term in the second sum. The second term in the third sum is the Chern-Simons term, it is responsible for the chiral- ity of the protein chain. The third term is a Proca mass term and the last term can also be related to the Abelian Higgs Model, and it is also chiral. As explained in [10] this energy functional is essentially unique, and in partic- ular it can be related to a gauge invariant (supercurrent) version of the energy of 1+1 dimensional lattice Abelian Higgs Model. In three space dimensions this model is also known as the Ginzburg-Landau Model of conven- tional superconductivity [2]. Note that in (4) there is no reference to the specifics of the interactions involving amino acids such as hydrophobic, hydrophilic, long-range Coulomb, van der Waals, saturating hydrogen bonds etc. interactions that are presumed to drive the folding pro- cess. The only explicit long-range force present in (4) is the nearest neighbor interaction described by the first term. Moreover, as it stands (4) depends only on six site-independent, homogeneous parameters. There is no direct reference whatsoever to the underlying in general highly inhomogeneous amino acid structure of a protein. We argue that this becomes possible since (4) supports solitons that describe the common secondary structural motifs such as α-helix/β-strand - loop - α-helix/β-strand as solutions to its classical equations of motion. Further- more, even though the actual numerical values of the pa- rameters are certainly motif dependent and for long loops that constitute bound states of several solitons one might need to introduce more than six parameters, we expect there to be wide universality so that a given soliton with its relatively few parameters describes a general class of homologous motifs. Consequently only a relatively small set of parameters are needed to provide soliton templates for structure prediction. In fact, we propose that solitons are the mathematical manifestation of the experimental observation, that the number of different protein folds is surprisingly limited. The presence of solitons could then be the reason for the success of bioinformatics based ho- mology modeling in predicting native folds [4].In order to quantitatively disclose the soliton solution of (4) we start by observing that the first two sums in (4) can be interpreted as a discrete version of the energy of the 1+1 dimensional double well λφ4 model that is known to sup- port the topological kink-soliton. In the continuum limit the kink has the analytic form [1], [2], κ(x) = m · tanh[m c · (x − x0)] . √ We can try to estimate the parameters m and c for each of the two solitons in the Figure 1(b) by a least square fitting where we use this continuum soliton to approximate the exact soliton solution of the discrete equations of motion. We consider here explicitly only the first soliton of 1VII, located between (PDB index) sites 49-54. Using the sites 46-56 we find the following least square fit κ(x) ≈ 1.4627 · tanh[2.0816(x − 52.597)] . (5) In order to construct τ (x) we solve for its equation of motion in (4). The result is τ (x) ≈ −2.4068 · 1 − 0.4689 · κ2(x) 1 − 0.4619 · κ2(x) (6) In Figure 3 we show how the data in Figure 1(b) is de- scribed by the approximate soliton profile (5), (6). When we construct the ensuing discrete curve in the three di- mensional space by solving (2) with for κi and τi given by (5) and (6), we reproduce the first loop of 1VII with a surprisingly good RMSD accuracy of ∼1.43 A for the PDB indices 46-56 which is quite remarkable, taking into account the simplicity of our approximation. In order to construct a more accurate description of 1VII, we resort to a numerical construction of a soli- ton solution to the equations of motion if (4). We use simulated annealing that involves a Monte Carlo energy 3 FIG. 3: The PDB data for the first α-helix - loop - α-helix motif in 1VII, on the left κi and on the right τi, together with the least square approximations (5) and (6) (the blue solid lines). minimization of the energy functional (cid:110)(cid:18) ∂E F = −β1 · N(cid:88) (cid:118)(cid:117)(cid:117)(cid:116) 1 N(cid:88) −β2 · ∂κi i=1 N i=1 (cid:19)2 (cid:18) ∂E ∂τi + (cid:19)2(cid:111) (7) rPDB(i) − rsoliton(i)2 with a simultaneous cooling of the two (inverse) tem- peratures β1 and β2. Here the first sum vanishes when we have a solution to the classical difference equation of motion of (4), and the second sum computes the RMSD distance between the ith α-carbon of the solution and the protein we wish to construct. The second term in (7) acts like a chemical potential that selects the param- eters in (4) so that we arrive at a soliton solution that corresponds to the given protein. We have numerically constructed the classical solutions of (4) that describe the secondary structural motifs in proteins with PDB codes 1VII, 2RB8 and 3EBX. The first one has three α-helices separated by loops, while the second and third have β-strand-loop-β-strand mo- tifs; Both cases can be described equally by (4), the only difference is that in the case of β-strands the two minima of the (classical) potential in (4) are located at (κ, τ ) ≈ (±1, π). In each of the proteins that we have studied we have routinely been able to reproduce the sec- ondary structural motifs as classical soliton solutions to the equations of motion for (4) in terms of only six param- eters and with an overall RMSD accuracy of around 0.7 A per motif which is essentially the experimental accuracy in X-ray crystallography and NMR; in our simulations the first sum in (7) decreases typically by around ten or- ders of magnitude indicating that the final configuration is a solution, essentially within numerical accuracy. Con- sequently at least in these proteins the secondary struc- tural motifs can be viewed as solitons of the model (4), within experimental accuracy. Since the motifs that we have considered are quite generic in PDB data, we have very little doubt that our results will continue to persist whenever we have loops that connect α-helices and/or β-strands. And as long as the loops are not very long and do not describe bound states of several solitons there does not appear to be any need to introduce more than six parameters. Work is now in progress to systemati- cally construct and classify the solitons that describe the secondary structural motifs in a large class of biologically active proteins. We have also made tentative attempts to use our soli- tons to reconstruct entire proteins, by naively joining the solitons that describe the secondary structural motifs at their ends. In the case of 1VII we have been able to re- produce in this manner the entire protein as a classical soliton with an overall RMSD accuracy of around 1.2 A and the result is shown in Figure 4. Even though the FIG. 4: The helix-loop-helix-loop-helix structure of the 1VII protein (green) together with its reconstruction in terms of two solitons (purple). The RMSD distance between the two configurations is ≈ 1.2 A. accuracy we obtain is very good, the loss of accuracy from ∼ 0.7 A to ∼ 1.2 A when we combine the two soli- tons suggests that we can still substantially improve the method of assembling an entire folded protein from its solitons. Work is now in progress to develop more effi- cient methods for assembling entire proteins from their solitons. In conclusion, we have proposed that the common sec- ondary structural motifs that describe loops connecting α-helices and/or β-strands can be interpreted as topo- logical solitons, with the α-helices and β-sheets viewed as ground states that are interpolated by the loops as solitons. Entire proteins can then be assembled simply 4 by combining these solitons together one after another. We have also presented a model that allows us to fold proteins in terms of its solitons within experimental accu- racy. In its simplest form that we have considered here, the model has only six site independent but in general motif dependent parameters. This appears to be suffi- cient to describe loops that are not too long. This ob- servation that all the details and complexities of amino acids and their interactions can be summarized in so sim- ple terms suggests the existence of wide universality in protein folding, and it can be viewed as a mathematically precise formulation of the experimental observation that the number of protein conformations is far more limited than the number of different amino acid combinations. Finally, we leave it as a future challenge to expand the model so that it incorporates an order parameter that describes the local orientation of the amino acids along the α-carbon backbone. Our research is supported by grants from the Swedish Research Council (VR). We thank Martin Lundgren for discussions. ∗ On leave of absence from ITEP, Moscow, Russia; Elec- tronic address: [email protected] † Electronic address: [email protected] ‡ Electronic address: [email protected] [1] T. Dauxois and M. Peyrard, Physics of Solitons (Cam- bridge University Press, Cambridge, 2006) [2] N. Manton and P. Sutcliffe, Topological Solitons (Cam- bridge University Press, Cambridge, 2004) [3] A.S. Davydov, Journ. Theor. Biology 38, 559 (1973). [4] K.A. Dill, S. Banu Ozkan, M.S. Shell and T.R. Weikl, The Protein Folding Problem, Annual Review of Biophysics 37, 289 (2008) [5] G. Jayachandran, V. Vishal and V.S. Pane, J. Chem. Phys. 124, 164902 (2006) [6] K. Huang, Lectures On Statistical Physics And Protein Folding (World Scientific Publishing Co. Pte. Ltd. Singa- pore, 2005) [7] H.M. Berman, K. Henrick, H. Nakamura and J.L. Markley, Nucl. Acids Research 35 (Database issue) D301-3 (2007) [8] P.J. Flory, Statistical Mechanics of Chain Molecules (Wi- ley, New York, 1969) [9] A. J. Niemi, Phys. Rev. D 67, 106004 (2003) [arXiv:hep- th/0206227]. [10] U.H. Danielsson, M. Lundgren and A.J. Niemi, e-print arXiv:0902.2920
1204.5585
1
1204
2012-04-25T08:33:29
The dynamics of the DNA denaturation transition
[ "physics.bio-ph", "cond-mat.mes-hall", "cond-mat.soft" ]
The dynamics of the DNA denaturation is studied using the Peyrard-Bishop-Dauxois model. The denaturation rate of double stranded polymers decreases exponentially as function of length below the denaturation temperature. Above Tc, the rate shows a minimum, but then increases as function of length. We also examine the influence of sequence and solvent friction. Molecules having the same number of weak and strong base-pairs can have significantly different opening rates depending on the order of base-pairs.
physics.bio-ph
physics
epl draft The dynamics of the DNA denaturation transition Titus S. van Erp1 and Michel Peyrard2 1 Centrum voor Oppervlaktechemie en Katalyse, KU Leuven, Kasteelpark Arenberg 23, B-3001 Leuven, Belgium 2 Laboratoire de Physique, Ecole Normale Sup´erieure de Lyon, 46 all´ee d'Italie, 69364 Lyon Cedex 07, France 2 1 0 2 r p A 5 2 ] h p - o i b PACS 87.15.H- -- Dynamics of biomolecules PACS 82.20.Pm -- Rate constants, reaction cross sections, and activation energies PACS 87.14.gk -- DNA Abstract -- The dynamics of the DNA denaturation is studied using the Peyrard-Bishop-Dauxois model. The denaturation rate of double stranded polymers decreases exponentially as function of length below the denaturation temperature. Above Tc, the rate shows a minimum, but then increases as function of length. We also examine the influence of sequence and solvent friction. Molecules having the same number of weak and strong base-pairs can have significantly different opening rates depending on the order of base-pairs. . s c i s y h p [ 1 v 5 8 5 5 . 4 0 2 1 : v i X r a For decades, experimental and theoretical scientists have been fascinated by the thermal DNA denatura- tion [1]. It is biologically relevant since the opening of the double helix in an important step for the transcrip- tion of the genetic code [2] but it is also becoming im- portant for nanotechnology as DNA is now used for its self-assembly properties [3], to create nanodevices [4] or to design molecular memories [5]. The different AT and GC base-pairs (bps) are intra-connected by two and three hydrogen-bonds, respectively. Denaturation experiments, in which UV absorbance is measured as function of tem- perature, show discrete steps associated to the sequential opening of soft and stiff regions in the DNA sequence. These regions mainly differ by their density of weak AT- versus strong GC-bps, but also the specific order is im- portant [6]. On the other hand, large homogeneous chains denature in a single step within a very small tempera- ture interval, a remarkable realization of an effectively one-dimensional phase transition [7]. Theoretically, the statistics of DNA have been modeled via Ising type models [8] in which each bp is given a value 0 or 1 depending on whether it is open or closed. This approach allows the study of extremely large sequences but cannot be used to study dynamics. This has been the principle motivation for Peyrard, Bishop and Daux- ois to develop a continuous model (PBD) [9]. The PBD model is computationally somewhat more expensive than the Ising type models, but still quite efficient due to its mesoscopic character. As a result, the PBD model is very well suited to study dynamics of reasonably long DNA molecules at timescales that are not reachable with full- atom simulations [10]. The PBD model, therefore, meets our requirements that we need to study the dynamics of the DNA denaturation transition. In specific, we want to understand how the rate of full denaturation depends on several parameters such as length, sequence, solvent fric- tion, and temperature. The PBD model has revealed a rich spectrum of dy- namical phenomena such a the existence of nonlinear lo- calized modes [11]. Ironically, most studies on the PBD model that apply specific types of dynamics, such as Nose- Hoover, Brownian motion, and Langevin (sometimes even using a configurational dependent friction parameter [12]), have reported on equilibrium properties like the fraction of open bp's at a given temperature. These properties, in principle, do not depend on the dynamics [13]. So far, the study on actual dynamics has been limited to the local opening of DNA [15] or denaturation at elevated temper- atures [14]. For applications such as nanodevices [4] on molecular memories [5], the timing becomes important. The dynamics of full denaturation is difficult to access using standard MD as it is a rare event on the timescale that is achievable by molecular simulations. Transition interface sampling (TIS) [16] and the more recent Replica Exchange TIS [17] are powerful techniques to beat this timescale problem. However, a very efficient implementa- tion of the reactive flux (RF) method [18] can treat the PBD model even more efficiently [17]. The RF method starts from the Transition State Theory (TST) expression, but corrects for correlated recrossings via a dynamical p-1 Titus S. van Erp1 and Michel Peyrard2 factor, the transmission coefficient. Unlike TST, the RF method provides an exact expression for the rate constant that is independent to the choice of reaction coordinate (RC). In Ref. [17], we showed how the free energy and the transmission coefficient can be calculated very efficiently by exploiting specific peculiarities of the PBD model. Us- ing this approach, we can achieve rates far below 10−15 ns−1 or, in other words, relaxation times of several days or even years. We are not aware of any mesoscopic model- ing approach that is able to reach such timescales. In this letter, we apply this method to determine the dependence of the denaturation rate of double stranded DNA as func- tion of its length, the temperature and solvent friction, the content of weak AT versus strong CG bp's, and their specific order. The PBD describes the DNA molecule as an one- dimensional chain of effective atom compounds yielding the relative base-pair separations yi from the ground state positions. The total potential energy U for an N base-pair i=2 Vi(yi) + DNA chain is given by U ({yi}) = V1(y1) + PN W (yi, yi−1) with Vi(yi) = Di(cid:16)e−aiyi − 1(cid:17)2 W (yi, yi−1) = 1 2 K(cid:16)1 + ρe−α(yi+yi−1)(cid:17)(yi − yi−1)2 (1) The first term Vi is the Morse potential describing the hydrogen bond interaction between bases on opposite strands. Di and ai determine the depth and width of this potential for the AT and GC base-pairs. Note that the PBD model does not distinguish between A- and T- nor between G- and C-bases. The second term W is the stacking interaction. The ρ-term makes that the effective strength of the stacking interaction drops from K(1 + ρ) down to K whenever either yi or yi−1 becomes signifi- cantly larger than 1/α. This effect mimics the decrease of overlap between π- electrons when one of two neighboring bases move out of stack and it is thanks to this that the sharp phase transition in long homogeneous chains can be reproduced. Our results are primary focused on the parameter set of Campa and Giansanti [19] with K = 0.025 eV/A2, ρ = 2, α = 0.35 A−1, DAT = 0.05 eV, DGC = 0.075 eV, aAT = 4.2 A−1, aGC = 6.9 A−1. However, we will also shortly investigate the more recent parameter set by Theodorakopoulos [20]: K = 0.00045 eV/A2, ρ = 50, α = 0.2 A−1, DAT = 0.1255 eV, DGC = 0.1655 eV, aAT = 4.2 A−1, aGC = 6.9 A−1. Particular of this data-set is the very high ρ value which should reflect the large dif- ference in persistence length of single stranded and double stranded DNA [21]. λ ≡ min[{yi}] was chosen as reaction coordinate RC [17] and y0 = 1 A as the opening threshold. Henceforth, yi > y0 implies that base-pair i is open and λ > y0 that the complete molecule is denatured. The RF method expresses the overall reaction rate as an equilibrium probability density to be at a surface on the barrier (here defined by λ({yi}) = y0), under the condition that the system is at the reactant side of this surface, times a dynamical transmission factor. k = P (λ = y0λ ≤ y0) × R with R = D λθ( λ)hb 0,y0Eλ=y0 (2) The dynamical factor R is called is the unnormalized transmission coefficient. The brackets with subscript de- note an ensemble average that is constrained on the sur- face λ({yi}) = y0. It is calculated by releasing dynamical trajectories forward and backward in time starting from a proper equilibrium ensemble of configuration points on this surface and Maxwell-Boltzmann distributed veloci- ties. hb 0,y0 is a binary function that is 1 if the backward trajectory from such a point evolves to the minimum of the Morse potential (λ = 0 A) without recrossing the ini- tial surface (λ = y0). Otherwise it is 0. Since this surface is at the frontier of what is considered to be the product state, no forward trajectory is needed as in Ref. [23], since λ({yi}) = y0 and λ > 0 implies that the product state will be entered within an infinitesimal small time step. This procedure is sketched in fig. 1. Assuming hb 0,y0 = 1 whenever λ > 0 would result into the TST approximation of the reaction rate. In practice, hb 0,y0 = 0 for the vast majority of trajectories though its average value remains measurable for this RC due to the flat plateau of the Morse potential. The time duration of these trajectories are much shorter than the actual re- laxation time that relates to the long time evolution of the system having relatively small oscillations. Therefore, even though the very large fluctuations of these short tra- jectories are at the limit of what the PBD can describe accurately, their short duration, with respect to the ac- tual relaxation time, make quantitative rate calculations reliable. Also the fact that we stop our trajectories when- ever yi > y0 for all i is reasonable. Besides that the PBD model isn't too accurate beyond this point, it is expected that the chance of reclosing is even smaller in a more accu- rate three-dimensional model. We can, therefore, assume that the system is committed to the denatured state be- yond this point and it is therefore sufficient to stop the trajectory whenever the y0 transition surface is crossed. Still, it is fair to say that our approach and RC will prob- ably not work for more elaborated potentials that contain explicit potential barriers for reclosing [24] since it will result in R ≪ 1. The equilibrium density is related to the free energy via F (λ) = −kBT ln P (λ) with kB the Boltzmann constant and T the temperature in Kelvin. The surface does not necessarily have to be at the local maximum of the free energy profile, the transition state (TS), though this is generally the most efficient choice since it maximizes the transmission R. In our case the surface λ({yi}) = y0 is slightly beyond the TS at the beginning of the denatu- ration state or product state region. This has, however, no effect on the final results since the transmission co- efficient and probability density are like communicating p-2 0.12 0.10 ) V e ( V 0.08 0.06 0.04 0.02 0.00 -1 λ y 0 0 1 2 y(A) 3 4 5 Fig. 1: (color online) The above figure illustrates an example configuration (solid red spheres) at the surface λ({yi}) = y0 that was generated by our algorithm for a homogeneous N = 8 AT-chain. It has one particle that is exactly at y0 while yi > y0 for all others. The arrows envision the corresponding Maxwellian velocities. Each of these sampled phases points will give a single value x that is either zero or positive. If λ (which is simply the velocity of the particle that is exactly on the surface) equals less than zero, then x = 0. If λ > 0, all velocities of the system will be reversed and the dynamics are propagated using the Langevin dynamics. These dynamics are continued until one particle reaches the well (yi < 0 for one particle or λ < 0) or until all particles move beyond the initial surface (yi > y0 for all particles or λ > y0). In the last case x is assigned zero again. In the first case x equals the initial velocity λ. The final configuration is shown by the green open circles. A particle at the end of the chain has moved inside the well of the Morse potential (hence x = λ for this case). This event implies a commitment to the natured state; this particle will quickly pull the others inside the well and a rapid recrossing with the y0 surface is highly unlikely after this point. The unnormalized transmission coefficient is the average of the sampled values x: R = ¯x. vessels; a too high value for P (or too low value of ∆F ) is compensated by a lower transmission R. Therefore, the calculated rate values are insensitive to the exact location of the transmission surface. Because of the first neighbor character of the PBD model, the free energy or probability density P (λ = y0λ ≤ y0) can be computed very efficiently by means of an iter- ative numerical integration scheme [17]. P = Pi R dy1 . . . dyN δ(yi − y0)Qj6=i θ(yj − y0)e−βU({yi}) (3) R dy1 . . . dyN (1 − Qk θ(yk − y0))e−βU({yi}) Now, as the integrals of Eq. (3) are all of a special facto- rial form [22], we can apply the direct integration method of Ref. [22] using an integration step of dy = 0.05 A and integration boundaries yi−yi−1 < d = p2 ln τ/βK and −1/aAT lnh p ln τ/βDAT + 1i < yi < y0 + √N d. The tolerance τ is set to 10−40 which implies that all contri- butions of e−βV ({yi}) lower than this value are neglected. The very low value of K in the parameter set of [20] The dynamics of the DNA denaturation transition with respect to [19] implies significant larger integration boundaries and a 50 times higher computational cost. In the next step, we need to generate a representative set of configurations on the surface λ({yi}) = y0. Also this can be achieved very effectively for this particular model and RC. We make use of the fact that ensemble constraint to this surface is identical to a freely mov- ing chain on a translational invariant potential U ′ with U ′({yi}) ≡ U ({yi− λ({yi}) + y0}) [22]. Therefore, we gen- erate the required surface points by running a MD simu- lation using potential U ′, save every 1000th time step to dissolve correlations. Then, we shift these configurations back to the surface λ({yi}) = y0. From each point, we release a backward trajectory using normal potential U and determine λθ( λ)hb 0,y0. We averaged over 106 trajecto- ries using a time step of 1 fs and bp masses of 300 amu. The temperature was controlled using Langevin dynam- ics with a friction coefficient of γ = 50 ps−1 unless stated otherwise. It is important to stress that factor P is purely a ther- independent of γ or masses) modynamic property (i. e. while R depends on the full dynamics. This methodol- ogy allows us for the first time to examine the theoretical PBD denaturation rates of long DNA molecules as func- tion of sequence, solvent friction, temperature, and model parameters. Fig. 2 shows the calculated denaturation rates of homo- T=300, 305, 310, ..., 350 K -2 10 -6 10 -10 10 0 80 160 ) s n 1 ( k / 3.0 2.0 1.0 κ 0.0 0.04 0.03 0.02 0 20 40 60 80 100 120 140 160 180 200 N Fig. 2: (color online) Top: Denaturation rate of homogeneous AT chains as function of the number of bps at different tem- peratures. Inset shows the same results in log-scale. Bottom: The corresponding transmission coefficients. geneous A/T-DNA sequences of different lengths and tem- peratures. Below the critical denaturation temperature (Tc ≈ 325K) the denaturation rate of the DNA molecule is exponentially decreasing as function of its length. Above Tc, there is an initial decrease as function of the number of bps N , but then starts to increase again. Both the depth and the position of the minimum decrease as function of temperature. This feature is mainly an effect of equilib- rium statistical physics (factor P ) rather then dynamics (factor R) as we can conclude from the lower panel of fig. 2. This panel shows the normalized transmission coef- ficient κ ≡ R/D λθ( λ)E = √2πβmR which is equivalent to p-3 Titus S. van Erp1 and Michel Peyrard2 the true rate constant divided by the transition state the- ory (TST) value. The results show that, although κ is far smaller than one, it is more or less constant with respect of temperature and chain length. Only for short polymers N < 20 there is a noticeable upturn of κ. Henceforth, despite that TST overestimates the rate by more than a factor 40, the relative TST rates are approximately cor- rect. An increase of denaturation rate as function of its length is remarkable even for T > Tc. For macroscopic systems having two phases A and B, crossing the phase transi- tion temperature coincides with a discontinuous jump of the equilibrium constant from zero to infinity. Since the equilibrium constant is related to the ratio of forward and backward rate constants, kA→B/kB→A goes from zero to infinity in the limit N → ∞ when T crosses Tc. Still, this does not imply that the absolute values of any of the two rates should increase as function of N . In fact, the decreasing denaturation time τ ∼ 1/k as function of N is in contrast to the power-law scaling τ ∝ N 2.57 of a non- equilibrium case study [14]. To examine the effect of temperature on the equilib- rium statistics, fig. 3 shows the free energy F (λ′) = 0.5 0.4 0.3 0.2 0.1 0.0 -0.1 0.08 0.04 0.00 ) V e ( F N=20,40,...,100 T=300 K T=350 K N=20,40,...,100 -0.04 -0.4 -0.2 0.0 0.2 0.4 o λ(A) 0.6 0.8 1.0 Fig. 3: (color online) The free energy profile as function of the RC for homogeneous A/T chains of different lengths and temperatures T =300 K (top) and T =350 K (bottom). − ln(cid:2)P (λ = λ′λ < y0)/P 0(cid:3) /β with P 0 = 1/A being an arbitrary reference density. At temperature T = 300 K the free energy barrier is clearly increasing as function of chain length while it is slightly decreasing at T = 350 K. The longer polymers profit from the larger entropy gain at high T . In fig. 4, we examined the effect of heterogeneities in the DNA sequences at room temperature conditions (T = 300K). The red and blue curve represent the denatu- ration rate for the homogeneous A/T and G/C-sequences. The green curve in the middle corresponds to sequences where the first half is purely A/T and the other half G/C. All curves become straight lines for sufficiently large N which implies an exponential dependence on the chain length. Linear fits in the log-plot reveal that k ≈ abN with a(1/ns), b = 0.240, 0.897 for the A/T- and 0.436, 0.678 for the G/C-chain. The mixed AG-sequence has 0/40 10/30 #A/#G 20/20 1.0 -4 10 -8 10 ) s n / 1 ( k -12 10 -16 10 0 40 30/10 40/0 -2 10 AA -4 10 -6 10 k ( 1 / n s ) AG GG 80 120 160 200 N -8 10 Fig. 4: (color online) Denaturation rate versus the number bps for homogeneous A/T- (red line) and G/C-chains (blue) and a sequence having N/2 consecutive A's followed by N/2 consecutive G's (green). Crosses correspond to the parameter set of Ref. [19], triangles correspond to Ref. [20]. Blue line indicates the rate (right and top axis) for a mixed 40 bp chain having M consecutive A- followed by 40-M G-bps. Dashed black lines are best linear fits in the log-plot. a = 0.346 and b = 0.778. The latter value is very close to pb(A) × b(G) = 0.780 which suggest k ∝ b(A)NA b(G)NG as a general rule for a mixed sequence where NA and NG are the number of A/T- or G/C-bps. The results based on the alternative parameter set of Ref. [20] show the same trend but predicts denaturation rates that are orders of magnitude lower. The linear fits give a, b = 0.039, 0.639 for A/T, 0.048, 0.240 for GC-, and 0.0249, 0.415 for the mixed chain. The latter is again very close to pb(A) × b(G) which corroborates with the above. In fig. 4 we also show the denaturation rate for a polymer of fixed N = 40 length as function of the A,G-content using parameter set [19]. Like before, the order of the polymer is such that the A- and G-bps are placed consecutively at oppo- site sides. Making the linear fit in the log-plot shows that k ≈ 7.06(1/ns)1.31NA which is again in excellent agree- ment with the above rule since b(A)/b(G) = 1.32. In Fig. 5 we examined the influence of the friction con- stant γ. As the dynamics do not change the equilibrium distribution they can only have an effect on κ. In agree- ment with previous finding, we see that the transmission coefficient is hardly affected by the polymer length, the type of bps and temperature. In conclusion, κ is mainly determined by the friction coefficient γ. At high friction we notice a standard Kramer's behavior [25] κ ∝ 1/γ which is somewhat surprising given the exotic nature of the RC that depends on all yi coordinates in a discontin- uous way. Contrary to Kramer's theory, κ converges to values in the range 0.38 − 0.48 and does not approach 1 in the limit γ → 0. Finally, we also examined the influence of the order of the sequence on the denaturation rate. In fig. 6 we show p-4 0.6 0.5 0.4 κ 0.3 0.2 0.1 AT GC 300 K 350 K 20 40 1 0.1 0.01 0.001 0.01 0.1 1 10 100 1000 0.0 0.0 100. γ(1/ps) 200. 300. Fig. 5: (color online) Transmission coefficient versus the fric- tion coefficient γ for a homogeneous A- (solid line) or G- (dashed line) at temperatures T = 300 K (spheres) or 350 K (triangles) consisting of 20 (green) or 40 (blue) bps. Inset shows the same results in a log-log plot. Dashed black line corresponds to the fit κ(γ) = 1.3 ps−1/γ. normalized denaturation rates as function of N for se- quences that all contain 50% A- and 50% G-bases. The computed rate constant where normalized by the values of fig. 4. Hence, all values are relative to a chain having all A's and all G's at one side of the polymer. We call this the AAGG-sequence after the N = 4 polymer having this characteristics. Besides the AAGG-sequence, we ex- i n a h c G G A A o t e v i t l a e r k 1.3 1.0 0.7 0 10 20 30 40 AGAG 3.0 2.6 2.2 1.8 1.4 1.0 AGGA GAAG 0 20 40 60 100 120 140 80 N Fig. 6: (color online) Relative denaturation rates versus chain length for DNA sequences having 50% A- and 50% G-bps. The curves only differ in the ordering of bps. All values are normal- ized by the chain having all A's at one side and the G's at the other side ("AAGG" after the N=4 bp chain). The alternating sequence AGAG (red) is shown together with the sequences having an A- (blue) or G-block (green) in the middle, respec- tively GAAG and AGGA. Full circles with error-bars show the results of the full calculation while the solid line is the result when transmission coefficients κ would be considered to be identical for all sequences. Inset shows the same results for the alternative parameter set of Ref. [20]. The dynamics of the DNA denaturation transition in two ways i) using the full rates, ii) using the proba- bilities densities P alone. The second approach basically assumes that all transmission coefficients are identical for all sequences which seems to be right considering previous conclusions and has the advantage that its result is not affected by statistical uncertainties. Fig. 6 shows that the AGAG sequence gives significant faster denaturation compared to the AAGG-chain and this effect increases with chain length. The GAAG-chain is second best being approximately 13% faster than AAGG for N > 20 without further increase as function of chain length. Conversely, the AGGA-sequence is about 10% slower than AAGG. Naturally, the bases at the end of the polymer are the most easy to open. Apparently, the best way to profit from this effect is to put the bases at the end that are difficult to open otherwise, but an even distribution of the G-bases is by far the best. The results of the alternative parameter set, however, (inset in fig. 6) seems to give a somewhat different con- clusion. The statistical error-bars are much larger for this parameter set due to the very weak coupling K which increases the accessible phase-space. Still, AGGA and GAAG seem to be reversed regardless method and i) or ii) is applied. The AGGA seem to denature even faster than the AGAG-chain though this might be just a statistical fluctuation. If we consider the method ii), we see that the red curve starts to grow as function of N and surpasses the AGGA result at N = 40. Therefore, we believe that the alternating sequence has the fastest denaturation irrespec- tive of the parameter set for sufficiently large N . Besides their interest for studies involving real devices, those re- sults on the effect of the sequence are important because they provide experimentally testable data [26] that dis- criminate between two parameter sets that give very simi- lar equilibrium properties. This offers a unique possibility to validate model parameters. In conclusion, we have utilized an innovative approach to study the denaturation rate as function of different pa- rameters. This has resulted in several predictions which can be tested experimentally. DNA models, giving similar results regarding statistics, can differ fundamentally when denaturation are concerned. Therefore, our methodology in combination with experiments provides an additional dimension to probe the validity of DNA models and to improve them. This will be a significant step forward in understanding the sequence-dependent dynamics of DNA which play a crucial role in several biological processes and new DNA-based devices for which the time response is important. ∗ ∗ ∗ amined the alternating sequence AGAG, the "weak-block in the middle" sequence GAAG, and the "strong-block in the middle" sequence AGGA. The results were computed TSvE acknowledges the Methusalem funding (CASAS) by the Flemish government. p-5 Titus S. van Erp1 and Michel Peyrard2 REFERENCES [1] Thomas R., Gene, 135 (1993) 77. [2] Benham C. J., Proc. Natl. Acad. Sci. USA, 90 (1993) 2999. [3] Goodman R. P, Schaap A. T., Tardin C. F., Erben C. M., Berry R. M., Schmidt C. F. and Turberfield A. J., Science, 310 (2005) 1661-1665 [4] Komiya K., Yamamura M. and Rose J. A., Nucleic Acids Res., 38 (2010) 4539. [5] Takinoue M. and Suyama A., Chem-Bio Informatics J., 4 (2004) 93; Takinoue M. and Suyama A, Small, 2 (2006) 1244. [6] Dornberger U., Leijon M. and Fritzsche H., J. Biol. Chem., 274 (1999) 6957. [7] Inman R. B. and Baldwin R. L., J. Mol. Biol., 8 (1964) 452; Peyrard M. and Dauxois T. , Math. Comp. Sim., 40 (1996) 305. [8] Poland D. and Scheraga H. A., J. Chem. Phys., 45 (1996) 1456. [9] Dauxois T., Peyrard M. and Bishop A. R., Phys. Rev. E, 47 (1993) R44. [10] V´arnai P. and Zakrzewska K., Nucleic Acids Res., 32 (2004) 4269. [11] Dauxois T., Peyrard M. and Willis C. R., Physca D, 57 (1992) 267; Cuevas J., Archilla J. F. R., Gaididei Y. B. and Romero F. R., Physica D, 163 (2002) 106. [12] Das T. and Chakraborty S., EPL, 83 (2008) 48003. [13] van Erp T. S., Cuesta-Lopez S., Hagmann J.-G. and Peyrard M., EPL, 85 (2009) 68003. [14] Baiesi M., Barkema G. T., Carlon E. and Panja D., J. Chem. Phys., 133 (2010) 154907. [15] Alexandrov B. S., Gelev V., Yoo S. W., Bishop A. R., Rasmussen K. O., Kim O. and Usheva A., PLoS Comput. Biol., 5 (2009) e1000313. [16] van Erp T. S., Moroni D. and Bolhuis P. G., J. Chem. Phys., 118 (2003) 7762. [17] van Erp T. S., Phys. Rev. Lett., 98 (2007) 268301. [18] Frenkel D. and Smit B., Understanding molecular sim- ulation, 2nd ed. (Academic Press, San Diego, CA) 2002. [19] Campa A. and Giansanti A., Phys. Rev. E, 58 (1998) 3585. [20] Theodorakopoulos N., Phys. Rev. E, 82 (2010) 021905. [21] Baumann C., Smith S., Bloomfield V. and Busta- mante C., Proc. Natl. Acad. Sci. USA, 94 (1997) 6185. [22] van Erp T. S., Cuesta-Lop`ez S., Hagmann J.-G. and Peyrard M., Phys. Rev. Lett., 95 (2005) 218104. [23] van Erp T. S., J. Chem. Phys., 125 (2006) 174106. [24] Peyrard M., Cuesta-Lopez S. and James G., Nonlin- earity, 21 (2008) T91. [25] Hanggi P., Talkner P. and Borkovec M., Rev. Mod. Phys., 62 (1990) 251. [26] Bonnet G., Krichevsky O. and Libchaber A., Proc. Natl. Acad. Sci. USA, 95 (1998) 8602; Altan-Bonnet G., Libchaber A. and Krichevsky O., Phys. Rev. Lett., 90 (2003) 138101. p-6
1706.09110
1
1706
2017-06-28T03:04:22
Improving the glial differentiation of human Schwann-like adipose-derived stem cells with graphene oxide substrates
[ "physics.bio-ph" ]
There is urgent clinical need to improve the clinical outcome of peripheral nerve injury. Many efforts are directed towards the fabrication of bioengineered conduits, which could deliver stem cells to the site of injury to promote and guide peripheral nerve regeneration. The aim of this study is to assess if graphene and related nanomaterials can be useful in the fabrication of such conduits. A comparison is made between GO and reduced GO substrates. Our results show that the graphene substrates are highly biocompatible, and the reduced GO substrates are more effective in ncreasing the gene expression of the biomolecules involved in the regeneration process compared to the other substrates studied.
physics.bio-ph
physics
Improving the glial differentiation of human Schwann-like adipose-derived stem cells with graphene oxide substrates Andrea Francesco Verre,1 Alessandro Faroni,2 Maria Iliut,1 Claudio Silva,1,3 Cristopher Muryn,4 Adam J Reid 2,5 and Aravind Vijayaraghavan1,* 1 School of Materials and National Graphene Institute, University of Manchester, Manchester M13 9PL, UK 2 Blond McIndoe Laboratories, Division of Cell Matrix Biology and Regenerative Medicine, School of Biological Sciences, Faculty of Biology Medicine and Health, University of Manchester, Manchester Academic Health Science Centre, Manchester M13 9PL 3 Department of Fundamental Chemistry, Institute of Chemistry, University of São Paulo, São Paulo, Brazil. 4 School of Chemistry, University of Manchester, Manchester M13 9PL, UK 5 Department of Plastic Surgery & Burns, University Hospitals of South Manchester, Manchester Academic Health Science Centre, Manchester * Corresponding author: [email protected] Keywords: adipose stem cells, graphene, glial differentiation Abstract There is urgent clinical need to improve the clinical outcome of peripheral nerve injury. Many efforts are directed towards the fabrication of bioengineered conduits, which could deliver stem cells to the site of injury to promote and guide peripheral nerve regeneration. The aim of this study is to assess if graphene and related nanomaterials can be useful in the fabrication of such conduits. A comparison is made between GO and reduced GO substrates. Our results show that the graphene substrates are highly biocompatible, and the reduced GO substrates are more effective in increasing the gene expression of the biomolecules involved in the regeneration process compared to the other substrates studied. 1. Introduction Schwann cells (SC) are key cellular elements in assisting the regeneration of peripheral nerve after injury. SC switch from a myelinating to repair phenotype which results in increased expression of extracellular matrix (ECM) proteins, neurotrophins and growth factors; furthermore, SC undergo profound morphological changes which result in upregulation of filament cytoskeletal proteins such as nestin and actin [1,2]. Despite a clear need for novel therapies, use of SC as a clinical intervention for peripheral nerve injury (PNI) is problematic due to the necessity of harvesting a functional nerve and the limited expansion capacity of SC. As a clinically viable alternative, mesenchymal stem cells and adipose-derived mesenchymal stem cells (ASCs) have been differentiated in vitro towards a Schwann-like cells phenotype [3,4]. These differentiated adipose stem cells (dASCs) express glial markers such as glial fibrillary acidic protein (GFAP), S100 and p75 [4]; express myelin protein [5] and myelin structures when in co-culture with neurons [6,7]; and when implanted in bioengineered conduits to repair murine peripheral nerve gap in vivo, dASC have demonstrated promotion of nerve regeneration, reduction of muscle atrophy, increased nerve conduction velocity and higher rates of myelination [8-11]. Graphene and related nanomaterials can play an important role in the fabrication of bioengineered nerve conduits for the treatment of peripheral nerve injuries. Although the biocompatibility of these materials for in vivo studies depends on many variables such as the thickness, the lateral size of the flakes, the level of hydrophilicity and the extent of functionalization [12], it can be stated that when these materials were used as coatings on surfaces to support stem cell growth, the extent of cytotoxicity was limited and enhanced stem cell differentiation was reported [13-17]. Graphene and related nanomaterials were found to be effective in positively modulating axonal outgrowth and nerve regeneration in vitro [18-21]. Thus far, researchers have been exploring the effect of graphene and related nanomaterials on the neurite outgrowth, but there have not been studies regarding the effect of these materials in supporting the growth of dASCs. The aim of this study is then the biological characterization of graphene oxide (GO) and reduced GO (rGO) coated coverslips and verify if these materials are able to sustain the differentiated dASCs phenotype, which is rapidly lost following withdrawal of growth factors. 2. Materials and Methods 2.1 Graphene based materials synthesis, substrates preparation and characterization Graphite oxide was synthesized by a modified Hummer`s method [22,23] and exfoliated down to constituent monolayers to yield GO. Glass coverslips were washed in a sonication bath with Decon® 90 for 15 mins followed by deionised (DI) water for another 15 mins and finally with isopropanol for 15 mins. After the coverslips were dried, they were treated for 5 mins in oxygen plasma to increase the hydrophilicity of the coverslips. GO dispersions were then spin-coated on the glass coverslips at the concentration of 2 mg/ml at 2500 rpm, 250 rpm/sec acceleration for 2 mins. To obtain rGO coverslips, GO coverslips were kept for three days at 180 °C in vacuum. Atomic force microscopy (AFM) measurements were carried using a Bruker FastScan microscope in tapping mode. Raman spectroscopy was performed using a Renishaw inVia Raman microscope with 532 nm laser excitation. X-ray photoelectron spectroscopy (XPS) spectra of drop-casted substrates were recorded with a SPECS NAP-XPS system employing a monochromatic Al Kα source (1486.6 eV). Before the biological experiments, the coverslips were sterilised by immersion in pure ethanol then washed in deionised water and then left to dry under the hood. 2.2 Human Adipose stem cells harvesting and differentiation ASCs were isolated according to a previously reported protocol [4]. Human abdominal, subcutaneous adipose tissue was harvested from three female surgical patients undergoing reconstructive surgery at the University Hospital of South Manchester, UK. All patients were fully consented and procedures approved by the National Research Ethics Committee, UK (NRES 13/SC/0499). Adipose tissue biopsies were minced by a razor blade and dissociated by an enzymatic treatment of 0.2 % (w/v) of collagenase I (Life Technologies, Paisley, UK) for 60 minutes at 37 °C under constant agitation. The digested tissue was then filtered through a vacuum-assisted 100 µm nylon mesh (Merck Millipore, Watford, UK). An equal volume of stem cell growth medium containing a-minimum essential Eagle's medium (aMEM) (Sigma- Aldrich, Poole, UK), 10% (v/v) foetal bovine serum (FBS) (LabTech, Uckfield, UK), 2 mM L-glutamine (GE Healthcare UK, Little Chalfont, UK), and 1% v/v) penicillin–streptomycin was added. The solution were centrifuged at 300 g for 10 minutes and the resulting pellet was suspended in 1 mL of Red Blood Cell Lysis Buffer (Sigma-Aldrich) for 1 min, and 20 mL of aMEM was added to arrest lysis. The mixture was centrifuged at 300 g for 10 min, and the resulting pellet was resuspended in aMEM and plated in T75 flasks for cell culture. Cells were routinely characterised for the expression of stem cell surface markers as per [24] The differentiation of ASC towards dASCs was performed following a previously reported protocol [4]. Briefly, ASCs at the passage 1-2 at 30% of confluence were treated with 1 mM -mercaptoethanol (sigma-Aldrich) for 24 hours, then with 35 ng/mL of all-trans-retinoic acid for 72 hours. After this initial treatment, ASCs were treated by 5 ng/mL of platelet-derived growth factor (Peprotech EC, London,UK), 10 ng/mL basic fibroblast growth factor (Peprotech EC), 14 μM of forskolin (Sigma-Aldrich) and 192 ng/mL glial growth factor (GGF-2) (Acorda Therapeutics, Ardsley, NY, USA). ASCs were kept under these conditions for 2 weeks, replacing media every 72 hours and passaging when confluence was reached. 2.3 Cell Proliferation and Live/Dead assays To assess cell proliferation rate by the (3-(4,5-dimethylthiazol-2-yl)-5-(3- carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium) (MTS) assay , dASCs cells were plated at the concentration of 5,000 cells per coverslip in triplicate on the sterilised graphene- coated coverslips. The coverslips were put in ultra-low adherence 24-well cell plates to avoid attachment of the cells to the tissue culture plastic underneath the coverslips. At days 1, 4 and 7 after seeding the cell medium was aspirated and cells were washed in phosphate buffered saline (PBS). After the washing step, the cells were incubated in 20% (v/v) CellTiter 96 Aqueous One Solution Cell Proliferation Assay (Promega, Southampton, UK), diluted in phenol-free DMEM (Sigma-Aldrich) for 90 mins in the dark at 37 °C. After the incubation, the absorbance at 490 nm was recorded using an Asys UVM-340 microplate reader/spectrophotometer (Biochrom, Cambridge, UK). For the viability assays, ASC cells were plated at the concentration of 25,000 cells per coverslips in triplicate in a cell medium containing α-MEM and 1% (v/v) P/S without growth factors. After 48 hours, the medium was aspirated and the cells were washed in PBS. Calcein-AM fluorescein and ethidium homodimer- 1 (Eth-D1) purchased from LIVE/DEAD® Viability/Citoxicity Kit (Molecular Probes, Invitrogen, UK) were added at the concentration of 0.5 and 2 μg/ml respectively in PBS and left to react for 15 minutes at 37º C. After this step, images were taken using a fluorescence inverted microscope (Olympus IX51, Japan) under 4X magnification. Data from MTS experiment were expressed were expressed as absorbance at 490 nm ± SE of the mean (n = 3), while data from viability assay were expressed as percentage of live cells measured by dividing the average green stained area by the average of the whole (Live + Dead) stained area, measured with Image J software (version 1.48) multiplied by 100. 2.4 Quantitative real-time polymerase chain reaction (qRT-PCR) For gene expression studies, cells were seeded as above at a concentration of 50,000 cells per coverslips in triplicate and the RNA was extracted after 48 hours of cellular growth on the different coverslips. RNA was extracted using the RNeasy Plus Mini Kit (Qiagen) following the instruction of the manufacturer. The concentration of the RNA was quantified at the NanoDrop ND-100 (Thermo Fisher Scientific, Waltham, MA, USA) spectrophotometer. 1 µg of each sample were reverse transcribed using the RT2 First Strand Kit (Qiagen) following the instruction of the manufacturer. DNA elimination steps were included in both RNA extraction and cDNA synthesis to prevent downstream genomic DNA amplification. qRT-PCR was performed with RT2 SYBR Green qPCR Mastermix (Qiagen) and a Corbett Rotor Gene 6000 Qiagen), by the use of the following protocol: hot start for 10 min at 95 °C, followed by 40 cycles of 15 s at 95 °C, annealing for 30 s at 55 °C, and extension for 30 s at 72 °C. To verify the specificity of the reactions, a melting curve was obtained with the following protocol: 95 °C for 1 min, 65 °C for 2 min, and a gradual temperature increase from 65 °C to 95 °C (2 °C/min). Data were normalized for the housekeeping gene, and the ΔΔCt method was used to determine the fold changes in gene expression with glass coverslips as controls. The primer assays were obtained from QIAGEN as reported in the literature. [24] 2.5 Statistical analysis Statistical significance of the studies was evaluated by the use of GraphPad Prism 6.0 (GraphPad Software, La Jolla, CA, USA) using a one-way ANOVA test followed by Dunnett's multiple comparison test using glass as a control sample. Level of significance was expressed as P-values. 3. Results and Discussion 3.1 Substrates Characterization Optical microscopy and AFM showed the uniformity of thin film coverage on the substrate and no empty areas were observable on the substrates as shown in Fig. 1a-d. Raman spectroscopy is a useful tool to detect the presence of graphene oxide on any surfaces. The typical spectrum of GO is composed by two peaks: the D peak at ~1350 cm-1 and the G peak at ~1586 cm1 The intensity ratio (ID/IG) between these two peaks is employed to characterize GO dispersions. The measured ID/IG of GO was 0.92 while upon thermal reduction we noticed a decreased ID/IG down to 0.84 (Fig. 1e-f). XPS C1s spectra of all the different substrates can be deconvoluted into 6 components: C=C (sp2 carbon) at 284.6 eV, C-C (sp3 carbon) at 285.1 eV, C-OH at 286 eV, C-O-C at 286.9 eV, C=O at 287.7 eV and HO-C=O at 288.8 eV (Fig.1g) [25,26]. Successful reduction is confirmed by the C1s spectrum of rGO: we noticed a clear decrease in all the oxygen functionalities with the exception of hydroxyl group, a decrease in sp3 carbon and increased sp2 carbon as shown in Fig.1h. As previously reported [14], graphene-based materials are efficient materials for the fabrication of artificial scaffolds in regenerative medicine. In fact, graphene and related nanomaterials are characterised by ultra-high specific surface area and by the ability of binding stem cells growth inducers both covalently and non- covalently acting as a pre-concentration platform for growth factors and other biomolecules present in the differentiation medium. Starting form this observation, human dASCs were cultured on the substrates to assess cell proliferation and the effect of the substrate on the gene expression of glial markers. 3.2 Proliferation of dASCs on graphene substrates. To assess dASCs proliferation rate on the different substrates studied we used MTS assay. We selected three time-points on day 1, day 4 and day 7 after seeding. As it can be seen on Fig. 2a, the proliferation rate of dASCs on both GO and rGO substrates was comparable to glass coverslips used as controls with values only marginally lower at each time points. To assess the biocompatibility of the substrates and confirm that the slightly reduced proliferation was not due to GO/rGO cytotoxicity we performed a live/dead viability assay after 48 hours of cellular growth on the different substrates. Indeed at each time point, we measured that 99.89 ± 0.01 % of cells were alive on GO substrates, 99.80 ± 0.02 % of cells were alive on rGO coverslips and finally 99.87 ± 0.02 % of cells were alive on glass substrates. We can therefore conclude that although the proliferation rate was marginally slower on the rGO and GO coverslips, the amount of live cells was very high and comparable in all the different substrates studied. We then decided to study the gene expression of crucial proteins and growth factors which are key features of the differentiated state of dASC and are involved in the peripheral nerve regeneration process. 3.3 Gene Expression Studies Key molecules involved in the regeneration process are neurotrophins. These growth factors are involved in neuronal survival, development and functionality [27,28,29]. We decided to focus our attention studying brain-derived neurotrophic factor (BDNF), nerve growth factor (NGF) and glial-derived neurotrophic factor (GDNF). Also another group of protein molecules involved in the regeneration process are intermediate filament proteins such as nestin and vimentin, which are strictly related to the profound morphological changes associated with SC response to injury. Nestin is a protein, which is involved in the axonal growth and is normally up regulated after nerve injury [30,31,32,33]. Moreover a recent study highlighted that nestin-positive hair- follicle pluripotent stem cells were able to promote peripheral nerve regeneration [34]. Interestingly, a bigger proportion of ASC were found to express nestin filaments compared to bone marrow mesenchymal stem cells [4]. Vimentin is reported to be up regulated during peripheral nerve regeneration and higher expression levels of this protein have been reported to augment the peripheral nerve regeneration [35]. Lastly we decided to investigate also the gene expression of neurotrophins receptor such as TrkB, TrkC and Ret as a measure of growth factor responsiveness of dASCs. GDNF expression is increased on rGO coverslips (1.64 ± 0.06 vs glass, p< 0.01, n=3) and no statistical difference is observed between GO coverslips and the glass controls. The expression of BDNF is increased on rGO and GO coverslips (1.53 ± 0.08, p  0.01 vs glass, n=3 on rGO substrates) and (1.68 ± 0.01, p<0.001 vs glass, n=3 on GO substrates). NGF expression increased on rGO coverslips (1.67 ± 0.14, p-value  0.01) while GO substrates behaved as glass controls. (Fig. 3 a-c) Nestin expression is increased on both rGO and GO substrates (1.44 ± 0.05, p<0.01 vs glass, n=3 on rGO substrates) and (1.31 ± 0.11, p<0.05 vs glass, n=3 on GO substrates). Vimentin expression is increased on both rGO and GO substrates (1.66 ± 0.04, p 0.001 vs glass, n=3 on rGO substrates) and (1.67 ± 0.01, p<0.001 vs glass, n=3 on GO substrates. (Fig. 3 d-e) TrkC expression is increased on rGO substrates (2.86 ± 0.29, p 0.01 vs glass, n=3) while no statistical difference is observed between GO substrates and the glass controls. The same trend is followed by the gene expression of TrkB receptor. The expression of this gene is increased on rGO substrates (1.85 ± 0.02, p<0.01 vs glass, n=3). The expression of Ret receptor is marginally increased on rGO substrates although not statistically significant due to the high variability of the rGO substrates. The expression of this gene on GO shows the same behaviour as the glass controls. (Fig. 3f-h). Faroni et al. [24] studied the gene expression changes when ASCs are differentiated towards Schwann-like dASCs. The gene expression of GDNF, BDNF, TrkC, Ret, nestin and vimentin markers was reported to be upregulated after the differentiation protocol compared to non- differentiated ASC. The expression of NGF and TrkB was reported to be downregulated after the differentiation protocol compared to non-differentiated ASC, although the level of NGF protein was found to be increased in the dASC phenotype. The development of a protocol that permanently differentiates ASC cells into dASC is crucial to implement a stem-cell based therapy strategy for peripheral nerve regeneration. dASC cells were found able to express glial markers and to promote nerve regeneration, myelination and to increase the speed of the conduction in the nerve [8-11]. The main obstacle in the clinical translation of dASC is the maintenance of the differentiated phenotype. Faroni et al [24] proved that after the withdrawal of the growth factors, dASC started to decrease the expression of glial markers and to reverse into ASC phenotype. There is consequently the need to develop better differentiation protocols and to test new materials that help maintaining the dASC phenotype even after the withdrawal of the growth factors for efficient delivery of stem cell therapies in vivo. Graphene and related nanomaterials have been widely reported as suitable material to support stem cell growth and differentiation [13-17]. In this study, the proliferation rate and the biocompatibility of these substrates were studied by two different assays and we can conclude although there is a slower proliferation rate of dASC on rGO and GO substrates, the amount of live cells is comparable between all the different substrates studies indicating that GO and rGO substrates do not cause cytotoxicity after 48 hours. Importantly, the analysis of gene expression of important glial markers increased after 48 hours of cellular growth on GO and rGO substrates. The expression of neurotrophins and their receptors is statistically increased on rGO substrates and to a lesser extent on GO substrates. Moreover, the expression of intermediate filament proteins such as nestin and vimentin is statistically increased on both and rGO substrates. Park et al. reported increased the neuronal differentiation of neural stem cells on graphene substrates together with decreased expression of glial cells [36]. The contrast between our observations and previous results from Park et al. can be explained by the presence of laminin on the surface of the graphene coverslips. Laminin is a protein of the extracellular matrix which is reported to increase neuronal differentiation of embryonic and neural stem cells [37,38]. Previous results on ASC differentiation on GO substrates [13] showed enhanced osteogenesis, adipogenesis, and epithelial genesis but no experiment was conducted on rGO or graphene substrates. Our study point out the increased glial differentiation especially on rGO substrates. This is very interesting as specific properties of rGO can be exploitable in the clinical translation of dASC. This is especially important for electrical conductivity as rGO compared to GO is not insulating and allows the possibility of electrically stimulate stem cells even in the presence of neuronal co-culture for peripheral nerve regeneration therapeutic strategy. 4. Conclusions Our results confirm the biocompatibility of the graphene-based substrates and show increased expression of neurotrophins and filament proteins mainly on rGO and GO substrates. These results strongly positions rGO and GO coatings to be used as functional surfaces to increase glial differentiation of ASC at earlier stage. As the initial results on 48 hours are encouraging, further studies need to be conducted to establish if the gene expression of these markers will be increased at longer time points as the entire differentiation protocol required 2 weeks of treatment. Funding AV and AFV acknowledge funding from the Engineering and Physical Sciences research Council (EPSRC) grants EP/G03737X/1 and EP/K016946/1. CS acknowledges funding from Brazilian agency FAPESP (grant 2014/05048-4). AF and AJR are supported by the Hargreaves and Ball Trust, the National Institute for Health Research (II-LA-0313-20003), the Academy of Medical Sciences and the Manchester Regenerative Medicine Network (MaRMN). Acknowledgments Thank you to Mr. Jonathan Duncan and Miss Siobhan O'Ceallaigh (Consultant Plastic Surgeons), and their patients at the University Hospital of South Manchester for donation of adipose tissue, and to Acorda Therapeutics Inc, for kindly providing GGF-2. Figure 1: (a, b) AFM topography image of GO, rGO substrates respectively; (c, d) Optical images of GO and rGO substrates respectively; (e, f) Raman spectra of GO and rGO substrates respectively. (g, h) XPS C1s spectra of GO and rGO coated substrates respectively Fig.2: (a) MTS proliferation assay of dASC cells grown on rGO (purple), GO (red) and glass (blue) substrates at day 1, at day 4 and day 7 time-points; **** p<0.0001 ,*** p<0.001, ** p<0.01 , * P<0.05; (b),(c) and (d) dASC cells after 48 hours of cellular growth on glass, GO and rGO respectively. Calcein AM positive cells are alive cells and they are visible in the image due to green fluorescence. 99.89 ± 0.01 % of cells on GO substrates, 99.80 ± 0.02 % of cells rGO coverslips and 99.87 ± 0.02 % of cells on glass substrates were alive. Scale bar: 100 µm Figure 3: Gene expression after 48 hours of dASCs cellular growth on glass, GO and rGO substrates for the following glial markers:.(a) increased expression of GDNF on rGO substrates; b) increased expression of BDNF on both GO and rGO substrates ; c) increased expression of NGF on rGO; d) increased expression of nestin on both GO and rGO substrates; e) increased expression of vimentin on both GO and rGO substrates; f) increased expression of TrkC receptor on rGO substrates ; g) increased expression of Ret receptor on rGO substrates; h) increased expression of TrkB receptor on rGO substrates. ,*** p<0.001, ** p<0.01 , * p<0.05 and ns= non-significant. Expreriments were performed in triplicate. References: (1) 1. Son YJ, Thompson WJ. 1995 Schwann cell processes guide regeneration of peripheral axons. Neuron 14, 125-132. (doi:10.1016/0896-6273(95)90246-5) 2. Jessen, KR, Mirsky R. 1991 Schwann cell precursors and their development. Glia 4, 2, 185-194 (doi: 10.1002/glia.440040210) 3. Faroni A, Rothwell SW, Grolla AA, Terenghi G, Magnaghi V, Verkhratsky A. 2013 Differentiation of adipose-derived stem cells into Schwann cell phenotype induces expression of P2X receptors that control cell death. Cell Death Dis. 4, e743. (doi: 10.1038/cddis.2013.268) 4. Kingham PJ, Kalbermatten DF, Mahay D, Armstrong SJ, Wiberg M, Terenghi G. 2007 Adipose-derived stem cells differentiate into a Schwann cell phenotype and promote neurite outgrowth in vitro Exp. Neurol. 207, 267-274. (doi: 10.1016/j.expneurol.2007.06.029) 5. Tomita K, Madura T, Mantovani C, Terenghi G. 2012 Differentiated adipose-derived stem cells promote myelination and enhance functional recovery in a rat model of chronic denervation. J. Neurosci. Res. 90, 1392-1402. (doi:10.1002/jnr.23002) 6. Xu Y, Liu L, Li Y, Zhou C, Xiong F, Liu Z, Gu R, Hou X, Zhang C. 2008 Myelin- forming ability of Schwann cell-like cells induced from rat adipose-derived stem cells in vitro. Brain Res. 1239, 49-55. (doi: 10.1016/j.brainres.2008.08.088) 7. Mantovani C, Mahay D, Makam VS, Kingham PJ, Terenghi G, Shawcross SG, Wiberg M. 2010 Bone marrow- and adipose-derived stem cells show expression of myelin mRNAs and proteins. Regen Med. 5, 3, 403-410. (doi: 10.2217/rme.10.15) 8. Georgiou M, Golding JP, Loughlin AJ, Kingham PJ, Phillips JB. 2015 Engineered neural tissue with aligned, differentiated adipose-derived stem cells promotes peripheral nerve regeneration across a critical sized defect in rat sciatic nerve. Biomaterials 37, 242-251. (doi:10.1016/j.biomaterials.2014.10.009) 9. di Summa PG, Kalbermatten DF, Pralong E, Raffoul W, Kingham PJ, Terenghi G. 2011 Long-term in vivo regeneration of peripheral nerves through bioengineered nerve grafts. Neuroscience 181, 278-291. (doi:10.1016/j.neuroscience.2011.02.052) 10. di Summa PG, Kingham PJ, Raffoul W, Wiberg M, Terenghi G, Kalbermatten DF. 2010 Adipose-derived stem cells enhance peripheral nerve regeneration. J Plast Reconstr Aesthet Surg. 63, 1544-1552. (doi:10.1016/j.bjps.2009.09.012) 11. di Summa PG, Kingham PJ, Campisi CC, Raffoul W, Kalbermatten DF. 2014 Collagen (NeuraGen®) nerve conduits and stem cells for peripheral nerve gap repair. Neurosci. Lett. 572, 26-31. (doi:10.1016/j.neulet.2014.04.029) 12. Bussy C, Ali-Boucetta H, Kostarelos K. Safety Considerations for Graphene: Lessons Learnt from Carbon Nanotubes. Acc. Chem. Res. 46, 3, 692-701. (doi: 10.1021/ar300199e) 13. Kim J, Choi KS, Kim Y, Lim KT, Seonwoo H, Park Y, Kim DH, Choung PH, Cho CS, Kim SY, Choung YH, Chung JH. 2013 Bioactive effects of graphene oxide cell culture substratum on structure and function of human adipose-derived stem cells. J Biomed Mater Res A. 101, 3520-3530. (doi: : 10.1002/jbm.a.34659) 14. Lee WC, Lim CH, Shi H, Tang LA, Wang Y, Lim CT, Loh, KP. 2011 Origin of Enhanced Stem Cell Growth and Differentiation on Graphene and Graphene Oxide. ACS Nano 5, 7334- 7341. (doi:10.1021/nn202190c) 15. Yoo J, Kim J, Baek S, Park Y, Im H, Kim J. 2014 Cell reprogramming into the pluripotent state using graphene based substrates. Biomaterials 35, 8321-8329. (doi: 10.1016/j.biomaterials.2014.05.096) 16. Garcia-Alegria E, Iliut M, Stefanska M, Silva C., Heeg S, Kimber SJ, Kouskoff V, Lacaud G, Vijayaraghavan A, Batta K. 2016 Graphene Oxide promotes embryonic stem cell differentiation to haematopoietic lineage. Sci Rep. 6, 25917. (doi: 10.1038/srep25917) 17. Nayak TR, Andersen H, Makam VS, Khaw C, Bae S, Xu X, Ee PL, Ahn JH, Hong B H, Pastorin G, Özyilmaz B. 2011 Graphene for Controlled and Accelerated Osteogenic Differentiation of Human Mesenchymal Stem Cells. ACS Nano 5, 4670-4678. (doi: 10.1021/nn200500h) 18. Li N, Zhang X, Song Q, Su R, Zhang Q, Kong T, Liu L, Jin G, Tang M, Cheng G. 2011 The promotion of neurite sprouting and outgrowth of mouse hippocampal cells in culture by graphene substrates. Biomaterials 32, 9374-9382. (doi: 10.1016/j.biomaterials.2011.08.065) 19. Tang M, Song Q, Li N, Jiang Z, Huang R, Cheng G. 2013 Enhancement of electrical signaling in neural networks on graphene films. Biomaterials 34, 6402-6411. (doi:10.1016/j.biomaterials.2013.05.024) 20. Tu Q, Pang L, Wang L, Zhang Y, Zhang R, Wang J. 2013 Biomimetic Choline-Like Graphene Oxide Composites for Neurite Sprouting and Outgrowth. ACS Appl Mater Interfaces. 5, 13188-13197. (doi: 10.1021/am4042004) 21. Tu Q, Pang L, Chen Y, Zhang Y, Zhang R, Lu B, Wang J. 2014 Effects of surface charges of graphene oxide on neuronal outgrowth and branching. Analyst 139, 105-115. (doi: 10.1039/c3an01796f) 22. Marcano DC, Kosynkin DV, Berlin JM, Sinitskii A, Sun Z, Slesarev A, Alemany LB, Lu W, Tour JM. 2010 Improved Synthesis of Graphene Oxide. ACS Nano 4, 4806-4814. (doi: 10.1021/nn1006368) 23. Hummers WS, Offeman RE. 1958 Preparation of Graphitic Oxide. J. Am. Chem. Soc. 80, 1339-1339. (doi: 10.1021/ja01539a017) 24. Faroni A, Smith RJP, Lu L, Reid AJ. 2016 Human Schwann-like cells derived from adipose-derived mesenchymal stem cells rapidly de-differentiate in the absence of stimulating medium. Eur J Neurosci. 43, 417-430. (doi: 10.1111/ejn.13055) 25. Michio K, Hikaru T, Kazuto H, Shinsuke M, Chikako O Asami F, Takaaki T, Yasumichi M. 2013 Analysis of Reduced Graphene Oxides by X-ray Photoelectron Spectroscopy and Electrochemical Capacitance. Chem. Lett. 42, 924-926. (doi: http://dx.doi.org/10.1246/cl.130152) 26. Ganguly A, Sharma S, Papakonstantinou P, Hamilton J. 2011 Probing the Thermal Deoxygenation of Graphene Oxide Using High-Resolution In Situ X-ray-Based Spectroscopies. J. Phys. Chem. C 115, 17009-17019. (doi: 10.1021/jp203741y) 27. Hempstead BL. 2006 Dissecting the Diverse Actions of Pro- and Mature Neurotrophins. Curr. Alzheimer Res 3, 19-24. (doi: https://doi.org/10.2174/156720506775697061) 28. Reichardt LF. 2006 Neurotrophin-regulated signalling pathways. Philos Trans R Soc Lond B Biol Sci. 361, 1545. (doi: 10.1098/rstb.2006.1894) 29. Lentz SI, Knudson CM, Korsmeyer SJ, Snider WD. 1999 Neurotrophins Support the Development of Diverse Sensory Axon Morphologies. J Neurosci. 19, 1038-1048. 30. Sahin Kaya S, Mahmood A, Li Y, Yavuz E, Chopp M. 1999 Expression of nestin after traumatic brain injury in rat brain. Brain Res. 840, 153-157. (doi: http://dx.doi.org/10.1016/S0006-8993(99)01757-6) 31. Frisén J, Johansson CB, Török C, Risling M, Lendahl U. 1995 Rapid, widespread, and longlasting induction of nestin contributes to the generation of glial scar tissue after CNS injury. J Cell Biol. 131, 453-464. (doi: 10.1083/jcb.131.2.453) 32. Huaser S, Widera D, Qunneis F, Müller J, Zander C, Greiner C, Lüningschrör P, Heimann P, Schwarze H, Ebmeyer J, Sudhoff H, Araúzo-Bravo MJ, Greber B, Zaehres H, Schöler H, Kaltschmidt C. 2011 Isolation of Novel Multipotent Neural Crest-Derived Stem Cells from Adult Human Inferior Turbinate. Stem Cells Dev 21, 5 742-756. (doi: 10.1089/scd.2011.0419) 33. Faroni A, Terenghi G, Magnaghi V. 2012 Expression of Functional γ-Aminobutyric Acid Type A Receptors in Schwann-Like Adult Stem Cells. J Mol Neurosci. 47, 3, 619–630 (doi: 10.1007/s12031-011-9698-9) 34. Amoh Y, Aki R, Hamada Y, Niiyama S, Eshima K, Kawahara K, Sato, Tani Y, Hoffman RM, Katsuoka K. 2012 Nestin-positive hair follicle pluripotent stem cells can promote regeneration of impinged peripheral nerve injury. J Dermatol. 39, 33-38. (doi: 10.1111/j.1346-8138.2011.01413.x) 35. Perlson E, Hanz S, Ben-Yaakov K, Segal-Ruder Y, Seger R, Fainzilber M. 2005 Vimentin-Dependent Spatial Translocation of an Activated MAP Kinase in Injured Nerve. Neuron 45, 715-726. (doi: 10.1016/j.neuron.2005.01.023) 36. Park SY, Park J, Sim SH, Sung MG, Kim KS, Hong BH, et al. 2011 Enhanced Differentiation of Human Neural Stem Cells into Neurons on Graphene. Adv Healthc Mater. ,23, 36, H263-H7 (doi: 10.1002/adma.201101503) 37. Ma W, Tavakoli T, Derby E, Serebryakova Y, Rao MS, Mattson MP. 2008 Cell- extracellular matrix interactions regulate neural differentiation of human embryonic stem cells. BMC Developmental Biology , 8, 1 , 90 (DOI: 10.1186/1471-213X-8-90) 38. Wilkinson AE, Kobelt LJ, Leipzig ND. Immobilized ECM molecules and the effects of concentration and surface type on the control of NSC differentiation. Journal of Biomedical Materials Research Part A, 102, 10, 3419-28. (DOI: 10.1002/jbm.a.35001)
1505.05803
3
1505
2015-11-15T20:54:17
Particle diffusion in active fluids is non-monotonic in size
[ "physics.bio-ph", "cond-mat.soft" ]
We experimentally investigate the effect of particle size on the motion of passive polystyrene spheres in suspensions of Escherichia coli. Using particles covering a range of sizes from 0.6 to 39 microns, we probe particle dynamics at both short and long time scales. In all cases, the particles exhibit super-diffusive ballistic behavior at short times before eventually transitioning to diffusive behavior. Surprisingly, we find a regime in which larger particles can diffuse faster than smaller particles: the particle long-time effective diffusivity exhibits a peak in particle size, which is a deviation from classical thermal diffusion. We also find that the active contribution to particle diffusion is controlled by a dimensionless parameter, the Peclet number. A minimal model qualitatively explains the existence of the effective diffusivity peak and its dependence on bacterial concentration. Our results have broad implications on characterizing active fluids using concepts drawn from classical thermodynamics.
physics.bio-ph
physics
Particle diffusion in active fluids is non-monotonic in size Alison E. Patteson,a Arvind Gopinath,a,b Prashant K. Purohit,a and Paulo E. Arratiaa Received Xth XXXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XX First published on the web Xth XXXXXXXXXX 200X DOI: 10.1039/b000000x We experimentally investigate the effect of particle size on the motion of passive polystyrene spheres in suspensions of Es- cherichia coli. Using particles covering a range of sizes from 0.6 to 39 microns, we probe particle dynamics at both short and long time scales. In all cases, the particles exhibit super-diffusive ballistic behavior at short times before eventually transitioning to diffusive behavior. Surprisingly, we find a regime in which larger particles can diffuse faster than smaller particles: the particle long-time effective diffusivity exhibits a peak in particle size, which is a deviation from classical thermal diffusion. We also find that the active contribution to particle diffusion is controlled by a dimensionless parameter, the P´eclet number. A minimal model qualitatively explains the existence of the effective diffusivity peak and its dependence on bacterial concentration. Our results have broad implications on characterizing active fluids using concepts drawn from classical thermodynamics. 1 Introduction The diffusion of molecules and particles in a fluid is a process that permeates many aspects of our lives including fog forma- tion in rain or snow 1, cellular respiration 2, and chemical dis- tillation processes 3. At equilibrium, the diffusion of colloidal particles in a fluid is driven by thermal motion and damped by viscous resistance 4. In non-equilibrium systems, fluctua- tions are no longer only thermal and the link between these fluctuations and particle dynamics remain elusive 5. Much ef- fort has been devoted to understanding particle dynamics in non-equilibrium systems, such as glassy materials and sheared granular matter 6. A non-equilibrium system of emerging in- terest is active matter. Active matter includes active fluids, that is, fluids that contain self-propelling particles, such as motile microorganisms 7,8, catalytic colloids 9,10 and molecular mo- tors 11. These particles inject energy, generate mechanical stresses, and create flows within the fluid medium even in the absence of external forcing 12,13. Consequently, active fluids display fascinating phenomena not seen in passive fluids, such as spontaneous flows 8, anomalous shear viscosities 7,14, un- usual polymer swelling 15,16, and enhanced fluid mixing 17 -- 20. Active fluids also play important roles in varied biological and † Electronic Supplementary Information (ESI) available: [details of any supplementary information available should be included here]. See DOI: 10.1039/b000000x/ a Department of Mechanical Engineering & Applied Mechanics, University of Pennsylvania, Philadelphia, PA 19104. Fax: XX XXXX XXXX; Tel: XX XXXX XXXX; E-mail: [email protected] b School of Engineering, University of California Merced, Merced, CA 95343. ‡ Additional footnotes to the title and authors can be included e.g. 'Present address:' or 'These authors contributed equally to this work' as above using the symbols: ‡, §, and ¶. Please place the appropriate symbol next to the author's name and include a \footnotetext entry in the the correct place in the list. Fig. 1 (Color online) Trajectories of 2 µm particles (a) without bacteria and (b) with bacteria (c = 3× 109 cells/mL) for time interval 8 s. Trajectories of (c) 0.6 and (d) 16 µm particles (c = 3× 109 cells/mL). Scale bar is 20 µm. 1 -- 8 1 ecological settings, which include the contributions of suspen- sions of microorganisms to biofilm infections 21,22, biofouling of water-treatment systems 23, and biodegradation of environ- mental pollutants 24. The motion of passive particles in active fluids (e.g. suspen- sion of swimming microorganisms) can be used to investigate the non-equilibrium properties of such fluids. For example, at short times particle displacement distributions can exhibit ex- tended non-Gaussian tails as seen in experiments with algae C. reinhardtii 17,18. At long times, particles exhibit enhanced diffusivities Deff greater than D0, their thermal (Brownian) diffusivity 17 -- 20,25 -- 27. These traits are a signature of the non- equilibrium nature of active fluids; the deviation from equilib- rium also manifests in violations of the fluctuation dissipation theorem 28. In bacterial suspensions, the enhanced diffusivity Deff de- pends on the concentration c of bacteria. In their seminal work, Wu and Libchaber 25 experimentally found that Deff in- creased linearly with c in suspensions of E. coli. Subsequent studies 26,27,29 -- 32 have observed that this scaling holds at low concentrations and in the absence of collective motion. In this regime, Deff can be decomposed into additive components as 26,27,29 -- 32 where D0 and DA are the thermal Deff = D0 + DA and active diffusivities, respectively. It has been proposed that the active diffusivity DA is a consequence of advection due to far-field interactions with bacteria 26 and may even be higher near walls 26,27. While a majority of studies have focused on the role of bac- terial concentration c on particle diffusion, the role of particle diameter d remains unclear. In the absence of bacteria, the diffusivity of a sphere follows the Stokes-Einstein relation, D0 = kBT / f0, where kB is the Boltzmann constant, T is the temperature, and f0 = 3πµd is the Stokes friction factor 4 in a fluid of viscosity µ. In a bacterial suspension, this relation is no longer expected to be valid. Surprisingly, for large par- ticles (4.5 and 10 µm), Wu and Libchaber 25 suggested that Deff scales as 1/d, as in passive fluids. Recent theory and sim- ulation by Kasyap et al. 32 however do not support the 1/d scaling and instead predict a non-trivial dependence of Deff on particle size, including a peak in Deff. This non-monotonic dependence of Deff on particle size implies that measures of effective diffusivities 25, effective temperatures 12,33, and mo- mentum flux 26,27 intimately depend on the probe size and thus are not universal measures of activity. This has important im- plications for the common use of colloidal probes in gauging and characterizing the activity of living materials, such as sus- pensions of bacteria 26,27, biofilms 34,35, and the cytoskeletal network inside cells 36, as well as in understanding transport in these biophysical setting. Despite the ubiquity of passive particles in active environments, the effects of size on particle dynamics in active fluids has yet to be systematically investi- gated in experiments. 2 1 -- 8 Fig. 2 (Color online)(a) Mean-square displacements (MSD) of 2 µm particles versus time ∆t for varying bacterial concentration c. Dashed line is a fit to Langevin dynamics (eqn (1)). (b) MSD for varying particle diameter d versus time for bacterial concentration c = 3.0× 109 cells/mL. The MSD peaks at d = 2 µm. The cross-over time τ (arrows) increases with d. Solid lines are Langevin dynamics fits. In this manuscript, we experimentally investigate the effects of particle size d on the dynamics of passive particles in sus- pensions of Escherichia coli. Escherichia coli 37 are model organisms for bacterial studies and are rod-shaped cells with 3 to 4 flagella that bundle together as the cell swims forward at speed U approximately 10 µ/s. We change the particle size d from 0.6 µm to 39 µm, above and below the effective total length (L ≈ 7.6 µm) of the E. coli body and flagellar bun- dle. We find that Deff is non-monotonic in d, with a peak at 2 < d < 10 µm; this non-monotonicity is unlike the previ- ously found 1/d scaling 25 and suggests that larger particles can diffuse faster than smaller particles in active fluids. Fur- thermore, the existence and position of the peak can be tuned by varying the bacterial concentration c. The active diffusion DA = Deff− D0 is also a non-monotonic function of d and can be collapsed into a master curve when rescaled by the quan- tity cUL4 and plotted as a function of the P´eclet number Pe = UL/D0 (cf. Fig. 5(b)). This result suggests that the active contribution to particle diffusion can be encapsulated by an universal dimensionless dispersivity ¯DA that is set by the ratio of times for the particle to thermally diffuse a distance L and a bacterium to swim a distance L. 2 Experimental Methods Active fluids are prepared by suspending spherical polystyrene particles and swimming E. coli (wild-type K12 MG1655) in a buffer solution (67 mM NaCl in water). The E. coli are prepared by growing the cells to saturation (109 cells/mL) in culture media (LB broth, Sigma-Aldrich). The saturated culture is gently cleaned by centrifugation and resuspended in the buffer. The polystyrene particles (density ρ = 1.05 g/cm3) are cleaned by centrifugation and then suspended in the buffer-bacterial suspension, with a small amount of sur- factant (Tween 20, 0.03% by volume). The particle volume fractions φ are below 0.1% and thus considered dilute. The E. coli concentration c ranges from 0.75 to 7.5 ×109 cells/mL. These concentrations are also considered dilute, correspond- ing to volume fractions φ = cvb <1%, where vb is the vol- ume 26 of an E. coli cell body (1.4 µm3). A 2 µl drop of the bacteria-particle suspension is stretched into a fluid film using an adjustable wire frame 7,18,38 to a mea- sured thickness of approximately 100 µm. The film interfaces are stress-free, which minimizes velocity gradients transverse to the film. We do not observe any large scale collective be- havior in these films; the E. coli concentrations we use are below the values for which collective motion is typically ob- served 32 (≈ 1010 cells/ml ). Particles of different diameters (0.6 µm < d < 39 µm) are imaged in a quasi two-dimensional slice (10 µm depth of focus). We consider the effects of parti- cle sedimentation, interface deformation, and confinement on particle diffusion and find that they do not significantly af- fect our measurements of effective diffusivity in the presence of bacteria (for more details, please see SI-§I). Images are taken at 30 frames per second using a 10X objective (NA 0.45) and a CCD camera (Sony XCDSX90), which is high enough to observe correlated motion of the particles in the presence of bacteria (Fig. 2) but small enough to resolve spatial dis- placements. Particles less than 2 µm in diameter are imaged with fluorescence microscopy (red, 589 nm) to clearly visu- alize particles distinct from E. coli (2 µm long). We obtain the particle positions in two dimensions over time using par- ticle tracking methods 38,39. All experiments are performed at T0 = 22◦C. 3 Results and Discussion 3.1 Mean Square Displacements Representative trajectories of passive particles in the absence and presence of E. coli are shown in Fig. 1 for a time inter- val of 8 s. By comparing Fig. 1(a) (no E. coli) to Fig. 1(b) (c = 3× 109 cells/mL), we readily observe that the presence of bacteria enhances the magnitude of particle displacements compared to thermal equilibrium. Next, we compare sample trajectories of passive particles of different sizes d, below and above the E. coli total length L ≈ 7.6 µm. Figure 1(c) and 1(d) show the magnitude of particle displacement for d = 0.6 µm and d = 16 µm, respectively. Surprisingly, we find that the particle mean square displacements in the E. coli suspension are relatively similar for the two particle sizes even though the thermal diffusivity D0 of the 0.6 µm particle is 35 times larger than that of the 16 µm particle. The 16 µm particles also appear to be correlated for longer times than the 0.6 µm parti- cles. These observations point to a non-trivial dependence of particle diffusivity on d and are illustrated in sample videos included in the Supplementary Materials. To quantify the above observations, we measure the mean- squared displacement (MSD) of the passive particles for vary- ing E. coli concentration c (Fig. 2(a)) and particle size d (Fig. 2(b)). Here, we define the mean-squared particle displacement as MSD(∆t) = (cid:104)r(tR + ∆t)− r(tR)2(cid:105), where the brackets de- note an ensemble average over particles and reference times tR. For a particle executing a random walk in two dimensions, the MSD exhibits a characteristic cross-over time τ, corre- sponding to the transition from an initially ballistic regime for ∆t (cid:28) τ to a diffusive regime with MSD ∼ 4Deff∆t for ∆t (cid:29) τ. Figure 2(a) shows the MSD data for 2 µm particles at vary- ing bacterial concentrations c. In the absence of bacteria (c = 0 cells/mL), the fluid is at equilibrium and Deff = D0. For this case, we are unable to capture the crossover from ballistic to diffusive dynamics due to the lack of resolution: for col- loidal particles in water, for example, cross-over times are on the order of nanoseconds and challenging to measure 40. Experimentally, the dynamics of passive particles at equilib- rium are thus generally diffusive at all observable time scales. We fit the MSD data for the d = 2 µm case (with no bacte- ria) to the expression MSD = 4D0∆t, and find that D0 ≈ 0.2 µm2/s. This matches the theoretically predicted value from the Stokes-Einstein relation; the agreement ((cid:79)) can be visu- ally inspected in Fig. 3(a). In the presence of E. coli, the MSD curves exhibit a ballis- tic to diffusive transition, and we find that the cross-over time τ increases with c. For ∆t (cid:29) τ, the MSD ∼ ∆t with a long- time slope that increases with bacteria concentration c. Ad- ditionally, the distribution of particle displacements follows a Gaussian distribution (see SI-§II A for details and measures of the non-Gaussian parameter). These features, MSD ∼ ∆t and Gaussian displacements, indicate that the long-time dynam- ics of the particles in the presence of E. coli is diffusive and can be captured by a physically meaningful effective diffusion coefficient Deff. We next turn our attention to the effects of particle size. For varying particle diameter d at a fixed bacterial concentration 1 -- 8 3 Fig. 3 (Color online) (a) Effective particle diffusivities Deff versus particle diameter d at varying c. The dashed line is particle thermal diffusivity D0. (b) The crossover-time τ increases with d, scaling as approximately dn, where 1/2 (cid:46) n (cid:46) 1. This is not the scaling in passive fluids 41 where τ ∝ d2. (c = 3× 109 cells/ml), the MSD curves also exhibit a ballistic to diffusive transition, as shown in Fig. 2(b). Surprisingly, we find a non-monotonic behavior with d. For example, the MSD curve for the 2 µm case sits higher than the 39 µm case and the 0.6 µm case. This trend is not consistent with classical diffu- sion in which MSD curves are expected to decrease monoton- ically with d (D0 ∝ 1/d). We also observe that the cross-over time τ increases monotonically with d. As we will discuss later in the manuscript, the cross-over time scaling with d also deviates from classical diffusion. 3.2 Diffusivity and Cross-over Times We now estimate the effective diffusivities Deff and cross-over times τ of the passive particles in our bacterial suspensions. To obtain Deff and τ, we fit the MSD data shown in Fig. 2 to the MSD expression attained from the generalized Langevin equation 41, that is MSD(∆t) = 4Deff∆t 1− e− ∆t τ . (1) (cid:16) (cid:16) 1− τ ∆t (cid:17)(cid:17) Equation 1 has been used previously to interpret the diffu- sion of bacteria 16 as well as the diffusion of particles in films with bacteria 35. In the limit of zero bacterial con- centration, DA = 0 and eqn (1) reduces to the formal solu- tion to the Langevin equation for passive fluids, MSD(∆t) = (cid:0)1− e−∆t/τ0(cid:1)(cid:1) with τ0 = τ(c = 0). For more 4D0∆t(cid:0)1− τ0 ∆t details on the choice of our model, see SI-§III. Figure 3 shows the long-time particle diffusivity Deff (Fig. 3(a)) and the cross-over time τ (Fig. 3(b)) as a function of d for bacterial concentrations c = 0.75, 1.5, 3.0 and 7.5× 109 cells/mL. We find that for all values of d and c considered here, Deff is larger than the Stokes-Einstein values D0 at equi- librium (dashed line). For the smallest particle diameter case 4 1 -- 8 (d = 0.6 µm), Deff nearly matches D0. This suggests that activity-enhanced transport of small (d (cid:46) 0.6 µm) particles or molecules such as oxygen, a nutrient for E. coli, may be entirely negligible 32. For more information, including figures illustrating the dependence of Deff on c and comparisons be- tween our measured effective diffusivities and previous exper- imental work, see sections SI-§IIB and SI-§IIC in the supple- mental materials. Figure 3(a) also reveals a striking feature, a peak Deff in d. Our data demonstrates that, remarkably, larger particles can diffuse faster than smaller particles in suspensions of bacteria. For example, at c = 7.5× 109 cells/mL ((cid:13)) the 2 µm parti- cle has an effective diffusivity of approximately 2.0 µm2/s, which is nearly twice as high as the effective diffusivity of the 0.86 µm particle, Deff =1.3 µm2/s. We also observe that the peak vanishes as c decreases. In fact, for the lowest bacterial concentration (c = 0.75×109 cells/mL), there is no peak: Deff decreases monotonically with d. Clearly, Deff does not scale as 1/d. Figure 3(b) shows the cross-over times τ characterizing the transition from ballistic to diffusive regimes as a function of particle size d for varying c. We find that the values of τ in- crease with d and c. We note that the variation of τ with c (SI- Fig. 3(b)) does not follow a linear form. Instead, the data sug- gests possible saturation of τ for suspensions of higher -- but still dilute -- concentrations. The cross-over time τ scales with particle diameter approximately as τ ∼ dn, where 1 (cid:46) n (cid:46) 1. 2 This cross-over time does not correspond to the inertial relax- ation of the particle. Therefore, this scaling does not follow the trend seen for passive particles at thermal equilibrium 41, where τ, being the Stokes relaxation time (order 1 ns), scales as m/ f0 ∝ d2 with particle mass m ∝ d3. Our data (Fig. 3(b)) highlights that in active fluids the super-diffusive motion of the passive particles cannot be ignored -- even for time scales Fig. 4 (Color online) (a) Distribution of 2 µm particle speeds p(v) follow a Maxwell-Boltzmann distribution (solid curves) with clear peaks that shift right as c increases. (b) The effective temperature Teff extracted by fitting p(v) data to eqn (2) ((cid:52)) matches those obtained from an extended Stokes-Einstein relation ((cid:13)). (c) The effective temperature Teff increases with d for varying c. with peak speeds vmax =(cid:112)2kBTeff/m, where m is the mass of the polystyrene particle. Fitting the p(v) data in the absence of bacteria (∇ in Fig. 4(a)) to eqn (2) yields Teff ≈ T0, as ex- pected. as large as a second -- and that the time scales over which diffusive motion is valid (∆t > τ) depends on the size d of the particle. Further implications of a particle size-dependent cross-over time will be discussed below. 3.3 Effective Temperature Our data so far suggests that particle dynamics in bacterial sus- pensions, while having an anomalous size-dependence (Figs. 1,2,3), maintain the characteristic super-diffusive to diffusive dynamics for passive fluids 41. The long-time diffusive behav- ior (Fig. 2) and enhancement in Deff (Fig. 3(a)), which is rooted in particle-bacteria encounters, suggest that the parti- cles behave as if they are suspended in a fluid with an effec- tively higher temperature. In order to further explore the concept of effective temper- ature in bacterial suspensions, we measure the distribution of particle speeds p(v) as a function of bacterial concentration c. The particle speed distributions determine the mean kinetic energy of the particles. If the distribution follows a Maxwell Boltzmann form -- as is always the case in fluids at equilibrium -- the mean kinetic energy is related to the thermodynamic temperature via the equipartition theorem. Such a relationship may not always exist for out-of-equilibrium fluids. Figure 4(a) shows p(v) for d = 2 µm case for a range of c. We define the particle speeds v over a time interval of 0.5 s. This time interval is greater than the crossover times for the 2 µm particles (cf. SI-Fig. 3) to ensure that a particle sam- ples multiple interactions with bacteria and exhibits diffusive behavior. In the absence of bacteria, the system is in thermal equilibrium and p(v) follows the two dimensional Maxwell- Boltzmann distribution, p(v) = vm(kBTeff)−1e −mv2 2kBTeff , (2) In the presence of bacteria, the particle speed distributions also follow a Maxwell-Boltzmann form, eqn (2), with peaks that shift toward higher values of v as the E. coli concentration c increases. Because our data follows the Maxwell-Boltzmann form (Fig. 4(a)) for all c, this indicates that there is no corre- lated motion at long times as is the case in swarming bacterial suspensions, for which the particle speed distribtuions exhibits an exponential decay 42. We also note that the power spectra of particle speeds (SI-§II D) are reasonably flat, consistent with white-noise forcing and an absence of correlated motion. Fig- ure 4(a) thus indicates that an effective temperature Teff can be defined from p(v) and is increasing with the bacterial concen- tration. To quantify this 'enhanced' temperature and the deviation from equilibrium behavior, we fit the p(v) data (Fig. 4(a)) to eqn (2). These fits allow us to obtain Teff as a function of c, as shown in Fig. 4(b) ((cid:52)). Also shown in Fig. 4(b) are the values extracted of Teff ((cid:13)) from an extended Stokes-Einstein relation, Teff = 3πµdDeff/kB, where Deff are the values from the MSD fits shown in Fig. 3(a) and µ is the viscosity of the solvent (µ ≈ 1 mPa s). We find that both estimates of Teff are higher than room temperature T0 and increase linearly with c. For a fluid at equilibrium, the temperatures from these two methods, MSD and p(v), are expected to be the same but not for systems that are out of equilibrium 5,41,43,44, such as the bacterial suspensions investigated here. The good agreement in Fig. 4(b) suggests that Teff may be a useful signature of bacterial activity with some analogies to equilibrium systems. Note that in defining the effective temperature via the general- ized Stokes-Einstein relation, we have assumed the viscosity to be a constant and independent of bacterial concentration, 1 -- 8 5 which may not be true when the bacterial concentrations are sufficiently high 45. Figure 4(b) suggests that an unchanging viscosity is a valid assumption for our system. (E. coli) as well as challenges in gauging activity using passive particles. 3.4 Active Diffusivity of Passive Particles in Bacterial Suspensions To explore the aforementioned dependence of DA on particle size, we plot DA = Deff − D0 with d. Here, D0 is the parti- cle thermal diffusivity (in the absence of bacteria) and is ob- tained from the Stokes-Einstein relation D0 = kBT0 3πµd . Indeed, as shown in Fig. 5(a), DA exhibits a non-monotonic depen- dence on d for all c. To understand the observed dependence of the active diffu- sivity on the particle size, we consider the relevant time scales in our particle/bacteria suspensions, namely: (i) the time for the particle to thermally diffuse a distance L equal to the total bacterial length L2/D0, (ii) the time for a bacterium to swim at a speed U for a distance L given by L/U, and (iii) the mean run time which is the inverse of the tumbling frequency ω−1 T . Di- mensionless analysis then suggests there are two independent time parameters: (i) the ratio of the first two above, which is the P´eclet number, Pe ≡ UL/D0, and (ii) τ∗ = U/ωT L, which is the ratio of the run length U/ωT to the bacterial length. It is these two parameters, Pe and τ∗, that govern the parti- cle dynamics in our experiments. When the P´eclet number is much less than one, then the thermal particle diffusion dom- inates and transport by the bacteria is ineffective. When the P´eclet number is much larger than one, then thermal diffusion is negligible and the transport is due to the convection from bacteria.The swimming speed U, tumble frequencies ω−1 T , and combined length of the cell body and flagellar bundle L are estimated from prior experiments 16,46 with E. coli as ap- proximately 10 µm/s, 1 s−1, and L ≈ 7.6 µm, respectively. Thus, in our experiments, the P´eclet number varies from ap- proximately 130 to 8600, via the particle bare diffusivity D0 (through the particle diameter). We note that one stain of bac- teria is used -- thus, the run length and bacteria size do not change in our experiments, and consequently, τ∗ ≈ 1.8 is a constant. In order to gain insight into the non-trivial dependence of DA on d, we scale out the concentration dependence by in- troducing the dimensionless active diffusivity DA = DA/cUL4 and plot it against the P´eclet number, Pe. Figure 5(b) shows that all the active diffusion DA versus d data shown in Figure 5(a) collapses into a single master curve, thereby indicating that DA is independent of c, at least for dilute suspensions in- vestigated here. We find that for Pe (cid:46) 103, the values of DA initially in- creases with increasing Pe (or particle size) and follows a scal- ing DA ∼ Pe2. The observed increase of DA with Pe may be due to the decreasing particle Brownian motion, which allows the particle's motion to be correlated with the bacterial veloc- Fig. 5 (Color online) (a) Active diffusivities DA = Deff − D0 are non-monotonic with particle size for varying concentrations of bacteria. (b) Scaled hydrodynamic diffusivity DA = DA/cUL4 collapses with P´eclet number Pe = UL/D0. The maximum DA occurs at PeA between 450 and 4500. For 100 < Pe < PeA, DA scales as Peα, where α ≈ 2. Next, we investigate the role of particle diameter in the suspension effective temperature. Figure 4(c) show the val- ues of Teff estimated from an extended Stokes-Einstein re- lation as a function of d for different values of c. Surpris- ingly, we find that Teff increases with particle size d, which is clearly different from thermally equilibrated systems where temperature does not depend on the probe size. We note that for the largest particle diameter, d = 39 µm, estimated Teff values are approximately 100 times greater than room temperature T0 = 295 K, consistent with previous reports 25. The dependence of Teff on particle size d (Fig. 5(a)) may be understood through the extended Stokes-Einstein relation, which yields Teff = Deff f0/kB = T0 + (3πµDA/kB)d, where T0 = (3πµD0/kB)d. If DA were independent of d, then Teff would be linear in d. However, a linear fit does not adequately capture the trend. This hints at a particle size dependent active diffusivity DA, which has been predicted in recent theory 32. This variation of Teff with d highlights the interplay between particle size and the properties of the self-propelling particles 6 1 -- 8 ity disturbances for longer times. In the limit of Pe → ∞, the particle's Brownian motion van- ishes and the particle displacements are dominated by the con- vective transport via bacteria-particle interactions. Thus DA is expected to be independent of Pe and depend only on the pa- rameter τ∗, defined here as the ratio of the run length U/ωT to the bacterial length L. Given that our experiments correspond to a constant τ∗ ≈ 1.8, our experimentally measured DA agree well with recent theoretical predictions 32 for very large P´eclet (as Pe → ∞) for τ∗ = 1 and τ∗ = 4, as shown in Fig. 5(b). An increase in the run time (or τ∗) would increase the asymptotic value of the scaled active diffusivity DA (see SI-§IV). An important feature of the data shown in Fig. 5(b) is the peak in DA at PeA ≈ 103. The existence of such a peak has been predicted in a recent theory/simulation investigation 32, and our data agrees at least qualitatively with such predic- tions. We note that the predicted peak in DA happens at a Pe ≈ O(10), which is smaller than our experimental values of (Pe ≈ 103). The appearance of the peak in DA in our data may be due to the weak but non-zero effects of Brownian motion, which allows particles to sample the bacterial velocity field in such a way that the mean square particle displacements and correlation times are higher compared to both the very high Pe (negligible Brownian motion) as well as the very low Pe (Brownian dominated) 32. This feature (i.e. peak in DA) is sur- prising because it suggests an optimum particle size for max- imum particle diffusivity that is coupled to the activity of the bacteria. 4 Maximum Particle Effective Diffusivity Deff Our data (Fig. 5(b)) shows that the dimensionless active dif- fusion DA collapses unto a universal curve with a peak in Pe for all bacterial concentrations. Our data also shows that, with the exception of the lowest bacterial concentration c, the parti- cle effective diffusivity Deff exhibits a peak in d, which varies with c (Fig. 3(a)). For instance, for c = 1.5× 109 cells/mL, the peak is at approximately 2 µm, while for larger concentra- tions (c = 7.5× 109 cells/mL), the peak (as obtained by fitting the data to a continuous function) shifts to higher values of d. This suggests that one can select the particle size which diffuses the most by tuning the bacteria concentration. In what follows (see also SI-§V), we provide a prediction, based on our experimentally-measured universal curve of DA with Pe (Fig. 5(b)) for the existence as well as the location of the peak of Deff in d. As noted before, the particle effec- tive diffusivity Deff can be described as the linear sum of the particle thermal diffusivity D0, which is independent of c and decreases with d, and the active diffusivity DA, which is linear in c and non-monotonic in d, through the particle-size depen- dent DA. Therefore, we can recast the effective diffusivity as Deff = D0 + (cL3) (UL)DA. (3) The criterion for the existence of a maximum Deff is obtained by taking the derivative of eqn (3) with respect to the P´eclet number and setting the derivative to zero. In order to estimate DA (and its slope), we fit the data in Fig. 5(b) near the peak in the range 200 < Pe < 4000 with a second order polynomial equation. We find that a peak exists in Deff if (cid:20) (cid:18) (cid:19)(cid:21) cL3 (cid:38) 0.4. (4) For the bacterial length used here L = 7.6 µm, this yields c ≈ 0.9 × 109 cells/mL, which is in quantitative agreement with the concentration range (0.75× 109 cells/mL < c < 1.5× 109 cells/mL) in which the peak in Deff emerges in our data (Fig. 3(a)). As described in SI-§V, we find that the location of the Deff peak in d here defined as dmax eff is given by 1 eff ≈ dA dmax 1− 5cL3 − 2 (5) where dA corresponds to the P´eclet number PeA ≈ 1000 at which DA is maximum. For the E. coli used here, dA = kT PeA/3πµUL ≈ 6 µm. Note that dmax is always less than or equal to dA and increases with c, consistent with our exper- imental observations (Fig. 3(a)). In general the criterion, eqn , eqn (5), will depend on τ∗. For suspensions of (4), and dmax eff bacteria, the universal curve of DA informs when and where a peak in the particle diffusivity occurs. eff 5 Conclusions In summary, we find that the effective particle diffusivity Deff and temperature Teff in suspensions of E. coli show strong de- viations from classical Brownian motion in the way they de- pend on particle size d. For example, Fig. 3(a) shows that Deff depends non-monotonically in d and includes a regime in which larger particles can diffuse faster than smaller parti- cles. The existence as well as the position of a Deff peak in d can be tuned by varying the bacterial concentration c. We also find that the cross-over time τ increases with particle size and scales as approximately dn, where 1/2 (cid:46) n (cid:46) 1, as shown in Fig. 3(b). Measures of Teff obtained from either an extended Stokes- Einstein relation or particle speed distributions seem to agree quite well (Fig. 4(b)). This is surprising since this kind of agreement is only expected for systems at equilibrium. The good agreement between the measurements suggests that Teff may be a useful signature of bacterial activity. However, un- like thermally equilibrated systems, Teff varies with size d (Fig. 4(c)). This non-trivial dependence of both Deff and Teff on particle size d implies that these common gauges of activ- ity are not universal measures. Nevertheless, our data suggest 1 -- 8 7 20 M. J. Kim and K. S. Breuer, Phys. Fluids, 16, L78 (2004). 21 C. Josenhans and S. Suerbaum, Int. J. Med. Microbiol., 291, 605-614 (2002). 22 J. W. Costerton, P. S. Stewart and E. P. Greenberg, Science, 284, 1318- 1322 (1999). 23 Bixler GD, Bhushan B. Biofouling: lessons from nature. Phil. Trans. R. Soc. A 2012;370:23812417. 24 D. L. Valentine, J. D. Kessler, M. C. Redmond, S. D. Mendes, M. B. Heintz, C. Farwell, L. Hu, F. S. Kinnaman, S. Yvon-Lewis, M. Du, E. W. Chan, F. G. Tigreros and C. J. Villanueva, Science, 330, 208 (2010). 25 X. L. Wu and A. Libchaber, Phys. Rev. Lett., 84, 3017 (2000). 26 A. Jepson, V. A. Martinez, J. Schwarz-Linek, A. Morozov, and W. C. K. Poon, Phys. Rev. E, 88, 041002 (2013). 27 G. Mino, T. E. Mallouk, T. Darnige, M. Hoyos, J. Dauchet, J. Dunstan, R. Soto, Y. Wang, A. Rousselet, and E. Cl´ement, Phys. Rev. Lett., 106, 048102 (2011). 28 D. T. N. Chen, A. W. C. Lau, L. A. Hough, M. F. Islam, M. Goulian, T. C. Lubensky, and A. G. Yodh, Phys. Rev. Lett., 99, 148302 (2007). 29 D. O. Pushkin and J. M. Yeomans Phys. Rev. Lett., 111, 188101 (2013). 30 P. T. Underhill, J. P. Hernandez-Ortiz, and Michael D. Graham, Phys. Rev. Lett., 100, 248101 (2008). 31 J. P. Hernandez-Ortiz, P. T. Underhill, and Michael D. Graham, J. Phys.: Condens. Matter, 21, 204107 (2009). 32 T. V. Kasyap, D. L. Koch, and M. Wu, Phys. Fluids, 26, 081901 (2014). 33 D. Loi, S. Mossa, and L. F. Cugliandolo, Phys. Rev. E, 77, 051111 (2008). 34 S. S. Rogers, C. van der Walle, and T. A. Waigh, Langmuir, 24, 13549- 13555 (2008). 35 L. Vaccari, D. B. Allan, N. Sharifi-Mood, A. R. Singh, R. L. Leheny and K. J. Stebe, Soft Matter, 11, 6062-6074 (2015). 36 ´E. Fodor, M. Guo, N. S. Gov, P. Visco, D. A. Weitz and F. van Wijland Euro Phys Lett, 110, 48005 (2015). 37 H. C. Berg, E. coli in Motion. Springer Science & Business Media, (2008). 38 B. Qin, A. Gopinath, J. Yang, J. P. Gollub and P. E. Arratia, Sci. Rep., 5, 9190 (2015). 39 J. Crocker, and D. Grier, J. Colloid. Interf. Sci, 179, 298 (1996). 40 R. Huang, I. Chavez, K. M. Taute, B. Luki´c, S. Jeney, M. G. Raizen and E. Florin, Nature Physics, 7, 576-580 (2011). 41 R. K. Pathria, Statistical Mechanics, (Butterworth, Oxford, 1996). 42 A. Sokolov, M. M. Apodaca, B. A. Grzybowski, and I. S. Aranson, Proc. Natl. Acad. Sci., 107, 969-974 (2010). 43 E. Ben-Isaac, Y. Park, G. Popescu, F. L. Brown, N. S. Gov and Y. Shokef, Phys. Rev. Lett., 106, 238103 (1996). 44 E. Ben-Isaac, ´E. Fodor, P. Visco, F. van Wigland and N. S. Gov, Phys. Rev. E, 92, 012716 (2015). 45 H. M. L´opez, J. Gachelin, C. Douarche, H. Auradou and E. Cl´ement, Proc. Natl. Acad. Sci., 107, 969-974 (2010). 46 T. L. Min, P. J. Mears, L. M. Chubiz, C. V. Rao, I. Golding and Y. R. Chemla, Nature Methods, 6, 831-835 (2009). that one can define optimal colloidal probes of activity for sus- pensions of bacteria, which correspond to Pe = UL/D0 ≈ 103. At these Pe values, ¯DA is maximized, which provides ample dynamical range (magnitude of the signal). Also at these Pe values, the cross-over time τ is still relatively small, which al- lows for adequate temporal resolution. Both of these features are important in using passive particles to characterize a spa- tially and temporally varying level of activity in materials. Our anomalous particle-size dependent results in active flu- ids has important implications for particle sorting in microflu- idic devices, drug delivery to combat microbial infections, re- suspension of impurities and the carbon cycle in geophysical settings populated by microorganisms. A natural next step would be to study the role of external fields such as gravity or shear in influencing particle transport in these active envi- ronments. Acknowledgements We would like to thank J. Crocker, S. Farhadi, N. Keim, D. Wong, E. Hunter, and E. Steager for fruitful discussions and B. Qin for help in the spectral analysis of data. This work was supported by NSF-DMR-1104705 and NSF-CBET-1437482. References 1 H. R. Pruppacher and J. D. Klett, Microphysics of Clouds and Precipita- tion, (Reidel, Dordrecht, 1978). 2 S. F. Perry and W. W. Burggren Integr. Comp. Biol., 47, 506-509 (2007). 3 A. G´orak and E. Sorensen, Distillation: Fundamentals and Principles, (Academic Press, 2014). 4 A. Einstein, Ann. Physik, 17, 549 (1905). 5 R. P. Ojha, P. A-. Lemieux, P. K. Dixon, A. J. Liu, A. J. and D. J. Durian Nature, 427, 521-523 (2004). 6 L. Cipelletti and L. Ramos J. Phys.: Condens. Matter, 17, R253-R285 (2005). 7 A. Sokolov and I.S. Aranson, Phys. Rev. Lett., 103, 148101 (2009). 8 H. P. Zhang, A. Be'er, E. L. Florin, and H. L. Swinney, Proc. Nat. Acad. Sci., 107, 13626 (2010). 9 D. Takagi, A. B. Braunschweig, J. Zhang, and M. J. Shelley, Phys. Rev. Lett., 110, 038301 (2013). 10 J. Palacci, S. Sacanna, A. P. Steinberg, D. J. Pine, and P. M. Chaikin, Science, 339, 936 (2013). 11 F. J. Ndlec, T. Surrey, A. C. Maggs and S. Leibler Nature, 389, 6648 (1997). 12 M. C. Marchetti, J. F. Joanny, S. Ramaswamy, T. B. Liverpool, J. Prost, M. Rao and R. Aditi Simha Rev. Mod. Phys, 85, 1143 (2013). 13 S. Ramaswamy Annu. Rev. Condens. Matter Phys., 1, 323-45 (2010). 14 M. Mussler, S. Rafai, P. Peyla, and C. Wagner, EPL, 101, 54004 (2013). 15 A. Kaiser and H. Lowen J. Chem Phys., 141, 044903 (2014). 16 A. E. Patteson, A. Gopinath, M. Goulian, and P. E. Arratia Sci. Rep, 5, 15761 (2015). 17 K. C. Leptos, J. S. Guasto, J. P. Gollub, A. I. Pesci and R. E. Goldstein, Phys. Rev. Lett., 103, 198103 (2009). 18 H. Kurtuldu, J. S. Guasto, K. A. Johnson and J. P. Gollub, Proc. Nat. Acad. Sci., 108, 10391 (2011). 19 J. L. Thiffeault and S. Childress, Phys. Lett. A, 374, 34873490 (2010). 8 1 -- 8
1801.00176
4
1801
2018-10-01T09:32:26
Fluctuations in active membranes
[ "physics.bio-ph" ]
Active contributions to fluctuations are a direct consequence of metabolic energy consumption in living cells. Such metabolic processes continuously create active forces, which deform the membrane to control motility, proliferation as well as homeostasis. Membrane fluctuations contain therefore valuable information on the nature of active forces, but classical analysis of membrane fluctuations has been primarily centered on purely thermal driving. This chapter provides an overview of relevant experimental and theoretical approaches to measure, analyze and model active membrane fluctuations. In the focus of the discussion remains the intrinsic problem that the sole fluctuation analysis may not be sufficient to separate active from thermal contributions, since the presence of activity may modify membrane mechanical properties themselves. By combining independent measurements of spontaneous fluctuations and mechanical response, it is possible to directly quantify time and energy-scales of the active contributions, allowing for a refinement of current theoretical descriptions of active membranes.
physics.bio-ph
physics
Fluctuations in active membranes Herv´e Turlier and Timo Betz Abstract Active contributions to fluctuations are a direct consequence of metabolic energy consumption in living cells. Such metabolic processes continuously create active forces, which deform the membrane to control motility, proliferation as well as homeostasis. Membrane fluctuations contain therefore valuable information on the nature of active forces, but classical analysis of membrane fluctuations has been primarily centered on purely thermal driving. This chapter provides an overview of relevant experimental and theoretical approaches to measure, analyze and model ac- tive membrane fluctuations. In the focus of the discussion remains the intrinsic prob- lem that the sole fluctuation analysis may not be sufficient to separate active from thermal contributions, since the presence of activity may modify membrane me- chanical properties themselves. By combining independent measurements of spon- taneous fluctuations and mechanical response, it is possible to directly quantify time and energy-scales of the active contributions, allowing for a refinement of current theoretical descriptions of active membranes. 1 Introduction Biological and bio-mimetic membranes consist in bilayers of phospholipids, which can embed various transmembrane or peripheral proteins. They constitute a selec- tively permeable barrier between distinct biological compartments, such as the cy- tosol and extracellular medium. The stability of lipid bilayers in water is the result of Herv´e Turlier Center for Interdisciplinary Research in Biology, Coll`ege de France, PSL Research University, CNRS UMR7241, Inserm U1050, 11 place Marcelin Berthelot, F-75005 Paris, France, e-mail: [email protected] Timo Betz Institute of Cell Biology, Center for Molecular Biology of Inflammation, Von-Esmarch-Str. 56, D-48149 Muenster, Germany, e-mail: [email protected] 1 2 Herv´e Turlier and Timo Betz an entropic effect, which combines non-covalent interactions between hydrophobic and hydrophilic parts of lipids. This leads to a large in-plane rigidity, making lipid bilayers almost incompressible: area strains of only 2-4% are generally enough to rupture a membrane. The non-covalent nature of interactions make phospholipid bilayers moreover tangentially fluid. Any tangential force on a lipid or embedded protein will lead to lateral flows balancing almost instantly any density gradient. Bending a lipid bilayer, in contrast, requires only to slightly displace the polar heads, which are separated by a distance of the order of 0.5-1 nm [1, 2]. The bending mod- ulus, generally denoted κ, is therefore not very large compared to thermal energy, of the order of a few tens of kBT , which explains why membrane bending modes are readily excited at ambient temperatures. Hence, biological membranes are continu- ously fluctuating as a result of the thermal agitation of the surrounding medium, and this movement is directly observable using standard microscopy. However, besides thermal agitation, non-equilibrium active forces, of intrinsic or extrinsic origins, may also contribute and enhance membrane fluctuations. If such active fluctuations have random, uncorrelated sources, it remains however complicated to determine by simple observation to which extent membrane undulations are driven by thermal or by non-equilibrium effects. One of the most prominent example of fluctuating biological membranes is the 'flickering' of red blood cells, already described in the 19th century [3]. Since its first observation, the origin of red blood cell flickering had been debated and only recently its possible active nature has been precisely investigated [4, 5, 7, 8]. Ini- tially, flickering was suggested to be passive, similar to the Brownian motion of microscopic particles [9, 10]. But in 1951, the amplitude of flickering was shown to be correlated to ion transport across the membrane [11], suggesting a possible active metabolic driving. However, this metabolic interpretation was soon revised as flick- ering was also observed in the absence of ATP in red blood cell ghosts [12]. After this finding, the pure passive origin of the flickering was generally accepted for more than 40 years [13], until in 1997 new experimental approaches revealed a change in the membrane fluctuations amplitude upon ATP starvation [4]. These new experi- ments, however, remained debated and a series of conflicting results were reported [14]. While most of the differences may be attributed to variations in preparation protocols, more and more indirect findings suggested that an active driving may contribute to the low frequency fluctuations spectrum [5, 15]. A conclusive experi- mental evidence for the active nature of red blood cell flickering was recently given by comparing directly flickering and mechanical response of the membrane [8]. The experimental observations could show that the flickering directly violates equilib- rium statistical mechanics, proving the presence of non-equilibrium active forces driving membrane movement. This is an emblematic case of scientific controversy that took about 125 years to be conclusively resolved. As this example attests, it remains difficult to evaluate to which extent active pro- cesses may contribute to membrane fluctuations. In general, active fluctuations are superimposing upon passive thermal fluctuations, and from an experimental point of view, active and passive fluctuations may share similar characteristics. Hence, ac- tive contributions might not be visible, especially if thermal agitation dominates the Fluctuations in active membranes 3 fluctuation spectrum at the specific membrane position, length scale or timescale of interest. In this view, to answer the question whether fluctuations are active or passive, one should always define the relevant time and length scales involved. From a mechanistic point of view, active membrane fluctuations originate from the conversion of metabolic energy into forces by proteins inserted in the bilayer, or connected to it (Figure 1a). One may define active membrane fluctuations as intrinsic, when they are produced by proteins directly embedded in the membrane, or extrinsic when activity originates from an independent structure tethered to the membrane, like the cytoskeleton. Active membrane fluctuations originating from ion pumps have been the focus of pioneering biophysics studies over the last decades, both from theoretical [16, 17, 18] and experimental perspectives [19, 20, 21, 22]. These studies show that the activity of pumps leads to significant modifications in the fluctuation properties of reconstituted vesicles, measured as changes in the fluctuation amplitude, in the effective membrane tension or in the excess surface area. Besides ATP or photon driven ion pumps, lipid transport systems such as flipases and flopases [23] may also contribute to active fluctuations, as well as membrane-fusion and fission of transport vesicles [24]. Active fluctuations can also originate from the interaction of the membrane with the underlying cytoskeleton, such as the spectrin network or the actomyosin cortex. In this case, the potential sources of active forces on the membrane are various: the proteins linking the membrane may change their binding affinity, or mechanical properties, upon phosphorylation, the cytoskeleton may exert tangential and normal forces on the bilayer under the action of molecular motors, or via polymerization of filaments. The aim of this chapter is to present methods used to measure membrane fluc- tuations, to analyze their active and passive components, and to show how active membrane fluctuations may be modeled theoretically. 2 Experimental observations of active membrane fluctuations and indications of activity A challenging experimental task is to show that the observed fluctuations are active by nature. This requires separating the active and passive contributions in the fluc- tuation spectrum. A typical approach proposed to detect active fluctuations was to remove the source of chemical energy, primarily ATP, and then to attribute possible differences to active processes. Unfortunately, biological membranes are complex systems, and the suppression of metabolic energy sources does not only remove active noise, but it may also change the mechanical properties of the membrane. For example, ion pumps do not only contribute to active fluctuations but are also important for maintaining the osmotic pressure in the cell. This is, in turn, a key ele- ment determining membrane tension, and therefore influencing thermal fluctuations 4 Herv´e Turlier and Timo Betz characteristics. It remains therefore generally challenging to disentangle fluctuations changes due to the suppression of activity from the ones due to passive mechanical variations. Another example is cytoskeletal softening or stiffening upon ATP deple- tion [25, 26]. In both cases, a simple removal of the energy source is not sufficient to pin down the contribution of active forces to the measured membrane fluctuations. In the following we will review different approaches that have been proposed to tackle this question in the context of membranes. Fig. 1 Introduction of the concepts of membrane fluctuations, projected area surface tension and of the different active processes driving membrane fluctuations. a) Thermal and active fluctuations buffer a significant fraction of the total membrane area, so that the projected area A is smaller than the real area A . Lateral pulling forces, effectively reduce the amplitudes of continuous fluctuations to extend the projected area. Hence these forces pull out the membrane reservoir that is stored in the fluctuations. The energy required to increase the projected area is used to define a membrane tension. It should be noted that this definition of tension depends on the entropic effect of thermally excited fluctuations. Biological membranes are typically connected to an underlying network of cytoskeletal elements such as F-actin or spectrin. b) Beside thermal agitation, membrane fluctuations can be driven by active, force generating processes such as ion pumps or lipid transporters activity or via mechanical coupling to the underlying cytoskeleton. 2.1 Micropipette aspiration: surface tension and excess area Surface tension is formally defined as the energy required to change surface area, but in lipid bilayers, the real surface can only be modified by a few percent before rupturing. However, a significant amount of membrane area is stored in the fluctua- tions, leading to clear difference between the total membrane area and its projected area (or apparent surface) (Figure 1a). In other words, the apparent surface is always smaller than the real surface of the membrane, and this difference is measured as the membrane excess area. When applying a tensile force to a membrane, the amplitude of the fluctuations is reduced and the apparent area increases, corresponding to a de- Fluctuations in active membranes 5 crease of excess area (Figure 1b). For a vesicle or a cell, where the volume and total membrane area are supposed constant, the excess area is on the contrary fixed. In this case, active forces are expected to increase the fluctuations amplitude compared to pure thermal driving, leading to an increase of tension (see section 4.2.4). Building on this physical reasoning, one of the first experiments showing that an active process can change membrane fluctuations was done in biomimetic lipo- somes, combining classical micropipette aspiration techniques with purified bac- teriorhodopsin proteins. Bacteriorhodopsin is a light-driven transmembrane pump, which transfers protons across the membrane when exposed to green-yellow light of wavelength around 566 nm [27]. Each time a proton is pumped, the membrane experiences a small active force. To measure this activity, micropipette aspiration was used. In this method the membrane area stored in fluctuations is measured by steadily increasing the aspiration pressure while following the changes in surface area. This change is calculated by following the length of the membrane tongue, while knowing the inner radius of the micropipette (Figure 2a). In passive mem- branes, equilibrium statistical mechanics allows to predict that the logarithm of the applied tension is a linear function of the excess area. For the equilibrium case, the slope of this curve is proportional to κ/kBT (see equation (6)). Intuitively a larger bending rigidity will indeed decrease the area stored in fluctuations, and a smaller excess area should hence be measured at the same stretching force. Conversely, increasing temperature will increase the area stored in fluctuations. When light sen- sitive ion pumps are activated by exposing the liposome to yellow-green light, the slope of the logarithm of tension vs. excess area curve is significantly changed, indicating a larger excess area stored in fluctuations. To explain the experimental result, the real temperature can be simply replaced by an increased effective tem- perature. Hence, the difference between the real and the effective temperature can be used to estimate the energy injected to drive active membrane fluctuations. In the case of bacteriorhodopsin the effective temperature was found to be about twice the real temperature [20]. In a further experiment using a calcium pump driven by ATP hydrolysis, the effective temperature was found to be in the same range [22]. In these experiments a clear dependence of the effective temperature on the pump concentration was observed. Micropipette aspiration experiments provide hence a clear hint that membrane fluctuations are enhanced by active processes. Yet the actual fluctuations are only measured indirectly from the excess area. Other methods provide more direct access to membrane fluctuations and to possible active contributions. 2.2 Image based contour analysis Video microscopy provides direct spatial and temporal access to membrane fluc- tuations. Brochard and Lennon have pioneered its use to determine relative thick- ness fluctuations in red blood cells, responsible for the flickering effect [13]. At the time, an equilibrium model was used to analyze the membrane fluctuation spectrum, 6 Herv´e Turlier and Timo Betz Fig. 2 Summary of the currently used techniques to study active and passive membrane fluc- tuations. a) Active fluctuations can be measured as an increase of area stored in the membrane fluctuations upon illumination with light (Data reproduced with permission from [20]) b) A further method to determine membrane fluctuations is based on RICM, where constructive and destructive interference from reflections at the glass and membrane surface are detected with a camera. c) More recently DPM has been introduce which measures thickness changes by exploiting the change in optical path along the light propagation through the object to be measured. (Inset adapted from [6], Figure 2) d) By exploiting the phase shift of light partially touching a membrane, the lateral fluctuations can be measured using an interferometric approach using a quadrant photodiode. e) Another new technique related to fluorescence correlation spectroscopy can measure height fluctu- ations of a membrane with excellent spatial and temporal resolution. (Data reproduced from [35]) f) Finally, direct membrane mechanics can be measured using optical tweezer based pulling on the membrane. Fluctuations in active membranes 7 however as only relative amplitudes were measured it was not possible to extract mechanical properties from this model. This seminal work triggered a series of ex- periments aiming at inferring mechanical properties of biological and biomimetic membranes from precise measurement of their fluctuation spectrum. In a further improvement, image processing algorithms were developed to determine the time dependent fluctuation amplitudes as a function of lateral modes in liposomes and red blood cells. This was used by Sackmann and coworkers to gain experimental access to membrane properties such as the bending modulus and, in later work, to infer an effective tension value in red blood cells [28, 29]. Over the following years, these techniques have been successfully refined, taking advantage of the rapid de- velopment in computer processing power and optical microscopy methods, paired with faster image acquisition methods [30]. However, in all these works, pure pas- sive membrane models were used to analyze the data, including the case of red blood cells. First approaches to study the fluctuation dynamics of active membranes were done again on the bacteriorhodopsin system, where a mode-dependent en- hancement of the fluctuations was confirmed [21]. More recently, optical tweez- ers have been used to systematically excite well defined fluctuation modes and to subsequently study the mode-dependent relaxation behavior of the membrane [31]. Mode-dependent studies of active membranes allow to study possible com- plex mode-couplings due to activity or nonlinear interactions, which have remained overlooked so far. This approach may turn to be essential to study large biologi- cal membrane fluctuations, where the cytoskeleton may be the dominant source of active forces. 2.3 Interferometric methods: edge and height fluctuations Video microscopy analysis of membrane movement still suffers from limited spatial and temporal resolution and from complex and long image processing. To overcome this, a number of alternative methods have been developed that exploit interferomet- ric approaches to gain sub-nm precision, with sometimes even µs time resolution. Here we will briefly discuss different approaches and their application to active membranes. Among the first interferometric techniques applied to measure mem- brane fluctuations is the Reflection Interference Contrast Microscopy (RICM, Fig- ure 2b) which allows to determine the distance between a membrane and the glass substrate with a resolution down to 1 nm [32]. Monochromatic light is reflected at the glass-medium interface as well as at the medium-membrane interface. Both reflected waves interfere on the camera either destructively or constructively, de- pending on the phase shift ∆λ of the light. The interference represents the distance between the membrane and the glass, and is used to detect height fluctuations of a membrane in close proximity to the glass. RICM has has been extensively used to study attachment phenomena both, in equilibrium and non-equilibrium situations [33, 34]. 8 Herv´e Turlier and Timo Betz A further interferometric method, that was used to detect active membrane fluc- tuations in red blood cells, is Diffraction Phase Microscopy (DPM, Figure 2c). Here the optical path of the light traveling through the object creates an interference pat- tern where the zero and first order diffracted beams are arranged to interfere on a camera chip. The final image has diffraction limited resolution in the image plane, and nm precise resolution of membrane height fluctuations. Combined with a fast camera this allows spatially resolved membrane height fluctuations with a time res- olution that essentially depends on the camera acquisition speed. Using this method it was recently shown that the positional probability distribution of the red blood cell membrane height has non-Gaussian contributions, that were proposed to re- sult from active processes [5]. However, also other possible explanation can give rise to such non-Gaussian behavior in a equilibrium situation, such as nonlinear force-displacement relations often found in biological systems. Interestingly, these additional contributions depend on the local curvature, with increased fluctuations at more curved regions. While RICM and DPM are sensitive to height fluctuations of passive and active membranes, another interferometric method was developed to determine the lateral fluctuations on a membrane at the rim of a cell or a liposome. The method is called time resolved membrane fluctuations spectroscopy and relies on a tight illumination of the membrane edge where part of the light penetrates the membrane and part is not interacting with the object (Figure 2d). Typically this is done by a focused laser beam that is precisely positioned at the membrane interface [7]. As the common objects of interest provide a higher internal refractive index than the medium, the photons that traverse the object acquire a phase delay with respect to the photons that do not interact with the object. All the light is collected by a high numerical aperture condenser and the back focal plane of this condenser is imaged on a posi- tion sensitive detector or a quadrant photodetector. A calibration curve relates the position of the membrane to the signal on the detector and allows to determine the membrane fluctuations with sub-nm and µs temporal precision. The advantage of this method is that it is very fast, it does not require complex post-processing and it can be applied to any membrane oriented parallel to the axial direction of the laser focus. This technique was successfully used to determine the fluctuation spectra of red blood cells, growing membrane blebs and biomimetic liposomes [7, 36, 37]. A similar technique derived from dark field microscopy was previously used to deter- mine the relative fluctuations amplitudes of membranes in the context of red blood cells [4]. In the absence of a calibration procedure, this initial approach was not able to determine the absolute membrane position, but the relative measurement gave the first insights into possible active effects on red blood cell membrane fluctuations. 2.4 Fluorescent detection of axial fluctuations Recently, a fluorescence based method has been used to determine height fluctu- ations of red blood cells and other cell types by combining a confocal imaging Fluctuations in active membranes 9 method with a fast detector. Called Dynamic Optical Displacement Spectroscopy (DODS) this technique is in principle closely related to fluorescence correlation spectroscopy, with the difference that not the lateral diffusion of membrane bound molecules, but the out-of-plane movement of the full membrane is measured. When a high concentration of fluorescent molecules is present in the membrane, it is not the number of molecules in the focus that dominates the signal, but the position of the membrane with respect to the focal plane. Since the detector is placed behind a confocal pinhole, only the fluorescent signal originating from the focal region is detected. The resulting fluorescence intensity as a function of the axial (z) position corresponds to the point spread function of the pinhole. Hence, the fluorescent in- tensity can be translated into the axial position of the membrane. This results in a axial precision below 20 nm and a temporal resolution of 20 µs. DODS success- fully verified the non-Gaussian behavior of the membrane fluctuations in red blood cells and the dependence of these fluctuation on the local curvature of the membrane [35]. In contrast to the DPM method introduced above, DODS has a superior time resolution similar to the time resolved membrane fluctuation spectroscopy. Further- more, it is not limited to samples providing a homogeneous optical density, but can, in principle, be used on any object. 2.5 Optical tweezers A direct method to probe active membranes is based on optical tweezers (Figure 2f). Here the momentum transfer of a highly focused laser generates a net force on arbitrary shaped particles if their refractive index is higher than the surround- ing medium. In a typical situation, micrometer sized beads are trapped, calibrated and then used to apply well defined forces on the object of interest. Initially, optical tweezers were used to measure the mechanical properties of red blood cells in the linear and nonlinear regimes [38, 39]. Many optical tweezer setups are also capable to detect the movement of single particles with very high precision. Here, the same methodology as for the time resolved membrane fluctuation spectroscopy is used, where a position sensitive detector measures asymmetries in the light deflected off the object in the beam path. The combination of these capabilities was used to de- termine both the mechanical response function of a red blood cell membrane, and the spontaneous fluctuation spectrum of the same cell. For this, polystyrene beads are attached to the membrane and serve simultaneously as handles to fix the cell in space, as probe particles that can apply precise forces and as position sensors to follow the membrane deformation. In a first experimental system two beads were attached to a red blood cell, and both the mechanical properties as well as the fluc- tuations were studied. However, in this approach the fluctuations were not fully free as the beads needed to be still optically trapped, and hence the free membrane fluctuations were partially suppressed by the trapping potential created by the laser [40]. To explore the full fluctuation spectrum, a more complex four beads system was developed [8]. Here, four beads are attached on opposing sides on the rim of 10 Herv´e Turlier and Timo Betz the red blood cell, and three of them are trapped by a tweezer and serve as handles to maintain the cell at a well-defined and stable position. The fourth bead is either forced to oscillate with a well-defined frequency, or simply allowed to move freely without any force to measure the free membrane fluctuations. In this setup the probe bead is not restricted in its movement by a laser trap. If the membrane fluctuations are purely thermal, equilibrium statistical mechanics connects the membrane fluctu- ation characteristics with the dissipative response of the membrane, and the tweezer setup was used to check this correspondence. The power of this approach is that it does not only provide clean experimental access to systematically study the active fluctuations and mechanical properties of a membrane, but it also results in precise frequency-dependent data. This permits comparison to detailed theoretical models, which include temporal characterization of the active membrane fluctuations. 3 Analysis of active membrane fluctuation data The different experimental methods described above provide various determina- tion of membrane fluctuations, ranging from their indirect assessment by the ex- cess area, to direct spatio-temporal measurements using video-microscopy. To give sense to these experimental data, analysis methods have been developed. They can be separated into static analysis, that presents typically time averages or histograms without consideration of time-dependent aspects, and on the other hand dynamic analysis concepts such as the auto-correlation function and the power spectral den- sities, which directly quantify temporal variations in membrane position. In general, dynamic analysis provides more extensive insight in fluctuations properties, but it requires stringent time resolution and a large number of data-points to produce sta- tistically relevant results. Analysis of fluctuations requires to introduce an adapted theoretical formalism for the interpretation of quantitative data. The Helfrich's physical framework has proven its broad relevance to biological membranes since its introduction in 1973 [41], and is briefly overviewed in the following for a membrane at equilibrium. 3.1 Theoretical framework for equilibrium membrane fluctuations 3.1.1 Helfrich's membrane Hamiltonian 2 (2C−C0)2 for the bend- Helfrich proposed a surface energy density of the form κ ing energy of a lipid bilayer, where C is the local mean curvature and C0 a spon- taneous curvature (that we will generally ignore in the following). Note that we generally disregard an additional Gaussian curvature term, which reduces to a con- stant when integrated over a closed, or periodic surface. Since stretching lipids from one-another represents a high energy cost at the bilayer level, one may generally Fluctuations in active membranes 11 consider the membrane total area constant. From a theoretical point of view, this area constraint is enforced via a Lagrange multiplier denoted σ, and which is in- terpreted physically as the membrane surface tension. The Helfrich energy H is written as the sum of these bending and tension contributions, integrated over the total membrane area A (cid:90) (cid:110)κ A 2 (cid:111) H = (2C−C0)2 + σ dA (1) 3.1.2 Static membrane fluctuations spectrum We consider here a quasi-planar membrane of bending modulus κ and surface ten- sion σ, as sketched in Figure 1a and we describe its shape in the Monge represen- tation, where the position vector of the bilayer mid-plane is measured by the height vector(cid:0)(x,y), h(x,y)(cid:1) =(cid:0)r, h(r)(cid:1). In the limit of small deformations(cid:12)(cid:12)∇h(r)(cid:12)(cid:12) (cid:28) 1 the Helfrich Hamiltonian (1) can be written, to a constant (cid:90) (cid:110)κ 2 (cid:2)∇2h(r)(cid:3)2 + σ 2 (cid:2)∇h(r)(cid:3)2(cid:111) H = d2r A , (2) where A denotes the projection of the membrane area A onto the plane (x,y). Static fluctuation modes To study fluctuations, it is more convenient to work in spatial Fourier space, and we consider hence a square piece of membrane with periodic boundary conditions. The A ∑q hq eiq·r where membrane height in real space can be decomposed as h(r) = 1 A drh(r)e−iq·r is the Fourier component for the wave-vector q of membrane hq ≡(cid:82) deformation. Inserting this expression in the Helfrich Hamiltonian (2) yields (cid:0)σq2 + κq4(cid:1)hq2 . H = 1 2 ∑ q (3) If the membrane is at thermal equilibrium, we can use the equipartition theorem, which assigns an average energy kBT /2 energy to each Fourier mode, yielding the static membrane fluctuation spectrum (cid:10)hq2(cid:11) = kBT σq2 + κq4 , (4) where (cid:104).(cid:105) denotes a statistical average. size L ∼ √ The modes of deformation are limited for a real membrane, which has a typical A and is composed of lipids of microscopic size a ∼ 0.5nm. These two length scales allow us to define the following macroscopic and microscopic modes cutoffs: qmin ≡ 2π/L and qmax ≡ 2π/a. In the reasonable limit a (cid:28) L, the fluctuation spectrum amplitude can be calculated by integrating the static fluctuation spectrum 12 (4) over all wave-vectors in the range(cid:2)qmin, qmax (cid:18) (cid:10)h2(cid:11) = (cid:3), which yields (cid:19) kBT 4πσ ln 1 + σ κq2 min Herv´e Turlier and Timo Betz (5) (cid:112)κ/σ = 2π q−1 From the equation (5), one readily observes that a characteristic length λσ = σ may be defined from the bending modulus and surface tension. Two membrane fluctuation regimes can be identified, depending on the amplitude of membrane wave-vector q = q relative to the intrinsic length scale qσ : • If q(cid:29) qσ , the fluctuations are controlled primarily by modes in q4, corresponding to the membrane bending elasticity term. In this regime, membrane fluctuations are dominated by the longest deformation wavelength, and their squared ampli- For a bilayer of typical size L ∼ 10 µm, bending modulus κ ∼ 10kBT and van- tude scales as(cid:10)h2(cid:11) ∼ kBT ishing tension, one would obtain(cid:112)(cid:104)h2(L)(cid:105) ≈ 3 µm. the fluctuation squared amplitude as(cid:10)h2(cid:11) ∼ kBT obtains an amplitude(cid:112)(cid:104)h2(L)(cid:105) ≈ 60nm. • If q (cid:28) qσ , the tension modes in q2 dominate fluctuations, and one can evaluate For typical lipid size a ∼ 0.5nm and membrane tension σ ≈ 10−5N.m−1, one (cid:16) σ 4πσ ln (cid:17) A π2 . κ q2 min 4πκ . Excess area and membrane tension We see above that the amplitude of thermal fluctuations is strongly reduced when the membrane is tensed. As we mentioned before, the membrane tension is in fact intrinsically related to the constraint of conserved membrane area. To formalize this relation one may introduce the area excess α, which measures the difference between projected area A and total membrane area A [42] (see Figure 1a) (cid:90) α = A − A A ∼ 1 2A (cid:17) dr ∇h(r)2 = kBT 8πκ ln A (cid:19) (cid:18) q2 4πκ ln(cid:0) L a max + σ /κ q2 min + σ /κ (6) (cid:1). For intermediate (cid:16) κ q2 8πκ ln For vanishing tension, this formula yields simply α = kBT tension values qmin (cid:28) qσ (cid:28) qmax, tension is directly related to the excess area as α = kBT . Knowing the membrane excess area, one can therefore calcu- late the membrane tension directly by inverting this relation. Note that for higher tension values one needs to consider an additional term σ Kc to the excess area to take into account of the lipids stretching, characterized by a bulk modulus Kc [43]. min σ Fluctuations in active membranes 3.1.3 Dynamic fluctuation spectrum Membrane Langevin dynamics 13 Biological or bio-mimetic membranes are embedded in aqueous solutions, which need to be displaced when the membrane deforms. Since inertia is negligible at this scale, fluid flow can be described by Stokes equations −∇p(r) + η∇2v(r) = −f(r) ∇· v = 0 (7) (8) where v, p and η are respectively the fluid velocity, pressure and viscosity, and f is a bulk force in the fluid. The viscous force opposing membrane displacement can be inferred by calculat- ing the flow generated by a point-like force f(r) = Fδ (r) and integrating this force along the membrane. The Stokes flow solution is related to the point-like force f through a Green's function Λ (r), called the Oseen tensor [44] and defined in real space as v(r) =(cid:82) Λ (r− r(cid:48))f(r(cid:48))d3r(cid:48). Following the derivation proposed in [44], the diagonal part of the Oseen tensor may be calculated as Λ (r) = 1 8πηr (9) To obtain the membrane dynamics, we note that normal velocities of the fluid and of the membrane should coincide at the membrane surface v = ∂ h ∂t , and that the membrane exerts an instantaneous elastic restoring force per area on the fluid f el(r,t) = − δ H δ h(r,t). By adding an additional white noise term in the force, to ac- count for thermal agitation, we obtain an overdamped Langevin dynamics for the membrane height h(r,t) (cid:90) (cid:26) ∂ h ∂t (r,t) = d3r(cid:48) Λ (r− r(cid:48)) − δ H δ h(r(cid:48),t) + ζ th(r(cid:48),t) (cid:27) (10) (11) The thermal noise term has zero mean and its correlations obey the fluctuation- In spatial Fourier space, this equation reads more concisely ∂ hq(t) ∂t = Λq (cid:105) q (t) (cid:104)−(cid:0)κq4 + σq2(cid:1)hq(t) + ζ th (cid:69) q(cid:48) (t(cid:48)) = 2kBTΛ−1 dissipation relation(cid:68) (12) where the Oseen tensor in Fourier space is given by Λq = Λq = 1/4ηq, with q ≡ q. q (t)ζ th ζ th q δ (q + q(cid:48))δ (t −t(cid:48)) 14 Power spectral density Herv´e Turlier and Timo Betz From the membrane dynamics (11), we can directly compute in temporal Fourier space the mode-dependent autocorrelation function for an equilibrium membrane where ωq is the typical membrane relaxation rate for the mode q, given by (13) (14) (15) (16a) (16b) 2kBTΛq ω2 + ω2 q = (cid:68)hq(ω)2(cid:69) (cid:0)κq4 + σq2(cid:1) = ωq ≡ Λq κq3 + σq 4η (cid:68)h(ω)2(cid:69) −→ 6π(2η2κ)1/3ω5/3 ω→∞ (cid:68)h(ω)2(cid:69) −→ ω→0 kBT kBT 2σω To obtain the equilibrium membrane fluctuation spectrum (or power spectral den- sity, noted PSD), one has to integrate the autocorrelation function (13) over all de- formation modes explored by the membrane (cid:68)h(ω)2(cid:69) = (cid:90) (cid:90) qmax dq (2π)2 2kBTΛq ω2 + ω2 q = 4ηkBT π dq qmin (4ηω)2 + (κq3 + σq)2 Supposing qmin ∼ 0 and qmax ∼ ∞, the PSD scales as ω−5/3 and ω−1 in the limit cases of, respectively, high and low frequency Fluctuation-dissipation relation To write the fluctuation-dissipation relation, we have to determine the mechanical response function of the membrane, defined in temporal Fourier space as χ(ω) ≡ h(ω)/F(ω), where F is an external driving force. Adding this external force to the mode-dependent Langevin dynamics (11), we obtain in temporal Fourier space (cid:104)−(cid:0)κq4 + σq2(cid:1)hq(ω) + Fq(ω) + ζ th (cid:105) − iωhq(ω) = Λq q (ω) (17) Supposing that the driving force F is much larger than thermal noise ζ th, the mode-dependent response function is obtained as χq(ω) = Λq−iω + ωq (18) The response function can be separated into real and dissipative parts χq(ω) = q(ω) + iχ(cid:48)(cid:48) χ(cid:48) q (ω), leading to Fluctuations in active membranes χ(cid:48) q(ω) = ωqΛq ω2 + ω2 q χ(cid:48)(cid:48) q (ω) = ωΛq ω2 + ω2 q 15 (19) By identification with the auto-correlation function (13), and after summing on the modes q, we deduce the fluctuation-dissipation relation C(ω) ≡(cid:68)h(ω)2(cid:69) = 2kBT ω χ(cid:48)(cid:48)(ω) (20) 3.2 Indirect fluctuation analysis In micropipette aspiration experiments, tension and excess areas constitute inherent average measures of the fluctuations over the full liposome area. As introduced in Figure 1 and in equation (6), the membrane fluctuations lead to a difference between the projected membrane area A, and the total membrane area A . Experimentally accessible is the change in excess area ∆α ≡ α0 − α, obtained by measuring the radius of the liposome and the variation of the length of the membrane tongue δ L while aspirating with a micropipette. Here α0 is the excess area at the minimal tension σ0 sufficient to suck up the liposome at the start of the experiment [20]. As long as the membrane fluctuations remain in the entropic regime, we have shown above that the excess area takes the form α = kBT , from which we easily deduce the excess area change upon aspiration (cid:16) κ q2 8πκ ln (cid:17) min σ (cid:18) σ (cid:19) σ0 ∆α = kBT 8πκ ln (21) The tension is controlled by the aspiration pressure and can be calculated using Laplace's law [20]. In the experimental procedure the tension is systematically in- creased and the resulting excess area change is plotted over the tension which is nor- malized to the initial tension. This leads to datasets such as presented in Figure 2a. In the case of active ion pumps in the membrane one finds systematically a reduced slope of the experimental curve [20]. As the slope depends on the ratio between the temperature T and the bending modulus κ an initial analysis suggests investigation of these two parameters. To describe changes in the slope of the experimental data, an effective temperature Teff is introduced. Since the physical reason for the excess area in passive systems is indeed a temperature dependent fluctuation, this intuitive modification provides a simple way to retain the analytical expressions. However, it should be mentioned that due to the spatial and temporal averaging any possible frequency dependence of Teff is not accessible. This is important in the sense that the concept of temperature has no time dependence. Such an approach reflects the underlying assumption of statistical mechanics that thermal noise is a delta corre- lated stochastic variable with no inherent timescale. However, a dynamic analysis of the membrane fluctuations shows that the effective energy Eeff = kBTeff may be a 16 Herv´e Turlier and Timo Betz Fig. 3 Summary of the analysis methods to quantify active and passive fluctuations. a) Us- ing the histogram of the measured membrane fluctuations allows to quantify possible deviations from the expected Gaussian probability distribution. Commonly such non-Gaussian behavior is interpreted to result from active processes. b) The detailed mode analysis of the fluctuation in membranes shows an increased of fluctuation amplitude and its characteristic mode dependence (Data reproduced with permission from [21]) c) If only time dependent data is available, a common approach is to use the autocorrelation function and fit it with an exponential function, or more com- plex functions to describe the active component. (Data reproduced from [35]) d) Another possible way to analyze time dependence membrane fluctuation is to determine the fluctuations spectrum. Membrane models can account for the different powerlaw observed in these measurements. e) A direct check of activity can be achieved by comparing the directly measured response functions (e.g. via optical tweezers) with the expected response function that is obtained using the free fluc- tuations spectrum and applying equilibrium physics. Possible differences as found in red blood cells directly show quantify the extend of activity. (Figures adapted from [8]) Time[s]zz9zz/z9zz4z9zz6z9zz8z9zRMembraneposition[nm]O6zO4zO/zz/z4z6zPosition[nm]ORzzzRzzFrequencyRzO4RzO3RzO/RzORRzzPosition[nm]ORzzzRzzFrequencyRzO4RzO3RzO/RzORRzzMembranepositionhistogramGaussianFitATPdepletedRBCNormalRBCActive,non-Gaussiana)q=1/Rq=2/Rq=4/Rc)ActivePassivehCμmDCsDzz95Rz -- 4Rz -- /Rzzz6R/×Rz -- 3dACFCμm/DCsDDataFitb)te)zz9/z94z96z98−/zzz/zzAppliedforceCfNDOscillatingforceanddisplacementatR9Hz−/zz−RzzzRzz/zzTimeCsDDisplacementCnmDRz−RRzzRzRRz/Rz3Rz3Rz4Rz5Rz6Frequency[Hz]Response[mHN]Cπf/kBT,freshχ",freshd) Fluctuations in active membranes 17 frequency-dependent quantity that is hence qualitatively different from the classical concept of temperature. 3.3 Static analysis In contrast to the micropipette experiment, the static analysis of the data is often based on time-dependent raw data, such as video images. To obtain better statistics, time dependent data is condensed into histograms of the position or, in the extreme case, into a mean value and the standard deviation (Figure 3a). This approach is very useful to reduce experimental noise if only a limited number of measurements are available. As the name points out, the static approach is fundamentally limited to static properties such as an elastic energy storage module that does not depend on timescales and that does not have inherent relaxation behavior. It should be men- tioned that even such static measurements are still restricted to certain timescales that are given by the total acquisition time and the sampling rate of the raw data. Typically it assumes that the total acquisition time is large enough for the system to explore all possible conformations. This is for example the case when a proba- bility distribution of visited membrane positions does not change shape for longer measurement times. In this view, an important point is that changes in friction or viscous properties lead to slower dynamics, which might require a critical check of recording timescales: at higher viscosity it can take much longer to explore all possible configurations. Regarding image analysis, the membrane position is commonly detected by im- age processing and the amplitude for the different modes is extracted using Fourier analysis or spherical harmonics decomposition. The average amplitude of the dif- ferent modes is then plotted as a function of the mode number (Figure 3b). Other possible analysis are cross-correlation curves, where distance-dependent correlation functions are calculated to understand the lateral length scale over which mechanical interaction are mediated by the membrane(Figure 3c). In the case of active fluctu- ations initiated by bacteriorhodopsin, the activity enhances primarily lower modes (i.e. large wavelength) [21] (see Figure 3b). This particular feature reflects in the time-domain, where low frequency fluctuations - corresponding to low modes - are also predominately enhanced by activity. Besides these direct observation of active fluctuations in model membranes, a static analysis was typically used to identify initial signs of active fluctuations in the red blood cell membrane. First measurements of relative membrane position showed relative changes in fluctuations upon ATP depletion, which was interpreted as a sign for activity. However, as the whole cell becomes stiffer upon depletion of intracellular ATP, a pure equilibrium interpretation may be proposed. Additional ex- perimental results like changes of the static fluctuations amplitude when the buffer viscosity was changed did indeed hint for an active process [4], however these ex- periments could not been reproduced. It is possible that insufficient recording time 18 Herv´e Turlier and Timo Betz may not allow the membrane to explore all possible conformations in this pioneer- ing experiment. More recently, the static analysis of membrane fluctuations was used to deter- mine possible deviation from a Gaussian probability distribution of the membrane fluctuations. Such non-Gaussian contributions can be explained by active processes that generate membrane configurations hardly reachable from pure thermal agita- tion (Figure 3a). A key feature of these non-Gaussian elements is that they are quite rare, which requires excellent statistics to be able to detect them. Using DRM a non- Gaussian parameter was indeed detected in the height fluctuations of red blood cells [5]. In the case of ATP depletion, this non-Gaussian behavior was reduced, which was interpreted as a sign for activity. These results have been later confirmed by DODS. While this was an intriguing result, such a non-Gaussian behavior cannot be a conclusive proof of active fluctuations, as any nonlinear behavior in the membrane could lead to similar results. 3.4 Dynamic analysis When a large number of time-dependent measurements with good temporal reso- lution is available, it becomes reasonable to exploit time as additional dimension. Temporal analysis is interesting as it provides access to dynamic variables such as processes related to energy dissipation and friction. These parameters are funda- mental to understand and model the mechanical processes involved in both passive and active fluctuations. This becomes evident when considering energy dissipation as the fundamental reason for thermal fluctuations. Briefly, thermal excitation of a membrane fluctuation implies an energy transfer from the thermal bath to the mem- brane which in turn requires a reduction of the bath's temperature. However, driving a movement by cooling a reservoir is in contradiction to the rules of thermody- namics. The fluctuation-dissipation theorem states that the same amount of energy is returned to the thermal bath by dissipating the energy stored in the movement, thus leading to a heating that exactly compensates the 'cooling' required to drive the fluctuations. Having this in mind, it becomes clear why the dissipation of active fluctuations gives a useful quantity to describe active processes, because it allows to measure the energy added to the system by the active process. Hence, to gain more information about the elastic but also dissipative properties of an active system, a time dependent analysis is vital. A common approach to look at time dependence is to study the autocorrelation function (ACF) of the position x(t), typically denoted < x(τ)x(0) >. Depending on the type of analysis, the ACF is often normalized by subtracting the mean square of x(t) and dividing by the variance of x(t). When applying these normalizations the ACF becomes somewhat intuitive, as it varies only between ±1, where a value of 1 marks a total correlation and -1 marks total anti-correlation. Often, an exponential decay of the ACF is found, which can be associated to a single relaxation process with timescale τr (Figure 3c). In contrast, many systems show powerlaw behavior in Fluctuations in active membranes 19 the relaxation, which corresponds to a complex relaxation scheme with many pro- cesses. In the context of membranes, the ACF of a single mode typically relaxes exponentially. Every fluctuation mode has a characteristic wavelength λ (typically reported: wavenumber q = 2π/λ ) that corresponds to a characteristic relaxation frequency ωq = (κq3 + σq)/(4η), as introduced in equation (14), where η is the viscosity of the surrounding buffer solution. If the fluctuations are measured at one single position, they present the superposition of all accessible modes and hence the ACF results in a more complex powerlaw relaxation behavior. The overall advantage of the ACF lies in the visual and analytical interpretation of data containing a low number (1 to 3) of relaxation processes that are well separated in timescales, ide- ally by at least one order of magnitude. Also, the ACF can be efficiently calculated directly from the input data using dedicated hardware. It should be mentioned that the ACF function plays a key role in fluorescence correlation spectroscopy (FCS), which is a powerful tool in the study of lateral movements of lipids or membrane bound proteins. A key advantage of the correlation function is to combine tempo- ral and spatial analysis by correlating two functions that represent the membrane movement at two different positions separated by a well defined distance. This two- point correlation can give information about timescales and distances over which mechanical forces can act on the membrane. Such analysis was used to give direct hints for active processes representing force dipoles with characteristic length scales [5]. Besides the ACF, a second analysis type commonly used is the power spectral density (PSD), which is accessed by computing C(ω) = x(ω)× x(ω)∗ , where x(ω) is the Fourier transform of the time dependent signal x(t), and p and s are the sample rate and number of datapoints respectively. The PSD refers to 'power', as it was originally used to get the electrical power of voltage measurements. Hence, strictly speaking the PSD of the membrane position is not a mechanical power, but reflects the fluctuation amplitude. Figure 3d shows a typical PSD calculated from transverse fluctuations of a membrane. Important characteristic features are that the low fre- quencies provide a large amplitudes while at high frequencies the amplitudes are small. This is qualitatively explained by friction which prevents large amplitudes at high frequencies. The quantitative behavior of the PSD can be easily described in the case of an equilibrium quasi-flat membrane, as we presented in the paragraph 3.1.3: in the high-frequency regime, the spectrum is dominated by the bending mod- ulus κ and shows a −5/3 power-law, while in the low-frequency regime, dominated by the tension σ, the exponent changes to −1 (see equation (16) and Figure 3d). A quantitative model can be fit to the PSD to directly determine the mechanical properties of the membrane. In principle, the ACF and the PSD are exchangeable, as they are related to each other by a Fourier transformation in the time domain. Advantages of the PSD are that single frequency noise sources, such as electronic 50 Hz noise show up as delta peaks and can be easily eliminated. The ACF however, allows to directly identify if a single or if multiple relaxation processes act on the system studied. p×s In both cases, activity typically shows up as an increase in fluctuation ampli- tudes and possible difference in the power-laws from the equilibrium case. How- 20 Herv´e Turlier and Timo Betz ever, without a direct measure of the mechanical properties it is difficult to get a model-independent measure of active fluctuations. 3.5 Direct measurement of the mechanical response function To experimentally prove the active nature of membrane fluctuations and to quan- tify the contribution of active forces to the fluctuations it is necessary to know the mechanical characteristics of the system. To measure mechanical properties, a well defined force is applied and its resulting deformation is measured. Since typical systems of interest such as cells and liposomes are marked by viscoelastic charac- teristics, it is important to measure time-dependent response functions. In principle this can be done by measuring the time dependent relaxation after application of a step force with an atomic force microscope, a magnetic tweezer, an optical tweezer or a calibrated micropipette. The time-resolution provided by step response is how- ever generally not satisfactory in practice, and frequency response, where the system is probed successively for different frequencies, is generally preferred. For biolog- ical membranes we are interested in the mechanical properties at the micrometer scale, and hence the corresponding methods are called microrheology. Classically microrheology is separated into active and passive methods. In active microrheol- ogy, an external force is applied to the system to measure its response, while in passive microrheology, the spontaneous fluctuations of the system are used to infer its mechanical properties. It should be mentioned here again that for active and in particular living systems, passive microrheology may not be used to infer mechani- cal response, unless the properties of the active process are known and integrated in the analysis. In the context of active membranes, optical tweezers based microrheology has been used to measure the mechanical response of living cells, such as red blood cells and other eukaryotic cells [45]. In these measurements, the beads are attached to the plasma membrane and an oscillating force is applied while the response of the sys- tem is measured by following the bead movement. The advantage of the oscillatory driving is that noise and additional fluctuations can be removed by selecting the driv- ing frequency f = ω/2π in the analysis. The response χ is then simply obtained by χ( f ) = x( f )/F( f ), where F( f ) is the oscillating driving force of frequency f . This experiment is repeated for driving forces of different oscillating period, to finally ob- tain the complex response function as a function of driving frequencies (Figure 3e). This complex response function can be separated into its real and imaginary part, corresponding to the elastic (χ(cid:48)( f )) and dissipative (χ(cid:48)(cid:48)( f )) response. This approach was used to determine the dissipative response in the red blood cell membrane which was a key element to quantify active membrane fluctuations [8]. Furthermore, the same approach allowed to determine the active fluctuations of granules in mouse oocytes cells [46] and of particles embedded in biomimetic actomyosin systems [47]. Fluctuations in active membranes 3.6 Test of fluctuation dissipation theorem 21 Combining the dynamic fluctuations analysis with mechanical response measure- ments allows to decouple the active part from the thermal driving in the membrane fluctuations. This presents a key element to evaluate the pertinence of theoretical models describing active fluctuations. Here we use the example of active fluctu- ations in the red blood cell membrane [8], but the measurement principle can be equally applied to other systems. In the case of red blood cells, four beads are at- tached to the membrane, and the red blood cell is held in space by trapping three handle beads. The free fluctuations are measured by decreasing the laser intensity on the fourth bead (probe bead) to a level where the optical detection system works reliably, but where the trapping force on the particle is negligible compared to ther- mal fluctuations. In this situation the movement of the bead can be assumed to be purely thermally driven, while being precisely recorded. The PSD C( f ) is calculated as described in section 3.4. In a second step of the experiment, the laser power on the probe bead is increased and the laser beam is moved in a sinusoidal fashion. The resulting oscillatory forces on the bead are recorded using the optical detection sys- tem. As described in section 3.5, the frequency-dependent response function χ( f ) is obtained. In an equilibrium situation, the free fluctuations and the dissipation are connected by the fluctuation dissipation theorem (20), that we rewrite here as function of the 2π f χ(cid:48)(cid:48)( f ), where χ(cid:48)(cid:48)( f ) is the dissipative part of the response frequency f : C( f ) = 2kBT function. From an analysis point of view, the response function expected from an equilibrium system is obtained by measuring the PSD as χ(cid:48)(cid:48)( f ) = C( f )π f /kBT . Plotting this function and the response function directly measured on the system provides two curves, which shall collapse if, and only if the system is at thermody- namic equilibrium (Figure 3e). Any discrepancy between the two curves is therefore a clear evidence for an additional, non-equilibrium process driving active fluctua- tions. By analogy with the fluctuation-dissipation relation, one may define the effec- tive energy and temperature as Eeff ≡ C( f )π f /χ(cid:48)(cid:48)( f ) ≡ kBTeff. It becomes clear that the effective temperature is, in general, timescale dependent, which is somewhat contradictory to the concept of temperature itself. By measuring χ(cid:48)(cid:48)( f ) and C( f ), the effective energy becomes directly experimentally accessible, which is an inter- esting method to investigate active fluctuations from a thermodynamic perspective, and to detect the frequency onset of active energy input in the system. For the red blood cell membrane, metabolic activity is found to only excite the membrane at short timescales: for frequencies higher than 10Hz, thermal fluctuations dominate the fluctuations, while at timescale lower than 100 ms, active contributions largely exceed thermal driving. Previous methods using the FDT, or passive microrheol- ogy, to infer the mechanical properties of red blood cell membrane have hence to be re-considered carefully. 22 4 Theoretical models of active membrane Herv´e Turlier and Timo Betz The advent of theoretical modeling of active membranes may surely be associated to the seminal work of Prost and Bruinsma in 1996 [16]. In this landmark paper, an hydrodynamic model predicts the influence of ion channels gating activity on mem- brane fluctuations. Since then, a large set of models were proposed to account for di- rect or indirect sources of active noise in biological membranes and to analyze their specific features. All these hydrodynamic models take essentially a similar general form, where active and thermal noise are uncorrelated and may therefore be split into two distinct contributions. A common structure for active membrane models is therefore proposed. Then we discuss intrinsically active membrane models, where the active sources of noise comes from processes directly embedded in the bilayer, and extrinsically active membrane models, where activity is cytoskeleton-based but may be transmitted to the bilayer through mechanical coupling. Note that we do not cover here electrokinetic active membrane models, which have been recently reviewed elsewhere [48]. 4.1 General structure of active membrane models Starting from the equilibrium Langevin dynamics in Fourier space (11), the height dynamics for an active membrane may in general be cast into the generic following form (22) ∂ hq(t) ∂t + ωqhq(t) = Λqζ th q (t) +Λqζ a q(t) where ωq is the mechanical relaxation frequency of the membrane for the mode q, ζ th q and ζ a q are thermal and active sources of noise, respectively, and Λq is a mode- dependent dissipative coefficient (e.g. the Oseen tensor for membranes surrounded by viscous fluids). In general, the membrane relaxation frequency takes the follow- ing form ωq = Λq δ H /δ hq(t), where H is the elastic energy of the membrane. The characteristics of the active noise source term ζ a q(t) are not constrained a priori and may hence depend on the source of active forces, but as soon as this term is non-vanishing, one shall expect a violation of statistical equilibrium. We will hence describe in the following several models for the active source of noise in the membrane. Note that equation (22) can be readily generalized to quasi-spherical membranes with the use of spherical harmonics instead of Fourier decomposition [49, 50, 51]. Fluctuations in active membranes 4.2 Intrinsically active membrane models 4.2.1 Ion channels shot noise activity 23 In the first active membrane model proposed by Prost and Bruinsma [16], active fluctuations originate from the shot-noise activity of ion channels freely diffusing in the membrane (see Figure 1b). By switching stochastically between open and closed states, under the action of metabolic energy, these channels produce an additional source of noise, described by a two-state variable Sk(t) = 1 if the ion channel k is active and 0 otherwise 1. In agreement with single ion-channel gating measurements, the shot-noise is assumed to be exponentially correlated in time g(t) = (cid:104)Sk(t)Sk(0)(cid:105)−(cid:104)Sk(cid:105)2 = g(0)e−t/τa, with a typical correlation time τa. f ∑k Sk(t)δ(cid:0)r− Rk(t)(cid:1), where Rk is the position of the ion channel k in the mem- By changing local osmolarity in the vicinity of the membrane, ion channels gen- erate a local fluid pressure variation across the membrane of the form δ p(r,t) ∝ brane plane, and f ∼ kBT /w is the typical force amplitude exerted on the membrane of thickness w. Supposing the membrane semi-permeable with a permeability de- noted λp, the Langevin equation for the membrane height in Fourier space takes a similar form as (22) ∂ hq(t) ∂t + ωqhq(t) = Λqζ th q (t) + f λp∑ k Sk(t)eiq·Rk(t) , (23) where ωq = κ 4η q3 + κλpq4 is the sum of the relaxation frequency for an imper- meable membrane with vanishing tension and an additional permeation term. The dissipative coefficient is the membrane permeability Λq = λp, and we can identify the active noise term as ζ a q(t) = f ∑k Sk(t)eiq·Rk(t). In the long wavelength limit, the model predicts that area density fluctuations of ion channels of active nature shall dominate the membrane fluctuation spectrum (cid:0)q−4 + ξ−1q−5(cid:1), where ξ is a length scale proportional to the (cid:10)hq2(cid:11) ∼q→0 diffusion coefficient of channels in the membrane and inversely proportional to the shot noise correlation. kBT κ This seminal work triggered a substantial theoretical interest for active mem- branes [8, 17, 19, 20, 50, 51, 53, 54, 55, 56, 57, 58, 59, 60], that we aim to briefly overview in the following. 1 Note that ion channels do not require, in general, metabolic energy consumption and their ac- tivity may hence be considered as passive. However, when a non-zero (electro)chemical potential difference is maintained across the membrane (generally through the action of ion pumps), their gating activity is expected to be of non-equilibrium character [52]. 24 Herv´e Turlier and Timo Betz 4.2.2 Active curvature coupling The first experimental realization of active membranes in vitro was done by reconsti- tuting the transmembrane proton pump bacteriorhodopsin in giant unilamellar vesi- cles [19, 20]. A new theoretical model, considering the coupling between pumps activity and membrane curvature was conjointly proposed by Ramaswamy, Toner and Prost, to explain the experimental results [17]. The new ingredient added to the original model of Prost and Bruinsma [16] is an intrinsic asymmetry in the shape of ion pumps, inducing their preference either for positive or for negative membrane curvature regions (see Figure 4a). Fig. 4 Curvature coupling of ion pumps of asymmetric shape. a) Asymmetric proteins are drawn to regions with curvature adapted to their shape. b) An instability may develop if the cur- vature produced by local pumping has the right sign to attract more pumps. (Figures reproduced from [17]) A signed density of proteins ψ(r,t) = n+(r,t)−n−(r,t), measuring the local dif- ference between proteins with preference for positive curvature relative to proteins with preference to negative curvature, is introduced, and is coupled to the membrane elasticity in equation (2) up to second order in the variables (cid:90) 1 2 d2r A (cid:110) κ(cid:2)∇2h(r)(cid:3)2 + σ(cid:2)∇h(r)(cid:3)2 (cid:111) H (h,ψ) = + µψ2(r)− 2Ξψ(r)∇2h(r) (24) where µ is the susceptibility for the imbalance between curvature positive and neg- ative proteins and Ξ is the curvature coupling coefficient. To close the problem, a conservation law for ψ(r,t) is needed: (cid:19) (cid:18)δ H δψ ∂ψ(r,t) ∂t ∼ Γ ∆ + ∇· ζ th ψ (25) where Γ = D/µ is a mobility coefficient, with D is the diffusion coefficient of pro- teins in the membrane. The activity of positive and negative pumps leads essentially to two additional forces normal to the membrane : a)b) Fluctuations in active membranes 25 • An active permeation term of the form λpFaψ(r,t) in Darcy's permeation equa- tion across the membrane, where Fa is the elementary force transmitted to the membrane by the transfer of a proton. (cid:2)δ (z− w↑)− δ (z + w↓)(cid:3)ψ(r,t) • An hydrodynamic active dipolar force density Fa in the force balance equation between the membrane and the surrounding fluid, where w↑ and w↓ are the distances of the center of mass of the ion pump relative to the membrane mid-plane, which are supposed distinct (see Figure 5). A numerical estimation of these two effects shows that the active permeation term may be omitted, in comparison to the active dipolar force, for the typical micron to submicron length scales relevant to biological membranes [20]. This set of equations results in two coupled Langevin dynamics for h and ψ, which can be solved to obtain the membrane fluctuations autocorrelation and, even- tually, an expression for the areal strain measured in micropipette experiments [20]: ∆α = α0 − α = kBTeff 8πκ ln σ σ0 (26) The model predicts that the effect of pumps activity on the areal strain can be cast simply into an effective temperature Teff, in agreement with experimental results (cid:18) Teff T = κ κ" 1 + Pa Pad2 − Ξd κ(cid:48)d (cid:19) (27) (w↑)2−(w↓)2 where κ(cid:48) and κ" are renormalized bending moduli and Pa = Fa is the work of active dipoles. This formula shows that ion pumps will give rise to active fluctuations in the membrane only in the presence of an asymmetry w↑ − w↓ (cid:54)= 0 in the protein configuration within the bilayer. 2w The coupling between mobile force centers and membrane curvature leads fur- thermore to a new and rich physical behavior, exhibiting localized instabilities and traveling waves [17]. For example, ion pumps with a preference for positive cur- vature regions may trigger an instability: by increasing local membrane curvature through their pumping activity, they will attract more pumps and amplify the curva- ture even greater, as sketched on Figure 4b. The model of Ramaswamy, Toner and Prost, does not consider, however, the random fluctuations in the pump activities (shot noise), introduced by Prost and Bruinsma [16]. In a subsequent paper, Lacoste and Lau addressed therefore simul- taneously the effects of curvature couplings and pump activity fluctuations. While their model predicts similar local instability and travelling wave behaviors, it shows that shot noise effects are essential to consider for dynamical measurements of mem- brane fluctuations, as they lead to different scaling laws at short time scales [59]. 26 Herv´e Turlier and Timo Betz Fig. 5 Asymmetric dipole model used for active proteins: the center of mass of the pumps is displaced above the bilayer midline and force center lies at distances w↑ and w↓ from this midline. The shape asymmetry of the pump is represented by asymmetric triangles. (Figure reproduced from [60]) 4.2.3 'Direct' vs. 'curvature force' and monopole vs. dipole If one neglects density fluctuation effects due to the diffusion of active proteins in the lipid bilayer, the active noise term can generally be assumed to be uncorrelated in space. In analogy with the shot-noise dynamics of ion channels, many models suppose active forces to be exponentially correlated in time, which is the property typically expected for a protein switching between "on" and "off" states [61] (cid:68) (cid:69) q(cid:48)(t(cid:48)) ζ a q(t)ζ a = Γ a q δq+q(cid:48)e−(t−t(cid:48))/τa kon and koff for a two-states metabolic process τa ∼(cid:0)kon + koff τa is a typical active timescale, which may be defined from "on" and "off" rates q is the (cid:1)−1, and Γ a amplitude of the active noise, which may depend on the fluctuation mode q. Supposing that the elasticity of the membrane reduces to the Helfrich energy, the active membrane fluctuation spectrum can be calculated from equations (12), (22) and (28): (cid:10)hq(ω)2(cid:11) = 2kBTΛq ω2 + ω2 q + 1 ω2 + ω2 q 2τa Λ 2 q Γ a q 1 + ω2τ2 a (28) (29) where the intrinsic membrane relaxation ωq was defined in equation (14). In [55], Gov made a important distinction between two possible types of active forces, depending whether the activity of proteins embedded in the membrane cou- ples or not to the membrane curvature: • For a so-called 'direct force', a random force of non-equilibrium origin is ap- plied locally and directly to the membrane [55]. In this case, the active noise Fluctuations in active membranes 27 term ζ a q can be considered as an instantaneous "kick" on the membrane. The mean squared amplitude of the active noise is then independent of the mode q of membrane deformation and reads (30) Fa is the magnitude of the active force exerted on the membrane and ρa ∼ N/A is the density of active proteins in the membrane. Γ a q = ρaF2 a • A 'curvature force' is on a contrary an active random contribution c(r,t) to the spontaneous curvature, which may be introduced by generalizing the bending energy of the membrane as follows(cid:82) ∇2h(r)− c(r,t) (cid:105)2(cid:27) [18]. (cid:26) (cid:104) A dr κ 2 the time correlation function for the curvature of a single protein i as(cid:10)ci As a result, an additional random force density term appears in the right-hand side of the Langevin equation for the membrane (22) in the form ζ a q(t) = −κq2cq(t). Supposing that each active protein i may switch metabolically be- tween positive and negative spontaneous curvatures +c0 and −c0, we can write 0e−(t−t(cid:48))/τa, where τa is, like previously, a typical active timescale for the switch- c2 ing process. Summing over the active proteins in the membrane, the mean squared amplitude of the active noise now depends explicitly on the mode q 0(t(cid:48))(cid:11) = 0(t)ci (cid:0)κc0b2q2(cid:1)2 Γ a q = ρa (31) where b2 is the typical surface area occupied by an active protein. 'Direct force' and 'curvature force' activities are hence predicted to produce differ- ent scaling behaviors for the membrane fluctuation spectrum in the limits of large and small wavelengths [55]. In subsequent works, the same author and colleagues generalized the model to diffusing active proteins coupling to membrane curvature [18], and characterized the fluctuations of a membrane where the proteins are con- sidered to be nucleators of actin filaments [57]. Other authors considered a gen- eralization where the membrane curvature may feedback onto the conformational transition kinetics of the active inclusions [54]. An important distinction has to be made between monopolar, dipolar or quadrupo- lar active contributions. 'Direct forces' are typically force monopoles, which sup- pose essentially that an extrinsic agent, like the cytoskeleton, can push or pull locally on the membrane. Indeed, the spatial integral of the force over the system {protein + membrane + solvent} should vanish by force balance. If an active protein exerts a force on the membrane and solvent, the latter have to exert an opposite force of same magnitude on the protein. This implies that the force density field of the pro- teins has zero monopole moment, unless the system is actually not isolated because extrinsic agents enter the force balance. In general, the first contribution of active membrane proteins is a force dipole [50], which can be idealized by two force cen- ters of opposite sign but equal magnitude embedded in the membrane (see Figure 28 Herv´e Turlier and Timo Betz 5). A dipolar contribution is also expected from a permeation force [51]. In the ab- sence of dipolar contribution, higher multipole moments may however contribute to the active membrane fluctuations, such as for 'curvature forces', which can actually be considered as quadrupoles [50, 51]. 4.2.4 Non-equilibrium fluctuations and excess area In the models described so far, the description of active fluctuations ignores the interplay between excess area and surface tension. As introduced by Seifert for pas- sive membranes [62], membrane tension shall be formally regarded as a Lagrange multiplier for the conservation of total membrane area, and it is therefore intimately related to the excess area. The excess area is, by definition, a function of the fluctu- ation amplitude, which is itself a function of the tension. For a quasi-planar mem- brane, we can deduce from equation (6) 1 2A ∑ q2(cid:68)hq(σ )2(cid:69) α = q (32) Setting the value of membrane excess area, the membrane tension is therefore controlled by the amplitude of fluctuations. The previous relation has, in general, to be inverted numerically, but analytic formula may been obtained from a pertubative approach for three different limit cases depending on the value of the dimensionless parameter kBT /κα. Generalizing this approach to active membranes, Loubet and colleagues show that, by increasing the fluctuations amplitude, the presence of activity in the mem- brane shall, in general, increase the bilayer surface tension [51]. Distinguishing short and long membrane relaxation times ω−1 relative to a typical active time scale τa, q they derived analytical formula for the bilayer tension as function of the excess area of active membranes with either monopolar, dipolar or quadrupolar types of active forces. It should be noted that the phenomenological finding that the fluctuations in- crease due to activity may also be interpreted as a reduction in tension. This can be illustrated by recalling that with the same pulling pressure more hidden membrane can be pulled out of a vesicle [21] when active forces enhance membrane fluctu- ation. This shows that the interpretation of the activity in the context of classical equilibrium approach may be ambiguous. Fluctuations in active membranes 4.3 Cytoskeleton-based active membranes 29 4.3.1 Renormalization of membrane properties by the cytoskeleton The cell cytoskeleton being tightly coupled to the lipid bilayer, biological mem- branes are generally composite materials, where the mechanics is a combination of bilayer and cytoskeleton mechanics. A perturbative approach of the problem is to start with the Helfrich description of membranes, and to study how the presence of a cytoskeleton may renormalize endogenous properties, such as membrane tension or bending rigidity, or may create new types of mechanical response, such as resis- tance to shear or confinement of membrane deformations. Most of the work on the subject has been applied to the red blood cell membrane, where the cytoskeleton is, in comparison to other cell types, a more simple structure, made of a triangular network of extensible spectrin filaments, anchored to the bilayer at junction points via transmembrane protein complexes. Generalizations of these concepts may open new perspectives to characterize membrane fluctuations in cell types possessing an actomyosin cytoskeleton. Membrane confinement by the cytoskeleton The spectrin cytoskeleton in red blood cells has been proposed by Gov and col- leagues to confine the bilayer fluctuations [63], an effect that can be rendered by the addition of an harmonic potential to the Helfrich energy (2) A The confinement term, of amplitude γ, constrains the mean squared amplitude of √ fluctuations to be equal to d2 = 1 γκ. This is equivalent to consider that the 8 kBT / spectrin cytoskeleton acts as a rigid plane at h = 0 that maintains the membrane at an average distance d via an harmonic restoring force. The static fluctuation spectrum of the confined membrane in Fourier space can be calculated at equilibrium as (33) (cid:90) 1 2 V = drγ h(r)2 (cid:68)hq2(cid:69) = kBT γ + σq2 + κq4 (34) The confinement parameter γ defines a new characteristic length λγ = (κ/γ)1/4, which determines the wavelength onset for membrane confinement. In their study, Gov et al. show that adding a confinement parameter of the order of γ ∼ 10−7J.m−4 allows for a better fit of the mode-dependent experimental fluctuation spectrum mea- sured by Zilker et al. in red blood cells [67], which displays an abrupt drop in fluc- tuations for wavelengths above 100nm (see Figure 6). It should be noted that such confinement potential has been introduced originally in the context on membrane adhesion to a surface [64], to render the combined 30 Herv´e Turlier and Timo Betz effect of electrostatic attraction and steric repulsion (by long glycocalyx chains) of membrane to the surface [65]. Effective membrane mechanics A subsequent model was proposed by Fournier et al. to explain this jump in the fluc- tuation spectrum [66]. In this model, the composite red blood cell membrane energy is supposed to be the sum of the Helfrich energy (2) and an elastic contribution from 2 Nk((cid:96)−(cid:96)0)2, supposed to be a perfect network of N the spectrin cytoskeleton Hel = 1 entropic springs of stiffness k and of actual and resting lengths respectively (cid:96) and (cid:96)0. The authors show that this additional term leads to a jump in the membrane tension at wavelengths larger than the mesh size (cid:96) (cid:18) (cid:19) ∆σ ∼ 1 2 gk 1− (cid:96)0 (cid:96) (35) where g characterizes the topology of the network. Fig. 6 Composite structure and static fluctuation spectrum of a quasi-planar red blood cell membrane. a) Sketch of a nearly planar red blood cell membrane of total area A , of coarse- grained area A(cid:96) and projected area A. The spectrin filaments are represented by linear springs of length (cid:96) anchored in the bilayer. b) Fit (plain line) of the static fluctuation spectrum of a human red blood cell measured by [67] with the model of Fournier et al. [66] predicting a tension jump at q−1 ≈ 125nm due to the spectrin cytoskeleton elasticity. The dashed line extrapolates the large wavelengths fit, without tension jump. (Figures adapted or reproduced from [66]) This model supposes fundamentally that the spectrin network is prestressed ((cid:96) (cid:54)= (cid:96)0) and the best fit of the experimental data from Zilker et al. [67] is obtained for a spectrin network under extension ∆σ = σ <−σ > ∼ 1.6× 10−6 N.m−1 and for a lipid bilayer under compression σ > = σ ∼ −0.8× 10−7 N.m−1. The authors sug- gest that the spectrin cytoskeleton may indeed lead to extra-folding of the bilayer, therefore regulating the membrane tension both directly at large scales compared to the mesh size, and indirectly at shorter scales, via the membrane area constraint. In a following paper, a more detailed calculation at higher orders predicts that the bending modulus is also renormalized by the presence of the spectrin cytoskeleton [68]. a)b) Fluctuations in active membranes 31 4.3.2 Models of active red blood cell membrane fluctuations Fig. 7 Active spectrin dissociation model for the red blood cell membrane. a) Schematic side view of the red blood cell composite membrane with fully connected spectrin filaments. b) Sketch of the dissociation of a filament, which would generate a direct normal force on the bilayer. (Figures reproduced from [56]) In successive papers [55, 56, 58], Gov & Safran proposed a first model for active fluctuations in the red blood cell membrane, where the activity is supposed to origi- nate from the spectrin cytoskeleton. Direct active forces ("kicks") on the membrane (see equation (30)) are suggested to be triggered by the detachment, under ATP hydrolysis, of spectrin filament ends from the lipid bilayer. In average, filament de- tachment is also predicted to soften the red blood cell membrane by decreasing the stiffness of the spectrin cytoskeleton, which is assumed to be naturally prestretched. Balancing the tension developed by a stretched filament with the energy needed to buckle the membrane, the authors predict a steady-state prestretch of the spectrin network of approximately 20% (see [56]). Assuming that detachment events are exponentially correlated in time, the non-thermal fluctuation spectrum originating from active direct forces are calculated in the same form as in equation (29). In a subsequent work, Auth, Safran & Gov calculated the entropic pressure exerted on the lipid bilayer by fluctuations of the spectrin filaments themselves [69] and they derived the fluctuation spectrum of coupled solid and fluid membranes maintained at fixed distance. They showed that this composite system may be described as a single polymerized membrane with renormalized bending rigidity [70]. The model of Gov & Safran assumes that active detachment of spectrin filaments may drive direct normal forces onto the membrane, but the precise microscopic mechanism of momentum transfer is not explicitly derived. In an alternative ap- proach [8], Turlier et al. recently proposed a new active model for the composite red blood cell membrane, where the mechanical coupling between the lipid bilayer and the spectrin cytoskeleton is precisely derived, and the specrin activity is described in a more generic manner. Since all metabolic events identified in the spectrin network, or in its anchoring proteins, have been associated with a decreased mechanical strength of the mem- brane it is supposed that any phosphorylation shall lead to a local decrease of the network shear modulus, the single parameter necessary to characterize the spectrin network mechanics. A simple two-state dynamics is assumed for simplicity, and is characterized by two transition rates kon and koff, defining the active timescale −1 and the fraction of active sites (cid:104)na(cid:105) = kon/ (kon + koff). The net- τa = (kon + koff) ƊR/R ~ 20%a)b) 32 Herv´e Turlier and Timo Betz Fig. 8 Active quasi-spherical model of the composite red blood cell membrane a) Schematic representation of the red blood cell membrane composed of a lipid bilayer and a regular trian- gular network of spring-like spectrin filaments. b) The composite membrane deformation can be separated in bending and stretching modes, illustrated for the spherical harmonic l = 3. c) The an- alytic model (solid lines) can reproduce the experimental fluctuations and response data (crosses). (Figures adapted from [8]) work shear modulus is then supposed to fluctuate around a mean value that decreases with metabolic activity (cid:104)µ(cid:105) = µ0 (1−(cid:104)na(cid:105)). In line with previous studies, the spec- trin cytoskeleton is idealized as a perfect triangular network of linear, prestressed springs (see Figure 8a). As the prestress is supposed to be isotropic, the discrete elastic energy of the network can be homogenized into a continuous isotropic elas- tic membrane. Its energy is expressed as a function of the incremental deformation from the prestressed state and takes an Hookean form with an additional prestress term. The mechanics of the prestressed network is characterized by an effective spec- trin tension S and an incremental shear modulus M, which are functions of the orig- inal shear modulus µ0, of the activity (cid:104)na(cid:105) and of a prestretch ratio. The membrane is supposed quasi-spherical and due to curvature, stretching and bending modes of deformations of the network are linearly coupled, as illustrated on Figure 8b for the mode l = 3. Since the lipid bilayer is tangentially fluid, the proteins anchor- ing the network to the bilayer may have non-zero sliding velocity relative to lipids, which, in turn, exert a drag force on the network. This drag force is found to be the dominant dissipative contribution slowing down membrane stretching at large wave- lengths. On the contrary, normal deformations are balanced by viscous forces from surrounding fluids. In this context, one has to calculate explicitly the lateral lipid pressure - or instantaneous surface tension -, which acts as a Lagrange multiplier for local bilayer incompressibility (to distinguish from the global surface tension, which acts as a Lagrange multiplier for the total bilayer area). It turns out, from the calculation, that the lateral lipid pressure cancels systematically direct normal forces that may originate from active fluctuations in the spectrin mechanics. Since fluctu- ations in the network shear modulus lead to active forces in both tangential and normal directions, an indirect source of active noise is however conserved in the normal direction thanks to the coupling between bending and stretching modes via the curvature. Decomposing the deformations in spherical harmonics modes (l, m), Fluctuations in active membranes 33 the membrane shape fluctuations appear classically as a sum of thermal and active contributions, expressed here with the membrane response function and the spectrin metabolic activity, respectively: Clm( f ) = 2kBT 2π f χ(cid:48)(cid:48) lm( f ) + 2(cid:104)na(cid:105) (1−(cid:104)na(cid:105))τa 1 + (2π f τa) (cid:12)(cid:12)Nlm( f )2(cid:12)(cid:12)2 (36) where f = ω/2π is the frequency, and Nlm( f ) captures the complex mode- and frequency-dependent propagation of tangential active noise into membrane shape fluctuations. The model predicts that spectrin-based active fluctuations should vanish for quasi-planar geometries, and it anticipates higher fluctuations in more curved re- gions of the red blood cell membrane, in agreement with recent spatial interfero- metric measurements [5]. It also shows clearly that a prestress in the network is the necessary ingredient for the emergence of spectrin-based active fluctuations, in agreement with previous hypotheses [56, 58]. It finally predicts that the network prestress may be maintained internally by an excess area of bilayer membrane, as suggested earlier by Fournier et al. [66]. Inverting the relation between excess area and bilayer tension (see section 4.2.4), a negative bilayer tension of the order of 10−7 N.m−1 is found by fitting experimental data. This analytical model reproduces fairly both the response function and active membrane fluctuations measured in the red blood cell membrane [8], as shown on Figure 8c. 5 Perspectives 5.1 A physiological role for active membrane fluctuations? An interesting and still overlooked question is the potential physiological role of active membrane fluctuations. The occurrence of active fluctuations does not sub- stantiate by itself any functional role in biology, and active noise may be simply a unavoidable byproduct of normal active processes in the cell. Yet, as noise is ubiq- uitous at this scale, cells may also control and take advantage of active noise, to actively facilitate or regulate other essential processes. It remains difficult, both ex- perimentally and theoretically, to discriminate the potential role played by active fluctuations, from the main purpose of the active process considered. While still un- proven, a number of possible physiological roles of membrane flickering have been suggested. As fluctuations increase membrane movement and fluidity, it should help mixing lipids and proteins within the membrane, an important property for cellular homeostasis (Figure 9)[71]. Additionally, the fluctuations might help larger mem- brane bound proteins to be transported laterally as it can help to overcome possible steric obstacles such as intracellular cytoskeletal components that might prevent or reduce lateral transport [71]. A further possible role lies in the interaction with other, intracellular as well as extracellular membranes. Depending on the function, 34 Herv´e Turlier and Timo Betz the fluctuations may help surface bound receptors to interact with other membranes, as the fluctuations allow to explore a larger region (Figure 9). On the other hand, active fluctuations may help to suppress nonspecific interaction by creating an ef- fective repulsive force when an obstacle enters the region of the fluctuations [72, 73]. Finally, it was observed that active fluctuations modify the effective tension that is measured on membranes [7], while on the other hand, membrane tension is known to be an important mechanical parameter for a number of cellular functions ranging from cell motility to endo/exocytosis and mechanosensing. Fig. 9 Possible biological functions of enhanced fluctuations related to membrane adhesion, repulsion and mixing. The fluctuations do create an effective potential barrier that can prevent uncontrolled membrane fusion. On the contrary, the large amplitudes of fluctuation may allow receptors and trans-membrane proteins to explore larger volumes, and thus facilitate the controlled binding and adhesion between membranes. Additionally, the active forces might increase lipid mixing and help protein transport in the membrane again the cortical cytoskeleton (red lines). 5.2 Active fluctuations in membrane adhesion processes Membrane adhesion is the physical process of interaction and attachment of a mem- brane to a surface, substrate or another membrane (Figure 9). In cells, membrane adhesion processes play essential roles in cell migration, cell-cell interaction or bilayer-cytoskeleton coupling and are therefore tightly regulated by various mem- brane proteins (cadherins, integrins, ERM proteins) [65]. Membrane adhesion is, in general, controlled by a competition between attractive forces at short range (stick- ers), repulsive forces at intermediate distances (spacers) and elastic stresses coming from membrane deformation [74, 75]. Depending on their amplitude, membrane fluctuations may play antagonist roles towards adhesion: moderate fluctuations may assist the nucleation [76] and the expansion of adhesion domains, but at higher am- plitudes fluctuations will compete with attractive forces and promote detachment of the membrane. The control of active fluctuations becomes in this context essential, but the role of activity in membrane adhesion has remained widely overlooked so far. Most of the theoretical models for membrane adhesion have considered mem- branes at equilibrium, which is justified when the focus is on biomimetic systems [77, 76]. To generalize these concepts to biological membranes, it will become crit- ical to evaluate experimentally and theoretically the influence of active fluctuations on adhesion processes [78]. EntropicrepulsionAdhesion Fluctuations in active membranes 5.3 Fluctuations of membranes with an actomyosin cytoskeleton 35 Most animal cell types possess a cortex of actomyosin, thin layer of actin fila- ments with embedded molecular motors, which lies beneath the membrane. The actomyosin cortex is tightly connected to the lipid bilayer via different proteins, the principal family being formed by the ERM (Ezrin-Radixin-Moesin) [79]. Regu- lated by phosphorylation, the connection between ERM and the cortex is dynamic and non-equilibrium by nature, providing a first possible source of active forces in the membrane. The active regulation of membrane-cortex adhesion is particularly important for the formation of blebs, membrane bulges originating from local cor- tex detachment [80, 37, 81]. But the cortex itself is notoriously very dynamic, with several processes requiring metabolic energy consumption, such as actin polymer- ization and myosin motor activity. All these processes constitute potential sources of active membrane fluctuations, covering a large spectrum of length and timescales. To disentangle and characterize precisely the various sources of activity in compos- ite bilayer-cortex membranes, a substantial mutual effort from biologists, physicists and theorists will be required over the next years, and precise numerical modeling will become critical [8, 82]. Acknowledgements H. Turlier acknowledges support from the CNRS/Inserm program ATIP- Avenir, from the Bettencourt-Schueller Foundation and from the Coll`ege de France. T. Betz is supported by the Deutsche Forschungsgemeinschaft (DFG), Cells-in-Motion Cluster of Excellence (EXC 1003-CiM), University of Munster, Germany. References 1. Campelo, F., Arnarez, C., Marrink, S.J., Kozlov, M.M.: Helfrich model of membrane bending: from Gibbs theory of liquid interfaces to membranes as thick anisotropic elastic layers. Adv. Colloid Interface Sci. 208, 25 -- 33 (2014) 2. Hochmuth, R.M., Evans, C.A., Wiles, H.C., McCown, J.T.: Mechanical measurement of red cell membrane thickness. Science, 220, 101 -- 102 (1983) 3. Browicz, T.: Further observation of motion phenomena on red blood cells in pathological states. Zbl. med. Wissen, 28, 625 -- 627 (1890) 4. Tuvia, S., Almagor, A., Bitler, A., Levin, S., Korenstein, R., Yedgar, S.: Cell membrane fluc- tuations are regulated by medium macroviscosity: evidence for a metabolic driving force. Proc. Natl. Acad. Sci. U.S.A., 94, 5045 -- 5049 (1997) 5. Park, Y., et al.: Metabolic remodeling of the human red blood cell membrane. Proc. Natl. Acad. Sci. U.S.A., 107, 1289 -- 1294 (2010) 6. Park, Y, et al.: Diffraction phase and fluorescence microscopy. Optics Express 14, 8263-8 (2006) 7. Betz, T., Lenz, M., Joanny, J.F., Sykes, C.: ATP-dependent mechanics of red blood cells. Proc. Natl. Acad. Sci. U.S.A., 106, 15320 -- 15325 (2009) 8. Turlier, H. et al.: Equilibrium physics breakdown reveals the active nature of red blood cell flickering. Nat. Phys., 12, 513 -- 519 (2016) 9. Cabot, R. C.: A Guide to the Clinical Examination of the Blood, p. 52. 4th ed. Longmans, Green & Co. London (1901) 36 Herv´e Turlier and Timo Betz 10. Pulvertaft, R.J.V.: Vibratory movement in the cytoplasm of erythrocytes. J. clin. Path. 2,281- 11. Blowers, R., Clarkson, E.M., Maizels, M.: Flicker phenomenon in human erythrocytes. J. 283 (1949) Physiol., 113, 228 (1951) 12. Parpart, A.K., Hoffman, J.H.: Flicker in erythrocytes.vibratory movements in the cytoplasm?. 13. Brochard, F., Lennon, J.F.: Frequency spectrum of the flicker phenomenon in erythrocytes. J. J. Cell. Comp. Physiol., 47, 295 -- 303 (1956) Phys. (Paris), 36,1035 -- 1047 (1975) 14. Evans, J., Gratzer, W., Mohandas, N., Parker, K., Sleep, J.: Fluctuations of the red blood cell membrane: relation to mechanical properties and lack of ATP dependence. Biophys. J., 94, 4134 -- 4144 (2008) 15. Rodr´ıguez-Garc´ıa, R., L´opez-Montero, I., Mell, M., Egea, G., Gov, N.S., Monroy, F.: Direct cytoskeleton forces cause membrane softening in red blood cells. Biophys. J., 108, 2794 -- 2806 (2015) 16. Prost, J., Bruinsma, R.: Shape fluctuations of active membranes. Europhys. Lett. 33, 321 -- 326 (1996) 17. Ramaswamy, S., Toner, J., Prost, J.: Nonequilibrium fluctuations, traveling waves, and insta- bilities in active membranes. Phys. Rev. Lett., 84, 3494 -- 3497 (2000) 18. Lin, L.C.L., Gov, N., Brown, F.L. . Nonequilibrium membrane fluctuations driven by active proteins. J. Chem. Phys., 124, 074903 (2006) 19. Manneville, J.B., Bassereau, P., L´evy, D., Prost, J. Activity of Transmembrane Proteins In- duces Magnification of Shape Fluctuations of Lipid Membranes. Phys. Rev. Lett. 82, 4356 -- 4359 (1999) 20. Manneville, J.B., Bassereau, P., Ramaswamy, S., and Prost, J. Active membrane fluctuations studied by micropipet aspiration. Phys. Rev. E 64, 021908 (2001) 21. Faris, M.E.A., Lacoste, D., P´ecr´eaux, J., Joanny, J.F., Prost, J., Bassereau, P.: Membrane ten- sion lowering induced by protein activity. Phys. Rev. Lett., 102, 038102 (2009) 22. Girard, P., Prost, J., Bassereau, P.: Passive or active fluctuations in membranes containing proteins. Phys. Rev. Lett., 94, 088102 (2005) 23. Hankins, H.M., Baldridge, R.D., Xu, P., Graham, T.R.: Role of flippases, scramblases and transfer proteins in phosphatidylserine subcellular distribution. Traffic, 16, 35 -- 47 (2015) 24. Rao M., Sarasij R. C. Active Fusion and Fission Processes on a Fluid Membrane Phys. Rev. Lett. 87, 128101 (2001) 25. Humphrey, D., Duggan, C., Saha, D., Smith, D., Ks, J.:Active fluidization of polymer net- works through molecular motors. Nature 416(6879), 413-6 (2002) 26. Koenderink, G.H., Dogic, Z., Nakamura, F., Bendix, P.M., MacKintosh, F.C., Hartwig, J.H., Stossel, T.P., Weitz, D.A.: An active biopolymer network controlled by molecular motors. Proc Natl Acad Sci USA., 106, 15192-15197 (2009) 27. Wickstrand, C., Dods, R., Royant, A., Neutze, R.: Bacteriorhodopsin: Would the real struc- tural intermediates please stand up?. Biochim. Biophys. Acta, Gen. Subj., 1850, 536 -- 553 (2015) 28. Zilker, A., Ziegler, M., Sackmann, E.: Spectral analysis of erythrocyte flickering in the 0.34µm−1 regime by microinterferometry combined with fast image processing. Phys. Rev. A, 46, 7998 (1992) 29. Strey, H., Peterson, M., Sackmann, E.: Measurement of erythrocyte membrane elasticity by flicker eigenmode decomposition. Biophys. J., 69, 478 (1995) 30. P´ecr´eaux, J., Dobereiner, H.G., Prost, J., Joanny, J.F., Bassereau, P.: Refined contour analysis of giant unilamellar vesicles. Europ. Phys. J. E, 13, 277 -- 290 (2004) 31. Brown, A.T., Kotar, J., Cicuta, P.: Active rheology of phospholipid vesicles. Phys. Rev. E, 84, 021930 (2011) 32. Radler, J., Sackmann, E.: Imaging optical thicknesses and separation distances of phospho- lipid vesicles at solid surfaces. J. Phys. II, 3, 727 -- 748 (1993) 33. Schmidt, D., Monzel, C., Bihr, T., Merkel, R., Seifert, U., Sengupta, K., Smith A.S.:Signature of a Nonharmonic Potential as Revealed from a Consistent Shape and Fluctuation Analysis of an Adherent Membrane. Phys. Rev. X 4, 021023, (2014) Fluctuations in active membranes 37 34. Monzel, C., Sengupta, K.: Measuring shape fluctuations in biological membranes. Journal of Physics D: Applied Physics, 49, 24 (2016) 35. Monzel, C., et al.: Measuring fast stochastic displacements of bio-membranes with dynamic optical displacement spectroscopy. Nat. Commun., 6 (2015) 36. Betz, T., Sykes, C.: Time resolved membrane fluctuation spectroscopy. Soft Matter, 8, 5317 -- 5326 (2012) 37. Peukes, J., Betz, T.: Direct measurement of the cortical tension during the growth of mem- brane blebs. Biophys. J., 107, 1810 -- 1820 (2014) 38. Henon, S., Lenormand, G., Richert, A., Gallet, F.: A new determination of the shear modu- lus of the human erythrocyte membrane using optical tweezers. Biophys. J., 76, 1145 -- 1151 (1999) 39. Mills, J.P., Qie, L., Dao, M., Lim, C.T., Suresh, S.: Nonlinear elastic and viscoelastic defor- mation of the human red blood cell with optical tweezers. Mol. Cell. Biomech. Tech Science Press, 1, 169 -- 180 (2004) 40. Yoon, Y.Z., Kotar, J., Brown, A.T., Cicuta, P.: Red blood cell dynamics: from spontaneous fluctuations to non-linear response. Soft Matter, 7, 2042 -- 2051 (2011) 41. Helfrich, W.: Elastic properties of lipid bilayers: theory and possible experiments. Z. Natur- forsch. C, 28, 693 -- 703 (1973). 42. Helfrich, W.S., Servuss, R.M.: Undulations, steric interaction and cohesion of fluid mem- branes. Il Nuovo Cimento D, 3, 137 -- 151 (1984) 43. Fournier, J.B., Ajdari, A., Peliti, L.. Effective-area elasticity and tension of micromanipulated membranes. Phys. Rev. Lett., 86, 4970 (2001) 44. Doi, M., Edwards, S.F.: The Theory of Polymer Dynamics. Oxford Science Publications, pp. 45. Schlosser, F., Rehfeldt, F., Schmidt, C.-F.: Force fluctuations in three-dimensional suspended fibroblasts. Phil. Trans. R. Soc. B, 370, 20140028 (2015) 46. Almonacid, M. et al.: Active diffusion positions the nucleus in mouse oocytes. Nat. Cell Biol. 88-89 (1988) 17, 470 -- 479 (2015) 47. Mizuno, D., Tardin, C., Schmidt, C. F., MacKintosh, F.C.: Nonequilibrium mechanics of ac- tive cytoskeletal networks. Science, 315, 370 -- 373 (2007) 48. Lacoste, D., Bassereau, P.: An Update on Active Membranes. In Liposomes, Lipid Bilayers and Model Membranes, CRC Press pp. 1 -- 18 (2014) 49. Milner, S.T., Safran, S.A.: Dynamical fluctuations of droplet microemulsions and vesicles. 50. Lomholt, M.A.: Fluctuation spectrum of quasispherical membranes with force-dipole activity. 51. Loubet, B., Seifert, U., Lomholt, M.A.: Effective tension and fluctuations in active mem- 52. Gadsby, D.C.: Ion channels versus ion pumps: the principal difference, in principle. Nat. Rev. Phys. Rev. A, 36, 4371 (1987) Phys. Rev. E, 73, 061914 (2006) branes. Phys. Rev. E, 85, 031913 (2012) Mol. Cell Biol., 10, 344 -- 352 (2009) Rev. Lett., 92, 168101 (2004) 53. Chen, H.Y.: Internal states of active inclusions and the dynamics of an active membrane. Phys. 54. Chen, H.-Y., Mikhailov, A.S.: Dynamics of biomembranes with active multiple-state inclu- sions. Phys. Rev. E, 81, 031901 -- 031911 (2010) 55. Gov, N.S.: Membrane Undulations Driven by Force Fluctuations of Active Proteins. Phys. Rev. Lett., 93, 268104 -- 268104 (2004) 56. Gov, N.S., Safran. S.A.: Red Blood Cell Membrane Fluctuations and Shape Controlled by ATP-Induced Cytoskeletal Defects. Biophys. J. 88, 1859 -- 1874 (2005) 57. Gov, N.S., Gopinathan, A.: Dynamics of membranes driven by actin polymerization. Biophys. 58. Gov, N.S.: Active elastic network: Cytoskeleton of the red blood cell. Phys. Rev. E, 75, 59. Lacoste, D., and Lau, A.W.C.: Dynamics of active membranes with internal noise. Europhys. J., 90, 454 -- 469 (2006) 011921 (2007) Lett., 70, 418 -- 424 (2005) 38 Herv´e Turlier and Timo Betz 60. Sankararaman, S., Menon, G.I., Sunil Kumar, P.B.: Two-component fluid membranes near repulsive walls: Linearized hydrodynamics of equilibrium and nonequilibrium states. Phys. Rev. E, 66, 031914 -- 031916 (2002) 61. C.W. Gardiner: Handbook of Stochastic Methods for physics, chemistry and the natural sci- 62. Seifert, U.: The concept of effective tension for fluctuating vesicles. Z. Phys. B, 97, 299 -- 309 ences. Springer, p. 77 (1985) (1995) 63. Gov, N., Zilman, A., Safran, S.: Cytoskeleton Confinement and Tension of Red Blood Cell Membranes. Phys. Rev. Lett., 90, 228101 (2003) 64. Radler, J.O., Feder, T.J., Strey, H.H., and Sackmann, E. Fluctuation analysis of tension- controlled undulation forces between giant vesicles and solid substrates. Phys. Rev. E, 51, 4526 (1995) 65. Sackmann, E., Smith, A.S.: Physics of cell adhesion: some lessons from cell-mimetic sys- tems. Soft matter, 10, 1644 -- 1659 (2014) 66. Fournier, J.-B., Lacoste, D., and Raphael, E.: Fluctuation Spectrum of Fluid Membranes Cou- pled to an Elastic Meshwork: Jump of the Effective Surface Tension at the Mesh Size. Phys. Rev. Lett., 92, 018102 -- 018104 (2004). 67. Zilker, A., Engelhardt, H., Sackmann, E.: Dynamic reflection interference contrast (RIC) mi- croscopy: a new method to study surface excitations of cells and to measure membrane bend- ing elastic moduli. J. Phys. (Paris), 48, 2139 -- 2151 (1987) 68. Dubus, C., Fournier, J.B.: A Gaussian model for the membrane of red blood cells with cy- toskeletal defects. Europhys. Lett., 75, 181 -- 187 (2007) 69. Auth, T., Safran, S.A., Gov, N.S.: Filament networks attached to membranes: cytoskeletal pressure and local bilayer deformation. New J. Phys. 9, 430 -- 430 (2007) 70. Auth, T., Safran, S.A., Gov, N.S.: Fluctuations of coupled fluid and solid membranes with application to red blood cells. Phys. Rev. E 76, 051910 -- 051918 (2007) 71. Lin, L., Brown, F.: Dynamics of pinned membranes with application to protein diffusion on the surface of red blood cells. Biophys. J., 86, 764 -- 780 (2004) 72. Evans, E.A., Parsegian, V.A.: Thermal-mechanical fluctuations enhance repulsion between bimolecular layers. Proc. Natl. Acad. Sci. U.S.A., 83, 7132 -- 7136 (1986) 73. Prost, J., Manneville, J. B., and Bruinsma, R. Fluctuation-magnification of non-equilibrium membranes near a wall. Europ. Phys. J. B, 1, 465 -- 480 (1998) 74. G.I. Bell: Physical basis of cell-cell adhesion. CRC Press, Boca Raton, p. 227 (1988) 75. E. Evans: Detailed mechanics of membrane-membrane adhesion and separation. I. Contin- uum of molecular cross-bridges, Biophys. J. 48, 175 -- 183 (1985) 76. Bihr, T., Seifert, U., Smith, A.S.: Nucleation of ligand-receptor domains in membrane adhe- sion. Phys. Rev. Lett., 109, 258101 (2012) 77. Bruinsma, R., Behrisch, A., Sackmann, E.: Adhesive switching of membranes: Experiment and theory. Phys. Rev. E 61, 4253 -- 4267 (2000). 78. Weikl, T.R., Asfaw, M., Krobath, H., R´ozycki, B., Lipowsky, R.: Adhesion of membranes via receptorligand complexes: Domain formation, binding cooperativity, and active processes. Soft Matter 5, 3213 -- 3224 (2009) 79. Fehon, R.G., McClatchey, A.I., and Bretscher, A.: Organizing the cell cortex: the role of ERM proteins. Nat. Rev. Mol. Cell Biol. 11, 276 -- 287 (2010) 80. Charras, G.T.: A short history of blebbing. J. Microsc. 231, 466 -- 478 (2008) 81. Alert, R., Casademunt, J.: Bleb Nucleation through Membrane Peeling. Phys. Rev. Lett., 116, 068101 (2016). 82. Fedosov, D.A., Caswell, B., Karniadakis, G.E.: A multiscale red blood cell model with accu- rate mechanics, rheology, and dynamics. Biophys. J. 98, 2215 -- 2225 (2010)
1609.08871
1
1609
2016-09-28T12:01:27
Plasmonic Photothermal Therapy in Third and Fourth Biological Windows
[ "physics.bio-ph", "physics.comp-ph", "physics.med-ph", "physics.optics" ]
The recently reported 3rd and 4th biological transparency windows located respectively at 1.6-1.9um and 2.1-2.3um promise deeper tissue penetration and reduced collateral photodamage, yet they haven't been utilized in photothermal therapy applications. Nanoparticle based plasmonic photothermal therapy poses a nontrivial optimization problem in which the light absorption efficiency of the nanoparticle has to be maximized subject to various constraints that are imposed by application environment. Upscaling the typical absorber-dominant nanoparticle designs (rod, sphere etc.) that operate in the 1st and 2nd transparency windows is not a viable option as their size gets prohibitively large for cell intrusion and they become scatterer-dominant. The present study addresses this issue and suggests a versatile approach for designing both lithography based and self-assembling absorber dominant nanostructures for the new transparency windows, while keeping their size relatively small. The proposed nanoparticles demonstrate up to 40% size reduction and 2-fold increase in absorption efficiency compared to the conventional nanobar design. The overall photothermal performance per nanoparticle in the 4th window is boosted up by 250% compared to the 2nd window.
physics.bio-ph
physics
Plasmonic Photothermal Therapy in Third and Fourth Biological Windows E. Doruk Onal∗, Kaan Guven Koc University September 29, 2016 Abstract The recently reported 3rd and 4th biological transparency windows located respectively at 1.6 − 1.9µm and 2.1−2.3µm promise deeper tissue penetration and reduced collateral photodamage, yet they haven't been utilized in photothermal therapy applications. Nanoparticle based plasmonic photothermal therapy poses a nontrivial optimization problem in which the light absorption efficiency of the nanoparticle has to be maximized subject to various constraints that are imposed by application environment. Upscaling the typical absorber-dominant nanoparticle designs (rod, sphere etc.) that operate in the 1st and 2nd transparency windows is not a vi- able option as their size gets prohibitively large for cell intrusion and they become scatterer-dominant. The present study addresses this issue and suggests a versatile approach for designing both lithography based and self-assembling absorber dominant nanostructures for the new transparency windows, while keeping their size relatively small. The proposed nanoparticles demonstrate up to 40% size reduction and 2-fold increase in ab- sorption efficiency compared to the conventional nanobar design. The overall photothermal performance per nanoparticle in the 4th window is boosted up by 250% compared to the 2nd window. ∗Corresponding author: [email protected] 1 Wavelength (nm)16001800200022002400NIR IV500400300200100Size(nm)00.250.500.751.00Heat Generated Per Particle (A.U.)Wavelength (nm)1600180020002200240000.250.500.751.00NIR IIIHeat Generated Per Particle (A.U.)500400300200100Size(nm)? I. Introduction Hyperthermia therapy is based on increasing the tempera- ture of a malignant tissue above its standard value(37◦C) to hinder cellular processes. In this context, incorporat- ing metallic nanoparticles (NP) that convert electromag- netic radiation into heat via plasmonic resonances has been widely investigated in the last decade and became known as plasmonic photothermal therapy (PPT). The optical response of a NP is characterized by its scat- tering (σScat) and absorption cross sections (σAbs). The heating power of a metallic NP under continuous wave il- lumination is related to the incident light intensity and the absorption cross section of the NP: P = I × σAbs. The efficiency of PPT is determined by these two parameters. The amount of light intensity reaching the NP is limited by the attenuation of human tissue. The absorption cross section depends on the size, shape and material of the NP. Among various materials, gold is the dominant choice for NPs especially in biological applications due to its chemical inertness, biocompatibility and ability to support localized surface plasmon resonance(LSPR). Several studies indicate that a rodlike design is the most efficient geometry for PPT applications [1, 2, 3]. The light penetration problem into the human body is overcome either by using fiber optics to transmit light through the body into tumors near intrabody cavities or by utilizing light sources at certain wavelengths where human body is most transparent. These are called the biological transparency windows and located in near-infrared (NIR) region of the spectrum. So far, PPT is experimentally demonstrated in NIR-I (700-950 nm)[4] & NIR-II (1000- 1350 nm)[5] which were discovered in 2001 and 2010 re- spectively. Recently, advancing photodetectors and optical instruments led to the discovery of new transparency win- dows at longer wavelengths: NIR-III (1600-1870 nm) and NIR-IV (2100-2300 nm) in 2014 and 2016 [6, 7]. Although radiation in NIR-I & II are successfully used in PPT ap- plications, utilizing NIR-III & IV assures better light pen- etration into deep tissue with less attenuation. To the best of our knowledge, there is no published study exploring the NPs that can operate in the NIR-III & NIR-IV for PPT applications. The objective of this article is to fill this gap by in- vestigating NP designs that can efficiently operate in these bands. As revealed in this study, upscaling the existing NP designs of NIR-I or NIR-II to the NIR-III and NIR-IV is not a feasible solution due to inefficiencies in cell intrusion and photothermal conversion. II. Simulation & Modeling In this work we studied several NP designs from solid and contour-shaped gold nanobars to self-assembling gold nanodisk- and nanoring chains. The electromagnetic simu- lations are performed by a commercially available Lumer- ical software under linearly polarized light along the long 2 axis of the NP. The frequency dependent complex dielectric function of gold is approximated by the Brendal-Bormann model which is shown to be in very good agreement with experimental observations in the studied wavelength range [8, 9, 10]. The refractive index of the environment is set to 1.40 which corresponds to that of living cells [11]. III. Gold Nanobars Therapy in NIR-III and NIR-IV for Photothermal The LSPR of a NP can be easily adjusted spectrally across the biological transparency windows by modifying its size or geometry. There is almost a linear relation between the NP length and its LSPR wavelength. However, simply scaling up the NP designs reported for NIR-I & II is not enough to adopt them for NIR-III & IV because of two fundamental drawbacks. Figure 1: The schematic of (a-c) solid, 50% contour and 70% contour nanobars and (d-f) their respective absorption efficiency (φAbs) as a function of the nanobar dimensions. Black lines indicate the regions where the nanobar is reso- nant in NIR-III or NIR-IV. The first drawback is that scaling alters the NP's dom- inant response character at its LSPR: Solid gold nano- bars (or nanospheres) that are good absorbers in NIR-I become scatterer when scaled to NIR-III or IV. We show this in Fig.1.d where the absorption efficiency coefficient (φAbs = σAbs/(σAbs + σScat)) of a solid gold nanobar NIR-IIINIR-IVNIR-IIINIR-IVNIR-IIINIR-IVEkLxLy0.5Lx0.5LyLxLyEkLxLy0.7Lx0.7LyEkabcdef0.20.30.40.50.60.40.50.60.70.80.50.60.70.80.90.20.30.40.50.60.40.50.60.70.80.50.60.70.80.9 Figure 2: Benchmarking the solid nanobar (triangle) and contour nanobars (star, circle) in terms of absorption efficiency (φAbs), nanobar length, and absorption cross section in NIR-III and NIR-IV. Contour nanobars provide better φAbs (a), smaller size (b), and crowd the top border of the maximum absorption cross section trend (c) across the spectra, making them suitable candidates for PPT applications. (Fig.1.a) is plotted as a function of its length and width. The majority of solid nanobars for NIR-III&IV are scat- terers (φAbs < 0.5). In a previous work, we proposed a contour nanobar design that enhances φAbs significantly, note how the colormapped φAbs shifts towards darker (i.e. higher) values in the entire plot region in Fig.1e&f with increasing contour size [12]. Even though the current research on the cell intrusion mechanism for NPs is not conclusive [13], decreasing the dimensions of NP would likely ease this process in addition to increasing the spatial resolution for thermal spot gen- eration. A comparison among Figure 1.d-f also highlights that the NIR-III&IV active regions shift towards smaller nanobar lengths with increasing contour size. Thus, the contour nanobar design aids in eliminating both of these two drawbacks simultaneously. Increasing the absorption efficiency (φAbs) is only one aspect as the absolute value of the absorption cross section (σAbs) must be taken into account in maximizing the heat generation. A recent experimental study of PPT in NIR-I employs small gold nanorods (L = 16 − 45nm) that have almost 100% absorption efficiency and very high cellular uptake [14]. However, due to the very small σAbs, in order to generate enough heat for cell ablation, high laser inten- sities around 12W/cm2 were required which is well above the healthy limit (1− 2W/cm2). The intensity could be re- duced by increasing the NP concentration but this causes further detrimental effect due to the increased cytotoxicity. Evidently, designing NP for PPT involves many trade- offs and requires a multidimensional optimization of ab- sorption efficiency, NP size and absorption cross section. Figure 2.a shows the contour nanobars on average achieve 100 − 200% improvement in absorption efficiency (φAbs ∼ 0.4 − 0.6) depending on the contour percentage. Second, the contour nanobars provide 15-40% reduction in size com- pared to solid nanobars as seen in Figure 2.b: 275-200nm vs. 350 nm in NIR III and 400-325nm vs 475nm in NIR IV, respectively. Regarding the absorption cross section, Fig- ure 2.c shows that the contour nanobars crowd along the maximum absorption cross section trendline. This implies more freedom in design parameters while keeping the ab- sorption cross section close to its maximum. Figure 2.c also highlights the advantages of working in longer wavelength transparency windows (i.e. NIR-III & IV) as the maximum absorption cross section linearly increases with resonance wavelength. Therefore, heat generation per particle is sig- nificantly higher in NIR-III (150%) and NIR-IV (250%) in comparison to NIR-II. Overall, longer wavelength trans- parency windows (NIR-III & IV) enhance the heat gener- ation for PPT in two respects. The same amount of heat can be delivered to tumor cells at lower NP concentrations and also the energy delivery to the NPs is achieved with lower energy photons which reduces the lateral heating in healthy tissue. As stated before, the relation between NP size and cel- lular uptake requires further research. However, cellular uptake for NPs that are smaller than 100 nm is well docu- mented for PPT applications in NIR-I & II [15, 14, 3, 1, 2]. In NIR-III & IV even the smallest contour nanobars exceed 100 nm (Fig. 1). Fortunately, the size limitation of mono- lithic nanobars can be bypassed by a bottom-up approach in which disk/ring shaped NPs with diameter smaller than 100nm can be first transferred individually through the cell membrane and then assembled into a chain to construct the 'nanobar' inside the cell. We investigate this approach in the next section. IV. Self-Assembling Nanoantenna Alter- natives in NIR-III and NIR-IV: Nano Disk and Ring Chains The self assembly of nanodisks via DNA or protein as- sistance is well documented [16, 17]. A nanodisk chain (NDC) is smaller than its monolithic nanobar counterpart but still resonates at the same wavelength [18]. Replacing nanodisks by nanorings would be a simple implementation of the contour design to the self-assembling chains. There are both top-down and bottom-up methods for fabricating these nanodisk and nanoring NPs [19, 20]. 3 1200140016001800200022002400100150200250300350400450500550NIR IVNIR III1200140016001800200022002400x10-1300.511.522.533.544.5NIR IVNIR III12001400160018002000220024000.10.20.30.40.50.60.70.80.91.0NIR IVNIR IIIAbsorption Efficiency (φAbs)Resonance Wavelength (nm)Resonance Wavelength (nm)Resonance Wavelength (nm)Nanobar Length (nm)Absorption Cross-section (m2)70% Cont NB50% Cont NBNanobar(NB)bca70% Cont NB50% Cont NBNanobar(NB)70% Cont NB50% Cont NBNanobar(NB) Figure 3: A comparison of the optical cross sections (absorption, scattering and extinction) of solid nanobar, contour nanobar, NDC, NRC, designed for NIR-IV and NIR-III respectively. The contour and ring design enhances the absorption efficiency significantly. For a comparison of the absorption properties of self-assembling nanodisk/nanoring-chains and monolithic solid/contour nanobars, we picked a sample from each with the same resonance wavelength centered in the NIR-IV. The scattering and absorption cross section spectra of these samples plotted respectively in Fig. 3.a-d indicate that the NRC is the smallest in size and has the highest absorption efficiency. If we were to set the size considerations aside, the best performing NP candidate among these would be the contour nanobar with highest absorption cross section among all (7% larger than the solid nanobar) and also pro- viding a moderate reduction in size (18% smaller than the solid nanobar). The rest of the optimization for the best performing NP is about managing the trade-offs. If the future experimen- tal evidence suggests that NPs smaller than 100 nm are the only way forward than self-assembling NP chains become the only viable option. The self-assembling structures such as NDCs come at a cost of 9% reduction in absorption cross section and also equivalently in heat generated per NP. On the positive side, the absorption efficiency is increased to 63% (Fig 3.c). Replacing NDC by NRC would be only feasible when further size reduction is mandatory. This would reduce σAbs significantly (51% compared to NDC) which effec- tively halves the heating power per NP. However, in pho- tothermal imaging applications, where the maximum ab- sorption efficiency of the NP might be of primary concern, NRCs perform better than NDCs in suppressing the scat- tering induced noise and interference. Figure 3 e-f shows that choosing NIR-III as the opera- tional window instead of NIR-IV provide further reduction in the proposed NP sizes, but also leads to smaller σAbs. It is reported that NIR-III can provide better transparency in certain tissue types compared to NIR-IV [7]. Whether the increased transparency of NIR-III can compensate for the reduction of σAbs requires further experimental data. It is also worth to point out that cytotoxic effects does not appear to be in linear relation with the NP size [13]. V. Conclusion The contour based monolithic nanobars and self-assembling NDC/NRCs presented in this study provide both enhanced absorption efficiency and size reduction, which are the pri- mary concerns in the optimization of NPs for PPT applica- tions. Combined with the increased transmission and the magnitude of absorption in NIR-III&IV, their utilization for PPT can be a viable choice. We should however point out that the relation between NP size and toxicity doc- umented thus far in the literature introduces a nonlinear constraint to this optimization problem: A particular study reports that while small (3-5 nm) and large (50-100) NPs are not toxic, the same dose of intermediate size (18-37nm) had lethal effects on mice, linked to major organ damage [13]. 4 Wavelength (nm)16001800200022002400Cross-sectionx10-1300.511.522.5NIR III192 nmΦAbs = 0.88ExtScatAbsWavelength (nm)16001800200022002400Cross-section00.511.522.5NIR III240 nmΦAbs = 0.73x10-13ExtScatAbsWavelength (nm)16001800200022002400NIR IV352 nmΦAbs = 0.63ExtScatAbsWavelength (nm)16001800200022002400Cross-section012345678910NIR IV256 nmΦAbs = 0.87x10-13ExtScatAbsWavelength (nm)16001800200022002400NIR IV360 nmΦAbs = 0.56ExtScatAbsWavelength (nm)16001800200022002400NIR IV440 nmΦAbs = 0.36ExtScatAbsNIR IVNIR IIICross-section012345678910x10-13Cross-section012345678910x10-13Cross-section012345678910x10-13abcdef References [1] L. M. Maestro, E. Camarillo, J. A. Sánchez- Gil, R. Rodríguez-Oliveros, J. Ramiro- Bargueño, a. J. Caamaño, F. Jaque, J. G. Solé, and D. Jaque, RSC Adv. 4(96), 54122–54129 (2014). [2] L. M. Maestro, P. Haro-Gonzaíez, A. Sachez- Iglesias, L. M. Liz-Marza, J. Sole, D. Jaque, P. Haro-González, A. Sánchez-Iglesias, L. M. Liz-Marzán, J. García Solé, and D. Jaque, Langmuir 30(6), 1650–1658 (2014). [3] M. A. Mackey, M. R. K. Ali, L. A. Austin, R. D. Near, and M. A. El-Sayed, The Journal of Phys- ical Chemistry B 118(5), 1319–1326 (2014). [4] R. Weissleder, Nature biotechnology 19(4), 316– 317 (2001). [5] A. M. Smith, M. C. Mancini, and S. Nie, Nature Nanotechnology 4(11), 710–711 (2010). [6] K. Salas-Ramirez, L. Shi, L. Zhang, and R. R. Alfano, Proceedings of the SPIE 8940(212), 89400V (2014). [7] L. Shi, L. A. Sordillo, A. Rodríguez- Contreras, and R. Alfano, Journal of Bio- photonics 9(1-2), 38–43 (2016). [8] A. D. Rakic, A. B. Djurišic, J. M. Elazar, and M. L. Majewski, Applied Optics 37(22), 5271 (1998). [9] P. Jahanshahi, M. Ghomeishi, and F. R. M. Adikan, The Scientific World Journal 2014 (2014). [10] R. Brendel and D. Bormann, Journal of Applied Physics 71(1), 1–6 (1992). [11] M. A. Calin, M. R. Calin, and C. Munteanu, The European Physical Journal Plus 129(6), 116 (2014). [12] E. D. Onal and K. Guven, Arxiv (2015). [13] A. M. Alkilany and C. J. Murphy, Journal of Nanoparticle Research 12(7), 2313–2333 (2010). [14] H. Jia, C. Fang, X. M. Zhu, Q. Ruan, Y. X. J. Wang, and J. Wang, Langmuir 31(26), 7418–7426 (2015). [15] M. F. Tsai, S. H. G. Chang, F. Y. Cheng, V. Shanmugam, Y. S. Cheng, C. H. Su, and C. S. Yeh, ACS Nano 7(6), 5330–5342 (2013). [16] K. L. Gurunatha, A. C. Fournier, A. Urvoas, M. Valerio-Lepiniec, V. Marchi, P. Minard, and E. Dujardin, ACS Nano 10(3), 3176–3185 (2016). [17] H. J. Yin, L. Liu, C. A. Shi, X. Zhang, M. Y. Lv, Y. M. Zhao, and H. J. Xu, Applied Physics Letters 107(19), 10–15 (2015). [18] Z. Li, S. Butun, and K. Aydin, ACS Photonics 1(3), 228–234 (2014). [19] T. Ozel, M. J. Ashley, G. R. Bourret, M. B. Ross, G. C. Schatz, and C. A. Mirkin, Nano Let- ters 15(8), 5273–5278 (2015). [20] H. J. Jang, S. Ham, J. A. I. Acapulco, Y. Song, S. Hong, K. L. Shuford, and S. Park, Journal of the American Chemical Society 136(50), 17674– 17680 (2014). 5
1708.00920
2
1708
2017-09-22T11:34:06
Role of Initial Coherence in Excitation Energy Transfer in Fenna-Matthews-Olson Complex
[ "physics.bio-ph" ]
We theoretically show that the initial coherence plays a crucial role in enhancing the speed of excitation energy transfer (EET) in Fenna-Matthews-Olson (FMO) complex. We choose a simplistic eight-level model considering all the bacateriochlorophyll-a sites in a monomer of FMO complex. We make a comparative numerical study of the EET, in terms of non-Markovian evolution of an initial coherent superposition state and a mixed state. A femto-second coherent laser pulse is suitably chosen to create the initial coherent superposition state. Such an initial state relaxes much faster than a mixed state thereby speeding up the EET. In this analysis, we have taken into account the relative orientation of the transition dipole moments of the bacateriochlorophyll-a sites and their relative excitation energies. Our results reveal that for 2D electronic spectroscopy experiments, the existing two-pathway model of energy transfer in FMO complex may not be suitable in our understanding of EET.
physics.bio-ph
physics
Role of Initial Coherence in Excitation Energy Transfer in Fenna-Matthews-Olson Complex Davinder Singh∗ and Shubhrangshu Dasgupta Department of Physics, Indian Institute of Technology Ropar, Rupnagar, Punjab - 140001, India (Dated: September 23, 2018) We theoretically show that the initial coherence plays a crucial role in enhancing the speed of excitation energy transfer (EET) in Fenna-Matthews-Olson (FMO) complex. We choose a simplistic eight-level model considering all the bacateriochlorophyll-a sites in a monomer of FMO complex. We make a comparative numerical study of the EET, in terms of non-Markovian evolution of an initial coherent superposition state and a mixed state. A femto-second coherent laser pulse is suitably chosen to create the initial coherent superposition state. Such an initial state relaxes much faster than a mixed state thereby speeding up the EET. In this analysis, we have taken into account the relative orientation of the transition dipole moments of the bacateriochlorophyll-a sites and their relative excitation energies. Our results reveal that for 2D electronic spectroscopy experiments, the existing two-pathway model of energy transfer in FMO complex may not be suitable in our understanding of EET. I. INTRODUCTION In bacterial photosynthesis, antenna complexes harvest the photon and transfer the captured energy to the reac- tion center (which stores it chemically) with almost near unity quantum efficiency [1 -- 3]. Resolving the function- ing of these complex biological systems could be useful for newer solar energy applications [4]. During the bacterial photosynthesis in, e.g., Chloro- bium Tepidum, a photon is captured in situ by a light-harvesting molecule of the chlorosome antenna and this excitation is then transferred to the Fenna- Matthews-Olson (FMO) complex. This complex is a trimer and works as a energy transmitting wire be- tween the chlorosome antenna and the reaction cen- ter [2, 5, 6]. Each monomer in this trimer consists of seven bacteriochlorophyll-a (BChla) molecules immersed in protein environment and transmits the excitation en- ergy quite independently from the other monomers due to weak Coulomb coupling among them [7, 8]. The BChla 1 and BChla 6 of each monomer are close to the base-plate of the chlorosome antenna and the BChla 3 and BChla 4 are close to reaction center[9]. Recent experimental observations, based on 2D elec- tronic spectroscopy (2-DES) [10 -- 12] on the isolated FMO complex, have revealed that the electronic excitation gets transferred in a wave-like manner [13 -- 15] rather than in- coherent hopping, between different BChla sites. These long lasting (∼ 600f s) oscillations were arguably ex- plained as reminiscence of coherent beats (or dynamical coherence). These coherence's were further explained as originated from electronic modes and it was argued that coherence avoids the energetic traps and enhances the ef- ficiency of EET[8, 16 -- 22]. The inhomogeniety of protein environment and its effect on coherence were also illus- trated to study the EET dynamics[23 -- 28]. Recently the ∗ [email protected] origin of dynamical oscillatory coherence was explained on the basis of vibronic modes of BChla sites [29 -- 34]. However so far the effect of initial coherence created by the coherent laser light as used in 2-DES experiments has not been explored. In the 2-DES experiments, the FMO complex is extracted out from the bacteria (namely, Chlorobium Tepidum). An ultra-short femtosecond laser pulse is used to excite it; thereby it creates a coherent superposition of several excited BChla sites. However in a realistic bac- terial EET, the incoherent excitation (due to sunlight) makes an incoherent mixture of the excitations of the BChla sites. Different initial conditions are expected to have a substantially varied effect on the temporal dynam- ics of EET [35]. In the most of the earlier theoretical models [8, 33, 36] to explain the observations of 2-DES experiments, either the BChla 1 or the BChla 6 has been considered as initially excited. The dynamics thereby follows two different pathways, initiated by excitation in BChla 1 and BChla 6 respectively to reach the destina- tion BChla's (namely BChla 3 and BChla 4). However, such initial conditions clearly do not match with that of the laser excited FMO complex or in-vivo absorption of excitation by BChla sites of monomer. In this paper, we suitably choose the initial condition and explicitly study the effect of initial coherence on the EET, specifically by comparing the EET for initial coherent and incoherent excitation. Through the study of the dynamics of co- herently initiated EET one would be able to mimic the observations of the 2-DES experiments, which are still not fully understood. We first describe how a superpo- sition of the excited BChla sites is created when a short pulse is applied. We compare its dynamics with that of mixed state that is formed by the incoherent absorption of light. Our numerical study has the following salient features: (i) We propose that the distribution of initial excita- tion is not dictated by the local minima of excited states energy rather by the relative orientation of 7 1 0 2 p e S 2 2 ] h p - o i b . s c i s y h p [ 2 v 0 2 9 0 0 . 8 0 7 1 : v i X r a the dipole moments and the polarization of laser pulse. This would govern the preferential distribu- tion of excitation among all the BChla sites. (ii) X-ray crystallography of the FMO complex reveals an inhomogeneous protein environment of differ- ent BChla sites. Further, various vibronic modes of BChla sites also effect the EET dynamics. We have incorporated these features by fitting the spec- tral density with experimentally observed phonon wing and vibronic peaks[37]. Moreover, following Wendling et al[38], we have used five different vi- bronic modes with large value of Franck-Condon factors instead of using only one or two modes as has been done in earlier approaches [29 -- 33]. (iii) We have carefully considered all the realistic param- eters for each BChla site, e.g., their excited state energy and mutual Coulomb couplings. The structure of the paper is as follows: In section II we describe the preparation of initial state. In the following section (Section III), we illustrate the EET dynamics for initial coherent superposition as well as initial incoherent mixed state in FMO complex. Finally in section IV we conclude with summary and an outlook. II. PREPARATION OF INITIAL STATE To understand the coherent distribution of excitation among the different BChla sites, each BChla site is mod- elled as a two-level system. We assume that the EET happens only in the single-excited BChla basis (see Fig. 1b), i.e., the probabilities of bi-exciton and other higher states are negligible compared to the single excitation. Such suppression of bi-exciton pathways is also prefer- able to avoid damage due to excess energy [39] and can be attributed to dipole blockade arising from the closely- spaced arrangement of the BChla sites[40]. Also note that, in the 2-DES experiments, the temporal dynamics of only the excitons have been observed [10 -- 12]. To excite the BChla sites in the monomer of the FMO complex, a linearly polarized pulse (cid:126)E(t) = ε(t)e−iωLt + h.c. of the central frequency ωL and the time-dependent amplitude ε(t) is applied. The Hamiltonian in the dipole- approximation can therefore be written as H = (cid:80)7 + (cid:80) − (cid:16) (cid:126)dj0 j(cid:105)(cid:104)0 + (cid:126)d0j 0(cid:105)(cid:104)j(cid:17) j=1 {0 0(cid:105)(cid:104)0 + j j(cid:105)(cid:104)j (cid:111) i>j ∆ij (i(cid:105)(cid:104)j + j(cid:105)(cid:104)i) . (cid:126)E (1) . Here j represents the unperturbed energy of the jth BChla site and ∆ij is the tunneling frequency between ith and jth BChla site. The transition dipole moment matrix element of the transition j(cid:105) ↔ 0(cid:105) is represented by (cid:126)dj0. Here we assume that the time-scale of the laser pulse is much shorter than the time-scale at which the 2 FIG. 1: (Color online)(a) Orientation of transition dipole moments of different BChla sites. This figure is created with VEGA-ZZ [41] using PDB entry 3EOJ. (b) Schematic illustration of interaction of Gaussian laser pulse with different BChla sites in a monomer of the FMO complex. Here j0 represents the transition frequency of the jth BChla site and θj0 is the angle between polarization of light pulse and the transition dipole moment of the jth BChla site. a superposition can be written as ψ(cid:105) = (cid:80)7 coupling of the bath modes become effective. The above pulse would create a general superposition of all the rel- evant states of the monomer. Considering a common ground state 0(cid:105) and seven excited states j (cid:54)= 0(cid:105), such j=0 cj(t)j(cid:105). In the interaction picture, this would evolve with time according to the following Schrodinger equation: dψ(cid:105) dt = − i  HI ψ(cid:105) . (2) where HI represents the interaction Hamiltonian. The probability amplitude cj for the excited state of the jth site therefore evolves according to the following equation: cj = −i2G(t)c0 cos θj0 cos ωLtei(j−0)t−(cid:88) i∆ijciei(j−i)t j (cid:54)= 0 . i (3) We choose the following Gaussian profile of the Rabi − (t − t0)2 ∆τ 2 frequency of the pulse: G(t) = ge . We assume that the transition dipole moments (cid:126)dj0 of the BChla sites have approximately the same magnitude d, such that dε  . However, these dipoles have different orien- g = tations in three dimension, designated by the angle θj0 it makes with the polarisation direction of the pulse. We 3 the presence of the pulse. In Fig. 3, we show that for suitable choices of the pulse parameters, one can pump all the population from the ground state to the excited states. In addition, as displayed in Fig. 4, the coher- ence between the excited states and the ground state van- ishes, while, the excited states build up certain nonzero coherence among themselves. It is particularly impor- tant here to note that almost after 60f s (when the laser pulse has already interacted with the BChlas), as illus- trated in Fig. 4e, the ground-to-excited state coherence j (cid:54)= 0) dephases as has been observed ex- (i.e. ρj0; perimentally by Panitchayangkoon et al. [14]. Clearly, a pulsed excitation creates a coherent superposition of all the excited states of the BChla sites, which then evolves under the action of Coulomb tunnelling and the bath modes, once they become effective. This is contrary to the assumption by Ishizaki and Fleming [8], that it is either the BChla 1 or the BChla 6, that absorbs the ex- citation first, which then propagates through a chain of BChla sites to reach the reaction center. We emphasise that the unequal distribution of excitation across differ- ent BChla sites, as illustrated in Fig. 3, is primarily due to their different relative orientation of the angle of tran- sition dipole moments with the pulse polarisation. On the other hand, according to Forster theory, one obtains an in-situ incoherent distribution of excitation at initial time (i.e t = 0) between baseplate BChla site and the FMO complex [42] as follows: ρ11 = 0.6444; ρ22 = 0.2; ρ33 = 0.0; ρ44 = 0.0222; ρ55 = 0.0578; ρ66 = 0.0378; ρ77 = 0.0378; ρij = 0.0; i (cid:54)= j. (4) III. EET DYNAMICS In our numerical simulation of EET, we consider the relevant complex network of BChla sites in the monomer of FMO complex is shown in Fig.5. We statrt with the total Hamiltonian, as [43, 44] H = HS + HB + HSB , (5) where the system Hamiltonian HS is given by (cid:19) (cid:18) (cid:88) HS = 2 i,j assume that the polarisation is parallel to the transition dipole moment of BChla 1 as illustrated in Fig. 1b, while the other dipoles are oriented at non-zero angles. To analyse the dynamics of interaction, the angles of tran- sition dipole moments are calculated using VEGA-ZZ as illustrated in Fig. 1a [41]. Further, different values of the transition frequencies of the different BChla sites, as in- dicated by Adolphs and Renger [7], have been considered in our numerical study. In the view of average transition frequency, we choose ωL = 12420cm−1. We consider that the laser pulse interacts with the FMO complex for a femtosecond time scale (∼ 40f s). However, the effect of tunneling rate ∆ij will be dom- inant at a time scale (ranging from 0.4ps to 3ps), much after the femtosecond pulse dies away. To eluci- date this effect, we numerically solve the set of Eqs.(3) for the probability amplitude of BChla 1 and BChla 2 only, which have the strongest Coulomb coupling among all the combinations of the BChla sites. We observe that for first 40 fs, when the laser pulse interacts with the BChla's, the dynamics remains unaffected by the Coulomb coupling (i.e ∆ij) as illustrated by lower panel of Fig.3. However, after the pulse dies out, the effect of ∆ij becomes domi- nant as it creates the oscillations of population between two BChla's. Based on this observation, we neglect the effect of Coulomb coupling ∆ij for the dynamics of full monomer FMO complex for the interval during which the laser pulse interacts. 1 ∆ij FIG. 2: (Color online) Temporal dynamics of interaction of laser pulse with two BChla sites (BChla 1 and BChla 2). (a) Profile of a Gaussian pulse of width 15 f s. (b) Time evolution of population distribution among two BChla sites. Here the solid line represents the absence of Coulomb coupling (i.e ∆21 = 0), while dotted line illustrate the dynamics in the presence of Coulomb coupling (i.e ∆21 = 87.7cm−1). We use normalized parameters g = 21.5, t0 = 0.8 and ∆τ 2 = 0.03 and choose ω0 = 100 cm−1 for normalization. ijσij z + ∆ijσij x . (6) z = gjei(cid:105)(cid:104)gjei − ejgi(cid:105)(cid:104)ejgi and σij Here σij x = ejgi(cid:105)(cid:104)gjei + gjei(cid:105)(cid:104)ejgi are the equivalent Pauli spin operators in two-site single excitation basis, ij = j − i is the energy difference between jth and ith BChla sites, and ∆ij represents the tunneling frequency between them as shown in Fig.5. The bath Hamiltonian can be written as HB = ωkij b † kij bkij . (7) (cid:88) Next, we study the dynamics of all the BChla sites in kij the contributions from those vibronic modes correspond- ing to non-negligible values of Franck-Condon factors as originally proposed by Singh and Dasgupta [37]. 4 Here we also introduce the spectral function, as fol- lows, that has been obtained by fitting with the experi- mental data (obtained using fluorescence line narrowing spectroscopy by Wendling et al.[38]) Jij(ω) = Kijω − ω ωcij + e − (ω−ωl)2 2d2 . Kle (cid:18) ω (cid:19)−1/2 ωcij (cid:88) l (12) The above form of spectral density consists of the contri- butions of vibrational motion, arising from, for exam- ple, the environmental phonons with the Huang-Rhys factor Kij (with Kij = g2 ) and the vibronic modes kij with the Huang-Rhys factor Kl. Here ωcij is the cutoff frequency and ωl represents the frequencies of the active vibronic modes, with only the dominant Franck-Condon In our analysis, we choose ωl = 36 cm−1, 70 factors. cm−1, 173 cm−1, 185 cm−1, and 195 cm−1, the values of Kl equal to 40 times that of the corresponding Franck- Condon factors[38], and the width of the vibronic band as d2 = 18. As in the 2-DES experiments, we choose the exciton basis. These excitons are delocalised on different BChla sites with time-independent probabilities. Following Cho et al. [12], we expand the density operator ρ in the exci- tonic basis and numerically solve the Eq.(9) for following two different initial conditions: (a) the state ψ(t)(cid:105), as obtained followed by the initial excitation by the laser pulse (see Fig. 3 and 4). (b) the mixed state given by Eq.(4) obtained according to the Forster theory[42]. 31 , ρEx Here we plot only those four elements of density ma- trix for which the experimental information is available [13 -- 15]. For the Case (a), as illustrated by Fig.6a, the frequency of oscillations of the relevant density matrix el- ements in the excitonic basis, namely ρEx 21 and ρEx 22 , is almost the same as have been observed in the relevant experiments [14, 15]. More importantly, the time-scale till which the oscillatory nature persists, also matches with the experimental observations. In this case, the ex- citonic dynamical coherences ρEx 31 dephase in about 2 ps. Here it is interesting to note that even for the Case (b), excitonic population as well as coherence oscillates for about 700 fs as illustrated by Fig. 6a, but the time-scale of oscillations as well as the frequency of oscillations is considerably different from the experimen- tal observations [14, 15]. It implies that consideration of more realistic initial condition is necessary to mimic the experimental results exactly. 21 and ρEx Moreover, it is clear from the Fig. 6a that the dynam- ics of EET is quite fast for the case (a) as compared to the case (b). This is further confirmed by plotting the tem- poral profile of the mixedness (≡ Tr(ρ2)) of the density matrix ρ (Fig. 6b). For the coherently prepared initial FIG. 3: (Color online) (a) Profile of Gaussian pulse of width 15 fs. (b) Population dynamics of all the eight states of a monomer of the FMO complex, as an effect of a Gaussian pulse. We use normalized parameters g = 13.6873, t0 = 0.8 and ∆τ 2 = 0.03. † Here, bkij = bkj − bki and b are the anni- kij hilation and the creation operators, respectively, for the kijth bath mode, where bki represents the kth bath mode, local to ith BChla site. † − b ki = b † kj The system-bath interaction can be described by the spin-boson Hamiltonian as (cid:88) (cid:88) i,j kij HSB =  2 σij z gkij (bkij + b † kij ) , (8) where gkij is the electron-phonon coupling constant. To study the dynamics of EET, we use the following non-Markovian master equation for the system density matrix ρ [37, 45]: (cid:8)(cid:0)σij z − ρσij (cid:1) Dij(t) z ρ(cid:1) D∗ z ρ(cid:1) Uij(t) (cid:1) U∗ ij(t)(cid:9) . z σij z ij(t) z ρσij z − σij z σij z − σij z σij z − ρσij z σij z (9) ρ = − i [HS, ρ] i,j + 1 4 z ρσij z ρσij z ρσij (cid:80) + (cid:0)σij + (cid:0)σij + (cid:0)σij dt(cid:48)(cid:90) ∞ (cid:90) t dt(cid:48)(cid:90) ∞ (cid:90) t 0 0 The time-dependent coefficients in Eq.(9) represent the system-bath correlations and are given by Dij(t) = Uij(t) = dωJij(ω)¯n(ω, T )e−iω(t−t(cid:48)) , (10) dωJij(ω)[¯n(ω, T ) + 1]e−iω(t−t(cid:48)) ,(11) 0 0 where ¯n(ω, T ) represents the average number of phonons and Jij(ω) is the spectral density function considering 2 ¯n and U (t) → γ state [Case (a)] the system achieves the steady state at around 2.5 ps while for the initial incoherent distribution of excitation, the system is far away from the steady state even at 3 ps. To explore the reason behind such temporal behavior, we compare the dynamics of these two differ- ent initial conditions using Markovian as well as non- Markovian master equation. Note that, in Markovian master equation the time-dependent correlation function of Eq.(10) and (11) become time-independent de-phasing rates γ i.e D(t) → γ 2 . From our dy- namical study we observe that the system attains steady state almost at the same time for both the initial con- ditions, in both the Markovian (Fig.7) as well as non- Markovian evolution (Fig. 6b). It indicates that the speed up of coherently initiated dynamics is not due to non-Markovianity. This implies that the exchange of in- formation between system and environment (as it hap- pens in the non-Markovian dynamics), which preserves the coherence for long time, has no effect on the addi- tional speed up of EET. Rather as the initial coherent superposition state relaxes more quickly than the mixed incoherent state, it is the initial coherence that plays the most crucial role in speeding up the EET dynamics. To further elucidate the role of initial coherence we compared the dynamics of EET using two different ini- tial coherent distributions; in first case we assumed the polarization of laser pulse parallel to transition dipole moment of BChla 1 and in the second case to that of BChla 6. From Fig.8 it is evident that in the latter case, the speed up of EET is more as compared to that in the former case. It implies that different initial coherent distributions leads to varied speed-up of EET in FMO complex. Similar observations of coherent control were investigated in detail for several molecular processes, e.g. photo-dissociation, bio-molecular reactions, by Shapiro, Brumer and their co-workers [46]. To get the further detail of role of initial coherence, we study the dynamics of EET by modifying the off-diagonal elements of the density matrix as discussed below. The evolution of the off-diagonal elements of the density ma- 5 trix is governed by the equations of the form: ρij = Aρii + B(t)ρij . (13) Here the ad-hoc removal of the B(t) term would refer to a case, when there is no initial coherence. Comparing the dynamics in the presence of B(t) and in the absence of B(t) (see Fig.9), we find that the system attains steady state much earlier in the presence of B(t) than in the absence of B(t). IV. CONCLUSIONS In conclusion, we have shown how the initial prepa- ration of the collective state of the monomer can influ- ence the EET. Precisely speaking, we have adopted an equivalent eight-state model of the singly excited BChla sites in a monomer and considered their relative orienta- tion and energy differences. The dynamics of the EET is studied using master equation approach in the non- Markovian limit. It is found that the initial coherence of the state of the monomer takes a crucial role in the speed of EET. This study suggests that the 2-DES experiments, in which one excites the monomer with a coherent pulse, does not properly mimic a realistic EET, in which the initial excitation is incoherent in nature. This also puts the two-pathway model of EET in FMO complex in ques- tion, in which it is assumed that either the BChla 1 or the BChla 6 are initially excited. ACKNOWLEDGMENTS This work was supported by Department of Science and Technology (DST), Govt. of India, under the grant number SR/S2/LOP-0021/2012. One of us (D.S.) grate- fully thanks Paul Brumer, Jianshu Cao, Birgitta Whaley, Theodore Goodson III and Tomas Mancal for insightful discussions at QuEBS 2017, Jerusalem. [1] R. K. Chain and D. I. Arnon, PNAS 74, 3377 (1977). [2] H. V. Amerongen, L. Valkunas, and R. Grondelle, Pho- tosynthetic Excitons (World Scientific, Singapore, 2000). [3] R. E. Blankenship, Molecular Mechanism of Photosyn- thesis (World Scientific, London, 2002). [4] G. J. Meyer, J.Phys. Chem. Lett. 2, 1965 (2011). [5] R. E. Fenna and B. W. Matthews, Nature 258, 573 (1975). [6] M. T. W. Milder, B. Bruggemann, R. v. Grondelle, and J. L. Herek, Photosynth Res 104, 257 (2010). [7] J. Adolphs and T. Renger, Biophys. J. 91, 2778 (2006). [8] A. Ishizaki and G. R. Fleming, PNAS 106, 17255 (2009). [9] J. Wen, H. Zhang, M. L. Gross, and R. E. Blankenship, PNAS 106, 6134 (2009). [11] T. Brixner, J. Stenger, H. M. Vaswani, M. Cho, R. E. and G. R. Fleming, Nature 434, 625 Blankenship, (2005). [12] M. Cho, H. M. Vaswani, T. Brixner, J. Stenger, and G. R. Fleming, J.Phys. Chem. B 109, 10542 (2005). [13] G. S. Engel, T. R. Calhoun, E. L. Read, T.-K. Ahn, T. Mancal, Y.-C. Cheng, R. E. Blankenship, and G. R. Fleming, Nature 446, 782 (2007). [14] G. Panitchayangkoon, D. Hayes, K. A. Fransted, J. R. Caram, E. Harel, J. Wen, R. E. Blankenship, and G. S. Engel, PNAS 107, 12766 (2010). [15] G. Panitchayangkoon, D. V. Voronine, D. Abramavicius, J. R. Caram, N. H. C. Lewis, S. Mukamel, and G. S. Engel, PNAS 108, 20908 (2011). [10] T. Brixner, T. Mancal, I. V. Stiopkin, and G. R. Flem- [16] P. Nalbach, D. Braun, and M. Thorwart, Phys. Rev. E ing, J. Chem. Phys. 121, 4221 (2004). 84, 041926 (2011). 6 [17] B. Palmieri, D. Abramavicius, and S. Mukamel, The [32] N. Christensson, H. F. Kauffmann, T. Pullerits, and Journal of Chemical Physics 130, 204512 (2009). T. Mancal, J. Phys. Chem. B 116, 7449 (2012). [18] P. Bhattacharyya and K. L. Sebastian, Phys. Rev. E 87, 062712 (2013). [19] B.-q. Ai and S.-L. Zhu, Phys. Rev. E 86, 061917 (2012). [20] A. Shabani, M. Mohseni, H. Rabitz, and S. Lloyd, Phys. Rev. E 86, 011915 (2012). [33] A. W. Chin, J. Prior, R. Rosenbach, F. Caycedo-Soler, S. F. Huelga, and M. B. Plenio, Nat Phys 9, 113 (2013). [34] N. Killoran, S. F. Huelga, and M. B. Plenio, The Journal of Chemical Physics 143, 155102 (2015). [35] J. Cao, The Journal of Physical Chemistry B 110, 19040 [21] G.-Y. Chen, N. Lambert, C.-M. Li, Y.-N. Chen, and (2006). F. Nori, Phys. Rev. E 88, 032120 (2013). [36] J. Prior, A. W. Chin, S. F. Huelga, and M. B. Plenio, [22] S. Hoyer, A. Ishizaki, and K. B. Whaley, Phys. Rev. E Phys. Rev. Lett. 105, 050404 (2010). 86, 041911 (2012). [37] D. Singh and S. Dasgupta, J. Phys. Chem. B 121, 1290 [23] E. Rivera, D. Montemayor, M. Masia, and D. F. Coker, (2017). J. Phys. Chem. B 117, 5510 (2013). [24] M. Mohseni, P. Rebentrost, S. Lloyd, and A. Aspuru- Guzik, The Journal of Chemical Physics 129, 174106 (2008). [25] F. Caruso, A. W. Chin, A. Datta, S. F. Huelga, and M. B. Plenio, The Journal of Chemical Physics 131, 105106 (2009). [38] M. Wendling, T. Pullerits, M. A. Przyjalgowski, S. I. E. Vulto, T. J. Aartsma, R. van Grondelle, and H. van Amerongen, J. Phys. Chem. B 104, 5825 (2000). [39] H. Dong, S.-W. Li, Z. Yi, G. S. Agarwal, and M. O. Scully, arXiv:1608.04364v2 (2017). [40] M. Weidemuller, Nat Phys 5, 91 (2009). [41] A. Pedretti, L. Villa, and G. Vistoli, J Comput Aided [26] A. Eisfeld and J. S. Briggs, Phys. Rev. E 85, 046118 Mol Des 18, 167 (2004). (2012). [42] M. Schmidt am Busch, F. Muh, M. El-Amine Madjet, [27] C. A. Mujica-Martinez, P. Nalbach, and M. Thorwart, and T. Renger, J. Phys. Chem. Lett. 2, 93 (2011). Phys. Rev. E 88, 062719 (2013). [43] L. A. Pachon and P. Brumer, J.Physl Chem. Lett. 2, 2728 [28] M. Sarovar, Y.-C. Cheng, and K. B. Whaley, Phys. Rev. (2011). E 83, 011906 (2011). [44] J. B. Gilmore and R. H. McKenzie, Chem. Phys. Lett. [29] A. Chenu, N. Christensson, H. F. Kauffmann, (2013), Scientific Reports and T. Mancal, 10.1038/srep02029. 3 421, 266 (2006). [45] H. J. Carmichael, Statistical Methods in Quantum optics 1 (Springer, 2002). [30] C. Kreisbeck, T. Kramer, and A. Aspuru-Guzik, J. Phys. [46] M. Shapiro and P. Brumer, Quantum Control in Molec- Chem. B 117, 9380 (2013). ular Processes (Wiley-VCH, 2012). [31] R. Tempelaar, T. L. C. Jansen, and J. Knoester, J. Phys. Chem. B 118, 12865 (2014). 7 FIG. 4: (Color online) Temporal dynamics of coherences between different BChla sites due to interaction with the coherent laser pulse. FIG. 5: (Color online) Schematic illustration of the active channels of EET in the monomer of FMO complex. The superscript '(cid:63)' indicates that the molecules are in excited state. Here ij = j − i is the energy difference between jth and ith BChla sites, and ∆ij represents the tunneling frequency between them. 8 FIG. 6: (Color online) Temporal dynamics of (a) the population and coherence and (b) the mixedness Tr(ρ2) in the excitonic basis at a cryogenic temperature of 77 K. Thick lines show the dynamics for the initial condition obtained by the interaction of laser pulse with FMO complex, while dotted lines illustrate the dynamics for the initial condition of incoherent superposition determined using Forster theory. 9 FIG. 7: (Color online)Dynamics of trace of square of density matrix at cryogenic temperature 77 K with Markovian mater equation. Here thick line represents the coherently initiated dynamics while dotted line represents the incoherently initiated dynamics. We choose de-phasing rate γ = 26cm−1. FIG. 8: (Color online)Time evolution of trace of square of density matrix at cryogenic temperature 77 K. Solid line represents the initial condition when the polarization of laser pulse is parallel to transition dipole moment of BChla 1. Dashed line illustrate the initial condition when the polarization of laser pulse is considered to be parallel to transition dipole moment of BChla 6. 10 FIG. 9: (Color online)Temporal dynamics of trace of square of density matrix at cryogenic temperature 77 K. Solid line represents the presence of B(t), while dashed line represents the absence of B(t).
1208.6187
1
1208
2012-08-27T12:50:27
Enhancement of coherent energy transfer by disorder and temperature in light harvesting processes
[ "physics.bio-ph", "physics.chem-ph", "quant-ph" ]
We investigate the influence of static disorder and thermal excitations on excitonic energy transport in the light-harvesting apparatus of photosynthetic systems by solving the Schr\"{o}dinger equation and taking into account the coherent hoppings of excitons, the rates of exciton creation and annihilation in antennas and reaction centers, and the coupling to thermally excited phonons. The antennas and reaction centers are modeled, respectively, as the sources and drains which provide the channels for creation and annihilation of excitons. Phonon modes below a maximum frequency are coupled to the excitons that are continuously created in the antennas and depleted in the reaction centers, and the phonon population in these modes obeys the Bose-Einstein distribution at a given temperature. It is found that the energy transport is not only robust against the static disorder and the thermal noise, but it can also be enhanced by increasing the randomness and temperature in most parameter regimes. Relevance of our work to the highly efficient energy transport in photosynthetic systems is discussed.
physics.bio-ph
physics
Enhancement of coherent energy transfer by disorder and temperature in light harvesting processes Shi-Jie Xiong,1, 2, ∗ Ye Xiong,3 and Yang Zhao1, † 1School of Materials Science and Engineering, Nanyang Technological University, Singapore 639798, Singapore 2National Laboratory of Solid State Microstructures and Department of Physics, Nanjing University, Nanjing, China 3Department of Physics, Nanjing Normal University, Nanjing, China (Dated: July 23, 2018) Abstract We investigate the influence of static disorder and thermal excitations on excitonic energy trans- port in the light-harvesting apparatus of photosynthetic systems by solving the Schrodinger equa- tion and taking into account the coherent hoppings of excitons, the rates of exciton creation and annihilation in antennas and reaction centers, and the coupling to thermally excited phonons. The antennas and reaction centers are modeled, respectively, as the sources and drains which provide the channels for creation and annihilation of excitons. Phonon modes below a maximum frequency are coupled to the excitons that are continuously created in the antennas and depleted in the reaction centers, and the phonon population in these modes obeys the Bose-Einstein distribution at a given temperature. It is found that the energy transport is not only robust against the static disorder and the thermal noise, but it can also be enhanced by increasing the randomness and tem- perature in most parameter regimes. Relevance of our work to the highly efficient energy transport in photosynthetic systems is discussed. PACS numbers: 31.15.xk, 63.20.kk, 71.35.Aa 2 1 0 2 g u A 7 2 ] h p - o i b . s c i s y h p [ 1 v 7 8 1 6 . 8 0 2 1 : v i X r a ∗Electronic address: [email protected] †Electronic address: [email protected] 1 I. INTRODUCTION Photosynthesis is an essential means to obtain energy for all life forms on Earth. Although the detailed structures of photosynthetic systems are complicated and species-dependent, it is now believed that solar photons are absorbed to produce electronic excited states (called excitons) in molecular chromophores that are found in light-harvesting (or antenna) complexes. Many antenna complexes, together with reaction center complexes which can convert the energy stored in the excitons to chemical energy via biochemical reactions, form a transportation network for excitons. The photoexcitation process starts with the exciton creation in the antenna complexes, which is followed by exciton transfer across the pigment network, and ends with exciton trapping by the reaction center complexes. This energy transport process is completed on a 10 - 100 picosecond timescale [1 -- 6]. In recent years the high efficiency and quantum coherence of the exciton transfer processes amidst a noisy environment have atrracted great interest [7 -- 20]. The mechanism for such coherent transport has been extensively studied with various theoretical models [21 -- 26]. The exciton transfer from antennas to reaction centers is often modeled by the semiclassical Forster theory which considers incoherent hoppings between sites [21, 22]. To account for coherence in the energy transport processes, a microscopic description is provided by the Redfield theory where the exciton dynamics is solved from a master equation in a reduced space of excitons in the weak phonon coupling and Born-Markov approximation [23]. Based on this theory, Silbey et al. developed an approach for the diffusion of excitons [24, 25]; Haken and Strobl used a stochastic model [27]; and Kenkre and Knox established a generalized master equation formalism [28, 29] to investigate the coherent and incoherent aspects of the excitonic transfer. To make the models more realistic, Zhang et al. introduced a modified Redfield equation to include static disorder in the exciton system [30] which was later used to simulate the energy transfer dynamics in light-harvesting complexes of green plants [31]. Silbey and co-workers proposed a generalized theory which includes coherent transport effects only within donors and acceptors while treating interactions between them with the standard Forster model [32 -- 34]. A comparison between Forster, Redfield, and other models is given by Yang and Fleming in Ref. [35]. With a master equation approach in the Born-Markov and secular approximations, Mohseni et al. carried out computer simulations and found that the energy transport efficiency can be enhanced by modulating environmental noise [36]. 2 In the energy transfer process in photosynthetic systems, excitons are successionally cre- ated in the antennas and then trapped in the reaction centers, generating an uninterrupted energy flow from the antennas to the reaction centers. This energy flow is significantly af- fected not only by the coherence and decoherence effects during the exciton transfers across the pigment network, but also by the creation and annihilation of the excitons in the an- tennas and the reaction centers, respectively. In fact, if the excitons are viewed as energy carriers in the photosynthetic processes, the creation and annihilation of them in the anten- nas and the reaction centers play roles of source and drain of carriers, respectively. Thus, the widely used source-network-drain models describing continuous current of electron trans- port in quantum systems may also be applicable to treat the energy flow in photosynthetic systems if the creation and annihilation of excitons can be properly addressed. This work is aimed to describe the excitonic energy transport in photosynthetic systems with a full quantum-mechanical source-network-drain model. We include on an equal footing effects of scattering, disorder, exciton-phonon interactions, thermal excitations in a noisy environment, and efficiencies of antennas and reaction centers. The key issues are how to model exciton creations and annihilations with a suitable pair of source and drain, and how to include the thermal (temperature) effect in such a quantum-mechanical treatment. We use incoming (source) channels and outgoing (drain) channels for the creation and annihilation of excitons in antennas and reaction centers, and those channels are semi-infinite with the ends connected to different sites of a pigment network so that the exciton current from the source to the drain can be sustained. Considering that this process is taking place in a thermal bath, we include interactions between excitons and bath phonons in the network. With these key ingredients captured by a Hermitian Hamiltonian, the wave functions of exciton and phonons can be directly solved via the Schrodinger equation without invoking any classical or semiclassical approximations. Structural fluctuations in realistic systems are modeled with adding disorders to control parameters in the Hamiltonian. Exciton creation and annihilation rates can be adjusted by parameter-tuning of incoming and outgoing channels. After obtaining the wave functions, the exciton current, which reflects the efficiency of the photosynthesis, is calculated via the extended Landauer-Buttiker formula which works in the Hilbert space for many-body states [37 -- 39]. The numerical results show that the creation and annihilation rates in antennas and reaction centers and their adaptability have crucial effects on the total efficiency and the degree of coherence in the network. It is interesting 3 that the efficiency may be enhanced by introducing randomness and thermal excitations in many parameter regimes. At ambient temperature the efficiency may reach its maximum or saturation. This can help explain why photosynthesis is so efficient at room temperature and in noisy environments. The paper is organized as follows. In the next section we present the basic model and formalism. In the third section we discuss the results in the absence of interactions with phonons and analyze the effect of static randomness. In section IV we include the interaction with phonons and investigate the effect of thermal excitations on the energy transport. A brief summary is given in section V. II. MODEL AND FORMALISM We consider the following Hamiltonian for the exciton transport in a photosynthetic energy-transfer system: H0 = Hc + Hs + Hd, (1) where Hc is the Hamiltonian of a pigment network for exciton transport, such as the the light harvesting LH1 and LH2 photosynthetic complexes in a membrane of Rsp. photometricum [40, 41] and the Fenna-Mathews-Olsen (FMO) protein complex in green sulfur bacteria [42], connecting the antennas with the reaction centers. Hc can be written as Hc =Xi ǫia† i,0ai,0 +Xi6=j Ji,ja† i,0aj,0, (2) where a† i,0 is the creation operator of exciton on site i, ǫi is the site energy for excitons, and Ji,j j,i as the Hamiltonian is the hopping integral for excitons from site j to site i. One has Ji,j = J ∗ is Hermitian. Here a site may label a single chromophore or a cluster of chromophores. For the former, Ji,j originates from nearest-neighbor transfer integrals and Forster dipole-dipole interactions between well-separated pigments, both of which are strongly dependent on the distance between the pigments [15, 21, 43 -- 45]. For the latter, ǫi is the resonance level to host an exciton in the ith cluster of chromophores, while Ji,j labels the effective transfer integral between ith and jth clusters with a distance dependence that differs from that in the Forster theory [32]. Hs (Hd) is the tight-binding Hamiltonian for the source (drain), describing a semi-infinite virtual chain connected to a site in the pigment network as a path 4 of the successional creation (annihilation) of excitons in an antenna (reaction center) by radiation (biochemical reaction), ∞ Hs(d) = Xi∈S(D) Xn=1 [gi(a† i,nai,n−1 + a† i,n−1ai,n) + ǫia† i,nai,n], (3) where a semi-infinite chain is labeled by the index i of a network site to which it is connected, S (D) is the set of chains belonging to source (drain), n denotes the position on the chain counting from the pigment network, and gi is the nearest-neighbor hopping integral in the ith virtual chain. Thus the tight-binding Hamiltonian of the ith chain yields a transition exciton band of width 4gi centered at ǫi, and both gi and ǫi vary from chain to chain as the transition bands depend on detailed configurations. The widths of transition bands in the virtual chains mimic the energy uncertainty of excitons caused by the timescales of the photon-exciton conversion events in the antennas and charge separation events in the reaction centers. For systems relevant to photosynthesis [46, 47], gi varies from 1 meV to 10 meV, and ǫi from 1 eV to 2 eV. As the events occur in a quantized manner, these band widths are intrinsic parameters independent of actual light intensity in antennas or actual energy flow in reaction canters. In fact, the widths only set upper limits on the exciton currents flowing from the sources to the drains, and actual currents may be much smaller than those limits as they are also determined by light intensities and adsorption cross sections in antennas and by exciton trapping rates in reaction centers [40]. These additional physical ingredients, which can be viewed as external conditions that vary from time to time, are not included in the Hamiltonian, but their influences on exciton statistics in the sources and the drains will be taken into account in the formulation of the exciton current. The combined Hamiltonian may be illustrated by Fig. 1, where the networked sites corresponding to the antennas (reaction centers) are represented with green (magenta) circles, and the creation (annihilation) of excitons is represented with dashed incoming (dotted outgoing) arrows attached to corresponding sites. For a specific system such as FMO, a site in the pigment network may be connected to an antenna or a reaction center, and it is also possible that it is linked only to adjacent sites in the network. Detailed configurations in the source-network-drain model can be determined from a realistic photosynthetic system of interest. In this work, however, we are concerned with generic features of energy transfer in photosynthesis instead of effects of detailed configurations. We will therefore consider a network in two dimensions (2D) with closely packed sites with a distribution of control 5 FIG. 1: Schematic for a network of exciton transport. The antenna (reaction center) sites are represented with green (magenta) circles which are interconnected forming a network, and creation (annihilation) of excitons is represented with dashed incoming (dotted outgoing) arrows attached to corresponding sites. The couplings between sites, represented with brown links in the network, correspond to hopping integrals Ji,j in Hamiltonian Hc, while the incoming and outgoing channels expressed by the arrows depict virtual semi-infinite chains in Hamiltonians Hs and Hd. parameters. Static disorder erases structural details but the salient features remain after configurational averages. The 2D network to be studied here resembles, for example, the RC-LH1-LH2 photosynthetic complexes in a membrane of Rsp. photometricum. In the photosynthetic process the excitons are constantly flowing into the network via the incoming channels. The incident exciton with energy E in an incoming channel i (i ∈ S) can be expressed by a plane wave propagating along channel i towards the network ∞ ψin i (E) = exp[−iki(E)n]a† i,n0i, (4) where ki(E) = arccos[(E − ǫi)/2gi] is the wave vector and E is distributed within the interval [ǫi − 2gi, ǫi + 2gi]. The coefficients in the summation are set to unity implying that the rate of incident flow is 4gi/h if the transition band in the ith chain is occupied by excitons. The Xn=0 6 incoming wave propagates around in the network and then either exits via the outgoing channels or gets reflected to the incoming channels. The outgoing part and reflected part of the wave function are also expressed with plane waves in corresponding channels: ψout ij (E) = ∞ Xn=0 ∞ and tij(E) exp[ikj(E)n]a† j,n0i for j ∈ D, (5) (6) ψref i,i′(E) = rii′(E) exp[iki′(E)n]a† i′,n0i for i′ ∈ S, Xn=0 where tij and rii′ are transmission and reflection amplitudes from incident channel i to channel j (∈ D) and to channel i′ (∈ S), respectively. The complete form of the wave function for exciton with energy E and entering via channel i is ψi(E) = ψin i (E) +Xj∈D ψout ij (E) +Xi′∈S ψref ii′ (E). (7) Note that the wave function is normalized according to the incident rate. The transmission and reflection amplitudes tij(E) and rii′(E) can be solved from the Schodinger equation with the transfer-matrix or Green function technique, H0ψi(E) = Eψi(E). (8) The total exciton current through the system, representing the total efficiency of the pho- tosynthesis, can be calculated by the Landauer-Buttiker formula summing over all incident and output channels and integrating over energy [37] as: I = 1 h Xi∈S,j∈DZEij dE(cid:20)tij(E)2ηij(E)vj(E) vi(E) − tji(E)2ηji(E)vi(E) vj(E) (cid:21) . (9) Here tij(E)2 is the transmission probability from channel i to channel j for given energy E, vi(E) is the exciton velocity in the ith chain, and the ratio vj(E)/vi(E) accounts for the difference in transport rate between the input and output channels. The integration range Eij is the region of E where both ki(E) and kj(E) are real. The first and second terms in the integrand correspond to the currents from source to drain (positive) and from drain to source (negative), respectively. The sign of the total current is controlled by ηij(E), the probability that the state of energy E in the ith chain has exciton to emit while the state in the jth chain is empty and available to adsorb the exciton. So we can express ηij(E) as ηij(E) = pi(E)qj(E), where pi(E) (qj(E)) is probability of state in the ith (jth) chain being 7 occupied (vacant). As the excitons are created (adsorbed) only in source (drain) chains, pi(E) (qj(E)) is nonzero only when i ∈ S (j ∈ D). This guarantees that the second term in the above formula is always zero and the current is positive (from source to drain). pi(E) as a function of energy is determined by the light intensity and adsorption cross section in antennas, while qj(E) depends on the reopening timescale in reaction centers [40]. Although pi(E) and qj(E) set a further limitation on the energy transport, their effect can be taken into account simply by introducing a prefactor to the current. For the sake of simplicity, we will adopt a simple form of ηij(E): Then the current can be calculated as ηij(E) =  h Xi∈S,j∈DZEij 1 I = 1, for i ∈ S and j ∈ D, 0, otherwise. dEtij(E)2gj sin kj(E) gi sin ki(E) , where gj sin kj(E)/gi sin ki(E) is the ratio of velocities in channels i and j. Taking into account the configurational fluctuations in the network, the antennas, and the reaction centers, we introduce disorder in ǫi and Jij. Since the inter-site coupling in the network decreases rapidly as the inter-site distance increases, only couplings between nearest neighbors (NN) are kept. The parameter distributions can then be written as (10) (11) (12) (13) 1/ws(d), for i ∈ S(D) and ǫs(d) − ws(d)/2 ≤ ǫi ≤ ǫs(d) + ws(d)/2, 0, otherwise, 1/W, for i, j are NN sites and J0 − W/2 ≤ Jij ≤ J0 + W/2, 0, otherwise. P (ǫi) =  P (Jij) =  Here ws(d) and W quantify disorder in ǫi for i ∈ S(D) and in Jij for NN sites, respectively, while ǫs(d) and J0 are their averages. On the other hand, in order to reduce the number of parameters, we set gi = gs for i ∈ S and gi = gd for i ∈ D. III. EFFECT OF STATIC DISORDER IN THE ABSENCE OF INTERACTION WITH PHONONS In this section we investigate general features of energy transfers from the source to the drain by taking into account static disorder in the pigment network which reflects configu- rational randomness of the photosynthetic system. We carry out numerical calculations for 8 a network of 400 sites closely packed in a 2D plane. At this stage we do not include interac- tions with phonons and other thermal excitations, and focus instead on how the efficiency of energy transfer is affected by configurations and disorder of the network. Also randomly distributed in this network are the source sites connected to the antennas and the drain sites connected to the reaction centers. The density of the drain sites is denoted by p. In order to reduce the number of parameters in our model, it is assumed that each site in the network is connected either to an antenna or a reaction center, so the probabilities of a site being source and drain are 1 − p and p, respectively. By simply removing some incoming or outgoing channels, our calculation can be easily extended to include cases where some sites in the pigment network are connected to neither antennas nor reaction centers. Figure 2 displays the network transmission spectrum, i.e., the total probability of energy transfer from the incoming photons to the reaction centers as a function of the photon energy, for two different amplitudes of drain disorder. The spectrum has a global resonance peak of the source-network-drain system. It is apparent that the width of the resonance peak is rather small, reflecting the narrow transition bands in the source and drain channels. Detuning in the transition bands between the source and the drain rapidly reduces both the peak height and width. In the resonant case, ǫs = ǫd, the transmission spectrum is almost unaffected by the size of the energetic fluctuations in the drain, as can be seen by comparing the central peaks in Figs. 2(a) and 2(b). On the contrary, in the detuning case, the energetic disorder helps enhance energy transfers by increasing both the height and width of the resonance peak. It is interesting to note that even though the transition bands in the source and drain channels are in complete detuning, the transmission peak is only partially suppressed, but does not vanish. This means that such a pigment network structure with multiple antennas and reaction centers is a highly robust system for photosynthesis. Now we investigate how the total exciton current, an indicator of the global energy transfer efficiency of the system, is affected when the structure and control parameters of the pigment network are changed. In Fig. 3, we show the exciton current as a function of the reaction center density p for various values of the trapping rate gd which controls the width of transition bands in reaction centers. For a given set of gs and gd, there exists a critical reaction-center density pc that maximizes the exciton current, i.e., at pc, the excitons created in antennas can be most efficiently transported to the reaction centers. Upon increasing the transition band width gd, pc is decreased but the exciton current is increased, because for 9 s=1.493eV s=1.5eV s=1.507eV 16 (a) 8 0 20 10 t i n e c i f f e o C n o s s m s n a r T i i (b) 0 1.490 1.495 1.500 1.505 1.510 Energy (eV) FIG. 2: Total transmission coefficient as a function of the photon energy. (a) wd = 0.002eV, (b) wd = 0.006eV. Other parameters are: ǫd = 1.5eV, ws = 0.001eV, gd = 0.003eV, gs = 0.001eV, J0 = 0.01eV, W = 0.005eV, and p = 0.3. For a given set of parameters, there is only a single peak corresponding to the global resonance of the entire source-network-drain system. 10 gd=0.002eV gd=0.003eV gd=0.004eV gd=0.006eV gd=0.008eV gd=0.01eV 0.10 0.08 0.06 0.04 / ) h V e ( t n e r r u C n o t i c x E l a t o T 0.02 0.0 0.3 0.6 p 0.9 FIG. 3: Total exciton current as a function of drain density p for different values of gd. ǫs = ǫd, wd = 0.002eV, ws = 0.001eV, gs = 0.001eV, J0 = 0.01eV, and W = 0.005eV. For all combinations of gs and gd the curve of exciton current exhibits a single maximum at pc, pointing to an optimal p for energy transfer efficiency. a larger reaction rate, a given number of excitons can be collected by a smaller number of reaction centers. With the given form of ηij(E) in Eq. (10), the exciton creation and trapping rates in the antennas and the reaction centers may be controlled by gs and gd, respectively. Thus results obtained here could be used to explain the fact that some purple bacteria develop more expansive antenna systems when cultured under low light conditions, as a single maximum in the exciton current for a given set of gs and gd indicates that the concentration p may be adapted to the environment for optimal exciton transfer efficiency [48]. Our finding is also in qualitative agreement with the work of Fassioli et al. using a master equation approach, in which the efficiency can be optimized by changing the number ratio of LH2 and LH1 complexes (cf. Fig. 4 of Ref. [40]). Nevertheless, the maxima in Fig. 3 are not sharp peaks, pointing again to the wide range of structural suitability for the network to conduct photosynthetic energy transport. To investigate how the energy transfer efficiency is affected by configurational disorder in 11 our model, we plot in Fig. 4 the dependence of the total exciton current on the amplitudes of disorder in the pigment network, the sources, and the drains. It is interesting to note that the exciton transfer efficiency can be substantially enhanced by increasing the energetic disorder in the antennas and the reaction centers. The major efficiency bottlenecks are the low creation and annihilation rates in the antennas and reaction centers that lead to a very narrow transmission resonance centered at the exciton energies as shown in Fig. 2. Although the restriction of the narrow band on the efficiency can be partially lifted by increasing the number of channels, there is still a limitation on the maximum current through the drain. By introducing the disorder in the exciton energy, the resonant energies spread over a wider range, and consequently, the transmission resonance is widened and the exciton current increases. In the work of Plenio and co-workers [49], the efficiency can be enhanced by increasing the environment noise in some parameter regimes which correspond to J12 of about 100 cm−1 and a full width at half maximum of the spectral density of the noise from 10 cm−1 to 50 cm−1 (cf. Fig. 4 of Ref. [49]). This is comparable to J0 ∼ 0.01 eV and ws(d) from 0.0012 eV to 0.0062 eV in Fig. 4. Computer simulations carried out by Mohseni et al. also suggest that random noise in the environment might actually enhance the efficiency of the energy transfer in photosynthesis rather than degrade it [36]. On the other hand, the transfer rate in the network is usually much larger than the creation and annihilation rates in the antennas and in the reaction centers, i.e., J0 ≫ gs and J0 ≫ gd, so the excitonic quantum coherence can be easily preserved during the exciton transfer across the network, and the transfer efficiency has a very weak dependence on the network disorder W (i.e., the variance in the NN coupling in the network), as shown by the flat I(W ) curve in Fig. 4. This helps explain why a noisy environment can still preserve quantum coherence and the randomness in antennas and reaction centers even enhances the energy transfer efficiency. IV. EFFECT OF THERMALLY EXCITED PHONONS Previously, the effect of phonons on the energy transfer process was excluded in the discussion. However, in realistic photosynthetic systems, thermally excited phonons are always present and are believed to play important roles. In this section we investigate the effects of phonons and thermal excitations on the energy transport efficiency. The phonon- 12 0.090 0.085 0.080 0.075 0.070 0.065 0.060 / ) h V e ( ) W ( I , ) d w = s w ( I I(ws=wd), s= d I(ws=wd), d- s=2.5meV I(ws=wd), d- s=4meV I(W) 0.000 0.002 0.004 0.006 ws=wd, W (eV) 0.008 FIG. 4: Dependence of the total exciton current as a function of the degrees of disorder wd (= ws) and W . For I(ws = wd), W = 5meV. For I(W ), wd = 2meV, ws = 1meV, and ǫs = ǫd. Other parameters are: p = 0.3, ǫs = 1.5eV, gd = 0.003eV, gs = 0.001eV, and J0 = 0.01eV. The exciton current is enhanced by energetic fluctuations in the antennas and the reaction centers, but it is almost unaffected by disorder in the NN hopping integral of the pigment network. related Hamiltonian can be written as [11, 50, 51] Hph =Xk ¯hωkb† kbk +Xk Xi λk,ia† i,0ai,0(b† k + bk), (14) where bk (b† k) is the phonon annihilation (creation) operator for phonon mode k, ωk is the corresponding phonon frequency, and λk,i labels the exciton-phonon interaction strength at site i for phonon mode k. A wave function in the phonon subspace can be written as where ϕmi = X{nk} cm,{nk}φ{nk}i, (b† k)nk √nk! 0i φ{nk}i =Yk 13 (15) with {nk} being a set of phonon numbers in all modes, and cm,{nk} is the coefficient in the superposition. In the coupled exciton-phonon system, the wave function in Eq. (7) can be extended to include the phonon Hilbert subspace Ψi,{nk}(E) = Ψin i,{nk}(E) +Xj∈DX{n ′ k} Ψout i,{nk};j,{n ′ k}(E) +Xj∈S X{n ′ k} Ψref i,{nk};j,{n ′ k}(E), (16) where and Ψin i,{nk}(E) = ∞ Xn=0 exp[−iki,{nk}(E)n]a† i,n0i ⊗ φ{nk}i, (17) Ψout(ref) i,{nk};j,{n ′ k} (E) = ∞ Xn=0 t(r)i,{nk};j,{n ′ k}(E) exp[ikj,{n ′ k}(E)n]a† j,n0i ⊗ φ{n ′ k}i for j ∈ D(S). (18) Here Ψi,{nk}(E) is the complete wave function for an exciton entering from channel i clothed by a phonon cloud φ{nk}i, and E now labels the total energy of the exciton and companying phonons. The first, second, and third terms in Eq. (16) are the incident, transmitted and reflected parts of the wave function in different channels and with different phonon states. Given the total energy and the phonon energies the wave vector is calculated as ki,{nk}(E) = ′ k}(E) and ri,{nk};j,{n arccos[(E − ǫi −Pk nk¯hωk)/2gi]. For an exciton-phonon composite with total energy, E ti,{nk};j,{n k}(E) are transmission and reflection amplitudes from channel i with phonon state φ{nk}i to channel j with phonon state φ{n k}i, respectively, which can be determined from the Schrodinger equation ′ ′ (H0 + Hph)Ψi,{nk}(E) = EΨi,{nk}(E) (19) with a much-enlarged Hilbert space. By substituting Eqs. (17) and (18) into Eq. (14), one may obtain terms such as cj,1 ∞ Xn=0 exp[ikj,1n]a† j,n! ⊗ b† 10i + cj,1 ∞ Xn=0 exp[ikj,2n]a† j,n! ⊗ b† 20i in output channel j, where kj,1(2) = arccos[(E − ǫj − ¯hω1(2))/2gj], and cj,1(2) is the coefficient related to the corresponding transmission amplitudes. Obviously these terms cannot be factorized if kj,1 6= kj,2. Thus, in general, the exciton and phonon states are entangled in an output channel even if the two are not in the incoming states. As the creation and annihilation of excitons in the antennas and the reaction centers is very slow, it is reasonable to assume that when an exciton is created in an antenna, the 14 phonon equilibrium distribution in the network has recovered from the perturbation by the movement of the previous exciton. Thus the probability of the phonon state being φ{nk}i which accompanies an incident exciton is P{nk}(T ) = 1 Z exp[−Xk nk¯hωk/kBT ], (20) where the partition function is defined as Z = X{nk} exp[−Xk nk¯hωk/kBT ]. Therefore, by including the interactions with phonons, the total exciton current at temper- ature T can be calculated as I(T ) = 1 h Xi∈S,j∈D X{nk},{n k}ZE ′ ′ i,{nk }; j,{n k dE } P{nk}(T )ti,{nk}; j,{n ′ k}(E)2gj sin kj,{n gi sin ki,{nk}(E) ′ k}(E) , (21) where the integration range Ei,{nk}; j,{n ki,{nk}(E) and kj,{n k}(E) are real. ′ ′ k} is the span of the total energy E within which both Because of computational constraints, only a restricted number of phonon states are included. In the calculation, we adopt a uniform phonon frequency distribution in interval [0, ωc], which is similar to a simplified form of spectral density in Ref. [47]. We adopt ten equally spaced modes in this interval, and the phonon states with energies much greater than kBT are not included. The interaction strengths λk,i are assumed to be independent of the modes, i.e., λk,i = λ0. Calculated temperature dependence of the exciton current is plotted in Fig. 5 for various amplitudes of energetic disorder in the sources and the drains, and in Fig. 6 for various strengths of exciton-phonon coupling. It is interesting to note that, similar to the aforementioned disorder effect, in most cases by increasing the temperature the exciton current is enhanced, and eventually reaches saturation at room temperature. This means that the thermal excitations may actually favor energy transport in photosynthetic systems. The opposite trend, that the current decreases with increasing temperature, only occurs in a very narrow range of coupling strength, i.e., 6 meV ≤ λ0 ≤ 7 meV. This result is in agreement with that obtained by Mohseni et al. [36] who has shown that an effective interplay between free Hamiltonian evolution and thermal fluctuations in the environment leads to a substantial increase in energy transfer efficiency (from about 70% to 99%). Using a phonon mode of 180 cm−1 (∼ 22meV) and an exciton-phonon coupling strength of ∼ 10 15 0.018 0.016 0.014 0.012 0.010 / ) h V e ( ) T ( I wd=16meV, ws=16meV wd=13meV, ws=13meV wd=10meV, ws=10meV wd=5meV, ws=3meV wd=2meV, ws=1meV 0.008 0.00 0.01 0.02 0.03 0.04 kBT (eV) FIG. 5: Temperature dependence of the total exciton current for various amplitudes of energetic disorder in the sources and the drains. Other parameters are: p = 0.3, ǫs = 1.5eV, ǫd − ǫs = 2meV, gd = 1meV, gs = 0.5meV, J0 = 0.01eV, W = 5meV, ωc = 60meV, and λ0 = 1meV. For a given value of exciton-phonon coupling strength, e.g., λ0 = 1meV, the exciton current increases with the increasing temperature for all amplitudes of energetic disorder. meV, Plenio and coworkers showed that the coupling to phonons may significantly increase energy transfer efficiency (cf. Fig. 7 of Ref. [14]). Similar values of the exciton-phonon coupling strength can also be found in Fig. 6 where an enhancement of the exciton current with increasing temperature is shown. The phonon modes which interact with excitons can play the role of scatterers for the exciton transport, impeding the energy transmission, but they can also provide alternative paths for exciton transfers, therefore promoting energy transport. Our results show that at low temperatures the latter effect is dominant in most parameter regimes, but at high temperatures the two effects are in balance, yielding a temperature-independent exciton current. 16 0.024 0.022 0.020 0.018 0.016 0.014 0.012 0.010 / ) h V e ( ) T ( I 0.008 0.00 0.01 0=1meV 0=4meV 0=6meV 0=8meV 0=10meV 0.02 kBT (eV) 0=3meV 0=5meV 0=7meV 0=9meV 0.03 0.04 FIG. 6: Temperature dependence of the total exciton current for various strengths of exciton-phonon coupling. Other parameters are: p = 0.3, ǫs = 1.5eV, ǫd − ǫs = 2meV, wd = ws = 10meV, gd = 1meV, gs = 0.5meV, J0 = 0.01eV, W = 5meV, and ωc = 60meV. The exciton current is found to increase with the temperature for all exciton-phonon coupling strength except the narrow bracket from 6 meV to 7 meV. V. CONCLUSIONS In this work a source-network-drain exciton Hamiltonian is introduced to investigate the energy transport processes in photosynthetic systems. The excitons are electronic energy carriers, and the creation and annihilation of the excitons in antennas and reaction cen- ters are described by source and drain channels, respectively. The creation and annihilation rates in antennas and reaction centers are characterized by the band widths in corresponding channels and by the statistical factors of excitons depending on light intensity, adsorption cross section in antennas and the reopening timescale in reaction centers. The static dis- order in antennas and reaction centers is taken into account by introducing randomness of corresponding parameters in the Hamiltonian, while the effects of thermally excited phonons are described by exciton-phonon interactions. As the total Hamiltonian is Hermitian and the Schrodinger equation is solved for the combined wave function of excitons and phonons, 17 our treatments here are fully quantum mechanical and the effect of coherence is included accordingly. The exciton current, an efficiency indicator of the energy transport, can be calculated using the Landauer-Buttiker formula. Our results show that the exciton current may be enhanced by increasing the static disorder and increasing the temperature in many parameter regimes, suggesting that the coherent energy transport in light-harvesting sys- tems is robust against most environmental noises, a conclusion that is in agreement with previous findings in the literature. Furthermore, the current approach provides a novel, flex- ible platform to investigate complicated interplays among various configurational, quantum and thermal factors in the energy transport processes of natural and artificial photosynthetic systems. Previously, many theoretical efforts were based upon solving the phenomenological Lind- blad master equations with built-in dissipation. In this work we are interested in the steady state in which excitons flow continuously to the reaction centers, and our results are not affected by initial conditions and transient behavior. It is also possible to find the steady- state density matrix for a single-particle Lindblad equation. We note that the exciton life time is in the order of nanoseconds while it takes hundreds of femtoseconds for an exciton to be transferred to a reaction center. This huge difference in timescale indicates that exciton decay can be neglected when considering excitonic energy transfers from antennas to reac- tion centers, therefore lending support to the Landauer-Buttiker scheme adopted here. Our approach assumes the single-exciton picture in which the Schrodinger equation is solved, and the applicability of this basic assumption to photosynthetic systems is justified by slow injec- tion and fast transport of excitons in the pigment network. In a master-equation approach, energy transfer efficiency is measured by quantities such as the exciton trapping probability, while in the Landauer-Buttiker scheme it is described by the exciton current. Furthermore, the total exciton current also provides a global description of the photosynthetic system with multi-channel inputs and outputs and interferences among them. Acknowledgments Support from the Singapore National Research Foundation through the Competitive Re- search Programme (CRP) under Project No. NRF-CRP5-2009-04 is gratefully acknowledged. This work is also supported in part by the State Key Programs for Basic Research of China 18 (Grant No. 2011CB922102), and by National Foundation of Natural Science in China Grant No. 61076094. [1] R. van Grondelle, J. P. Dekker, T. Gillbro, and V. Sundstrom, Biochim. Biophys. Acta 1187, 1 (1994). [2] H. van Amerongen, L. Valkunas, and R. van Grondelle, Photosynthetic Excitons (World Sci- entific, 2000). [3] V. Sundstrom, T. Pullerits, and R. van Grondelle, J. Phys. Chem. B 103, 2327 (1999). [4] A. R. Grossman, D. Bhaya, K. E. Apt, and D. M. Kehoe, Annu. Rev. Genetics 29, 231 (1995). [5] Y. C. Cheng and G. R. Fleming, Annu. Rev. Phys. Chem. 60, 241 (2009). [6] R. J. Cogdell, A. Gall, and J. Kohler, Quarterly Rev. Biophys. 39, 227 (2006). [7] Y. Zhao, T. Meier, W. M. Zhang, V. Chernyak, and S. Mukamel, J. Phys. Chem. B 103, 3954 (1999). [8] T. Meier, Y. Zhao, V. Chernyak, and S. Mukamel, J. Chem. Phys. 107, 3876 (1997). [9] G. C. Yang, N. Wu, T. Chen, K.W. Sun, and Y. Zhao, J. Phys. Chem. C 116, 3747 (2012). [10] N. Wu, K. W. Sun, Z. Chang, and Y. Zhao, J. Chem. Phys. 136, 124513 (2012). [11] J. Ye et al., J. Chem. Phys. 136, 245104 (2012). [12] J. Moix, Y. Zhao, and J. Cao, Phys. Rev. B 85, 115412 (2012). [13] E. Anisimovas and A. Matulis, Phys. Rev. A 75, 022104 (2007). [14] A. W. Chin, A. Datta, F. Caruso, S. F. Huelga, and M. B. Plenio, New J. Phys. 12, 065002 (2010). [15] C. Olbrich et al., J. Phys. Chem. Lett. 2, 1771 (2011). [16] G. S. Engel et al., Nature 446, 782 (2007). [17] H. Lee, Y. C. Cheng, and G. R. Fleming, Science 316, 1462 (2007). [18] G. Panitchayangkoon et al. Proc. Natl Acad. Sci. USA 107, 12766 (2010). [19] E. Collini et al. Nature 463, 644 (2010). [20] P. Ball, The dawn of quantum biology. Nature 474, 272 (2011). [21] T. Forster, in Modern Quantum Chemistry, Istanbul Lectures, edited by O. Sinanoglu (Aca- demic, New York, 1965), Vol. 3, pp. 93-137. [22] G. D. Scholes, Annu. Rev. Phys. Chem. 54, 57 (2003). 19 [23] A. G. Redfield, Adv. Magn. Reson. 1, 1 (1965). [24] M. Grover and R. Silbey, J. Chem. Phys. 54, 4843 (1971) [25] S. Rackovsky and R. Silbey, Mol. Phys. 25, 61 (1973). [26] J. Sun, B. Luo, and Y. Zhao, Phys. Rev. B 82, 014305 (2010). [27] H. Haken and G. Strobl, Z. Phys. 262, 135 (1973). [28] V. M. Kenkre and R. S. Knox, Phys. Rev. Lett. 33, 803 (1974). [29] V. M. Kenkre and P. Reineker, Exciton Dynamics in Molecular Crystals and Aggregates (Springer, Berlin, 1982). [30] W. M. Zhang, T. Meier, V. Chernyak, and S. Mukamel, J. Chem. Phys. 108, 7763 (1998). [31] V. I. Novoderezhkin, M. A. Palacios, H. van Amerongen, and R. van Grondelle, J. Phys. Chem. B 108, 10363 (2004). [32] S. Jang, M. D. Newton, and R. J. Silbey, Phys. Rev. Lett. 92, 218301 (2004). [33] Y. C. Cheng and R. J. Silbey, Phys. Rev. Lett. 96, 028103 (2006). [34] S. Jang, M. D. Newton, and R. J. Silbey, J. Phys. Chem. B 111, 6807 (2007). [35] M. Yang and G. R. Fleming, Chem. Phys. 275, 355 (2002). [36] M. Mohseni, P. Rebentrost, S. Lloyd, and A. Aspuru-Guzik, J. Chem. Phys. 129, 174106 (2008). [37] R. Landauer, IBM J. Res. Dev. 1, 223 (1957); M. Buttiker, Phys. Rev. Lett. 57, 1761 (1986). [38] S. J. Xiong and Y. Xiong, Phys. Rev. Lett. 83, 1407 (1999). [39] Y. Xiong and S. J. Xiong, J. Comp. Phys. 231, 1197 (2012). [40] F. Fassioli, A. Olaya-Castro, S. Scheuring, J. N. Sturgis, and N. F. Johnson, Biophysical Journal, 97, 2464 (2009). [41] E. Harel, J. Chem. Phys. 136, 174104 (2012). [42] M. Cho, H. M. Vaswani, T. Brixner, J. Stenger, and G. R. Fleming, J. Phys. Chem. B 109, 10542 (2005). [43] V. M. Agranovich and M. D. Galanin, Electronic Excitation Energy Transfer in Condensed Matter, (North-Holland, Amsterdam, 1982). [44] Y. Zhao, M. F. Ng, G. H. Chen, Phys. Rev. E, 69, 032902, (2004). [45] M. S. am Busch, F. Muh, M. E. Madjet, and T. Renger, J. Phys. Chem. Lett. 2, 93 (2011). [46] A. Olyaya-Castro, C. F. Lee, F. F. Olsen, and N. F. Johnson, Phys. Rev. B 78, 085115 (2008). [47] C. Olbrich and U. Kleinekathofer, J. Phys. Chem. B 114, 12427 (2010). 20 [48] S. Scheuring and J. N. Sturgis, Science 309, 484 (2005). [49] A. W. Chin, S. F. Huelga, and M. B. Plenio, Phil. Trans. R. Soc. A 370, 3638 (2012). [50] T. Holstein, Ann. Phys. 8, 325 (1959); 8, 343 (1959). [51] G. D. Mahan, Many Particle Physics, 3rd ed. (Kluwer, 2000). 21
1804.05786
1
1804
2018-04-16T16:55:57
Periodic spiking by a pair of ionic channels
[ "physics.bio-ph" ]
Neuronal cells present periodic trains of localized voltage spikes involving a large amount of different ionic channels. A relevant question is whether this is a cooperative effect or it could also be an intrinsic property of individual channels. Here we use a Langevin formulation for the stochastic dynamics of a pair of Na and K ionic channels. These two channels are simple gated pore models where a minimum set of degrees of freedom follow standard statistical physics. The whole system is totally autonomous without any external energy input, except for the chemical energy of the different ionic concentrations across the membrane. As a result it is shown that a unique pair of different ionic channels can sustain membrane potential periodic spikes. The spikes are due to the interaction between the membrane potential, the ionic flows and the dynamics of the internal parts (gates) of each channel structures. The spike involves a series of dynamical steps being the more relevant one the leak of Na ions. Missing spike events are caused by the altered functioning of specific model parts. The time dependent spike structure is comparable with experimental data.
physics.bio-ph
physics
Periodic spiking by a pair of ionic channels ✩ L. Ram´ırez -- Piscina1 Departament de F´ısica, Universitat Polit`ecnica de Catalunya, Avinguda Doctor Maran´on 44, 08028 Barcelona, Spain J.M. Sancho Departament de F´ısica de la Mat`eria Condensada, Universitat de Barcelona, Universitat de Barcelona Institute of Complex Systems (UBICS), Mart´ı i Franqu´es 1, 08028 Barcelona, Spain Abstract Neuronal cells present periodic trains of localized voltage spikes involving a large amount of different ionic channels. A relevant question is whether this is a cooperative effect or it could also be an intrinsic property of individual channels. Here we use a Langevin formulation for the stochastic dynamics of a pair of Na and K ionic channels. These two channels are simple gated pore models where a minimum set of degrees of freedom follow standard statistical physics. The whole system is totally autonomous without any external energy input, except for the chemical energy of the different ionic concentrations across the membrane. As a result it is shown that a unique pair of different ionic channels can sustain membrane potential periodic spikes. The spikes are due to the interaction between the membrane potential, the ionic flows and the dynamics of the internal parts (gates) of each channel structures. The spike involves a series of dynamical steps being the more relevant one the leak of Na ions. Missing spike events are caused by the altered functioning of specific model parts. The time dependent spike structure is comparable with experimental data. Keywords: Langevin equations; nonlinear oscillations; periodic firing; channel gating 1. Introduction Neurons exhibit a great variety of firing electrical pat- terns. It is recognized that the neuronal electrical activity depends, besides from the synaptic inputs, on its electro- physiological membrane properties of the specific type of neuron [1, 2]. These membrane properties ultimately de- pends on physical processes, such as the movement of ions through the molecular channels, the membrane potential dynamics, and the gating dynamics of the channels. These processes are indeed complex, involving a large hierarchy of biomolecular structures from the atomic to the cellular and multicellular scales. However, from the point of view of physical modeling, it is interesting to explain these processes by using a re- duced formulation, with only a minimum of relevant phys- ical mechanisms and variables. One could then address for instance the question of whether the observed firing behavior could appear in a very simple device or it is in- stead necessary a large biological complexity or a whole ✩ c(cid:13)2018. This manuscript version is made available under the CC-BY-NC-ND 4.0 license http://creativecommons.org/licenses/by- nc-nd/4.0/ Email address: [email protected] (L. Ram´ırez -- Piscina) 1Corresponding author collectivity of channels [2, 3]. Also, due the the large va- riety and complexity of real neural outputs, it is funda- mental the recognition of the different fundamental firing patterns. Then one could in principle try to reproduce the most basic response patterns by using only the neces- sary ingredients in a unified framework, and to study the changes of these basic patterns when varying the physical parameters. Once this aim is achieved, it is open the pos- sibility of introducing more elements in order to reproduce more complex patterns. This objective implies the choice of a simple theoretical scenario as starting approach. Most theoretical approaches for action potential dynam- ics are based on the classic Hodking-Huxley [4, 5] frame- work. It originally consists of deterministic equations for the dynamics of membrane permeability. At the level of individual channels other computational approaches con- sider microscopic details at the atomic scale by means of molecular dynamics simulations [6]. Also mechanical mod- els for the gating dynamics for a K channel have been proposed [7]. In the last years there has been an increas- ing interest on the role of channel noise in neural firing patterns [8, 9, 10, 11, 12, 13, 14]. Fluctuations have usu- ally been modeled by using either master equations for the gate states [15, 16], or by including stochastic terms into the membrane conductivities [17]. Also the diffusion of ions inside the channel has been considered [18, 19]. Preprint submitted to Physica A November 9, 2018 Recently, a semi-microscopic approach for the stochastic dynamics of individual molecular channels was formulated [20, 21]. This approach uses only some relevant physical mechanisms acting on a minimum model. Variables repre- senting the relevant degrees of freedom (ion positions and gate states) interact through a single energy functional, and a Langevin dynamics is then constructed by follow- ing standard rules from statistical physics. It is worth to remark the physical consistency of the resulting formula- tion. On the one hand the magnitude of the fluctuations verifies the fluctuation-dissipation theorem. On the other hand the working of the channel is autonomous, with the energy source being the chemical energy from the differ- ent ionic concentrations inside and outside the cell. The resulting model was able to reproduce the basic properties of Na and K channels [21] and the excitable properties of a single Na channel in the presence of K leak [20]. Here we will show how this approach, applied to a unique pair of channels following known physical mech- anisms acting on the membrane, is able to generate a pe- riodic firing pattern of electrical activity. We will find the effects of the external control parameters, and answer questions on how to control the stability of the periodicity or where are the sources of misfunctioning. Figure 1: Picture of the pair Na and K channel models with their respective gates Y1, Y2, Y3, ionic concentrations [N a+], [K +], ionic fluxes (arrows), and membrane potential ∆V . The structure of this paper is as follows. In the next section we will show how a couple of gating pores, model- ing Na and K channels, exhibit periodic firing with a well controlled period. Next we will study the changes in the re- sponse pattern by changes of different channel parameters and external conditions. We end with some conclusions. Some specific details of the approach are presented in the Appendix. 2. The periodic spiking minimum device We have employed a minimum device consisting of a couple formed by two channel models, as a rude simplifi- 2 cation of Na and K voltage-gated ionic channels, as pic- tured in Fig. 1. Each model is a simple physical pore with active gates controlled by the membrane potential. The Na-like channel has two gates, whereas the K-like chan- nel has a single gate. This physical structure has been designed [20, 21] to follow experimental observations on individual Na and K channels (see for instance Chapter 4 in Ref. [2]). The physical variables of the model are the ion positions xi, the gate coordinates Yj, and the membrane potential ∆V . The Yi are bistable variables such that the value 0/1 corresponds to the corresponding gate being closed/open. Gates Y1, Y2 of the Na channel are activating and deacti- vating (i.e. open and close with membrane depolarization) respectively. According to experimental evidence, they are mediated by different parts of the channel structure [2, 22]. Gate Y3 of the K gate is activating. These gates (Y1, Y2, Y3) can be related to the activating (m(∆V ), n(∆V )) and deactivating (h(∆V )) functions appearing in the H -- H for- mulation [4, 5]. We have considered a single degree of freedom (Y vari- able) for each gate. Generalization to more Y variables to account for the existence of several voltage-sensing do- mains in real channels is straightforward, but it has not been considered here in order to keep the formulation of the model to a minimum. The control parameters are the out cell ionic concentrations with fixed intra cell con- centrations. Thus we implicitly neglect changes in bulk concentrations originated by the small local flow from the pair of channels. The model is defined by an energy functional U describ- ing the interaction between the different physical variables. The explicit construction of this functional is described in the Appendix. Following basic statistical physics the vari- ables xi, Yj follow a brownian stochastic dynamics [20]: γx xi = −∂xiU (xi, Yj, ∆V ) + ξi(t), Yj = −∂Yj U (xi, Yj, ∆V ) + ξYj (t), γYj where thermal noises fulfill hξa(t)ξb(t′)i = 2γa kBT δa,b δ(t − t′), (1) (2) (3) and γa are the corresponding frictions. Note that by con- struction the model verifies fluctuation-dissipation theo- rem. An analogous formulation for the coupling between ion and channel state, consistent with statistical physics, was already used in Ref. [18] to describe the stochastic behavior of singly occupied ion channels. The numerical simulation of these equations, by a first order algorithm, allows to record the state of the gates and the position of the ions. The ionic concentrations in the bulk are implemented as boundary conditions at both ends of the channel for the Langevin dynamics of ions [23]. Finally, the dynamics of the membrane potential follows the classical capacitor equation CM d∆V dt = −Xi Ii, (4) ) V m ( V ∆ 50 0 -50 -100 50 75 100 t (ms) 125 150 Figure 2: Train of spikes with good periodicity and the same quali- tative structure. [Kout] = 0.149 M. where CM is the membrane capacity assumed to be con- stant and the r.h.s term includes all the ionic fluxes either across the channel or membrane leaks. Letting the system evolve according the dynamical equa- tions, without any external perturbation, we observe that the membrane potential present periodic pulses or spikes. We show a temporal evolution of this quantity in Fig. 2, which is very similar to those presented in some experi- ments, as it is seen for instance in Chapter 17 in Ref. [2]. The working of the system during a few spikes, in this range of parameters, can be seen in Figs. 3, 4, where the evolution of the three gate variables and the membrane potential are shown. Regime Stand-by Pulse Pulse Refractory Refractory Refractory Error Y2 Error Y2 Error Y3 step a b c d1 d2 d3 b′ c′ = d1 a′ = d3 Y1 0 1 1 1 0 0 1 1 0 Y2 1 1 1 0 0 1 0 0 1 Y3 0 0 1 1 1 1 0 1 1 Table 1: Table with the different regimes, steps and gate states, during the pulse. For the cases of missing spikes, the altered gate steps are also specified in boldface. These figures reveal that any successful pulse involves three different regimes : stand-by, the spike event and the refractory state. These physical regimens are composed of steps (Table 1): stand-by from the membrane at the K-Nernst potential which corresponds to the dynamical depolarizing (step a), the spike (steps b and c) and the restoring of the K-Nernst potential (step d). Being more explicit the description of each step is: a).- This is the main dynamical process to trigger the pulse and to control the spike period. This step has the largest time span where the Na gate Y1 is closed and the 3 2 Y , 1 Y 3 Y Y1 Y2 Y3 1 0.8 0.6 0.4 0.2 0 0.8 0.6 0.4 0.2 0 50 0 a a ) V m ( V ∆ -50 -100 30 32 34 36 38 t (ms) 40 42 44 46 Figure 3: Temporal evolution of Na and K channel variables during three typical pulses. It can be seen that the dynamical step (a) is the main factor controlling the periodicity of the train of spikes. (Top) Na gate variables Y1 (black line) and Y2 (red line) versus time; (Middle) gate variable Y3 of the K channel; (bottom) membrane action potential. 1 0.8 0.6 0.4 0.2 0 0.8 0.6 0.4 0.2 0 50 0 -50 -100 2 Y , Y 1 3 Y ) V m ( V ∆ 38.8 Y1 Y2 Y3 a b c 39 d1 39.2 t (ms) d2 d3 a 39.4 39.6 Figure 4: Magnified time evolution during a single typical pulse (mid- dle peak of Fig. 3). The shorter dynamical steps (b, c, d) specified in Table 1 are shown in more detail. (Top) Gates variables Y1 (black line) and Y2 (red line) of the Na channel versus time; (Middle) gate variable Y3 of the K channel; (bottom) membrane action potential. 2 Y , 1 Y 3 Y 1 0.8 0.6 0.4 0.2 0 0.8 0.6 0.4 0.2 0 50 0 -50 V ∆ -100 784 788 792 796 800 t (ms) 804 808 812 Figure 5: Example of gate dynamics during a missing pulse close to t = 803 ms due to an altered Na gates sequence (Y2 closes before opening of Y1). Color code as in previous figures. 2 Y , 1 Y 3 Y 1 0.8 0.6 0.4 0.2 0 0.8 0.6 0.4 0.2 0 50 0 -50 V ∆ -100 954 956 958 960 962 964 t (ms) 966 968 970 972 Figure 6: Example of gate dynamics during a missing pulse close to t = 962 ms due to altered K gate opening (Y3 opens before Y1). Color code as in previous figures. 50 0 ) V m ( V ∆ -50 -100 0 5 10 15 t (ms) 20 25 30 Figure 7: Time evolution of the membrane potential between two consecutive pulses for four values of the Y1 barrier height considered in Table 2. Times are shifted in order to place the first pulse of each case at t = 0 and in this way to highlight the corresponding period. The results illustrate how the period increases with barrier height. Black line: Vd(Y 1) = 8 kB T, red line: Vd(Y 1) = 9 kB T, blue line Vd(Y 1) = 9.85 kB T, and green line: Vd(Y 1) = 10 kB T. Y2 gate is open. The K gate Y3 is just closed and the mem- brane potential is the K Nernst value ∆V ∼ −92 mV (See Fig. 3 at t = 34 ms). Now the Y1 gate, although closed, has nevertheless a small leak of Na ions. In Fig. 3 we see how this leak increases the action potential depolarizing the membrane. An analogous phase, the so called pace- maker depolarization, is found in Na channels of Purkinje cells [2]. b) and c).- The pulse. This regime is the shortest time interval, which is magnified in Fig. 4. The Na leak increas- ing ∆V triggers the sequence of events in the Na channel that originates first the opening of gate Y1, which permits a larger flux of Na ions, depolarizing the membrane by increasing ∆V even further forming the pulse (b). Very quickly the gate Y3 of the K channel also opens and gate Y2 of the Na channel closes. Then the flux of K ions to- wards the exterior of cell leads the membrane potential to reset its polarization, leading to more negative values of ∆V (c). d).- Na gate Y2 closes and the membrane potential re- laxes towards the K-Nernst potential because during all this step the Na channel is closed and the K channel is open. In this regime the Na channel presents three dif- ferent configurations of closed channel. d1: Y1 open and Y2 closed, d2: with Y1 and Y2 both closed, and d3 with Y1 closed and Y2 open. Then the Na channel rests in stand-by regime, until the K gate Y3 closes and the depolarization starts a new cycle. In this dynamics some cycles are observed to be missing because of the inherent stochasticity of the gate dynam- ics. In the first example the Na door Y2 closes before the completion of the depolarization of the action potential. This misses a possible spike at t = 803 ms (Fig. 5). In the second example the K door Y3 opens reducing the mem- brane potential without a pulse at t = 962 ms (Fig. 6). The system goes back to the state d3. These two cases of 4 missed spikes disturb the regularity of the train of spikes, increasing the mean and the variance of the period. All these steps are summarized in Table 1. 3. Effect of the Na ions leak The pulse of the action potential is produced by the syn- chronized opening and closing of gates of the Na channel, and it is completely similar to the process described for the model of excitable membrane of Ref. [20]. The period- icity of the oscillatory behavior is controlled by the slow leak of the Na ions when both channels are closed, which introduces its time scale into the process. This slow leak could be produced by other processes or channels in more complex devices. In our case the period of the oscillations should directly be related to the rate at which ions can cross the potential barrier of the closed gate Y1. As it is well known the time scale for crossing a barrier is given by the Arrhenius (or Kramers) law, according to which T ∼ exp Vd/kBT for Vd ≫ kBT , where Vd is the height of the barrier (in this case that of the Y1 gate). To test this prediction we have performed simulations with differ- ent values of Vd(Y1). Results of mean periods, measured as the time intervals between membrane potential peaks, are presented in Table 2. Vd(Y1) (kBT ) Number of oscillations Mean period hT i (ms) Std dev. σT (ms) 8 501 4.03 2.77 9 274 7.33 3.33 9.85 115 17.3 7.2 10 94 20.9 10.2 Table 2: Oscillation periods for different Y1 barrier heights, calcu- lated for a simulation time span of 2025 ms. Note that according to the Kramers' law, by increas- ing the barrier height in 1 kBT the crossing time should increase in a factor e = 2.718... for Vd ≫ kBT . Ac- cording the results in Table 2: T9kB T/T8kB T = 1.8, T10kB T/T9kB T = 2.8. While the order of magnitude is correct, the result is better for the larger barrier case as expected. In Fig. 7 four examples with different periods are plotted. 4. Effect of external cell concentrations For a specific Na-K pair of channels, their internal pa- rameter values, including concentrations, are fixed, so the most viable possibility of some external control is to change the external (out cell) Na and K concentrations. Na+ out concentration: We expect that the height of the spike follows the Na concentration because the Na Nernst potential also increases as ln[ρN a(out)/ρN a(in)]. Moreover the period of the oscillations is decreased be- cause of the leak increase. Numerical simulation for differ- ent Na external concentrations were conducted to confirm [Naout] = 0.415 M [Naout] = 0.498 M [Naout] = 0.622 M 0.25 0.2 ) T ( P 0.15 0.1 0.05 0 0 2 4 6 10 8 12 T (ms) 14 16 18 20 Figure 8: Distribution P (T ) of periods for Naout concentrations of Table 3, which are indicated in the inset. Periods are shorter for larger Naout concentration, due to an increase of Na leak. ρN a(out) (M) Number of oscillations Mean period hT i (ms) Std dev. σT (ms) 0.415 220 9.10 5.38 0.498 274 7.33 3.33 0.622 356 5.65 2.14 Table 3: Mean and variance of spike periods for different external Na concentrations, calculated for a simulation time span of 2025 ms. this prediction. The distribution of periods can be seen in Fig. 8 and mean and variance values in Table 3. K+ out concentration: We have seen in simulations that the period is rather insensitive to moderate variations of the external values of K concentration. When increasing it further a new effect appears. For higher values of Kout con- centration the difference with the internal concentration is lower, and then with the channel K open the membrane polarization is lower (i.e. the potential reach smaller neg- ative values), as it can be seen in Fig. 9. As a result the K channel has less tendency to close. Then, for larger time spans, the system remains in a state with the Na channel closed and the K channel open. In such state it is nec- essary a larger fluctuation to induce the closing of the K channel and initiate the depolarization of the membrane that will produce the spike (Fig.9b). This behavior can be seen in more detail in Fig. 10 for a [Kout] = 0.415 M. Note that the periodicity of the spikes, given until now by the time scale of the Na leak, is now broken. Instead, to this time it is now added the waiting time for the stochastic closing of the K channel. This introduces a new source of stochasticity. For even larger Kout concentrations the effect is stronger and the polarization of the membrane is weaker. As a con- sequence there are very few events of closing of the K gate. In this case some peaks of the membrane potential can be observed due to strong fluctuations, but now not neces- sarily associated with closings of the K channel (Fig. 9c). This is then an excitable regime controlled by fluctuations. In this regime the apparition of peaks is expected to be a stochastic poisson process, with an exponential distribu- 5 (a) (b) (c) 50 0 ) V m ( V ∆ -50 -100 225 250 275 225 250 275 t (ms) 225 250 275 Figure 9: Three typical time evolutions of membrane potential for different Kout concentrations. Periodicity is lost for higher Kout concentration. [Kout] = (a) 0.208 M, (b) 0.415 M, (c) 0.830 M. The red dashed line indicates the Nernst potential corresponding to each concentration. 2 Y , 1 Y 3 Y 1 0.8 0.6 0.4 0.2 0 0.8 0.6 0.4 0.2 0 50 0 ) V m ( V ∆ -50 -100 1000 1025 1050 t (ms) 1075 1100 Figure 10: Gate dynamics for the case (b) of Fig. 9 ([Kout] = 0.415 M. Periodicity is lost and spikes appear at random times as it corre- sponds to an excitable regime. Color code as in Fig. 3 2 1.5 1 0.5 ) 〉 Τ 〈 T ( p / [Kout] = 0.104 M [Kout] = 0.208 M [Kout] = 0.311 M [Kout] = 0.830 M exp(-T/〈Τ〉) 0 0 1 2 T/〈Τ〉 3 4 Figure 11: Scaled distribution of periods, for different values of the external Kout concentration, indicated in the inset. Continuous lines: simulations of cases of Table 4. The dashed line is the exponential distribution, which is characteristic of an excitable regime dominated by fluctuations. tion of waiting times, and then a standard deviation equal- ing the mean value. In our case, for the lower Kout con- centrations the standard deviations are consistently lower than the mean values, indicating the presence of a well de- fined period. However for the larger concentrations both values are very similar, as seen in Table 4, indicating a poisson process. The period distributions, plotted in Fig. 11, clearly show the two limiting behaviors from an oscil- latory regimen to an excitable one. ρout K (M) Nosc hT i (ms) σT (ms) 0.0415 1086 7.45 3.08 0.104 1115 7.25 3.35 0.311 1079 7.50 3.68 0.21 766 10.6 6.30 0.415 324 24.8 19.3 0.83 159 49.9 52.5 Table 4: Spike periods for different external K concentrations, cal- culated for a simulation time span of 8100 ms. Other data from Tables A1 and A2. 5. Conclusions We have shown that the periodic spiking activity of the action potential can be modeled by an unique pair of Na and K ionic channels, by using a minimal semi-microscopic approach [20]. The emergence of regular spikes is caused by the interaction between the channel gates, ionic flows and the membrane potential, within a unified physical framework. The system is autonomous and it enters in the oscillatory regime without any kind of perturbation or energy supply. The only source of energy is the chemical energy of the ionic concentration differences between both sides of the membrane. This is assumed to be controlled by other cell mechanisms, such as molecular ionic pumps, not considered here. The dynamical patterns of the spike events are very similar to those observed in experiments, with a clear stand-by regimen of membrane depolarization and a very narrow pulse. Given the simplicity of the approach it has been possi- ble to establish the relevant steps and involved variables in the periodic firing. Special attention has been paid to the external Na and K concentrations. In particular the K concentration has shown to be responsible for the transi- tion from an oscillatory regime, with a well defined period, to an excitable regime, with stochastic waiting times. The period is controlled by the Na leak, which could be changed to model different neurons by using an internal parameter such as Vd(Y1). In our simulations we have obtained ex- amples of spike intervals from 2 ms to 25 ms, but results are in principle not limited by these values. The approach uses basic physical processes, described with a minimum of degrees of freedom, implemented by using standard formalisms of statistical physics. This al- lows for further possibilities of including other physical elements and mechanisms, which should permit to cover additional specific firing patterns. They could further be combined to deal with more complex neural activity. 6 Acknowledgments This work was supported by the Ministerio de Econo- mia y Competividad (Spain) and FEDER (European Union), under projects FIS2015-66503-C01-P2/P3 and by the Generalitat de Catalunya Projects 2009SGR14 and 2014SGR878. γ Y1 Y2 Y3 1000 4000 4000 Vd V0 Q kB T kB T e 7 7 7 9 10 8 φref mV -35 -35 -15 a 0.2 0.2 0.2 b 7 9 7 xc nm 1.0 3.0 3.0 +12 -8 +10 Appendix: Model details The total potential energy of the system is [20] U (xi, Yj, ∆V ) = Xi Vi(xi, ∆V ) +Xj V (Yj , ∆V ) +Xi,j (A-1) VI (Yj, xi). This potential has three different terms corresponding to the interactions between elements. The first one is the potential seen by the ions inside the channel Vi(xi, ∆V ) = q∆V L (xi − L), 0 < xi < L, (A-2) where ∆V is the potential difference between both sides of the membrane, and L is the length of the channel. The gate variables Yj evolve with the potential V (Y, ∆V ) = V0(cid:2)−a ln(Y (1 − Y )) − b(Y − 0.5)2(cid:3) + Q(∆V − φref)Y, (A-3) where the first part represents the bistable internal struc- ture. With a ≪ b it is a simple potential with two well minima at the closed (Yj ∼ 0) and open (Yj ∼ 1) states. In the last term the parameter Q is the effective charge of the gate sensor and φref is the reference potential that determines the ∆V value at which both states are equally probable. The values for the different parameters (Table A2) are characteristic of each specific channel and are cho- sen to enter into the experimental scales. γA particle friction γB particle friction KBT L channel length A channel section cA 0 (in) cB 0 (in) CM effective capacity 2 µs meV/nm2 200 µs meV/nm2 25 meV 4 nm 4 nm2 4.15 mM 8.30 M 1.25 charges/mV Table A1: Physical fixed control parameter values used in the simu- lations. Both ions have a positive charge q = +1 e. Finally, the effect of the collisions of ions with the gate is modeled by the potential energy VI (Y, xi) = Vdf (Y ) exp(cid:18)− (xi − xc)2 2σ2 (cid:19) , (A-4) where xc is the position of the center of the gate inside the channel, and σ is its width. The height of the barrier is 7 Table A2: Gates parameters of the Na and K channels. Units for γ are µs meV/nm2 and σ = 0.283 nm. modulated by the function f (Y ), depending on the state (open or close) of the gate, with a maximum value Vd. For the envelope modulating function the expression f (Y ) = (1 + cos πY )/2 is used which has the values f (0) = 1 for the close state, and f (1) = 0 for the open state. The parameters of the model are indicated in Tables A1-A2. References [1] B. Hille, Ion Channels of Excitable Membranes. Sinauer, 3rd ed., 2001. [2] C. Hammond, Cellular and Molecular Neurophysiology. Aca- demic Press, 4th ed., 2015. [3] J. A. Fraser and C. L.-H. Huang, "Quantitative techniques for steady-state calculation and dynamic integrated modelling of membrane potential and intracellular ion concentrations," Progress in biophysics and molecular biology, vol. 94, no. 3, pp. 336 -- 372, 2007. [4] A. L. Hodgkin and A. F. Huxley, "A quantitative description of membrane current and its application to conduction and ex- citation in nerve," The Journal of Physiology, vol. 117, no. 4, pp. 500 -- 544, 1952. [5] E. M. Izhikevich., Dynamical Systems in Neuroscience. Cam- bridge: The MIT press, 2010. [6] S. Furini and C. Domene, "K+ and Na+ conduction in selective and nonselective ion channels via molecular dynamics simula- tions," Biophysical journal, vol. 105, no. 8, pp. 1737 -- 1745, 2013. [7] A. Wawrzkiewicz-Ja lowiecka, P. Borys, and Z. J. Grzywna, "Im- pact of geometry changes in the channel pore by the gating movements on the channels conductance," Biochimica et Bio- physica Acta (BBA)-Biomembranes, vol. 1859, no. 3, pp. 446 -- 458, 2017. [8] H. C. Tuckwell and F. Y. Wan, "Time to first spike in stochas- tic Hodgkin -- Huxley systems," Physica A: Statistical Mechanics and its Applications, vol. 351, no. 2, pp. 427 -- 438, 2005. [9] M. Ozer, M. Perc, and M. Uzuntarla, "Controlling the sponta- neous spiking regularity via channel blocking on newman-watts networks of hodgkin-huxley neurons," EPL (Europhysics Let- ters), vol. 86, no. 4, p. 40008, 2009. [10] M. Ozer, M. Uzuntarla, M. Perc, and L. J. Graham, "Spike latency and jitter of neuronal membrane patches with stochas- tic Hodgkin -- Huxley channels," Journal of Theoretical Biology, vol. 261, no. 1, pp. 83 -- 92, 2009. [11] X. Sun, J. Lei, M. Perc, Q. Lu, and S. Lv, "Effects of channel noise on firing coherence of small-world Hodgkin -- Huxley neu- ronal networks," The European Physical Journal B-Condensed Matter and Complex Systems, vol. 79, no. 1, pp. 61 -- 66, 2011. [12] D. Guo, S. Wu, M. Chen, M. Perc, Y. Zhang, J. Ma, Y. Cui, P. Xu, Y. Xia, and D. Yao, "Regulation of irregular neu- ronal firing by autaptic transmission," Scientific Reports, vol. 6, p. 26096, 2016. [13] H. Yu, R. F. Gal´an, J. Wang, Y. Cao, and J. Liu, "Stochastic resonance, coherence resonance, and spike timing reliability of Hodgkin -- Huxley neurons with ion-channel noise," Physica A: Statistical Mechanics and its Applications, vol. 471, pp. 263 -- 275, 2017. [14] B. Maisel and K. Lindenberg, "Channel noise effects on first spike latency of a stochastic hodgkin-huxley neuron," Physical Review E, vol. 95, no. 2, p. 022414, 2017. [15] I. Goychuk and P. Hanggi, "The role of conformational diffusion in ion channel gating," Physica A: Statistical Mechanics and its Applications, vol. 325, no. 1, pp. 9 -- 18, 2003. [16] J. R. Groff, H. DeRemigio, and G. D. Smith, "Markov chain models of ion channels and calcium release sites," in Stochastic methods in neuroscience (C. Laing and G. J. Lord, eds.), pp. 29 -- 64, Oxford University Press, 2009. [17] J. H. Goldwyn and E. Shea-Brown, "The what and where of adding channel noise to the Hodgkin-Huxley equations," PLoS Comput. Biol, vol. 7, no. 11, p. e1002247, 2011. [18] K. Lee and W. Sung, "Ion transport and channel transition in biomembranes," Physica A: Statistical Mechanics and its Ap- plications, vol. 315, no. 1, pp. 79 -- 97, 2002. [19] K. Pawe lek, J. J. Kozak, and Z. Grzywna, "Asynchronous mo- tion and transport of in the KcsA selectivity filter," Physica A: Statistical Mechanics and its Applications, vol. 389, no. 16, pp. 3013 -- 3022, 2010. [20] L. Ram´ırez-Piscina and J. M. Sancho, "Molecular Na-channel excitability from statistical physics," EPL (Europhysics Let- ters), vol. 108, no. 5, p. 50008, 2014. [21] L. Ram´ırez-Piscina and J. M. Sancho, "Physical properties of voltage gated pores," The European Physical Journal B, vol. 91, p. 10, Jan 2018. [22] W. A. Catterall, "Voltage-gated sodium channels at 60: struc- ture, function and pathophysiology," The Journal of physiology, vol. 590, no. 11, pp. 2577 -- 2589, 2012. [23] L. Ram´ırez-Piscina, "Fixed-density boundary conditions in overdamped langevin simulations of diffusion in channels," In preparation. 8
1205.0453
1
1205
2012-05-02T15:04:51
Biased swimming cells do not disperse in pipes as tracers: a population model based on microscale behaviour
[ "physics.bio-ph", "physics.flu-dyn", "q-bio.CB" ]
There is much current interest in modelling suspensions of algae and other micro-organisms for biotechnological exploitation, and many bioreactors are of tubular design. Using generalized Taylor dispersion theory, we develop a population-level swimming-advection-diffusion model for suspensions of micro-organisms in a vertical pipe flow. In particular, a combination of gravitational and viscous torques acting on individual cells can affect their swimming behaviour, which is termed gyrotaxis. This typically leads to local cell drift and diffusion in a suspension of cells. In a flow in a pipe, small amounts of radial drift across streamlines can have a major impact on the effective axial drift and diffusion of the cells. We present a Galerkin method to calculate the local mean swimming velocity and diffusion tensor based on local shear for arbitrary flow rates. This method is validated with asymptotic results obtained in the limits of weak and strong shear. We solve the resultant swimming-advection-diffusion equation using numerical methods for the case of imposed Poiseuille flow and investigate how the flow modifies the dispersion of active swimmers from that of passive scalars. We establish that generalized Taylor dispersion theory predicts an enhancement of gyrotactic focussing in pipe flow with increasing shear strength, in contrast to earlier models. We also show that biased swimming cells may behave very differently to passive tracers, drifting axially at up to twice the rate and diffusing much less.
physics.bio-ph
physics
Biased swimming cells do not disperse in pipes as tracers Biased swimming cells do not disperse in pipes as tracers: a population model based on microscale behaviour R. N. Bearon,1, a) M. A. Bees,2 and O. A. Croze2 1)Department of Mathematical Sciences, University of Liverpool, Peach Street, Liverpool, L69 7ZL, UK 2)School of Mathematics and Statistics, University of Glasgow, Glasgow, Scotland, G12 8QW, UK (Dated: 23 November 2018) There is much current interest in modelling suspensions of algae and other micro- organisms for biotechnological exploitation, and many bioreactors are of tubular design. Using generalized Taylor dispersion theory, we develop a population-level swimming-advection-diffusion model for suspensions of micro-organisms in a vertical pipe flow. In particular, a combination of gravitational and viscous torques acting on individual cells can affect their swimming behaviour, which is termed gyrotaxis. This typically leads to local cell drift and diffusion in a suspension of cells. In a flow in a pipe, small amounts of radial drift across streamlines can have a major impact on the effective axial drift and diffusion of the cells. We present a Galerkin method to calculate the local mean swimming velocity and diffusion tensor based on local shear for arbitrary flow rates. This method is validated with asymptotic results obtained in the limits of weak and strong shear. We solve the resultant swimming-advection- diffusion equation using numerical methods for the case of imposed Poiseuille flow and investigate how the flow modifies the dispersion of active swimmers from that of passive scalars. We establish that generalized Taylor dispersion theory predicts an enhancement of gyrotactic focussing in pipe flow with increasing shear strength, in contrast to earlier models. We also show that biased swimming cells may behave very differently to passive tracers, drifting axially at up to twice the rate and diffusing much less. PACS numbers: 47.63.Gd Swimming microorganisms; 47.57.E- Suspensions Keywords: Algae; dispersion; swimming; pipe flow a)Corresponding author [email protected] 1 Biased swimming cells do not disperse in pipes as tracers I. INTRODUCTION Swimming micro-organisms, such as algae and bacteria, have their own agenda; selec- tive pressures lead cells to adopt strategies to optimize a combination of environmental conditions, such as illumination, nutrients or the exchange of genetic material. This can sig- nificantly impact the behaviour of suspensions of swimming micro-organisms, particularly in flows where biased motion across streamlines can lead to rapid transport. For example, various algae are gravitactic, that is they swim upwards on average in still fluid which can be beneficial for reaching regions of optimal light. For some species this is due to being bottom-heavy - the centre of gravity for these cells is offset from the centre of buoyancy, and the combination of the effects of gravity with the buoyancy force gives rise to a gravitational torque which serves to reorient the cell allowing it to swim upwards - whereas in others sedimentary torques lead to similar behaviour1. However, in shear flow the cells may be re- oriented from the vertical due to viscous torques2. For a vertical pipe containing downwelling fluid, gravitactic cells can accumulate near the centre3, a phenomenon known as gyrotactic focussing. As recently predicted theoretically by Bees and Croze 4, such a modification of the spatial distribution of algae in tubes alters significantly the effective axial dispersion of the cells. There is much current interest in employing micro-organisms for biotechnological pur- poses, from the production of biofuels5,6, such as hydrogen, biomass or lipids, to high-value products, such as β-carotene. Cells are grown either extensively on low value land or in- tensively to optimize growth. Intensive culture systems typically consist of arrays of tubes (vertical, horizontal or helical) and aim to maximize light and nutrient uptake. Bioreactors may be pumped or bubbled, in turbulent or laminar regimes. However, energy input may be energy wasted; efficient bioreactor designs might aim to make use of the swimming motion of the cells themselves, or accommodate the fact that swimming micro-organisms (where drift across streamlines is more important than axial motion) and nutrients are likely to drift and diffuse at different rates along the tubes. In a still fluid, the swimming behaviour of individual gyrotactic phytoplankton has been usefully described as a biased random walk: the cell orientation is assumed to be a ran- dom variable that undergoes diffusion with drift7. At the population-level the dynamics can be modelled with a swimming-diffusion equation for the cell concentration, where the cells 2 Biased swimming cells do not disperse in pipes as tracers swim in a preferred direction at a mean velocity and diffuse with an anisotropic diffusion ten- sor that represents the random component of swimming8. Extending such population-level models to incorporate the effects of ambient flow is non-trivial. Although the orientation distribution and resultant mean swimming velocity of such cells in unbounded homogeneous shear flow has previously been computed9,10, the resultant diffusion tensor is more compli- cated. For homogeneous shear flow, subject to certain constraints on the form of the flow, Hill and Bees 11 and Manela and Frankel 12 calculated expressions for the diffusion tensor using the theory of generalized Taylor dispersion (GTD). Because of its account of shear- induced correlations in cell position, GTD is a more rational account over earlier approaches based on an orientation only description using a Fokker-Planck equation and diffusion tensor estimate (FP)13. Bearon, Hazel, and Thorn 14 compared two-dimensional individual-based simulations of swimming micro-organisms with swimming-advection-diffusion models for the whole popu- lation in situations where the flow is not homogeneous, that is in flows in which the cells can experience a range of shear environments. Using GTD theory to calculate local expressions for the mean swimming direction and diffusion coefficients, the results of the individual and population models were generally in good agreement and were able to successfully predict the phenomena of gyrotactic focussing. However, this work was restricted to two-dimensions; both the swimming motions and velocity field were confined to a vertical plane. Here, we consider axisymmetric pipe flow, which locally can be described by planar shear, and consider swimming motions which are allowed to be fully three-dimensional. First, we develop a population-level swimming-advection-diffusion model where the mean swimming velocity and diffusion tensor are based on the local shear. Next, a Galerkin method is presented for calculating the mean swimming velocity and diffusion tensor based on the local shear, and asymptotic results are obtained in the limits of weak and strong shear. The resultant swimming-advection-diffusion equation is then solved numerically for the case of imposed Poiseuille flow. We contrast the GTD results with the FP approach. Finally, we investigate how the flow modifies qualitatively and quantitatively the dispersion of active swimmers from that of a passive scalar. This paper represents an important link study that will facilitate the comparison of the exact long-time theoretical results of Bees and Croze 4 and the forthcoming experimental results by the authors on the transient dynamics. 3 Biased swimming cells do not disperse in pipes as tracers II. MATHEMATICAL MODEL A. Vertical pipe flow Consider axisymmetric fluid flow with velocity u through a vertical tube of circular cross- section, radius a, with axis parallel to the z-axis pointing in the downwards direction, such that u = u(r)ez = U (1 + χ(r/a))ez. (1) Here, U is the mean flow speed, U χ is the variation of the flow speed relative to the mean, r is the radial distance from the centre of the tube and (er, eψ, ez) are right-handed orthonormal unit vectors that define the cylindrical co-ordinates. For flow subject to a uniform pressure gradient and no-slip boundary conditions on the walls, we have simple Poiseuille flow, χ(r) = 1− 2r2. In the fully coupled problem, where the negative buoyancy of the cells modifies the flow, χ(r) must be determined, as in Bees and Croze 4. A population-level model for gyrotactic micro-organisms in homogeneous shear flow has previously been derived based on generalized Taylor dispersion theory11,12 (GTD). Specifi- cally, for particular types of flow and on timescales long compared to 1/dr, where dr is the rotational diffusivity due to the intrinsic randomness in cell swimming, the cell concentration n(x, t) was shown to satisfy a swimming-advection-diffusion equation of the form (cid:20) (cid:21) D.∇xn + ∇x. ∂n ∂t (u + Vsq) n − V 2 s dr = 0, (2) where Vs is the constant cell swimming speed, and q and D are the non-dimensional mean cell swimming direction and diffusion tensor, respectively. Explicit expressions for q and D as a function of the local shear strength will be given in section II B. Furthermore, Bearon, Hazel, and Thorn 14 show that this population-level approach is a good approxi- mation for flow fields more general than homogeneous shear. Therefore, we shall use (2) to describe the cell concentration in a pipe flow with non-homogeneous shear. To solve the swimming-advection-diffusion equation numerically, it is convenient to non-dimensionalize lengths based on the pipe radius, a, and non-dimensionalize time on a2dr/V 2 s , a characteris- tic timescale for diffusion across the pipe. This reveals two non-dimensional parameters in the problem: the P´eclet number which is given by P e = U adr 2 , Vs 4 (3) Biased swimming cells do not disperse in pipes as tracers and β, the ratio of pipe radius to a typical correlation length-scale of the random walk in the absence of bias, defined as adr Vs s d−1 tion (2) in non-dimensional form thus becomes An alternative interpretation of β = aVs/(V 2 β = . (4) r ) is as a 'swimming P´eclet number'. Equa- + ∇x. [(P e[1 + χ(r)]ez + βq)n − D.∇xn] = 0. ∂n ∂t (5) B. Generalized Taylor dispersion The shear in the pipe flow given by (1) can be locally described as a simple shear flow. Specifically, consider a Taylor expansion of the flow field near some reference point R0 which is at radial position r = R0 u(R) ≈ u(R0) + (R − R0).er χ(cid:48)(R0/a)ez. U a (6) We consider local co-ordinates relative to an origin located at R0 such that k is pointing vertically upwards and (i, j, k) form a right-handed orthonormal set of unit vectors so that i = er, j = −eψ, k = −ez. (7) Defining the local position co-ordinate, R − R0 = ξi + ηj + ζk, the flow field can then be written locally as simple shear, such that u(R) = u(R0) + Gξk, (8) where the shear strength G is given by − U a χ(cid:48). With this choice of co-ordinates, the velocity gradient tensor, G, defined such that u(R) = u(R0) + (R − R0).G, has the simple form Gij = Gδi1δj3. The mean swimming direction, q, and non-dimensional diffusion tensor D can be written as integrals over cell orientation, p, in the form11,12 (cid:90) (cid:90) q = D = p p pf (p)dp, [bp + 2σ f (p) 5 bb. G]symdp. (9) (10) Biased swimming cells do not disperse in pipes as tracers Here []sym denotes the symmetric part of the tensor, G = ik, and σ is a non-dimensional measure of the shear, defined as σ = G 2dr = − P e 2β2 χ(cid:48). (11) We note that σ varies with r because the shear varies across the radius of the tube. However, in the theory of GTD the shear is assumed locally homogeneous, and so we calculate local expressions for the mean swimming and diffusion based on the local value of σ. The equilibrium orientation, f (p), and vector b(p) satisfy11,12 Lf = 0, Lb − 2σb. G = f (p)(p − q), subject to the integral constraints(cid:90) (cid:90) bdp = 0. f dp = 1, Here, the linear operator L for a spherical swimming cell is defined by Lf = ∇p.((λ(k − (k.p)p) − σj ∧ p)f − ∇pf ), p p (12) (13) (14) (15) the gyrotactic bias in swimming direction is represented by the non-dimensional parameter λ = 1 2drB , (16) and B = µα⊥/2hρg is the gyrotactic reorientation time scale, where h is the distance between an average cell's centre-of-mass and centre-of-buoyancy, α⊥ is the dimensionless resistance coefficient for rotation about an axis perpendicular to p, µ and ρ are the fluid viscosity and density respectively, and g is the magnitude of the gravitational force. To summarize, the non-dimensional mean swimming velocity and diffusion tensor are given as functions of two non-dimensional parameters: λ, which only depends on proper- ties of the cell, and σ, which quantifies the strength of the shear. See equations (9-16). Furthermore, for the pipe flow considered in the previous section, we can express σ as a simple function of the non-dimensional parameters P e, the global P´eclet number, and β, the swimming P´eclet number, and the shear profile χ(cid:48)(r) (Eq. 11). The solution of the governing non-dimensional swimming-advection-diffusion equation (Eq. 5) can therefore be determined by specifying the three non-dimensional parameters λ, P e and β and the non-dimensional flow profile χ(r). 6 Biased swimming cells do not disperse in pipes as tracers When explicit calculations are presented in this paper we have assumed that the flow is Poiseuille, χ(r) = 1 − 2r2, and take λ = 2.2 so as to compare with previous work9,11 based on the algal species C. augustae (wrongly identified as C. nivalis15). In Bearon, Hazel, and Thorn 14, good agreement was found in planar pipe flow between individual based simulations and the population-level model with β = 10 (where the reciprocal of β was defined as  = 0.1 therein). Motivated by the pipe dimensions in experiments currently in progress, we also consider β = 2.34. Note that β represents the ratio of pipe radius to the correlation length- scale of the random walk in the absence of bias. Therefore, when modelling a random walk as a diffusion process, β should be sufficiently large14. However, we hypothesize that this restriction may be relaxed in the case of gyrotactic cells that are well-focussed by the flow along the axis of the tube and only suffer rare collisions with the wall. C. Calculation of mean swimming velocity and diffusion Hill and Bees 11 demonstrate that the GTD equations (12) and (13) for f and b, re- spectively, can in general be solved by expanding in spherical harmonics using a Galerkin method. The method is summarized for the flow employed in this paper in appendix A. To simplify the numerical solution of the swimming-advection-diffusion equation in pipe flow, we fit the rather complex algebraic expressions in σ obtained using the Galerkin method for the mean swimming direction and diffusion tensor with the simpler curves qr(σ) = −σP (σ; ar, br), G (σ) = P (σ; arr Drr G , brr G ), Drz qz(σ) = −P (σ; az, bz), G (σ) = −σP (σ; arz G , brz G ), Dzz G (σ) = P (σ; azz G , bzz G ), (18) (17) where P (σ; a, b) = a0 + a2σ2 + a4σ4 1 + b2σ2 + b4σ4 . (19) The choice of a and b coefficients is described in table I with reference to asymptotic results presented below. Please refer to appendix E for the coefficients of fits to the full Galerkin solution. We also consider results using the simpler estimate for the diffusion tensor, which we describe as the Fokker-Planck approximation (or FP), discussed in detail in appendix D. In figure 1 we see that these simple functions are good approximations for the exact solutions for the mean swimming and diffusion, and it is evident how shear can significantly 7 Biased swimming cells do not disperse in pipes as tracers TABLE I. In order to obtain the simplest functional fits whilst ensuring the asymptotic results are satisfied, the a coefficients are as specified. The free parameters, b2, b4 are obtained through least squared optimization of each fit of velocity or diffusion component against σ. Because we are unable to obtain easily the coefficient of the O(σ) correction to Drz G , in addition we allow arz 0,G to vary. The fit coefficients for λ = 2.2 are given explicitly in appendix E. The subscript G highlights that the results are for generalized Taylor dispersion (GTD), and ** indicates that the parameter is fitted. ar az arr G azz G arz G a0 J1 λ K1 J1 λ2 L1 λ ** a2 2λ 3 br 4 4λ 3 bz 4 d1brr 4,G d4bzz 4,G + d3bzz 2,G d2brz 4,G a4 0 0 0 d3bzz 4,G 0 affect the mean swimming and diffusion. Furthermore, we note how that the calculation of diffusion via the GTD method is qualitatively different to that calculated via the sim- pler Fokker-Planck method (FP). In particular, we note that the components of diffusion approach zero in the limit of large shear using the GTD method, whereas they approach a finite non-zero limit via the FP method (see Hill and Bees 11). To provide confidence in the results from the Galerkin method, we have obtained asymp- totic expressions for σ (cid:28) 1 and σ (cid:29) 1, as described in appendices B and C. Specifically, for σ (cid:28) 1, the mean swimming direction with respect to coordinates (er, eψ, ez) correct to O(σ) is given by q = −( σ λ J1, 0, K1)T , (20) where the quantities J1 and K1 are specified functions of λ, coinciding with the results of Pedley and Kessler 13 using the FP model. However, calculation of the diffusion tensor from (10) reveals the new result that at leading order the diffusion tensor is diagonal with horizontal component Drr = J1 λ , where L1 is also a specified power series in λ (appendix B 2). For λ = 2.2 we have that K1 = 0.57, J1 = 0.45 and L1 = 0.11, see appendix B. There is an O(σ) correction to the off-diagonal term Drz, but λ2 , and vertical component Dzz = L1 8 Biased swimming cells do not disperse in pipes as tracers FIG. 1. Mean swimming and diffusion coefficients as a function of shear, σ, for λ = 2.2. Points are calculated using the Galerkin method, solid lines are functional fits described in the text, and dashed lines are asymptotic results. For diffusion calculations, black lines are for GTD, whereas red (grey) lines indicate the FP estimate. the second term in the definition of the diffusion tensor in (10) does not allow for a simple closed form expression for these components. (This is in contrast to expressions obtained by Pedley and Kessler 13 using the simpler orientation-only FP model with a diffusion estimate proportional to the variance of p.) For σ (cid:29) 1, as in Bees, Hill, and Pedley 9, the mean swimming direction correct to O(1/σ2) is q = −( 2λ 3σ , 0, 4λ 3σ2 )T . (21) Here, using GTD theory, we have the new result that the non-zero coefficients of the diffusion 9 Biased swimming cells do not disperse in pipes as tracers tensor are D =  d1 σ2 0 − d2 σ 0 1 6 − d5 σ2 − d2 σ 0 0 d3 + d4 σ2  , (22) where the quantities d1, d2, d3, d4, d5 are polynomials in λ, given in appendix C. For λ = 2.2, we find that d1 = 0.68, d2 = 0.0060, d3 = 0.0020, d4 = 5.9, d5 = 1.3. The asymptotic results are presented in figure 1, indicating excellent agreement with results from the Galerkin method and demonstrating correspondence with the functional fits described above. III. POPULATION-LEVEL NUMERICAL SIMULATIONS A. Numerical Methods The governing swimming-advection-diffusion equation is solved using a spatially adaptive finite element method as described in Bearon, Hazel, and Thorn 14. The cell concentration, n, is approximated using standard Lagrangian quadratic finite elements and the time derivative is approximated using an implicit second-order, backward difference scheme. The subsequent discrete linear system is assembled using the C++ library oomph-lib16 and solved by a direct solver, SuperLU17. In unsteady simulations, a fixed time-step of dt = 10−3 is used. The results were validated by repeating selected simulations with smaller error tolerances and timesteps. B. Steady gyrotactic focussing First, we seek an equilibrium solution n(r) of equation (5) which represents gyrotactic focussing of cells towards the centre of the pipe. Imposing zero flux on the pipe wall, at r = 1, we have that which we can integrate to obtain n = n0 exp βqrn − Drr dn dr = 0, (cid:19) , (cid:18)(cid:90) βqr Drr dr 10 (23) (24) Biased swimming cells do not disperse in pipes as tracers where the radial components of the mean swimming direction and diffusion tensor, qr and Drr, respectively, are functions of the local shear. In particular, if we take simple Poiseuille flow, χ(r) = 1 − 2r2, we have that σ, the non-dimensional measure of the shear, is given by β2 r. For σ (cid:28) 1, at leading order we have that qr/Drr = −σλ = − 2P eλ σ = − P e β2 r from which we predict the Gaussian distribution 2β2 χ(cid:48) = 2P e (cid:19) (cid:18) −P eλ β r2 n = n0 exp . (25) As demonstrated in figure 1(f), the leading order asymptotic solution qr/Drr = −σλ is an excellent approximation for σ = O(1). It is important to note that the GTD and FP methods yield a qualitative difference in the behaviour of qr/Drr. Example calculations of the equilibrium solution (Eq. 24) are shown in figure 2. For the given values of P e and β, we see that cells undergo gyrotactic focussing, and that the distribution predicted by GTD theory can be well-approximated by the Gaussian distribution (Eq. 25) but shows a marked difference to that predicted by FP theory. Furthermore, whereas we see that GTD predicts an enhancement of gyrotactic focussing with increasing shear strength, the FP approximation predicts a reduction in gyrotactic focussing with increasing shear strength at sufficiently large shear. C. Vertical dispersion Bees and Croze 4 investigated how the average axial dispersion was modified for gyrotac- tic organisms compared with a passive solute. Specifically, using the method of moments and the FP approach they obtained long-time expressions for the vertical drift relative to the mean flow and the effective axial swimming diffusivity as a function of P e and a gyro- tactic parameter. Here, we perform a similar calculation, using simulations and the GTD calculations for the diffusion tensor. We solve numerically the swimming-advection-diffusion equation (5) with initial condition n(r, z, 0) = n0 exp (cid:32) − (cid:18) z − 0.1L (cid:19)2 −(cid:16) r 0.01L 0.5 (cid:33) (cid:17)2 , (26) representing a Gaussian blob of cells centred at z = 0.1L, r = 0. For the simulation domain we take z ∈ (0, L), r ∈ (0, 1). Furthermore, we impose no-flux boundary conditions on the walls r = 1, symmetry around the centreline, and periodic boundary conditions in the 11 Biased swimming cells do not disperse in pipes as tracers FIG. 2. Equilibrium concentration (24) for P e = 20 (solid line) and P e = 50 (dashed line) for swimming parameter β = 2.34. Cell diffusion is calculated using GTD (black) and FP (red/grey) approaches. The dotted lines are the associated Gaussian distributions (25). The solutions are normalized so that there is unit total mass per unit length,(cid:82) 2πrn(r)dr = 1. vertical direction, but take L to be sufficiently large that boundary effects do not influence the vertical distribution. In the results presented in figures 3 and 4, we take P e = 50, L = 1200 and run the simulations for t ∈ [0, 8]. In figure 3 we see example plots of the early concentration distribution as a function of time for both gyrotactic cells and a passive solute. For the passive solute, we take D = 1 6I, and q = 0 in equation (5). As shown in appendix B, this is equivalent to considering mean swimming and diffusion in the absence of gyrotactic bias, λ = 0, and shear, σ = 0. Whereas the gyrotactic cells are focussed towards the centre of the pipe, the passive solute diffuses radially. Following Bees and Croze 4, we quantify dispersion in terms of cross-sectionally averaged axial moments of the concentration distribution. To compute the moments of the distribu- tion, we first translate to a reference frame moving with the mean flow, z = z − P e t. The 12 Biased swimming cells do not disperse in pipes as tracers FIG. 3. Concentration in region z ∈ (0, 600), r ∈ (0, 1) from t = 0 to t = 1 at intervals of δt = 0.1 with P e = 50. Upper plots are for gyrotactic cells with λ = 2.2, β = 10. Lower plots are equivalent results for a passive solute. The colour scale is based on the initial concentration distribution, with red representing the maximal initial concentration at the centre of the blob of cells, and blue zero concentration cross-sectional average, mp(t), of the pth axial moment, cp, is (dropping hats for clarity) (cid:90) (cid:90) cp(r, t) = zpn(r, z, t)dz, p = 0, 1, 2, mp(t) = 2 cp(r, t)rdr, p = 0, 1, 2. (27) (28) The mean and variance, m1 and m2 − m2 solution is normalised so that the total mass is unity, m0 = 1. 1, of the distribution are plotted in figure 4. The From the calculations of the m1 and m2, we then define the axial drift and effective axial 13 Biased swimming cells do not disperse in pipes as tracers FIG. 4. The mean and variance, m1 and m2 − m2 P e = 50. Upper plots are for gyrotactic cells with λ = 2.2, β = 10. Lower plots are equivalent 1 of the distribution as a function of time, t for results for a passive solute. Open circles are results from numerical simulation, solid lines are linear regressions for t ∈ [4, 8]. diffusion to be Λ0 = lim t→∞ De = lim t→∞ d dt 1 2 d dt m1, (m2 − m2 1). (29) (30) As depicted in figure 4, for P e = 50, performing a linear regression over the interval t ∈ [4, 8] we obtain Λ0 = 35.2 for the gyrotactic cells with parameters λ = 2.2, β = 10, compared to the long-time limit of Λ0 = 0 for a passive scalar predicted from classic Taylor dispersion theory. This occurs because gyrotactic cells are focussed towards the centre of the tube where the flow is fastest and, hence, they are transported more rapidly than the 14 Biased swimming cells do not disperse in pipes as tracers mean flow. Noting that D = 1 6I for the passive solute, for P e = 50 the classical Taylor dispersion result predicts that De = 1/6 + 6P e2/48 = 313, which compares well with the numerical calculation of De = 312. For the gyrotactic cells with parameters λ = 2.2, β = 10, we see a much reduced axial dispersion, with an estimate of De = 20.0. As discussed by Bees and Croze 4, this reduction in axial dispersion can be explained due to gyrotactic focussing: by self-concentrating towards the axis of the tube, cells undergo a much reduced sampling of radial space and thus sidestep classical shear-induced Taylor dispersion. Furthermore, preliminary calculations based on equations 6.1 and 6.2 of Bees and Croze 4 using the GTD values for the components of q and D give excellent agreement with these numerical computations. Specifically, the calculations yield Λ0 = 35.2 and De = 20.6 for gyrotactic cells with parameters λ = 2.2, β = 10. IV. DISCUSSION Here, we have considered the spatial distribution of gyrotactic algae in axisymmetric pipe flow. We have computed a population-level swimming-advection-diffusion model where the mean swimming velocity and diffusion tensor are based on the local shear using the theory of generalized Taylor dispersion. We have shown how shear modifies the mean swimming velocity and diffusion tensor and, furthermore, demonstrated how the diffusion tensor differs qualitatively from previous simpler models, such as the "Fokker-Plank" approach for which the diffusion tensor is estimated to be the product of the variance of the orientation distribu- tion and a correlation timescale. We have demonstrated that the shear-induced modification to mean swimming velocity and diffusion results in gyrotactic focussing and have quantified how the axial drift and diffusion of a population of cells is modified from that predicted for a passive scalar. In this paper, we only considered unidirectional coupling between flow field and cell concentration. However, actively swimming cells, that are typically denser than the fluid, will modify the flow field. In a dilute suspension, where direct cell-cell hydrodynamic coupling can be neglected, negatively buoyant cells modify the flow field from Poiseuille flow4, which results in a change in local shear and thus a modification of the mean swimming velocity and diffusion tensor. Furthermore, direct hydrodynamic interactions between cells, and stresses induced by the swimming motions, may also alter the flow field18. 15 Biased swimming cells do not disperse in pipes as tracers Work in progress by the authors aims to incorporate the population-level model derived here from generalized Taylor dispersion in Bees and Croze's4 modification of the classical Taylor-Aris theory in order to predict the axial drift and diffusion. Furthermore, both these predictions of long-time dispersion and the transient results presented in this paper will be compared with experimental observations of axial drift and diffusion in dyed suspensions of the alga Dunaliella salina in vertical tubes subject to imposed flow. Finally, work is in progress by the authors to use direct numerical simulations to study the dispersion of active swimmers in laminar and turbulent flows, comparing statistical measures of dispersion from simulations with analytical predictions using the GTD expressions derived in this paper. ACKNOWLEDGMENTS R.N.B. acknowledges assistance from A.L. Hazel to implement the C++ library oomph-lib. M.A.B. and O.A.C. gratefully acknowledge support from EPSRC (EP/D073398/1) and the Carnegie Trust. Appendix A: Galerkin method To implement the Galerkin method, we follow the approach of Hill and Bees 11 who considered the flow field u = Gζi. Here, for the flow field u = Gξk (see Sec. II A) we further extend the method to establish results for the full positive-definite diffusion tensor in (10). We parameterize cell orientation in terms of spherical-polar co-ordinates (θ, φ) p = sin θ cos φi + sin θ sin φj + cos θk. Note that the direction θ = 0 corresponds to cells directed vertically upwards. Equations (12) and (13) are solved by expanding f and bj, j = 1, 2, 3, in spherical harmonics: ∞(cid:88) ∞(cid:88) n=0 f = bj = n(cid:88) n(cid:88) m=0 Am n cos mφP m n (cos θ), (βm nj cos mφ + γm nj sin mφ)P m n (cos θ). (A1) (A2) n=0 m=0 16 Biased swimming cells do not disperse in pipes as tracers Defining  Am  equations (12) and (13) then yield ∞(cid:88) n(cid:88) n=0 m=0 (cid:8)n(n + 1)F m (cid:9) = −2λ cos θF m nj nj ≡ Rm F m nj(φ)P m n (cos θ), j = 0, 1, 2, 3, Rm nj = n cos(mφ), nj cos(mφ) + γm βm nj sin(mφ), j = 0 j = 1, 2, 3, (A3) (A4) nj + λ sin2 θRm njP m n (cid:48) + σ(cos φ sin θRm njP m n (cid:48) + cot θ sin φRm nj (cid:48) P m n ) 0, (cid:80)∞ (cid:80)∞ (cid:80)∞ n=0 n=0 n=0 (cid:80)n (cid:80)n m=0(sin θ cos φ − (4π/3)A1 (cid:80)n m=0 sin θ sin φF m n0 m=0 {2σF m 1)F m n0, n1 + (cos θ − (4π/3)A0 n0} , 1)F m j = 0, j = 1, j = 2, j = 3, (A5) where primes denote differentiation with respect to the dependent variable. Note that the normalization condition (Eq. 14) requires that A0 we calculate the mean swimming direction to be q = (4π/3)(A1 0j = 0, and from equation (9), 1)T . These equations can be simplified using identities for spherical harmonics so that inner products with other 0 = 1/4π, β0 1, 0, A0 harmonics can be calculated. Finally, the resulting equations can be approximated by trun- cating the above series solutions to give a set of simultaneous equations that may be solved for the coefficients Am n , βm nj and γm nj. The first term for the positive-definite diffusion tensor in equation (10), is given in part by equation (52) of Hill and Bees 11, which only depends on the first few terms in the expansion (i.e. βm 1j, γm 1j, for m = 0, 1). The second term cannot be written in such simple terms but can be approximated directly using all available coefficients. Appendix B: Small σ asymptotics The calculation of the mean swimming direction, q, is the same for both the generalized Taylor dispersion theory and the Fokker-Planck approach11. Hence, we follow Pedley and Kessler 13 to compute f for small vorticity case (note that their small parameter  is related to σ via σ = λ). 17 Biased swimming cells do not disperse in pipes as tracers 1. Calculation of equilibrium distribution, f , and mean swimming, q At leading order, σ = 0, (12) becomes −L0f = 1 sin θ ∂ ∂θ sin θ ∂f ∂θ (cid:18) (cid:19) + 1 sin2 θ ∂2f ∂φ2 + λ sin θ ∂ ∂θ (cid:0)sin2 θf(cid:1) = 0. Looking for a solution independent of φ, we obtain the von Mises distribution, f = f (0)(θ) = µeλ cos θ, (B1) where (14) yields the normalization constant µ = λ/4π sinh λ. From equation (9), the mean swimming velocity is computed to be (cid:90) 2π (cid:90) π q(0) = pf (0)(θ) sin θdθdφ = (0, 0, K1), where K1 = coth λ − 1/λ. For λ = 2.2 we have K1 = 0.57. 0 0 To find the O(σ) correction, put to obtain −L0f (1) = 1 sin θ ∂ ∂θ (cid:18) sin θ (cid:18) ∂f (1) ∂θ (cid:19) (cid:19) 1 sin θ d dθ sin θ dF dθ Defining x = cos θ, and letting f = f (0)(θ) + σf (1), + 1 sin2 θ ∂2f (1) ∂φ2 + λ sin θ ∂ ∂θ (cid:0)sin2 θf (1)(cid:1) = λ cos φ sin θf (0). (cid:0)sin2 θF(cid:1) = λ sin θf (0). (B2) Looking for a solution of the form f (1) = cos φF (θ), we obtain the ODE − F sin2 θ + λ sin θ d dθ F = −µg1(x), ∞(cid:88) n(cid:88) we obtain equation (3.4) of Pedley and Kessler 13, which has a power series solution g1(x) = λnAn(x), An(x) = an,rP 1 r (x), (B3) n=1 r=1 where the P 1 r are the associated Legendre functions and the coeffiicents an,r satisfy r + 2 (r + 1)(2r + 3) + an,r−1 (r − 1) r(2r − 1) + en+1,r r(r + 1) , where an+1,r = −an,r+1 en+1,r = (cid:90) 1 2r + 1 n!2r(r + 1) −1 18 (1 − x2)1/2xnP 1 r (x)dx. Biased swimming cells do not disperse in pipes as tracers The first order correction to the mean swimming is then given by (cid:90) 2π (cid:90) π q(1) = where pf (1) sin θdθdφ = 0 0 J1 = (cid:90) π (cid:90) 2π ∞(cid:88) 0 0 λµ l=0 4π 3 λ2l+1a2l+1,1. p cos φF (θ) sin θdθdφ = (−J1 λ , 0, 0), With λ = 2.2 a calculation using Maple provides J1 = 0.45. Thus the mean swimming correct to O(σ) with respect to i, j, k unit vectors is given by λ J1, 0, K1)T . Recalling the relationship between local and global co-ordinate vectors, q = (− σ i = er, j = −eψ, k = −ez, the mean swimming direction, with respect to er, eψ, ez is (see Eq. 20) q = −( σ λ J1, 0, K1)T . 2. Calculation of b, and diffusion tensor, D. (B4) (B5) At leading order, setting σ = 0 in equation (13) yields (cid:18) −L0b = 1 sin θ ∂ ∂θ By inspection, consider sin θ ∂b ∂θ (cid:0)sin2 θb(cid:1) = f (0)(K1k − p). + 1 sin2 θ ∂2b ∂φ2 + λ sin θ ∂ ∂θ bξ = BH(θ) cos φ bη = BH(θ) sin φ bζ = BV (θ). (cid:19) (cid:19) BH then satisfies the ODE (cid:18) 1 sin θ d dθ sin θ dBH dθ (cid:0)sin2 θBH (cid:1) = − sin θf (0). (B6) − BH sin2 θ + λ sin θ d dθ On comparing equations (B6) and (B2), we can write BH = − 1 λ F. From equation (10), the leading order expression for the non-dimensional horizontal com- ponent of diffusion (Dξξ = Dηη) can thus be written as (cid:90) 2π (cid:90) π 0 0 pξ cos φF (θ) sin θdθdφ = J1 λ2 .(B7) (cid:90) 2π (cid:90) π 0 0 Dξξ = pξ cos φBH(θ) sin θdθdφ = − 1 λ 19 Biased swimming cells do not disperse in pipes as tracers (cid:18) (cid:19) The function BV satisfies the ODE 1 sin θ d dθ sin θ dBV dθ + λ sin θ d dθ (cid:0)sin2 θBV (cid:1) = (K1 − cos θ)f (0). To solve this, as for F , we define x = cos θ, let and seek power series solutions BV = µ λ h1(x), ∞(cid:88) n(cid:88) n=1 r=1 λnBn(x), bn,rP 0 r (x). h1(x) = Bn(x) = By utilizing properties of Legendre polynomials we obtain the following recurrence rela- tionship for the bn,r: where bn+1,r = − bn,r+1 2r + 3 + bn,r−1 2r − 1 + fn+1,r r(r + 1) , fn+1,r = 2r + 1 2n! −1 (x − K1)xnP 0 r (x)dx. (cid:90) 1 The leading order expression for the non-dimensional vertical component of diffusion can (cid:90) 2π (cid:90) π thus be written as where Dζζ = 0 0 L1 = 4π 3 µ ∞(cid:88) n=1 λnbn,1. pζBV (θ) sin θdθdφ = L1 λ , (B8) For λ = 2.2 a computation employing Maple reveals that L1 = 0.11. The off-diagonal terms, Dξζ, etc., are all zero at leading order. When comparing with a passive solute, we note that if we set σ = 0, at leading order in λ we have that µ = 1/4π, J1 = 1 3 λ2a1,1, L1 = λb1,1. 1 3 Noting that a1,1 = b1,1 = 1/2, it is clear that in the limit of λ → 0 the diffusion tensor tends to the isotropic tensor I/6. This result can be obtained directly by considering equations (12-15), which have solution f = 1/4π, b = p/8π. 20 Biased swimming cells do not disperse in pipes as tracers Appendix C: Large σ asymptotics For the calculation in the limit of large σ, it is convenient to follow Manela and Frankel 12 and Brenner and Weissman 19 and define local co-ordinates so that the vorticity vector is in the direction of k: i = er, j = −ez, k = eψ. (C1) Defining the local position co-ordinate, R − R0 = ξi + ηj + ζ k, the flow field can then be written locally as simple shear: u(R) = u(R0) + Gξj, (C2) where the shear strength G is given as before by − U gradient tensor has the simple form Gij = Gδi1δj2. a χ(cid:48). For this flow field, the velocity Writing the orientation vector as p = sin θ cos φi + sin θ sin φj + cos θk we can write the governing equation (12) as Lf = σ ∂f ∂φ − Lsf = 0 (C3) where the linear operator independent of σ is given by Lsf = −λ ∂ ∂θ (cos θ sin θ sin φf ) + sin θ (cid:18) 1 (cid:19)(cid:19) (cid:18)cos φ sin θ f ∂ ∂φ + ∇2 pf and where in sphericals the Laplacian is given by ∇2 pf = 1 sin θ ∂ ∂θ sin θ ∂f ∂θ (cid:19) + 1 sin2 θ ∂2f ∂φ2 . For σ (cid:29) 1, we consider the following perturbation expansions for f and b: (cid:18) 1 σ 1 σ b(1) + 21 (cid:18) 1 (cid:19)2 (cid:19)2 (cid:18) 1 σ σ (cid:32) (cid:32) f = 1 4π b = 1 4π (cid:33) (cid:33) f (0) + f (1) + f (2) + . . . b(0) + b(2) + . . . . Biased swimming cells do not disperse in pipes as tracers 1. Calculation of equilibrium distribution, f , and mean swimming, q Substituting the expansion for f into equation (C3) we obtain an explicit iterative scheme for computing the expansion: ∂f (k+1) ∂φ = Lsf (k). ∂f (0) ∂φ = 0. (C4) At leading order: Hence f (0) = f (0)(θ) subject to(cid:82) π 0 f (0)(θ) sin θdθ = 2. At O( 1 σ ): ∂f (1) ∂φ = Lsf (0)(θ). Because f (1) must be periodic in φ with period 2π, integrating this equation with respect to 0 f (0)(θ) sin θdθ = 2, we obtain the solution φ from 0 to 2π gives: Excluding singular solutions, and given that (cid:82) π sin θ sin θ dθ 1 d dθ (cid:18) (cid:19) df (0)(θ) = 0. f (0) = 1. We can summarize the general iteration algorithm for the terms k ≥ 1: 1. Integrate equation (C4) (cid:90) φ 0 f (k+1) = Lsf (k)dφ + F (k+1)(θ). 2. Impose periodicity of f (k+2) & integral constraint (cid:90) 2π (cid:90) 2π (cid:90) π 0 Lsf (k+1)dφ = 0, f (k+1)dθdφ = 0, 0 0 to determine non-singular solutions for F (k+1)(θ). 22 Biased swimming cells do not disperse in pipes as tracers Specifically the first two terms in the expansions are given: f (1) = −2λ cos φ sin θ, f (2) = 2 3 λ2(1 − 3 cos2 θ) + 4λ sin θ sin φ + 3 2 λ2 cos 2φ sin2 θ. From equation (9), we thus can compute mean swimming at large σ correct to O(1/σ2): (cid:90) 2π (cid:90) 2π (cid:90) 2π φ=0 (cid:90) π (cid:90) π (cid:90) π φ=0 θ=0 θ=0 qξ = qη = qζ = f sin2 θ cos φdθdφ = − 2λ 3σ , f sin2 θ sin φdθdφ = 4λ 3σ2 , f cos θ sin θdθdφ = 0. (C5) (C6) (C7) (C8) (C9) When converting back to the global co-ordinates we note that φ=0 θ=0 i = er, j = −ez, k = eψ. and so with respect to er, eψ, ez unit vectors the mean swimming is given by q = −( 2λ 3σ , 0, 4λ 3σ2 )T . 2. Calculation of b, and diffusion tensor, D. We now apply similar techniques to calculate the vector field b. Substituting the expan- sion for b into equation (13) we obtain the following iterative scheme for computing the expansion: ∂b(k+1) ∂φ − 2b(k+1). G = (4π(p − q)f )(k) + Lsb(k). For the simple shear flow with Gij = δi1δj2, taking the dot product with i yields: =(cid:0)4π(sin θ cos φ − qξ)f(cid:1)(k) + Lsb(k) ξ ∂b(k+1) ξ ∂φ The method follows as for f : (C10) (C11) 1. Integrate equation (C11) (cid:90) φ (cid:0)4π(sin θ cos φ − qξ)f(cid:1)(k) b(k+1) ξ = 0 23 + Lsb(k) ξ dφ + B(k+1) ξ (θ). Biased swimming cells do not disperse in pipes as tracers 2. Impose periodicity of b(k+2) (cid:90) 2π 0 ξ & integral constraint (cid:0)4π(sin θ cos φ − qξ)f(cid:1)(k+1) (cid:90) π (cid:90) 2π + Lsb(k+1) ξ dφ = 0, b(k+1) ξ dθdφ = 0, to determine non-singular solutions for B(k+1) ξ (θ). 0 0 We obtain the following expressions: b(0) ξ = 0, ξ = λ(1 − 3 cos2 θ)/36 + sin θ sin φ, b(1) b(2) ξ = B(21) (θ) sin 2φ. (θ) cos φ + B(22) ξ ξ Taking the dot product of Eq. C10 with j yields: ∂b(k+1) η ∂φ = 2b(k+1) ξ + (4π(sin θ sin φ − qη)f )(k) + Lsb(k) η (C12) b(k+1) η = 2b(k+1) ξ + (4π(sin θ sin φ − qη)f )(k) + Lsb(k) η dφ + B(k+1) η (θ). 1. Integrate Eq. C12 (cid:90) φ 0 (cid:90) 2π 0 2. Impose periodicity of b(k+2) η & integral constraint 2b(k+2) ξ + (4π(sin θ sin φ − qη)f )(k+1) + Lsb(k+1) η dφ = 0, (cid:90) 2π (cid:90) π 0 0 b(k+1) η dθdφ = 0, to determine non-singular solutions for B(k+1) calculation of b(3) η ξ which requires calculation of f (3). These calculations were performed (θ). Note that calculation of b(2) η requires using Maple, and files are available from the authors on request. We obtain the following expressions: η = λ(1 − 3 cos2 θ)/108 b(0) b(1) η = B(11) (θ) cos φ η b(2) η = B(10) η (θ) + B(21) η (θ) sin φ + B(22) η (θ) cos 2φ 24 Biased swimming cells do not disperse in pipes as tracers Taking the dot product of equation (C10) with k yields: ∂b(k+1) ζ ∂φ = 4πf (k) cos φ + Lsb(k) ζ , (C13) which when combined with periodicity and the integral constrain yield the following expres- sions: cos θ 1 b(0) ζ = 2 ζ = −3 b(1) 4 b(2) ζ = B(10) ζ λ sin(2θ) cos φ (θ) + B(21) ζ (θ) sin φ + B(22) ζ (θ) cos 2φ From equation (10) we can now compute the diffusion tensor correct to O(1/σ2) : Dξξ = Dηη = Dζζ = Dξη = Dηξ = λ2) 1 270 1 σ2 (6 − 2λ2 243 + 2 3 1 σ2 ( λ2 + 2430 − 5 1 6 1 810σ 18σ2 λ2 λ2 − 41λ4 25515 ) (C14) (C15) (C16) (C17) all other entries are zero. When converting back to the global co-ordinates we note that i = er, j = −ez, k = eψ. and so with respect to er, eψ, ez unit vectors the diffusion tensor is given by D = where  d1 σ2 0 − d2 σ  , 0 1 6 − d5 σ2 − d2 σ 0 0 d3 + d4 σ2 λ2 − 41λ4 25515 d1 = d2 = 1 270 + 2 3 λ2 810 λ2 2430 d3 = d4 = 6 − 2λ2 243 d5 = 5 18 λ2. 25 (C18) (C19) (C20) (C21) (C22) (C23) (C24) Biased swimming cells do not disperse in pipes as tracers Appendix D: Fokker-Planck calculation of diffusion For the Fokker-Planck approximation,13 the diffusion tensor non-dimensionalised on V 2 s /dr is given by DF = τ dr (cid:90) (p − q)2f (p)dp, (D1) where τ is a directional correlation time estimated from experimental data. Although the quantity τ may vary with both λ and the shear, σ, for simplicity it is typically assumed to p be independent of the shear. Asymptotic results for the diffusion tensor for weak shear13 (σ (cid:28) 1) and strong shear9 (σ (cid:29) 1) are available. With this choice of non-dimensionalisation and using the notation of this paper, the σ (cid:28) 1 result correct to O(σ2) is given by Drr F = τ dr (D2) where K2 and J2 are specified functions of λ13.The σ (cid:29) 1 result correct to O(1/σ3) is given by F = τ drK2, F = τ dr λ σ, Dzz K1 λ , Drz J2 − K1J1 (cid:19) (cid:18)1 3 − λ2 5σ2 (cid:19) (cid:18)1 3 − 7λ2 45σ2 Drr F = τ dr , Drz F = 0, Dzz F = τ dr . (D3) To compare the Fokker-Planck approximation to the generalized Taylor method we choose τ so that the two alternative calculations for the horizontal component of the diffusion agree when the shear is zero. Specifically, when σ = 0 the generalized Taylor method yields Drr G = J1 λ2 and thus for the horizontal component of diffusion to agree we take τ dr = J1 λK1 For the specific gyrotactic bias λ = 2.2, this yields τ dr = 0.36. Taking this value of τ also . provides a value for the vertical component of diffusion, Dzz only a slight deviation from the generalized Taylor method, Dzz F = τ drK2 = 0.056, which is λ = 0.050. Clearly, by this careful choice of τ , the Fokker-Planck and generalized Taylor dispersion methods should G = L1 agree when the shear is weak. However, as the shear increases we expect the two theories to give diverging predictions because, for example, in the FP approach Drr F approaches 1 3τ dr, whereas in the GTD approach, Drr G tends to zero at large shear. As for the generalized Taylor method, we fit simple functions to the curves of diffusion against σ obtained with the Fokker-Planck method9,13. The specific functions were given by Drr F (σ) = P (σ; arr F , brr F ), Drz F (σ) = −σP (σ; arz F , brz F ) Dzz F (σ) = P (σ; azz F , bzz F ). (D4) where the rational function P (σ; a, b) is defined by equation 19 and the choice of a coefficients is described in table II. 26 Biased swimming cells do not disperse in pipes as tracers TABLE II. In order to obtain the simplest functional fits whilst ensuring the asymptotic results are satisfied, the a coefficients are as specified. arr F azz F arz F a0 J1 λ2 K2J1 K1λ (K1J1−J2)J1 K1λ2 a2 a4 − J1λ − 7J1λ 5K1 45K1 4,F + J1 brr 4,F + J1 bzz 3K1λ brr 3K1λ bzz 2,F 2,F 4,F J1 3K1λ brr J1 3K1λ brr 4,F 0 0 Appendix E: Coefficients of functional fits The fit coefficients for λ = 2.2 for the mean swimming and diffusion are given by: b2 b4 a0 a2 a4 0 0 ar 1.74 × 10−1 1.27 × 10−2 2.05 × 10−1 1.86 × 10−2 1.75 × 10−1 1.25 × 10−2 5.7 × 10−1 3.66 × 10−2 az 1.19 × 10−1 1.63 × 10−4 G 9.30 × 10−2 1.11 × 10−4 arr G 5.00 × 10−2 1.11 × 10−1 3.71 × 10−5 1.01 × 10−1 1.86 × 10−2 azz 2.81 × 10−1 2.62 × 10−2 G 9.17 × 10−2 1.56 × 10−4 arz F 9.30 × 10−2 5.73 × 10−4 1.85 × 10−3 4.96 × 10−2 1.54 × 10−2 arr F 5.60 × 10−2 3.23 × 10−2 1.70 × 10−5 2.70 × 10−1 1.42 × 10−4 azz F 1.58 × 10−2 9.61 × 10−2 7.88 × 10−2 arz 0 0 0 0 REFERENCES 1A. M. Roberts, "Mechanisms of gravitaxis in Chlamydomonas ," Biol. Bull. 210, 78 -- 80 (2006). 2T. J. Pedley and J. O. Kessler, "Hydrodynamic phenomena in suspensions of swimming microorganisms," Ann. Rev. Fluid Mech. 24, 313 -- 358 (1992). 3J. O. Kessler, "Hydrodynamic focusing of motile algal cells," Nature 313, 218 -- 220 (1985). 4M. A. Bees and O. A. Croze, "Dispersion of biased swimming micro-organisms in a fluid flowing through a tube ," Proc. Roy. Soc. Lon. Ser. B 466, 2057 -- 2077 (2010). 5A. Melis and T. Happe, "Hydrogen production. Green algae as a source of energy," Plant Physiol. 127, 740 -- 748 (2001). 6Y. Chisti, "Biodiesel from microalgae," Biotech. Adv. 25, 294 -- 306 (2007). 27 Biased swimming cells do not disperse in pipes as tracers 7N. A. Hill and D. P. Hader, "A biased random walk model for the trajectories of swimming micro-organisms," J. Theor. Biol. 186, 503 -- 526 (1997). 8R. N. Bearon and D. Grunbaum, "From individual behaviour to population models: A case study using swimming algae," J. Theor. Biol. 251, 679 -- 697 (2008). 9M. A. Bees, N. A. Hill, and T. J. Pedley, "Analytical approximations for the orientation distribution of small dipolar particles in steady shear flows," J. Math. Biol. 36, 269 -- 298 (1998). 10Y. Almog and I. Frankel, "Rheology of dilute suspensions of Brownian dipolar axisymmet- ric particles," J. Fluid. Mech. 366, 289 -- 310 (1998). 11N. A. Hill and M. A. Bees, "Taylor dispersion of gyrotactic swimming micro-organisms in a linear flow," Phys. Fluids 14, 2598 -- 2605 (2002). 12A. Manela and I. Frankel, "Generalized Taylor dispersion in suspensions of gyrotactic swimming micro-organisms," J. Fluid Mech. 490, 99 -- 127 (2003). 13T. J. Pedley and J. O. Kessler, "A new continuum model for suspensions of gyrotactic microorganisms," J. Fluid Mech. 212, 155 -- 182 (1990). 14R. N. Bearon, A. L. Hazel, and G. J. Thorn, "The spatial distribution of gyrotactic swimming micro-organisms in laminar flow fields," J. Fluid Mech. 680, 602 -- 635 (2011). 15O. A. Croze, E. E. Ashraf, and M. A. Bees, "Sheared bioconvection in a horizontal tube ," Phys. Biol. 7, 046001 (2010). 16M. Heil and A. L. Hazel, "oomph-lib -- An object-oriented multi-physics finite-element library in fluid structure interaction," in Lecture notes on computational science and en- gineering, edited by M. Schafer and H.-J. Bungartz (Springer-Verlag, 2006) pp. 19 -- 49. 17J. W. Demmel, S. C. Eisenstat, J. R. Gilbert, X. S. Li, and J. W. H. Liu, "A supernodal approach to sparse partial pivoting." SIAM J. Matrix Anal. App. 20, 720 -- 755 (1999). 18T. Ishikawa, "Suspension biomechanics of swimming microbes," J. Roy. Soc. Int. 6, 815 -- 834 (2009). 19H. Brenner and M. H. Weissman, "Rheology of a dilute suspension of dipolar spherical- particles in an external field .II. Effects of rotary Brownian motion," J. Colloid. Interf. Sci. 41, 499 -- 531 (1972). 20M. A. Bees and N. A. Hill, "Linear bioconvection in a suspension of randomly swimming, gyrotactic micro-organisms," Phys. Fluids 10, 1864 -- 1881 (1998). 28
1703.10409
2
1703
2017-04-15T09:07:08
First Passage Time in Computation by Tape-Copying Turing Machines: Slippage of Nascent Tape
[ "physics.bio-ph", "cond-mat.stat-mech", "physics.chem-ph", "q-bio.SC" ]
Transcription of the genetic message encoded chemically in the sequence of the DNA template is carried out by a molecular machine called RNA polymerase (RNAP). Backward or forward slippage of the nascent RNA with respect to the DNA template strand give rise to a transcript that is, respectively, longer or shorter than the corresponding template. We model a RNAP as a "Tape-copying Turing machine" (TCTM) where the DNA template is the input tape while the nascent RNA strand is the output tape. Although the TCTM always steps forward the process is assumed to be stochastic that has a probability of occurrence per unit time. The time taken by a TCTM for each single successful forward stepping on the input tape, during which the output tape suffers lengthening or shortening by $n$ units because of backward or forward slippage, is a random variable; we report some of the statistical characteristics of this time by using the formalism for calculation of the distributions of {\it first-passage time}. The results are likely to find applications in the analysis of experimental data on "programmed" transcriptional error caused by transcriptional slippage which is a mode of "recoding" of genetic information.
physics.bio-ph
physics
First Passage Time in Computation by Tape-Copying Turing Machines: Slippage of Nascent Tape Soumendu Ghosh,1 Shubhadeep Patra,2 and Debashish Chowdhury∗1 1Department of Physics, Indian Institute of Technology Kanpur, 208016 2ISERC, Visva-Bharati, Santiniketan 731235 Transcription of the genetic message encoded chemically in the sequence of the DNA template is carried out by a molecular machine called RNA polymerase (RNAP). Backward or forward slippage of the nascent RNA with respect to the DNA template strand give rise to a transcript that is, respectively, longer or shorter than the corresponding template. We model a RNAP as a "Tape- copying Turing machine" (TCTM) where the DNA template is the input tape while the nascent RNA strand is the output tape. Although the TCTM always steps forward the process is assumed to be stochastic that has a probability of occurrence per unit time. The time taken by a TCTM for each single successful forward stepping on the input tape, during which the output tape suffers lengthening or shortening by n units because of backward or forward slippage, is a random variable; we report some of the statistical characteristics of this time by using the formalism for calculation of the distributions of first-passage time. The results are likely to find applications in the analysis of experimental data on "programmed" transcriptional error caused by transcriptional slippage which is a mode of "recoding" of genetic information. I. INTRODUCTION Time taken by a system to reach one specific state for the first time starting from another specified initial state is defined as the corresponding first-passage time (FPT). In case of stochastic processes [1] FPT is a random variable and the distribution of First-passage times (DFPT) is used for quantitative statistical characterization of the transition. DFPT has been calculated for wide varieties of physical processes in nonliving as well as in living systems [2 -- 6]. Markov processes [7], which are memoryless stochastic processes, are often good approximations for real stochastic phenomena. For mathematical formulation of these processes one defines the probabilities of all the discrete states of the system and prescribes the rates for all the allowed transitions between these states; the time evolution of the probabilities are described mathematically in terms of master equations. Operations of almost all types of molecular machines and devices have been formulated in terms of master equations [8 -- 10]. DFPT calculated from the master equations for the kinetics of these machines often correspond to physical quantities that are experimentally measurable. For example, the duration of dwell of a molecular motor at a given position on its track is a FPT. The dwell time distribution of various types of molecular motors have received attention in the past [11 -- 17]. In this paper we calculate DFPT for a class of stochastic processes in a simplified theoretical model of operation of a molecular machine called RNA polymerase (RNAP) [18]. The biologically motivated, but highly simplified, model of molecular machine that we analyze here can be interpreted as a physical realization of a Turing machine [19] which is an idealized device conceptualized for abstract 'computation' [20]. In the simplest formulation, a finite set of discrete states are assigned to the machine. It can move forward and backward on a tape in discrete steps; the step size being the size of the boxes marked on the tape. Each box on the tape stores a digit. The head of the Turing machine can read the digit stored in the box at its current location. The Turing machine reads this 'input' and the result of its 'computation' is an 'output' digit and a concomitant transition of state of the machine according to the fixed set of rules (algorithm) prescribed in the beginning. A single DNA strand serves as the tape for a RNAP and the sequence of nucleotides are the analogues of sequence of digits stored on the tape. The biochemical and conformational states of the RNAP are the counterparts of the internal states of a Turing machine. However, in contrast to output digits of a Turing machine the output of the 'computation' (which biologists refer to as transcription) by the RNAP is another tape called RNA. Thus, a RNAP is a 'tape-copying Turing machine' (TCTM) [21 -- 23]. In order to adopt unambiguous terminology we refer to the tape on which input data are engraved as the 'input' tape while the incomplete output tape during ongoing elongation is also referred to as the 'nascent' tape. In our model the TCTM ia assumed to move always forward by a single step on the input tape. But its forward stepping is assumed to be a stochastic process with a probability of occurrence per unit time. Each forward stepping of the TCTM completes one step of the computation. Completion of L such steps of computation by the RNAP in ∗ Corresponding author; e-mail: [email protected] 7 1 0 2 r p A 5 1 ] h p - o i b . s c i s y h p [ 2 v 9 0 4 0 1 . 3 0 7 1 : v i X r a 2 a perfect error-free manner on an input DNA template tape of length L would result in an output RNA tape whose length (both in the units of nucleotides) is also L, i.e., exactly equal to that of the input DNA tape. But, as we explain in the next section, a shorter/longer RNA tape would result from an erroneous forward/backward 'slippage' of the nascent output tape at any step of the sequence of computations by the RNAP [24]. One of the fundamental questions is the dwell time of a RNAP at a nucleotide on the template DNA during which the nascent transcript associated with it suffers slippages causing it to become longer (or shorter) by n nucleotides. In the terminology of computation, this is equivalent to the dwell time of the reading tip of the TCTP at a single position on the input tape during which the nascent output tape becomes longer (or shorter) by n units because of its slippage. This time is intrinsically stochastic and can be formulated as a first-passage time; we calculate its statistical distribution in this paper. In the past, the dwell time distribution, which is a DFPT, has been calculated for simple models of RNAP [13]. However, those distributions characterize the stepping pattern of the RNAP, i.e., the statistics of the time taken by the Turing machine to perform each step of computation that produces an error-free output. In contrast, in this paper, we focus on the erroneous process of 'slippage' of the output RNA tape and calculate a new DFPT that characterize the statistical features of this slippage process. More precisely, the type of slippage we consider here take place only in so-called elongation stage of transcription when the nascent RNA tape gets elongated in each round of elementary computation by the RNAP machine. II. BIOLOGICAL MOTIVATION: SLIPPAGE OF NASCENT RNA The phenomenon of slippage of the nascent RNA during transcription (i.e., its synthesis by a RNAP machine), keeping the position of the machine fixed on the template DNA is depicted schematically in Fig.1. In the four-letter alphabet used for encoding genetic message, the letter 'A' on the DNA template is complementary to the letter 'U' on the RNA transcript. The grip of the RNAP on the template DNA as well as on the nascent RNA transcript are important for error-free normal transcription. Loosening of its grip on the template DNA can lead to its backward - or forward-slippage along the template DNA; the cause of this phenomenon has received lot of attention over the last decade [25]. In contrast, in this paper we focus on the consequence of lose grip of the RNAP on the nascent transcript that can cause backward or forward slippage of the transcript (see Fig.1) while the position of the RNAP on the template DNA remains unaltered. Such transcriptional errors are believed to be 'programmed', rather than random, and constitute one specific mode of 'recoding' of the genetic message [26]. III. THEORETICAL MODEL: SLIPPAGE OF NASCENT TAPE As stated in the introduction, this model is motivated by the biological phenomenon of RNA slippage during transcription which has been explained briefly in the preceding section. The input tape and the nascent tape in our theoretical model mimic the DNA template and nascent RNA respectively, while the tape-copying Turing machine represents the RNAP. We assume that the input tape passes through a channel inside the reader of the Turing machine. At any given instant of time, the head of the reader (a tip) can read only one of the boxes on the input tape in the segment located inside the channel. Similarly, a segment of the nascent tape, starting from its elongating tip, also remains in the same channel inside the reader. As the nascent tape elongates adding one box at each step of computation it emerges from the other end of the channel. We assume that the channel inside the reader covers several boxes on the input and nascent tapes at a time and that this channel constitutes a slippage prone tract. The slippage is known to occur on a homogeneous sequence of identical nucleotides on the template DNA although on a longer scale the sequence on the DNA is inhomogeneous. Since in this paper we focus exclusively on the slippage process we assume a homogeneous sequence on the input tape. For the sake of simplicity we denote the homogeneous sequence of digits on the template tape by a string of 0 (zero). We implement the complementary base-pairing between the template DNA and nascent RNA by the programming the Turing machine to perform the logical operation XOR (exclusive OR) which gives output 1 (one) whenever the input is 0 (zero) and vice-versa. The instantaneous position of the head of the reader of the TM is denoted by the integer index j; it takes a forward step from j to j + 1 after a successful error-free step of computation. The extra length of the nascent tape caused by the slippage is labelled by an integer index µ that can, in principle, be positive, negative or zero; µ = 0 if the nascent tape suffers no slippage or it suffers equal numbers of forward and backward slippages. In contrast, µ is positive (negative) in case backward (forward) slippage occurs more often than the reverse process (see Fig.4 where the model has been illustrated for the maximum value µ = 2). In other words, µ is the excess or deficiency of boxes ("extra" number of boxes) in the nascent tape as compared to the corresponding template. From now onwards, a TM associated with a nascent tape of "extra" length µ will be referred to as a TM in slippage state µ; a negative value of (a) 3 (b) (c) FIG. 1: A schematic diagram illustrating (a) error-free transcription, without slippage, by a RNAP, (b) backward slippage of nascent RNA, (c) forward slippage of nascent RNA. µ indicates shorter length of the nascent tape as compared to the template tape. The instantaneous state of the TM will be denoted by the pair j, µ. A. First-Passage Time Distributions and first two moments The time taken by the TM to reach, for the first time, the state j + 1, µ, from the state j, 0 is defined as the first-passage time Pµ(t) for the "slippage state" µ; since all the sites on the template are equivalent the site index is not explicitly written in the symbol for the first-passage time. Suppose, for the n-state model, P (n) µ (j, t) denotes the probability, at time t, that the head of the TM is located at the j-th site on its template tape and is in the slippage state µ. µ (t) of the first-passage times is known the corresponding mean first-passage (1) (2) In principle, if the full distribution P (n) times < tµ > can be computed from the definition < t(n) µ > = (cid:90) ∞ 0 t P (n) µ (t) dt while the randomness parameter can be obtained using the definition µ )2 > − < t(n) < (t(n) µ >2 r(n) µ = < t(n) µ >2 AAAAAAAAAAA UUUUUUUUUU U Transloca,on AAAAAAAAAAA UUUUUUUUUU U Incorpora,on UTP AAAAAAAAAAA UUUUUUUUUU U AAAAAAAAAAA UUUUUUUUUU U Incorpora+on UTP Backward Slippage (longer transcript) AAAAAAAAAAA UUUUUUUUUU AAAAAAAAAAA UUUUUUUUUU U UForward Slippage (shorter transcript) 4 (3) FIG. 2: A schematic diagram of the tape copying Turing machine. where (cid:90) ∞ 0 t2 P (n) µ (t) dt < (t(n) µ )2 > = However, most often the set of kinetic equations are too complicated to yield closed form analytical expression for P (n) In such situations the moments of the distribution of first passage time can still be obtained by taking µ (t). appropriate derivatives of P (n) µ (s) if the latter can be calculated in the s-space (Laplace space). For example, the normalised mean first-passage time can be obtained using (cid:12)(cid:12)(cid:12)s=0 (cid:12)(cid:12)(cid:12)s=0 − d µ (s) ds P (n) P (n) µ (s) < t(n) µ > = For the sake of simplicity, we write the master equations only for µ = 0,±1,±2 which correspond to the 5-state kinetic model in Fig. 4. In this case the master equations governing the time evolution of Pµ(j, t) can be written as a matrix differential equation. So, in compact form (for simplicity we omit the j indices) dP(t) dt P(t) = where and = A P(t).   P0(j, t) P+1(j, t) P+2(j, t) P−1(j, t) P−2(j, t) (4) (5) (6) Turing Machine Nascent Tape 0 0 1 0 1 1 0 0 Input Tape 1 1 0 1 0 0 1 1 1 1 1 0 0 (a) 5 (b) (c) FIG. 3: A schematic diagram of transcript slippage. Circle represents the head of the Turing machine. (a) translocation, (b) backward slippage and (c) forward slippage of Turing machine.  A = − (q0 + b+1 + f−1) f+1 b+1 0 f−1 0 −(q+1 + b+2 + f+1) b+2 0 0 0 f+2 −(q+2 + f+2) b−1 0 0 0 0 0 0 0 −(q−1 + b−1 + f−2) b−2 f−2 −(q−2 + b−2)  (7) For the calculation of the first-passage time, we impose the initial conditions Pµ(0) = 0 for all µ except P0 = 1. (8) (9) Carrying out Laplace transform L we get (cid:20) dP(t) (cid:21) = A L P(t) dt s P(s) − P(0) = A P(s) Transloca)on 0 1 Incorpora)on 1 1 1 0 1 0 1 Incorpora+on 1 1 1 0 1 Backward Slippage (longer tape) 0 1 1 0 1 1 Forward Slippage (shorter tape) 6 FIG. 4: A schematic diagram of the kinetic model for the special case where the maximum allowed value of µ is 2 (referred to as the 5-state model). b's and f's are the rates of backward and forward slippages, respectively. q's are the rates of translocation of the TM from j-th to j+1-th position on the input tape. and hence (10) where I is the identity matrix, L indicates the Laplace transform operator and P(s) is the Laplace transform of P(t), i.e., Pµ(s) is the Laplace transform of Pµ(t). After taking inverse Laplace transform of P(s), in principle, one would get P(s) = (sI − A)−1 P(0) P(t) = L−1 [ P(s) ] IV. RESULTS (11) For the sake of simplicity of graphical plots, we present the results here only for the cases, namely b+1 = b+2 = b−1 = b−2 = b , f+1 = f+2 = f−1 = f−2 = f and q0 = q+1 = q+2 = q−1 = q−2 = q. Moreover, from now onwards, the special case b = f will be referred to as the symmetric case while the more general situation b (cid:54)= f will be referred to as the asymmetric case. The analytical expressions that we derive for the 3-state model are given in the main text +1 +1 +2 +2 -2 -2 -1 -1 0 0 b+2 f+2 q0 j j+1 f+1 q+1 q+2 b+1 b-1 b-2 f-1 f-2 q-1 q-2 of this section because these are not only short but also display some systematic trends which can be exploited for physical interpretation. The analytical expressions of the results for 5-state and 7-state models are simple enough to be reproduced in the main text. For all the graphical plots we have taken q = 10 s−1. In each of the figures we draw the curves obtained from the theoretical expressions. 7 A. Results for the general n-state model in the symmetric case We get the generalised form of the probabilities in Laplace space, bµ(b + q + s)λ−µ + (λ − µ − 1)bµ+1(q + s)λ−µ−1 + 3 bµ+2(q + s)λ−µ−2 (cid:124) (cid:123)(cid:122) for λ≥µ (cid:124) (cid:125) (cid:26) (cid:124) for λ−µ≥2 (cid:123)(cid:122) (cid:125) (cid:124) (cid:123)(cid:122) forλ≥1 (cid:18)λ − µ − 1 (cid:19) λ − µ − 3 (cid:123)(cid:122) (cid:125) (cid:125) (cid:124) (cid:124) (cid:123)(cid:122) forλ≥2 for λ−µ≥3 + (λ − 1)nb2(q + s)λ−2 (cid:123)(cid:122) forλ≥3 (cid:125) P (n) µ (s) = (q + s) nbλ + (q + s)λ + nb(q + s)λ−1 (cid:125) +higher order terms (cid:27) +higher order terms (12) where n is the total number of states, µ denotes the slippage state and λ = µmax is the maximum possible slippage state for a given n, i.e., µ ≤ λ = µmax = (n − 1)/2. (cid:17)2 (cid:17)5 n q (cid:16) b (cid:16) b q n n + (nµ + 2) + µ + 1 (cid:17)2 (cid:16) b q + 2 (cid:16) b (cid:16) b (cid:17) (cid:16) b q q q (cid:17) (cid:17)4 (cid:16) b + 1 q n (cid:17)5 term1 − term2 term3 For example, = = = < t(3) < t(3) µ > 0 > < t(5) < t(5) µ > 0 > < t(7) < t(7) µ<λ > 0 > where + n(µ + 4) + (4nµ + 4n + 3) + (4nµ + n + µ + 10) + (nµ + 3µ + 6) (cid:17)3 (cid:16) b q (cid:17)3 (cid:16) b q (cid:17)4 (cid:16) b q (cid:17)2 (cid:16) b (cid:17)2 q (cid:16) b q (cid:17) (cid:16) b q + 3n + (4n + 3) + (2n + 9) + 7 + 1 (13) (14) + µ + 1 (cid:17) (cid:16) b q (cid:27) × + 4 × (15) (16) (17) (18) (19) (20) + 10n + 32n + (39n + 4) + (19n + 24) term1 = term2 = term3 = (2 − µ) + 1 (cid:17)6 (cid:17)6 (cid:17)6 q q n (cid:26) (cid:16) b (cid:110)(cid:16) b (cid:26) (cid:16) b (cid:110)(cid:16) b (cid:26) (cid:16) b (cid:110)(cid:16) b n n q q q + 1 q q q + (cid:16) b (cid:17)5 (cid:16) b (cid:17) (cid:16) b (cid:17)5 (cid:16) b (cid:17)5 (cid:16) b (cid:17) q q + q + 8n (cid:17)3−µ (cid:17)2−µ (cid:17)3−µ + 4n + 1 (3 − µ) + q (cid:16) b (cid:17)4 (cid:111) (cid:17)4 (cid:16) b (2 − µ)2(cid:111) (cid:17) (cid:16) b (cid:16) b (cid:17)4 (cid:111) q q + 10n q (2 − µ) q (cid:16) b (cid:17)3 (cid:17)3 (cid:16) b (cid:17)3 (cid:16) b q q q (cid:16) b (cid:17)2 (cid:17)2 (cid:16) b (cid:16) b q q (cid:17)2 + (3n + 20) (cid:16) b (cid:17) (cid:27) q (cid:16) b q (cid:17) (cid:27) × (cid:16) b (cid:17) q + 18n + (17n + 1) + (7n + 6) + (n + 5) + 1 + 2(5n + 2) + 3(n + 6) + 10 + 1 < t(7) < t(7) λ > 0 > = num den where num = den = (cid:17)6 (cid:17)λ (cid:17)6 (cid:17)λ (cid:26) (cid:110) (cid:26) (cid:110) 7 n 7 n q (cid:16) b (cid:16) b (cid:16) b (cid:16) b q q q + 120 q q + 56 + 126 (cid:16) b (cid:17)4 (cid:16) b (cid:17)5 (cid:17)2 + n(λ − 1)2(cid:16) b (cid:17)4 (cid:17)5 (cid:16) b (cid:16) b (cid:17)2 (cid:16) b + 70 + n(λ − 1) + n + 28 q q q (cid:16) b (cid:17) (cid:16) b + 74 q + nλ q (cid:16) b (cid:17) (cid:17)3 (cid:16) b (cid:111) q + 1 q q (cid:17)3 + (λ + 1) + 55 + 12 (cid:17)2 (cid:16) b (cid:111) (cid:17)2 (cid:16) b q q (cid:16) b (cid:17) (cid:16) b q q (cid:17) (cid:27) × + 1 (cid:27) × + 39 + 10 + 1 Interestingly, < t(n) µ > / < t(n) 0 > exhibits a maximum at a value of b/q that shifts with the variation of n. For example, for the 3-state model the maximum occurs at √ −1 + n − 1 n b q = = 0.138071 8 (21) (22) (23) (24) whereas for the 5-state model the maximum shifts to b q = 0.120274 FIG. 5: < t(n) µ > / < t(n) 0 > are plotted against b/q for the special value n = 7 (i.e., 7-state model). Note that the values of < t(n) µ > / < t(n) 0 > in the two limits b/q → 0 and b/q → ∞ reduce to very simple forms lim q → 0 b lim q →∞ b < t(n) < t(n) < t(n) < t(n) µ > 0 > µ > 0 > = µ + 1 = 1. (25) (26) which have clear physical meaning. In the limit b/q → ∞ the population in all the distinct states corresponding to different µ, for a given fixed j, get rapidly equilibrated (i.e., equally populated) and hence lead to < t(n) µ > / < <t+1>/<t0><t+2>/<t0><t+3>/<t0>��-���-�������������������/�<�μ>/<��> t(n) 0 >= 1 irrespective of µ. In the opposite limit time taken to add each extra unit (i.e., each extra slippage of the nascent tape) adds up causing linear increase of < t(n) 0 > with µ. Finally, the numerical values of b and q for wild type RNAP under physiological conditions [45 -- 49] are normally 0 >(cid:39) µ + 1. However, the rate of slippage, and hence such that b/q is of the order of 10−3 and hence < t(n) the ratio b/q, can increase as much as almost tenfold in mutant RNAP [50, 51]. But, even tenfold increase of b would not be adequate to observe a value of < t(n) 0 > that deviates significantly from the limiting value µ + 1. Therefore, we suggest that, in addition to using the mutant RNAP for budding yeast (Saccharomyces Cerevisiae), the RNAP should also be starved of nucleotides, monomeric subunits of the nascent transcript, so as to lower q to a value comparable to b. Under such experimental conditions the ratios < t(n) µ > / < t(n) 0 > can attain nontrivial values, different from µ + 1, that can be extracted from the exact expressions derived above. µ > / < t(n) µ > / < t(n) µ > / < t(n) 9 V. SUMMARY AND CONCLUSION Fidelity of transcription by a RNA polymerase (RNAP) motor is achieved by the correct positining of its active site at the tip of nascent mRNA. A backward or forward slippage of the nascent transcript with respect to the template DNA strand as well as RNAP motor results in a transcript that is, respectively, longer and shorter than the template. One of the fundamental questions is the time taken for successful transcription of one nucleotide on the template DNA by the RNAP during which the nascent RNA transcript associated with it becomes longer or shorter by n nucleotides because of the net effects of backward and forward slippages. In our model we have treated the RNAP as a Tape- copying Turing Machine (TM) where the DNA template strand is the input tape while the nascent is the output tape. In the terminology of the TM the quantity of interest here is the time taken to complete a single successful computation during which the output tape becomes longer or shorter by n units because of slippage. Identifying this time as a "first-passage time", we have derived exact analytical expression for the distribution of this first-passage time and hence the moments. In two limiting cases the analytical expressions reduce to very simple forms which have interesting physical interpre- tations. In the intermediate regime we discover an interesting trend of variation that may have important implications in the biological context of transcript slippage which is one of the modes of recoding of genetic information [26]. Appendix 11-state model: symmetric case P (11) 0 (s) = (b + q + s)5 + 10b4(q + s) + 25b3(q + s)2 + 18b2(q + s)3 + 4b(q + s)4 (q + s){11b5 + 55b4(q + s) + 77b3(q + s)2 + 44b2(q + s)3 + 11b(q + s)4 + (q + s)5} P (11) 1 (s) = b(b + q + s)4 + 6b4(q + s) + 9b3(q + s)2 + 3b2(q + s)3 (q + s){11b5 + 55b4(q + s) + 77b3(q + s)2 + 44b2(q + s)3 + 11b(q + s)4 + (q + s)5} P (11) 2 (s) = (q + s){11b5 + 55b4(q + s) + 77b3(q + s)2 + 44b2(q + s)3 + 11b(q + s)4 + (q + s)5} b2(b + q + s)3 + 3b4(q + s) + 2b3(q + s)2 P (11) 3 (s) = P (11) 4 (s) = P (11) 5 (s) = (q + s){11b5 + 55b4(q + s) + 77b3(q + s)2 + 44b2(q + s)3 + 11b(q + s)4 + (q + s)5} b3(b + q + s)2 + b4(q + s) (q + s){11b5 + 55b4(q + s) + 77b3(q + s)2 + 44b2(q + s)3 + 11b(q + s)4 + (q + s)5} b4(b + q + s) (q + s){11b5 + 55b4(q + s) + 77b3(q + s)2 + 44b2(q + s)3 + 11b(q + s)4 + (q + s)5} b5 (27) (28) (29) (30) (31) (32) Acknowledgements 10 This work is supported by "Prof. S. Sampath Chair" Professorship (DC) and a J.C. Bose National Fellowship (DC). [1] Bressloff P C 2014 Stochastic Processes in Cell Biology (Springer). [2] Redner S 2001 A Guide to First-Passage Processes (Cambridge University Press, Cambridge) [3] Redner S, Metzler R and Oshanin G 2014 (eds) First-passage Phenomena and their Applications (World Scientific Pub- lishing Co.). [4] Chou T and D'Orsogna M R, in ref.[3]. [5] Iyer-Biswas S and Zilman A 2016 Adv. Chem. Phys. 160 261. [6] A. Godec and R. Metzler, Scientific Reports 6, 20349 (2016). [7] Gillespie D T 1992 Markov Processes: An Introduction for Physical Scientists (Academic Press). [8] Chowdhury D, Biophys. J., 104, 2331 (2013). [9] Chowdhury, D, Phys. Rep., 529, 1 (2013). [10] Kolomeisky A B, Motor Proteins and Molecular Motors. (CRC Press, 2015). [11] Kolomeisky A B, Stukalin E B and Popov A A, Phys. Rev. E 71, 031902 (2005). [12] Bierbaum V and Lipowsky R, PloS One 8(2), e55366. (2013). [13] Tripathi T, Schutz G M and Chowdhury D, J. Stat. Mech: Theory and Expt. P08018 (2009). [14] Sharma A K and Chowdhury D, Phys. Biol. 8, 026005 (2011). [15] Sharma A K and Chowdhury D, Phys. Rev. E 86, 011913 (2012). [16] Sharma A K and Chowdhury D, J. Phys. Condens. Matter 25, 374105 (2013). [17] Kinz-Thompson C D, Sharma A K, Franck J, Gonzalez Jr. R L and Chowdhury D, J. Phys. Chem. B 119, 10999 (2015). [18] Buc H and Strick T, RNA polymerases as molecular motors, (Royal Soc. Chem., 2009). [19] Feynman R P, Feynman Lectures on Computation, (Addison-Wesley, 1996). [20] Minsky M L, Computation: Finite and Infinite Machines, (Prentice-Hall, 1967). [21] C. H. Bennett, Int. J. Theor. Phys. 21, 905 (1982). [22] Mooney R A, Artsimovitch and Landick R, J. Bacteriology 180, 3265 (1998). [23] Sharma A K and Chowdhury D, Biophys. Rev. Lett. 7, 135 (2012). [24] Anikin M, Molodtsov V, Temiakov D and McAllister W T 2010 in: ref.[26]. [25] Zhang J and Landick R, Trends in Biochem. Sci. 41, 293 (2016). [26] Atkins J F and Gesteland R F (eds.), Recoding: Expansion of Decoding Rules Enriches Gene Expression, (Springer, 2010). [27] Schutz G M 2001 in: Phase Transitions and Critical Phenomena, eds. Domb C and Lebowitz J L (Academic Press). [28] Derrida B 1998 Phys. Rep. 301 65. [29] Mallick K 2015 Physica A 418 17. [30] Kolomeisky A 2007 Phys. Rev. Lett. 98 1 [31] Berezhkovskii A M and Bezrukov S M 2005 Chem. Phys. 319 342 [32] Zilman A 2009 Biophys. J. 96 1235. [33] Zilman A, Pearson J, and Bel G 2009 Phys. Rev. Lett. 103 128103. [34] Berezhkovskii A M, Bezrukov S M, and Pustovoit M A 2002 J. Chem. Phys. 116 9952 [35] Muthukumar M 2011 Polymer Translocation (CRC Press, Boca raton, Fl) [36] Muthukumar M 2003 J. Chem. Phys. 118 5174 [37] Macdonald L E, Zhou Y and McAllister W T 1993 J. Mol. Biol. 232 1030 [38] Toulokhonov I and Landick R 2003 Mol. Cell 12 1125 [39] Guo H C and Roberts J W 1990 Biochemistry 29 10702 [40] Harley C B, Lawrie J, Boyer H W and Hedgpeth J 1990 Nucleic. Acids. Res. 18 547 [41] Liu C, Heath L S and Turnbough Jr. C L 1994 Genes Dev. 8 2904 [42] Kolakofsky D, Roux L, Garcin D and Ruigrok R W 2005 J. Gen. Virol. 86 1869 [43] Burch C L, Danaher R J and Stein D C 1997 J. Bacteriol. 179 982 [44] Wagner L A, Weiss R B, Driscoll R, Dunn D S, Gesteland R F 1990 Nucleic Acids Res. 18 3529 [45] Pal M and Luse D S, PNAS, 100, 5700 (2003). [46] Lin Ching-Ping et. al., PLOS ONE, 10, 1371 (2015). [47] Olspert A et. al., EMBO reports, 16, 8 (2015). [48] Olspert A et. al., Nucleic Acids Research, 44, 16 (2016). [49] Parks A R et. al., Nucleic Acids Research, 42, 9 (2014). [50] Strathern J et. al., The Journal of Biological Chemistry , 288, 4 (2013). [51] Zhou Y N et. al., The Journal of Biological Chemistry , 288, 4 (2013).
1501.06887
1
1501
2015-01-27T20:01:58
Structural transitions and energy landscape for Cowpea Chlorotic Mottle Virus capsid mechanics from nanomanipulation in vitro and in silico
[ "physics.bio-ph", "q-bio.BM" ]
Physical properties of capsids of plant and animal viruses are important factors in capsid self-assembly, survival of viruses in the extracellular environment, and their cell infectivity. Virus shells can have applications as nanocontainers and delivery vehicles in biotechnology and medicine. Combined AFM experiments and computational modeling on sub-second timescales of the indentation nanomechanics of Cowpea Chlorotic Mottle Virus (CCMV) capsid show that the capsid's physical properties are dynamic and local characteristics of the structure, which depend on the magnitude and geometry of mechanical input. Surprisingly, under large deformations the CCMV capsid transitions to the collapsed state without substantial local structural alterations. The enthalpy change in this deformation state dH = 11.5 - 12.8 MJ/mol is mostly due to large-amplitude out-of-plane excitations, which contribute to the capsid bending, and the entropy change TdS = 5.1 - 5.8 MJ/mol is mostly due to coherent in-plane rearrangements of protein chains, which result in the capsid stiffening. Dynamic coupling of these modes defines the extent of elasticity and reversibility of capsid mechanical deformation. This emerging picture illuminates how unique physico-chemical properties of protein nanoshells help define their structure and morphology, and determine their viruses' biological function.
physics.bio-ph
physics
Kononova et al. Mechanical Properties of CCMV Capsid Structural Transitions and Energy Landscape for Cowpea Chlorotic Mottle Virus Capsid mechanics from nanomanipulation in vitro and in silico* Olga Kononova,†‡ Joost Snijder,§ Melanie Brasch, ¶ Jeroen Cornelissen, ¶ Ruxandra I. Dima, Kenneth A. Marx, † Gijs J. L. Wuite,§ Wouter H. Roos§* and Valeri Barsegov†‡* †Department of Chemistry, University of Massachusetts, Lowell, MA 01854; ‡Moscow Institute of Physics and Technology, Moscow Region, Russia 141700; §Natuur- en Sterrenkunde and LaserLab, Vrije Universiteit, 1081 HV Amsterdam, The Netherlands; ¶Biomoleculaire Nanotechnology, Universiteit Twente, 7500 AE Enschede, The Netherlands; Department of Chemistry, University of Cincinnati, Cincinnati, OH 45221. Running Title: Mechanical properties of CCMV capsid *Corresponding authors: e-mail: [email protected], tel: 978-934-3661, fax: 978-94-3013 e-mail: [email protected], tel: +31 20 59 83974, fax: +31 20 59 87991 Keywords: Biophysics, GPU computing, Nanoindentation, Proteins, Viral Capsids ABSTRACT Physical properties of capsids of plant and animal viruses are important factors in capsid self- assembly, survival of viruses in the extracellular environment, and their cell infectivity. Virus shells can have applications as nanocontainers and delivery vehicles in biotechnology and medicine. Combined AFM experiments and computational modeling on sub-second timescales of the indentation nanomechanics of Cowpea Chlorotic Mottle Virus (CCMV) capsid show that the capsid’s physical properties are dynamic and local characteristics of the structure, which depend on the magnitude and geometry of mechanical input. Surprisingly, under large deformations the CCMV capsid transitions to the collapsed state without substantial local structural alterations. The enthalpy change in this deformation state ΔH = 11.5 – 12.8 MJ/mol is mostly due to large-amplitude out-of-plane excitations, which contribute to the capsid bending, and the entropy change TΔS = 5.1 - 5.8 MJ/mol is mostly due to coherent in-plane rearrangements of protein chains, which result in the capsid stiffening. Dynamic coupling of these modes defines the extent of elasticity and reversibility of capsid mechanical deformation. This emerging picture illuminates how unique physico-chemical properties of protein nanoshells help define their structure and morphology, and determine their viruses’ biological function. INTRODUCTION Hierarchical supramolecular systems that spontaneously assemble, disassemble and self-repair play fundamental roles in biology. Prime examples are viral capsids, which exhibit pronounced features such as shape alteration, viscoelasticity, and materials fatigue. Understanding the microscopic structural origin of the physico-chemical properties of these biological assemblies 1 Kononova et al. Mechanical Properties of CCMV Capsid and the mechanisms of their response to controlled mechanical inputs, remains a key research challenge. Beyond a fundamental understanding of these issues, there is a rapidly increasing interest in protein nanoshells for applications in biotechnology and nanomedicine. Recently, single-molecule techniques, such as Atomic Force Microscopy (AFM), have become available to explore physical properties of biological assemblies (1,2). This triggered tremendous research effort to characterize a variety of protein shells of plant and animal viruses, and bacteriophages. AFM based mechanical testing of viral capsids has now become the main tool to probe the physico-chemical and materials properties of viruses (3). For example, AFM deformation experiments yield information on the particle spring constant, reversibility of deformation, and forces required to distort capsid structures. A variety of viruses have been tested by this method including the bacteriophages Φ29, λ and HK97 (4-6), the human viruses Human Immunodeficiency Virus, Noro Virus, Hepatitis B Virus, Adeno Virus and Herpes Simplex Virus (7-13) and other eukaryotic cell infecting viruses like Minute Virus of Mice, Triatoma Virus and Cowpea Chlorotic Mottle Virus (14-16). These experiments reveal a truly amazing diversity of mechanical properties of viruses. Yet, due to their high complexity (~104-105 amino acid residues), experimental results are difficult to interpret without input from theoretical modeling. For biotechnological applications, it is essential to have full control over structure-based physical properties of virus shells, but in most instances a detailed knowledge of these properties is lacking. Viral capsids possess modular architectures, but strong capsomer intermolecular couplings modulate their properties. Consequently, the properties of the whole system (capsid) might not be given by the sum of the properties of its structural units (capsomers) (17-20). Under these circumstances, one cannot reconstruct the mechanical characteristics of the whole system using only information about the physical properties of its components. Biomolecular simulations have become indispensable for the theoretical exploration of the important dynamical properties and states of biological assemblies (21-23). Yet, the large temporal band width (ms-s) required limits the current theoretical capabilities. Theoretical studies employing triangulation of spherical surfaces and bead-spring models of stretching and bending have been used to probe the mechanical deformation and to test the mechanical limits of the shell (24,25). Yet, questions remain concerning structural details and dynamical aspects of these properties. How do discrete microscopic transitions give rise to the continuous mechanical response of a capsid at the macroscopic level? What are the structural rearrangements that govern the capsid’s transition from the elastic to plastic regime of the mechanical deformation? To address these questions, alternative detailed simulation approaches are needed. Here we feature a computational model, which is based upon the notion that the unique features associated with native topology, rather than atomic details, govern the physico-chemical properties of virus capsids. Our approach employs a topology based Self Organized Polymer (SOP) model (26,27), which provides an accurate description of the polypeptide chain (28-30), and high-performance computing accelerated on Graphics Processing Units (GPUs) (31,32). Here, we show that by combining AFM-based force measurements with accurate biomolecular simulations of forced indentation, we can obtain an in-depth understanding of the structural transitions and mechanisms of the mechanical deformation and the transition to the collapsed state in virus shells. On the theoretical side, the main challenge is to generate long (0.01 s – 0.1 s) dynamics of a virus particle using experimentally relevant force loads. Here, for the first time, we directly compare the results of experiments and simulations obtained under similar conditions 2 Kononova et al. Mechanical Properties of CCMV Capsid of force application for the Cowpea Chlorotic Mottle Virus (CCMV) used as a model system (15,33-38). CCMV is a member of the Bromoviridae, an important family of single stranded RNA plant viruses distributed worldwide that infect a range of hosts and are the cause of some major crop epidemics (39). The capsid of CCMV is an icosahedral protein shell (triangulation number T = 3) with an outer radius R = 13.2 nm and average shell thickness of 2.8 nm (3,40) consisting of 180 copies of a single 190 amino acid long protein. The shell comprises 60 trimer structural units and exhibits pentameric symmetry at the 12 vertices (pentamer capsomeres) and hexameric symmetry at the 20 faces (hexamer capsomeres) of the icosahedron (Fig. 1). The excellent agreement we demonstrated between experiments and simulations allowed us to describe the structural properties of the CCMV capsid, devoid of nucleic acids, to unprecedented detail, and to link the structural transitions in a protein shell with its mechanical properties. The insights into the dynamics of forced compression and structural collapse of this specific viral capsid provide a conceptual framework for describing the unique physico-chemical, thermodynamic and material properties of other virus particles. MATHERIALS AND METHODS Protein preparation: Purified capsid preparations of empty CCMV particles were obtained using the purification procedures described elsewhere (41). Briefly the procedure consists of isolation of CCMV particles from cowpea plants 13 days after infection (42,43). After UV/Vis spectroscopy characterization, Fast Protein Liquid Chromatography (FPLC) and running a sample on SDS-PAGE to determine the size of the monomers, the virions were examined by transmission electron microscopy. These measurements revealed that the CCMV particles had the expected size of ~28 nm in diameter. The buffer conditions for the imaging and nanoindentation experiments were 50 mM sodium acetate and 1 M sodium chloride (pH = 5.0). Atomic Force Microscopy: Hydrophobic glass slides were treated with an alkylsilane (4). The viral samples were kept under liquid conditions at all times; all the experiments were performed at room temperature. Capsid solutions were incubated for ~30 minutes on the hydrophobic glass slides prior to the indentation experiments. Olympus OMCL-RC800PSA rectangular, silicon nitride cantilevers (nominal tip radius < 20 nm and spring constant = 0.05 N/m) were calibrated in air yielding a spring constant of κ = 0.0524 ± 0.002 N/m. Viral imaging (44,45) and nanoindentation (3) were performed on a Nanotec Electronica AFM (Tres Cantos, Spain). Experimental nanoindentation measurements were carried out using the cantilever velocity vf = 0.6 and 6.0 μm/s. Additional measurements were performed at vf = 6.0×10-2 and 6.0×10-3 μm/s. The indentation data were analyzed using a home-written Labview program (National Instruments) as described previously (38). Self-Organized Polymer (SOP) Model: In the topology-based SOP model, each residue is described by a single interaction center (Cα-atom). The potential energy function of the protein conformation USOP specified in terms of the coordinates {ri} = r1, r2,…, rN (N is the total number of residues) is given by . The finite extensible nonlinear elastic (FENE) potential with the spring constant k = 14 N/m and the tolerance in the change of a covalent bond distance R0 = 2 Å describes the backbone 3 REPNBATTNBFENESOPUUUU20201120111log2RrrRk=U+ii,+ii,N=iFENE Kononova et al. Mechanical Properties of CCMV Capsid chain connectivity. The distance between residues i and i+1, is ri,i+1, and r0 i,i+1 is its value in the native structure. We used the Lennard-Jones potential to account for the non-covalent interactions that stabilize the native folded state. Here, the summation is performed over all the native contacts in the PDB structure; we assumed that if the non-covalently linked residues i and j (i-j > 2) are within the cut-off distance RC = 8 Å in the native state, then ∆ij =1, and is zero otherwise (RC = 8 Å is the cut-off distance for calculating the non-bonded forces). The value of εn quantifies the strength of the non-bonded interactions. We used εn = εinter = 1.29 kcal/mol and εn = εintra = 1.05 kcal/mol for the inter-chain contacts and intra-chain contacts (model parameterization is described in SM). The non-native interactions in the potential were treated as repulsive. Here, the summation is performed over all the non-native contacts with the distance > RC. A constraint is imposed on the bond angle formed by residues i, i+1, and i+2 by including the repulsive potential with parameters εr = 1 kcal/mol and σr = 3.8 Å, which determine the strength and the range of the repulsion. Six independent runs have been carried out for each simulation setup. Forced indentation simulations: Dynamic force measurements in silico were performed using the SOP model and Langevin simulations accelerated on a GPU. We used the CCMV virus capsid, empty of RNA molecules (Protein Data Bank (PDB) code 1CWP) (40), and a spherical tip of radius Rtip= 5, 10, and 20 nm to compress the capsid along the two-, three-, and five-fold symmetry axis. The tip-capsid interactions were modeled by the repulsive Lennard-Jones potential, V(ri) = ε[σ/(ri – Rtip)]6, where ri are the ith particle coordinates, ε = 4.18 kJ/mol, and σ =1.0 Å. We constrained the bottom portion of the CCMV by fixing five Cα-atoms along the perimeter. The tip exerted the time-dependent force f(t) = f(t)n in the direction n perpendicular to the surface of CCMV shell. The force magnitude f(t) = rft increased linearly in time t (force- ramp) with the force-loading rate rf = κvf (vf is the cantilever base velocity and κ is the cantilever spring constant). In the simulations of “forward indentation”, the spherical tip was moving towards the capsid with vf = 1.0 µm/s and vf = 25 µm/s (κ = 0.05 N/m). In the simulations of force-quenched retraction for vf = 1.0 µm/s, we reversed the direction of tip motion. In simulations, we control the piezo displacement Z (cantilever base), and the cantilever tip position X. The cantilever base (virtual particle) moves with constant velocity (vf), which ramps up a force f(t) applied to the capsid through the cantilever tip with the force-loading rate rf. The resisting force F from the capsid, which corresponds to the experimentally measured indentation force, can be calculated using the energy output. The capsid spring constant kcap can be obtained from an FX curve by calculating the slope kcap = dF/dX. Analysis of simulation output: Analysis of structures: To measure the extent of structural similarity between a given conformation and a reference state, we used the structure overlap 0 are the inter-particle distances between the i-th and j-th residues in the transient and native structure, respectively (β = 0.2 is the tolerance for the distance change). Analysis of energies: We analyzed the potential energy USOP, and utilized Umbrella Sampling simulations (see SM; 46,47) to estimate the Gibbs energy (ΔG), enthalpy (ΔH), and entropy (ΔS). Normal Mode Analysis was used to characterize the equilibrium vibrations (see SM) (48). We calculated the Hessian matrix for centers of mass of amino acid residues (HIJ). The eigenvalues {λI}, and eigenvectors {RI} . In Θ(x), Heaviside step function, rij(t) and rij 4 31,60120/2/jiijijijijijnATTNBrrrrU21,3,66)1(//jiijiijijrrijrrREPNBrrU)βrr(t)rΘ(NN=ξ(t)0ijijij01))1(2( Kononova et al. Mechanical Properties of CCMV Capsid obtained numerically were used to calculate the spectrum of normal frequencies and normal modes (qI–center-of-mass positions). In the Essential Dynamics approach (48), implemented in GROMACS (49), collective modes of motion describing the non- equilibrium displacements of amino acids are projected along the direction of global transition X (indentation depth), characterized by the displacements ΔX(t) = X(t) – X0 from equilibrium X0 (see SM). We diagonalized the covariance matrix C(t) = ΔX(t) ΔX(t)T = TΛTT to compute the matrix of eigenvalues Λ and the matrix of eigenvectors T. These were used to find the projections ΔX(t) on each eigenvector tI, PI(t) = tIΔX(t). RESULTS AFM Indentation Experiments: Before nanoindentation, AFM images of the capsid were recorded as depicted in Fig. 1A. Next, nanoindentation measurements were performed on the center of the CCMV capsid particle, and the corresponding force (F)–indentation (Z) curves (FZ curves) were recorded. The FZ curves quantify the mechanical response of the capsid (indentation force F) as a function of the piezo displacement (reaction coordinate Z). The FZ curves (Fig. 2) revealed that mechanical nanoindentation is a complex stochastic process, which might occur in a single step (“all-or-none” transition with a single force peak) or through multiple steps (several force peaks). To characterize the experimental FZ curves, we focused on the common features, which include an initial linear-like indentation behavior followed by a sharp drop in force (Fig. 2). Next, we performed a fit of a straight line to the initial region of each FZ curve to determine the capsid spring constant kcap, which quantifies the elastic compliance of the capsid, using the relationship 1/K = 1/κ + 1/kcap for the cantilever plus capsid setup. Here, K is the slope in the FZ curve (Fig. 2), and κ is the cantilever spring constant (see Materials and Methods). We found that the average spring constant of CCMV is kcap= 0.17 N/m at a loading rate vf = 0.6 μm/s and kcap= 0.14 N/m at vf = 6.0 μm/s (Table I). Additional experiments showed that kcap does not change much over four decades of vf (Fig. 3A) (38). The indentation force, where the linear-like regime in the FZ curve ends, corresponds to the critical force at which the mechanical failure of the capsid occurs. We analyzed the critical forces (F*) by extracting peak forces observed in the FZ curves, and the corresponding transition distances (Z*). The average critical force was determined to be F* = 0.71 and 0.72 nN for vf = 0.6 μm/s and vf = 6.0 μm/s, respectively, showing that critical force is not affected by loading rate in the regime of vf used (38). These experiments also showed a good agreement with previously published results (15), where it was found that kcap ≈ 0.15 N/m and F* ≈ 0.6 nN for comparable loading rates (0.02 µm/s – 2 µm/s range). Next, we considered whether mechanical deformation was reversible. We performed measurements for small forward indentation followed by backward movement of the AFM tip, which we refer to as “force-quenched retraction”. An example of such measurements clearly shows that there is a considerable difference between the mechanical response of the CCMV particle observed for small and large deformations (Fig. 3B). Whereas for large piezo displacements (Z > 35 nm) the deformation was irreversible with large hysteresis, for small displacements (Z < 10 nm) the deformation was completely reversible with almost no hysteresis. These results also agree with our previous findings (15). 5 IIλωIIJIqR=Q Kononova et al. Mechanical Properties of CCMV Capsid Forced Indentation in silico: We performed indentation simulations using a spherical tip of radius Rtip = 20 nm. The 250-fold computational acceleration on a GPU has enabled us to use experimentally relevant vf = 0.5 and 1.0 µm/s (κ= 0.05 N/m) and span the experimental 0.1–0.2 s timescale. To provide the basis for comparison of the results of experiments and simulations, we analyzed the indentation force F as a function of the piezo displacement Z (FZ curves; see Materials and Methods). The theoretical FZ curves (Fig. 2C and 2D) compare well with the experimental FZ profiles (Fig. 2A and 2B). The simulated FZ curves also exhibit single-step and multi-step transitions. The average values of kcap, F*, and Z* compare well with their experimental counterparts (Table I). We also analyzed the dependence of kcap on vf (including additionally vf = 5 and 25 µm/s) and found that as in experiment, kcap was insensitive to the variation of vf (Fig. 3A). Next, we performed simulations of force-quenched retraction, where we used the CCMV structures generated in the forward deformation runs for Z = 15, 25, 28, and 32 nm as initial conditions. Our simulations (Fig. 3C) agreed with experiments (Fig. 3B) in that indentation is fully reversible for small Z = 15 nm displacement but irreversible beyond the critical distance, Z > 25 nm. To summarize, we have obtained an almost quantitative agreement between the results of dynamic force measurements in vitro and in silico (Fig. 2), including our findings regarding the weak sensitivity of the capsid spring constant kcap on changes in the cantilever velocity vf (Fig. 3A) and our results of force-quenched retraction (Fig. 3B and 3C). Hence, the SOP model of CCMV provides an accurate description of the capsid mechanical properties, which validates our approach. The good agreement between the results of experiments and simulations allowed us to probe features of CCMV shell that are not accessible experimentally. Mechanical properties of CCMV depend on local symmetry: Next, we performed simulations of indentation under slow (vf = 1.0 µm/s) and fast (vf = 25 µm/s) force loading (Rtip = 20 nm). The capsid was indented at different points on its surface: at the symmetry axes of the hexamer capsomeres (three-fold symmetry), the pentamer capsomeres (five-fold symmetry), and at the interface between two hexamers (two-fold symmetry) as described in Fig. S1 in the Supporting Material (SM). The FX profiles for the two-fold symmetry axes are displayed in Fig. 4. The FX curves for three- and five-fold symmetry are presented in Fig. S2 and S3, respectively. In agreement with experiment (Fig. 2B), the mechanical reaction of the capsid is elastic up to X ≈ 3-5 nm (Z ≈ 8-10 nm; linear regime) and quasi-elastic up to X ≈ 8-11 nm (Z ≈ 22 - 25 nm; linear- like regime), regardless of the capsid orientation (panel A in Fig. 4, S2, and S3). The FX curves generated under fast force loading (vf = 25 µm/s) showed a slightly steeper slope. Fitting a straight line to the initial portion of FX curves (X < 3 nm) taken at vf = 1.0 µm/s yielded the spring constant kcap ≈ 0.11 N/m, 0.10 N/m, and 0.12 N/m for two-, three-, and five-fold symmetry, respectively (Table II). A more detailed analysis showed that kcap is significantly varying during simulated particle deformation: kcap varies from 0.06 - 0.14 N/m, 0.05 - 0.10 N/m, and 0.04 - 0.12 N/m for the two-, three-, and five-fold symmetry axes, respectively in the initial deformation regime. Fluctuations in kcap show systematic differences for the icosahedral symmetry axes (Fig. 4, S2, and S3). When probing along the two- and five-fold axes, the curves of kcap versus X show two maxima: the first maximum is at X ≈ 2-3 nm (for two- and five-fold symmetry), and the second maximum is at X ≈ 5-6 nm (two-fold symmetry) and 11-12 nm (five-fold symmetry). For the three-fold symmetry, kcap shows one broad skewed peak centered at X ≈ 5 nm. To explain this effect, we performed structural analysis of CCMV conformations, and found that at various 6 Kononova et al. Mechanical Properties of CCMV Capsid stages of mechanical deformation different types (pentamers and hexamers) and number of capsomers cooperate to withstand the mechanical stress (Fig. S1). The collapse transition occurs in the 11 - 15 nm range (Fig. 4, S2, and S3). For vf = 1.0 µm/s, the average critical forces from experiments and simulations agreed (F* = 0.71 nN; see Fig. 2 and Tables I and II). Here, the bottom portion of the shell becomes increasingly more flat (see snapshots in panel C in Fig. 4, S2, and S3), and the capsid undergoes a spontaneous shape change from a roughly spherical state to a non-spherical collapsed state, which is reflected in the sudden force drop and decrease of kcap to zero (panel B in Fig. 4, S2, and S3). Under fast loading (vf = 25 µm/s), force peaks were not detected (panel A in Fig. 4, S2, and S3). To quantify the extent of the structural collapse, we monitored the structure overlap ξ. In the transition regime, ξ decreased from ξ = 1 (native state) to ξ = 0.65 (collapsed state) for all symmetry types (panel C in Fig. 4, S2, and S3). Hence, notwithstanding the large scale transition, the capsid structure remained 65% similar to the native state. For faster vf = 25 µm/s, ξ decreased by <10%, indicating that fast force loading leaves the local arrangements of capsomers unaffected. At X > 15 nm, CCMV entered the post-collapse, second linear-like regime (panels A and B in Fig. 4, S2, and S3). Here, as the tip approached the solid surface, kcap increased sharply. Next, we reversed the direction of tip motion using the structures for the collapsed state obtained for Z = 15, 25, 28, and 32 nm (X = 5, 11, 15, and 19 nm) as initial conditions. In agreement with AFM data (Fig. 3B), the theoretical force-retraction curves for vf = 1.0 µm/s showed that the mechanical compression of CCMV was fully reversible in the elastic regime for X = 5 nm (no hysteresis), nearly reversible in the quasi-elastic regime for X = 11 nm (small hysteresis), but irreversible after the transition had occurred (X = 15 and 19 nm; panel A in Fig. 4, S2, and S3). We also analyzed the progress of CCMV shell restructuring by monitoring ξ as a function of time (panel C in Fig. 4, S2, and S3), and found that the CCMV shell recovered its original shape (ξ = 1) in the millisecond timescale. Thermodynamics of CCMV indentation: We evaluated the total work of indentation w by integrating the area under the FX curves. We repeated this procedure for the retraction curves to evaluate the reversible work wrev. Estimation of the relative difference (w - wrev)/w showed that in the elastic and quasi-elastic regime (X < 11 nm; Z < 25 nm) ~12% of w was dissipated. This agrees with the experimental finding that the fraction of energy returned upon retraction is ~90% (15). In the transition range (11 nm ≤ X ≤ 15 nm; 25 nm ≤ Z ≤ 30 nm), where the retraction curves showed large hysteresis especially for the three-fold symmetry, (w - wrev)/w ≈ 75%. Because wrev is equal to the Gibbs energy change, i.e., wrev = ΔG = ΔH - TΔS, where ΔH and ΔS are the enthalpy change and entropy change, we estimated ΔH and TΔS. The results for ΔH and TΔS for indentation along the two-, three-, and five-fold symmetry axes (Rtip = 20 nm; vf = 1.0 µm/s) are displayed, respectively, in Fig. 4, S2, and S3. In the linear regime and linear-like regime (X < 10-11 nm), ΔH and TΔS display a parabolic dependence on X and ΔH ≈ TΔS; in the (11-15 nm) transition range and in the post-collapse regime (X > 15 nm), ΔH and TΔS level off, each attaining a plateau, and ΔH > TΔS. The dependence of ΔH on X under fast force-loading conditions is more monotonic. The curves of ΔG, ΔH, and TΔS attain some constant values ΔGind, ΔHind, and TΔSind at X = 20 nm, which correspond to the Gibbs energy, enthalpy, and entropy of indentation (Table II). We also mapped the equilibrium energy landscape ΔG using the Umbrella Sampling simulations (see Experimental Procedures) and resolved the profiles of ΔH and TΔS (the inset in panel D in Fig. 4, S2, and S3). The equilibrium values of ΔGind, ΔHind, and TΔSind are accumulated in Table II. The thermodynamic functions indicate that mechanical compression of 7 Kononova et al. Mechanical Properties of CCMV Capsid the CCMV shell requires a considerable investment of energy, and that ΔGind, ΔHind, and TΔSind vary with the local symmetry under the tip. Mechanical response of CCMV depends on geometry of force application: We performed simulations of the nanoindentation of CCMV along the two-fold symmetry axis using a tip of smaller radius Rtip = 10 and 5 nm. The FX profiles, spring constant kcap, structure overlap ξ, and thermodynamic functions obtained for Rtip = 10 nm (Fig. S4) can be compared with the same quantities obtained for Rtip = 20 nm (Fig. 4). We present our findings for Rtip = 10 nm; results for Rtip = 5 nm show a similar tendency (data not shown). The FX curves for vf = 1.0 (and 25 µm/s) shows a less steep kcap, the collapse transition is less pronounced and starts sooner (X* ≈ 9 nm), and the critical force is lower (F* ≈ 0.6 nN) for indentation with a smaller tip (panels A and B in Fig. S4). The kcap versus X dependence shows two peaks, but the second peak at X ≈ 6 nm is weaker (compared to the results for Rtip = 20 nm). The overlap ξ decreased to 0.75, implying that in the collapsed state the CCMV shell remained ≈75% similar to the native state (panel C in Fig. S4). ΔH and TΔS show a familiar parabolic dependence on X (as for Rtip = 20 nm), but these level off at somewhat lower values (panel D in Fig. S4). Numerical estimates of kcap, ΔGind, ΔHind, and TΔSind obtained for Rtip = 10 and 5 nm (Table III) are directly proportional to the tip size as kcap, ΔGind, ΔHind, and ΔSind all decrease with Rtip. We also performed simulations using 10 and 5 nm tip but for the three- and five-fold symmetry axes, and arrived at the same conclusions (data not shown). Equilibrium and non-equilibrium dynamics of CCMV: To resolve the dynamic determinants underlying the mechanical response of CCMV, we first analyzed equilibrium properties of the capsid using Normal Mode Analysis (see Material and Methods and SM). We calculated the spectra of normal modes for Cα-atoms for a single hexamer, single pentamer, and for the whole CCMV particle (Fig. S5). Because the spectra for a pentamer and hexamer were identical, we only display the spectrum for a hexamer, which practically overlaps with the spectrum for the CCMV shell, implying that normal displacements of the CCMV shell and its constituents are similar, especially in the 50-500 cm-1-range of frequencies. Analysis of structures revealed that the more global modes of motion in the low-frequency part of the spectrum (≤ 50 cm-1) involve the “out-of-plane” expansion-contraction excitations and the “in- plane” concerted displacements of capsomers. The more local modes (100-250 cm-1 range) are small-amplitude displacements of the secondary structure elements. The high frequency 300-450 cm-1 end of the spectrum is dominated by the local vibrations of amino acids (Fig. S5). To describe the far-from-equilibrium mechanical response of CCMV, we employed the Essential Dynamics approach (see Materials and Methods and SM), which helps to single out the most important types of motion, showing the largest contribution in the direction of global transition (indentation depth X). Importantly, the essential dynamics modes should not be confused with the equilibrium normal modes or instantaneous normal modes. We examined separately the elastic regime (X ≤ 5 nm) and the transition regime (11 nm ≤ X ≤ 15 nm) using the simulation output for the two-fold symmetry (Fig. 4). We resolved the principal coordinates PI(t), collective variables describing the non-equilibrium dynamics of the system, and analyzed the relative displacement for each I-th mode given by the ratio of the average squared displacement to the total squared displacement . It turns out that the first two modes account for ~85% of dynamics of CCMV 8 22/I0II0IXXXX2I0IXX2I0IXX Kononova et al. Mechanical Properties of CCMV Capsid (“essential subspace”); the remaining modes are negligible in terms of the displacement amplitude. The CCMV dynamics in the essential subspace is displayed in Fig. 5. We characterized large-amplitude non-equilibrium displacements, which contribute the most to the mechanical compression of CCMV. In the elastic regime, the first mode (mode 1: 77% of dynamics) corresponds to the large-amplitude out-of-plane compression, which results in the capsid bending. Here, the top and bottom portions of the capsid become flat, while the capsid sides expand outward (Fig. 5A). The second mode (mode 2: 8% of dynamics) represents direct coupling of the in-plane displacements of capsomers and the out-of-plane capsid bending, for which the in-plane displacements and the out-of-plane bending occur at the same time. Here we observe that the arrangement of capsomers on the spherical surface change from the more ordered to the less ordered, and the capsid structure loses its near-spherical symmetry (Fig. 5A). In the transition regime, the in-plane and out-of-plane displacements are strongly coupled. That is, these displacements are neither purely in-plane nor out-of-plane excitations. The first mode (mode 1) represents the collapse transition, which is accompanied by the lateral translocation of the capsomers towards the tip-surface contact area (Fig. 5B). The second mode (mode 2) is dominated by the lateral translocation and twisting motions of hexamers and pentamers in the clock-wise and counter clock-wise direction around their symmetry axes, respectively (Fig. 5B). DISCUSSION By coupling dynamic force measurements in vitro and in silico, here we have directly compared, for the first time, the experimental data with simulation data for CCMV shell obtained under identical conditions of the mechanical force load. Excellent agreement between the experimental and simulation results validated our theoretical approach. Larger variation in the experimental FZ profiles are due to the fact that in experiments not only three different icosahedral orientations were probed (two-, three-, and five-fold symmetry) by the AFM tip, but also various intermediate orientations. This undertaking had enabled us to interpret the experimental forced indentation patterns in unprecedented detail with regards to the structural and thermodynamic changes in the CCMV capsid in response to external mechanical deformation. The main results are: 1) The physical properties of the CCMV shell are dynamic but local characteristics of the structure, and the mechanical response of the capsid depends not only on the symmetry of the local capsomer arrangement under the tip, but also on the indentation depth. 2) The mechanical characteristics of CCMV - the critical force and transition distance - depend on how rapidly the compressive force is increased. 3) The physical properties of the CCMV particle depend on the geometry of mechanical perturbation, as the mechanical response changes with tip size. 4) The extent to which the mechanical deformation of the CCMV shell can be retraced back reversibly depends on the indentation depth. 5) In the elastic regime of deformation, the “out-of-plane” excitations dominate the near-equilibrium displacements of capsomers, but these and “in-plane” modes are strongly coupled in the far-from-equilibrium transition range. 6) The entropic and enthalpic contributions are almost equally important for the capsid stiffening, whereas the capsid softening and transition to the collapsed state is driven mainly by the enthalpy change. Our conclusion about the local nature of physical properties also fits with previous modeling of Hepatitis B Virus, which showed that permanent deformation of the shell was due to 9 Kononova et al. Mechanical Properties of CCMV Capsid local rearrangements of the capsid proteins (21). That the CCMV shell displays multiple modes of mechanical resistance, which depends on the indentation depth, agrees well with recent studies, which showed two dynamic regimes to be responsible for the CCMV capsid stiffening and softening (34). The existence of multiple modes is reflected in the non-monotonic dependence and maxima of kcap as a function of X. The periods of mechanical resistance (stiffening), during which an increasingly larger portion of protein chains find themselves in the tip-shell contact area, are interrupted by the periods when the capsid yields to force (softening). The first peak of kcap at X ≈ 2-3 nm (Fig. 4 and S3), agrees with the previous results from finite element analysis (34). The second peak at X ≈ 6 nm and X ≈ 11 nm for the two-fold and five-fold symmetry described here correspond to the capsid softening beyond X ≈ 10 nm (34). The weak dependence of CCMV mechanical properties on the rate of change of compressive force is not unexpected, because the positions of transition states and barrier heights on the energy landscape depend upon how force is applied to the system. Under fast force loading (vf = 25 μm/s), the energy pumped into the system is more than sufficient to overcome the energy barrier, and the transition to the collapsed state is not well-pronounced (no force-drop in the FX curves in Fig. 4, S2, and S3). The force peaks are observed under slow loading (vf = 0.5 – 6 μm/s) because the amount of energy is comparable to the energy barrier for the collapse transition. The FZ curves for the CCMV capsid obtained using the finite element analysis (34) agree with our result for vf = 25 μm/s. This indicates that conditions of force application used in the finite element analysis correspond to the fast indentation case. The FZ profiles from the finite element analysis and our own results also agree in that, under fast force loading, differences in the mechanical response of CCMV for different symmetries disappear. In single-molecule manipulation on virus particles, mechanical force requires a physical contact between a system and a probe. Hence, their shape, size difference, and the direction of force become important factors. When a virus is indented by a plane (Rtip>> R – radius of a virus shell), all residues in the tip-capsid contact area are pushed in the same direction; when a virus is indented by a small sphere (Rtip≈ R), different residues are displaced in different directions. Our results show that the mechanical characteristics - FX profile, spring constant, critical force, and indentation depth (Fig. S4) - all change with probe size, and that ΔGind, ΔHind, and TΔSind are directly proportional to Rtip (Table III). A smaller tip means a smaller tip-capsid contact area, and, hence, weaker mechanical response and lower associated energy costs. In the elastic regime, quasi-elastic regime, and transition regime (Fig. 4, S2, and S3), the deformation is reversible for short X and almost reversible for longer X. In the post-collapse regime, the mechanical compression is irreversible. In fact, these same findings can be rationalized using our results from Essential Dynamics (Fig. 5). In the elastic regime and quasi- elastic regime, the first mode is dominated by the out-of-plane displacements of pentamers and hexamers. Hence, when a compressive force is quenched, as in the retraction experiments, the first mode provides a mechanism for capsid reshaping, and the amount of energy dissipated is small. In the transition range, the two most important modes represent strongly coupled out-of- plane and in-plane displacements. Here, the capsid is capable of restoring its original shape, but capsid restructuring comes at a cost of exciting additional degrees of freedom and, hence, a larger amount of dissipated energy. The question exists whether the property of a whole system can be represented by a sum of the properties of its structural elements (50). For the CCMV capsid dynamics at equilibrium, our results from Normal Mode analysis (37,51) provide the affirmative answer. The spectra of 10 Kononova et al. Mechanical Properties of CCMV Capsid eigenmodes for an isolated single pentamer or hexamer and for the whole capsid show only small differences at low frequencies (<50 cm-1), due to icosahedrally symmetric modes present in the full capsid, but practically overlap with that for the whole shell in the 50 – 500 cm-1-range (Fig. S5). This result is not surprising. The differences in the small-amplitude equilibrium fluctuations of residue positions for local modes are negligible for the penton, hexon, and full capsid. Of course, these modes represent collective motions, which correspond to penton, hexon, and full capsid decompositions; yet, when compared at the whole shell level, these motions in the penton and hexon units, and in the full capsid are nearly identical (Fig. S5). Hence, domain and capsomer interactions have little effect on equilibrium properties of CCMV. Under non-equilibrium conditions of mechanical deformation, the near-spherical symmetry of the capsid is broken and different capsomers start playing different roles. In this regime, we can no longer use a concept of equilibrium normal modes. We employed the Essential Dynamics approach to characterize large-amplitude displacements of capsomers under far-from-equilibrium conditions. Although in the linear regime the main mode of collective motions is dominated by the “out-of-plane” displacements, there are no pure “out-of-plane” and “in-plane” modes either in the elastic regime or in the transition range (Fig. 5). These coupled non-equilibrium essential modes of motion, which accompany the CCMV transition to the collapsed state, cannot be reconstructed using “a linear combination” of the out-of-plane modes and the in-plane modes. The concerted in-plane displacements mediate rearrangements of pentamers and hexamers on the CCMV surface, which leads to capsid stiffening reflected in the non-monotonic dependence of kcap (Figs. 4, S2, S3, and S4). These are exact results, i.e. we have arrived at these conclusions by analyzing the output from Essential Dynamics calculations. Similar findings have been reported by other research groups (52). We mapped the energy landscape for the mechanical deformation of the CCMV capsid (Fig. 4, S2, and S3). The similarity of non-equilibrium estimates of ΔGind, ΔHind, and ΔSind (from FX curves) and their equilibrium counterparts (from Umbrella Sampling) implies that slow force loading (vf = 1.0 µm/s) corresponds to near-equilibrium conditions of force application. Both the entropic and enthalpic contributions to ΔG (6.5-6.9 MJ/mol) are important: the entropy change TΔSind (5.1-5.8 MJ/mol) is roughly half the enthalpy change ΔHind (11.5-12.8 MJ/mol) for all three symmetry types (Table II). There are variations in the values of ΔGind, ΔHind, and TΔSind for different symmetries: these functions for five-fold symmetry differ by ~10% from the same functions for two- and three-fold symmetry (Table II). Hence, our findings stress the importance of any particular capsid’s discrete nature and local protein subunit(s)/capsomer symmetry when virus shells are tested mechanically. The potential energy of protein chains (USOP) sharply increases in the transition range where the capsid alters its shape from the convex to the concave (tip-indented convex down). These shape alterations are captured by the enthalpy change ΔH (Fig. 4). Compared to the elastic regime of CCMV deformation (X < 3-5 nm), where ΔH increases by ~3 MJ/mol, in the transition region (11 nm < X < 15 nm) ΔH increases three-fold to ~10 MJ/mol (Fig. 4D). Here, the large- amplitude out-of-plane displacements mediate the capsid bending inward. Hence, in the quasi- elastic regime before the collapse transition occurs the out-of-plane collective modes contribute mainly to the enthalpy change ΔH. Although small-amplitude in-plane displacements are coupled to the out-of-plane modes, the main effect from in-plane displacements is concerted transitions - displacements, translocations, and twisting, from the more ordered to the less ordered phase formed by protein chains (Fig. 5). Hence, the in-plane modes contribute mainly to the entropy 11 Kononova et al. Mechanical Properties of CCMV Capsid change TΔS, which increases two-fold from ~3 MJ/mol in the elastic regime to ~6 MJ/mol in the transition range (Fig. 4). The map of local potential energy for protein chains forming capsomers shows that there is more energy stored in pentamers than in hexamers in the elastic regime and in quasi-elastic regime (Fig. S6). This correlates well with the inhomogeneous stress distribution in CCMV capsid found earlier using other methods (53). However, this picture is more mixed in the transition range. Hence, under tension, the same protein chains forming capsomers play different roles in the energy distribution, which changes with the indentation depth. When the capsid is undergoing the global transition to the collapsed state, the average structure of the protein chains forming capsomers is affected, but to a limited extent. This is reflected in the small decrease of the structure overlap to ξ ≈ 0.6 for the slow force loading (vf = 1 μm/s). Under fast loading (vf = 25 μm/s), or for smaller tip (Rtip = 10 nm), the decrease in ξ is even smaller (Fig. 4, S2, and S3, and S4). This stands in contrast to mechanical protein unfolding where transitioning to the globally unfolded state occurs concomitant with the disruption of native interactions stabilizing the tertiary and secondary structures of the native fold. Hence, in the context of mechanical deformation of a capsid, force-induced spontaneous shape changing does not imply substantial structural transitions on the local scale. We have advanced a conceptual understanding of the unique physical properties of capsids - their dynamic and local structural features, and their dependence on the rate of change and geometry of external physical stimulus. We have resolved the origin of multiple modes of elastic compliance leading to mechanical stiffening and softening effects, and have characterized (ir)reversibility of the mechanical deformation of virions. We have described specific roles played by the in-plane and out-of plane non-equilibrium collective modes of the capsomers’ displacements and their connection to the thermodynamic functions. Because these properties are likely to be shared among different virion classes, the results of these studies are important to understand the nanomechanics of other protein shells. Furthermore, profiling the structural, dynamic, and thermodynamic characteristics of capsids can illuminate different aspects of their biological properties and function. Also, biotechnological applications of protein nanocontainers range from catalysis in constrained environments in organic synthesis to transport and delivery of substrates into cells in nanomedicine, and to building blocks in nanotechnology (54,55). Our combined in vitro and in silico techniques may become a widely used approach to reveal the structure-dynamics relationship for biologically derived nanoparticles. Acknowledgements: This work was supported by the Russian Ministry of Education and Science (Grant 14.A18.21.1239 to VB), by the “Physics of the genome” program grant from Fundamenteel Onderzoek der Materie (FOM) (to GJLW), and by the National Science Foundation (grant MCB-0845002 to RID). SUPPORTING CITATIONS: References (56,57) appear in the Supporting Material. REFERENCES: 1. Kasas, S., and G. Dietler. 2008. Probing nanomechanical properties from biomolecules to living cells. Pflug. Arch. Eur. J. Phys. 456:13-27. 12 Kononova et al. Mechanical Properties of CCMV Capsid 2. Engel, A., and D. J. Müller. 2000. Observing single biomolecules at work with the atomic force microscope. Nat. Struct. Biol. 7:715–718. 3. Roos, W. H., R. Bruinsma, and G. J. L. Wuite. 2010. Physical virology. Nat. Phys. 6:733- 743. 4. 5. Ivanovska, I. L., P. J. de Pablo, B. Ibarra, G. Sgalari, F. C. MacKintosh, J. L. Carrascosa, C. F. Schmidt, and G. J. L. Wuite. 2004. Bacteriophage capsids: Tough nanoshells with complex elastic properties. Proc. Natl. Acad. Sci. USA 101:7600-7605. Ivanovska, I. L., G. J. L. Wuite, B. Jönsson, and A. Evilevitch. 2007. Internal DNA pressure modifies stability of WT phage. Proc. Natl. Acad. Sci. USA 104:9603-9608. 6. Roos, W. H., I. Gertsman, E. R. May, C. L. Brooks 3rd, J. E. Johnson, and G. J. L. Wuite. 2012. Mechanics of bacteriophage maturation. Proc. Natl. Acad. Sci. USA 109:2342-2347. 7. Kol, M., Y. Shi, M. Tsvitov, D. Barlam, R. Z. Shneck, M. S. Kay, I. Rousso. 2007. A stiffness switch in human immunodeficiency virus. Biophys. J. 92:1777-1783. 8. Baclayon, M., G. K. Shoemaker, C. Uetrecht, S. E. Crawford, M. K. Estes, B. V. Prasad, A. J. R. Heck, G. J. L. Wuite, and W. H. Roos. 2011. Prestress strengthens the shell of Norwalk virus nanoparticles. Nano Lett. 11:4865-4869. 9. Liashkovich, I., W. Hafezi, J. E. Kühn, H. Oberleithner, A. Kramer, and V. Shahin. 2008. Exceptional mechanical and structural stability of HSV-1 unveiled with fluid atomic force microscopy. J. Cell. Sci. 121:2287-2292. 10. Pérez-BernÁ, A. J., A. Ortega-Esteban, R. Menendez-Conejero, D. C. Winkler, M. Menendez, A. C. Steven, S. J. Flint, P. J. de Pablo, and C. San Martin. 2012. The role of capsid maturation on adenovirus priming for sequential uncoating. J. Biol. Chem., 287:31582–31595. 11. Roos, W. H., K. Radtke, E. Kniesmeijer, H. Geertsema, B. Sodeik, and G. J. L. Wuite. 2009. Scaffold expulsion and genome packaging trigger stabilization of herpes simplex virus capsid. Proc. Natl. Acad. Sci. USA 106:9673-9678. 12. Roos, W. H., M. M. Gibbons, A. Arkhipov, C. Uetrecht, N. R. Watts, P. T. Wingfield, A. C. Steven, A. J. R. Heck, K. Schulten, W. S. Klug, and G. J. L. Wuite. 2010. Squeezing protein shells: how continuum elastic models, molecular dynamics simulation and experiments coalesce at the nanoscale. Biophys. J. 99:1175-1181. 13. Snijder, J., V. S. Reddy, E. R. May, W. H. Roos, G. R. Nemerow, and G. J. L. Wuite. 2013. Integrin and defensin modulate the mechanical properties of adenovirus. J. Virol. 87:2756. 14. Carrasco, C., A. Carreira, I. A. T. Schaap, P. A. Serena, J. Gómez-Herrero, M. G. Mateu, and P. J. de Pablo. 2006. DNA-mediated anisotropic mechanical reinforcement of a virus. Proc. Natl. Acad. Sci. USA 103:13706-13711. 13 Kononova et al. Mechanical Properties of CCMV Capsid 15. Michel, J. P., I. L. Ivanovska, M. M. Gibbons, W. S. Klug, C. M. Knobler, G. J. L. Wuite, and C. F. Schmidt. 2006. Nanoindentation studies of full and empty viral capsids and the effects of capsid protein mutations on elasticity and strength. Proc. Natl. Acad. Sci. USA 103:6184-6189. 16. Snijder, J., C. Uetrecht, R. Rose, R. Sanchez, G. Marti, J. Agirre, D. M. Guérin, G. J. L. Wuite, A. J. R. Heck, and W. H. Roos. 2013. Probing the biophysical interplay between a viral genome and its capsid. Nature Chemistry, 5:502-509. 17. Berendsen H. J. C. and S. Hayward. 2000. Collective dynamics in relation to function. Curr. Opin. Struct. Biol. 10:165-169 18. Bura, E., D. K. Klimov, and V. Barsegov. 2008. Analyzing forced unfolding of protein tandems by ordered variates, 2:Dependent unfolding times. Biophys. J. 94:2516-2528. 19. Joshi, H., F. Momin, K. E. Haines, and R. I. Dima. 2010. Exploring the contribution of collective motions to the dynamics of forced-unfolding in tubulin. Biophys. J. 98:657-666. 20. Lange, O. F. and H. Grubmüller H. 2008. Full correlation analysis of conformational protein dynamics. Proteins 70:1294-1312. 21. Arkhipov, A., W. H. Roos, G. J. L. Wuite, and K. Schulten. 2009. Elucidating the mechanism behind irreversible deformation of viral capsids. Biophys. J. 97:2061-2069. 22. Phelps, D. K., B. Speelman, and C. B. Post. 2000. Theoretical studies of viral capsid proteins. Curr. Opin. Struct. Biol. 10:170–173. 23. Zink, M., and H. Grubmüller. 2009. Mechanical properties of the icosahedral shell of southern bean mosaic virus: A molecular dynamics study. Biophys. J. 96:1350–1363. 24. Vliegenthart, G. A. and G. Gompper. 2006. Mechanical deformation of spherical viruses with icosahedral symmetry. Biophys. J. 91:834-841. 25. Buenemann, M. and P. Lenz. 2007. Mechanical limits of viral capsids. Proc. Natl. Acad. Sci. USA 104:9925-9930. 26. Hyeon, C., R. I. Dima, and D. Thirumalai. 2006. Pathway and kinetic barriers in mechanical unfolding and refolding of RNA and proteins. Structure 14:1633-1645. 27. Mickler, M., R. I. Dima, H. Dietz, C. Hyeon, D. Thirumalai, and M. Rief. 2007. Revealing the bifurcation in the unfolding pathways of GFP by using single-molecule experiments and simulations. Proc. Natl. Acad. Sci. USA 104:20268-20273. 28. Dima, R. I., and H. Joshi. 2008. Probing the origin of tubuline rigidity with molecular simulations. Proc. Natl. Acad. Sci. USA 105:15743-15748. 29. Lin, J. C., and D. Thirumalai. 2008. Relative stability of helices determines the folding landscape of adenine riboswitch aptamers. J. Am. Chem. Soc. 130:14080-14084. 14 Kononova et al. Mechanical Properties of CCMV Capsid 30. Zhmurov, A., A. E. X. Brown, R. I. Litvinov, R. I. Dima, J. W. Weisel, and V. Barsegov. 2011. Mechanism of fibrin(ogen) forced unfolding. Structure 19:1615-1624. 31. Zhmurov, A., R. I. Dima, Y. Kholodov, and V. Barsegov. 2010. SOP-GPU: accelerating biomolecular simulations in the centisecond timescale using graphic processors. Proteins 78:2984-2999. 32. Zhmurov, A., K. Rybnikov, Y. Kholodov, and V. Barsegov, V. 2011. Generation of random numbers on graphics processors: forced indentation in silico of the bacteriophage HK97. J. Phys. Chem. B 115:5278-5288. 33. Cieplak, M., and M. O. Robbins. 2010. Nanoindentation of virus capsids in a molecular model. J. Chem. Phys. 132:015101. 34. Gibbons, M. M., and W. S. Klug. 2008. Influence of nonuniform geometry on nanoindentation of viral capsid. Biophys. J. 95:3640-3649. 35. Johnson, J.E., and J. A. Speir. 1997. Quasi-equivalent viruses: a paradigm for protein assemblies. J. Mol. Biol. 269:665-675. 36. Johnson, J. M., J. Tang, Y. Nyame, D. Willits, M. J. Young, and A. Zlotnick. 2005. Regulating self-assembly of spherical oligomers. Nano Lett. 5:765-770. 37. May, E. R., A. Aggarwal, W. S. Klug, and C. L. Brooks 3rd. 2011. Viral capsid equilibrium dynamics reveals nonuniform elastic properties. Biophys. J. 100:L59-61. 38. Snijder, J., I. L. Ivanovska, M. Baclayon, W. H. Roos, and G. J. L. Wuite. 2012. Probing the impact of loading rate on the mechanical properties of viral nanoparticles. Micron 43:1343-1350. 39. Scott, S.W. 2006. Bromoviridae and allies. In: Encyclopedia of Life Sciences, John Wiley and Sons, Chichester. 40. Speir, J. A., S. Munshi, G. Wang, T. S. Baker, and J. E. Johnson. 1995. Structures of the native and swollen forms of cowpea chlorotic mottle virus determined by X-ray crystallography and cryo-electron microscopy. Structure 3:68-78. 41. Comellas-Aragonès, M., H. Engelkamp, V. I. Claessen, N. A. J. M. Sommerdijk, A. E. Rowan, P. C. M. Christianen, J. C. Maan, B. J. M. Verduin, J. J. L. M. Cornelissen, and R. J. M. Nolte. 2007. A virus-based single-enzyme nanoreactor. Nat. Nanotechnol. 10:635- 639. 42. Verduin, B. J. M. 1974. The preparation of CCMV-protein in connection with its association into a spherical-particle. FEBS Lett. 45:50-54. 43. Verduin, B. J. M. 1978. Degradation of cowpea chlorotic mottle virus ribonucleic acid in situ. J. Gen. Virol. 39:131-147. 15 Kononova et al. Mechanical Properties of CCMV Capsid 44. Baclayon, M., G. J. L. Wuite, and W. H. Roos, 2010. Imaging and manipulation of single viruses by atomic force microscopy. Soft Matter 6:5273-5285. 45. Kuznetsov, Y. G., and A. McPherson. 2011. Atomic force microscopy in imaging of viruses and virus-infected cells. Microbiol. Mol. Biol. Rev. 75:268-285. 46. Buch, I., S. Kashif Sadiq, and G. De Fabritiis. 2011. Optimized potential of mean force calculations for standard binding free energies. J. Chem. Theory Comput. 7:1765-1772. 47. Kumar, S., J. M. Rosenberg, D. Bouzida, R. H. Swendsen, and P. A. Kollman. 1992. The weighted histogram analysis method for free-energy calculations on biomolecules. I. The method. J. Comp. Chem. 13:1011–1021. 48. Hayward, S., and B. L. de Groot. 2008. Normal modes and Essential dynamics. Meth. Mol. Biol. 443:89-106. 49. Lindahl, E., B. Hess, and D. van der Spoel. 2001. GROMACS 3.0: a package for molecular simulation and trajectory analysis. J. Mol. Model. 7:306-317. 50. Zhmurov, A., R. I. Dima, and V. Barsegov. 2010. Order Statistics theory of unfolding of multimeric proteins. Biophys. J. 99:1959-1268. 51. Tama, F., and C. L. Brooks 3rd. 2005. Diversity and identity of mechanical properties of icosahedral viral capsids studied with elastic network normal mode analysis. J. Mol. Biol. 345:299-314. 52. Yang, L., G. Song, and R. L. Jernigan. 2007. How well can we understand large-scale protein motions using normal modes of elastic network models? Biophys. J. 93:920-929. 53. Zandi, R., and D. Reguera. 2005. Mechanical properties of viral capsids. Phys. Rev. E 72:021917. 54. Fischlechner, M., and E. Donath. 2007. Viruses as building blocks for materials and devices. Angew. Chem. Int. Ed. Engl. 46:3184–3193. 55. Ma, Y., R. J. Nolte, and J. J. Cornelissen. 2012 Virus-based nanocarriers for drug delivery. Adv. Drug. Deliv. Rev. 64:811-825. 56. Ferrara, P., J. Apostolakis, and A. Caflisch. 2002. Evaluation of a fast implicit solvent model for molecular dynamics simulations. Proteins 46:24-33. 57. Amadei, A., A. B. M. Linssen, and H. J. C. Berendsen. 1993. Essential Dynamics of proteins. Proteins: Struct., Funct., Bioinf. 17:412-425. 16 Kononova et al. Mechanical Properties of CCMV Capsid Figure 1. The CCMV capsid: Panel A: AFM image of a CCMV capsid (zmax = 30 nm, scale bar = 20 nm). Panel B: Structure of CCMV (PDB entry: 1CWP); protein domains forming pentamers are colored in blue, while the same protein domains in hexamers are shown in red and orange. Figure 2. Mechanical indentation of the CCMV capsid in vitro (panels A and B) and in silico along two-, three-, and five-fold symmetry axes (panels C and D). Show in different color are the most representative trajectories of forced indentation. The force (F)-distance (Z) profiles (FZ curves) obtained from experimental AFM measurements for vf = 0.6 and 6.0 µm/s are compared with the theoretical FZ curves obtained for vf = 0.5 and 1.0 µm/s (Rtip= 20 nm). The black dash- dotted control lines correspond to the cantilever deforming against the glass surface. The insets to panels C and D show the corresponding FX profiles. Due to the stochastic nature of the mechanical indentation, the values of critical force (F*, force peaks), transition distance (Z*) and indentation depth (X*) are varying. The FZ curves with a single (several) force peak represent single-step (multi-step) indentation transitions. Figure 3. Panel A: Log-linear plot of the spring constant of CCMV capsid kcap versus the cantilever velocity vf. The experimental data are compared with the results of simulations. Panel B and C: Reversible and irreversible deformation of the CCMV capsid obtained experimentally for vf = 6.0 µm/s (panel B), and theoretically for vf = 1.0 µm/s (panel C). Deforming the capsid with a small force of ~0.3 nN (experiment) and ~0.5 nN (simulations) resulted in the reversible mechanical deformation with no hysteresis. Increasing the force beyond ~0.5 nN led to the irreversible deformation - the forward indentation (solid curves) and backward retraction (dotted curves) do not follow the same path (hysteresis). Color denotation is presented in the graphs. Figure 4. Mechanical indentation in silico of CCMV capsid along the two-fold symmetry axis (see panel D and Fig. S1). Show in read and blue color are the two trajectories. The cantilever tip (Rtip= 20 nm) indents the capsid in the direction perpendicular to the capsid surface (vf = 1.0 µm/s). Results for the forward deformation and backward retraction are represented by the solid and dotted curves, respectively; results obtained for vf = 25 µm/s are shown for comparison (dashed black curve). Panel A: The FX curves. The grey line corresponds to the linear fit of initial elastic regime, i.e. X < 3-5 nm. Panel B: Capsid spring constant, kcap versus X. Panel C: Structure overlap ξ versus X: the inset shows the time-dependence of ξ for the backward retraction, which quantifies the progress of capsid restructuring. Panel D: The enthalpy change ΔH and entropy change TΔS from the FX curves generated for vf = 1.0 µm/s (the dashed curve of ΔH generated for vf = 25 µm/s is presented for comparison). The inset shows equilibrium energy change ΔG along the reaction coordinate X from Umbrella Sampling calculations. Also shown are the CCMV capsid structures (top view and profile) for different extents of indentation. The tip-capsid surface contact area shown in red color (see also Fig. S1). Figure 5. Dynamics of the CCMV shell in terms of the non-equilibrium displacement of pentamers (shown in blue) and hexamers (shown in red color). Displayed are the structures for the first two modes of the collective excitations (black arrows), projected along the reaction coordinate (large arrow) in the elastic regime (panel A) and transition regime (panel B) of mechanical compression. For each mode, the upper structure is the reference state. In the lower structure, we showed the type and amplitude of displacement by juxtaposing the representative conformation with the reference state (shown in gray color). 17 Kononova et al. Mechanical Properties of CCMV Capsid Figure 6. Surface map of the potential energy (color scale for USOP is in the graph) for three representative structures of the CCMV shell (top view) observed in the course of deformation in silico at Z = 3 nm (panel A), 18 nm (panel B), and 27 nm (panel C). The direction of motion of the tip is perpendicular to the CCMV surface as indicated. The map shows a gradual increase in the potential energy of proteins forming pentamers and hexamers as changes to the global structure occur. Table I: Mechanical properties of the CCMV capsid from indentation measurements in vitro and in silico: the average spring constant kcap, average critical force F*, and average transition distance Z*. These quantities were calculated by averaging over all FZ curves (all symmetry types). Experimental measurements were performed using the cantilever velocity vf = 0.6 µm/s and 6.0 µm/s; simulations were carried out using vf = 0.5 µm/s and 1.0 µm/s (Fig. 2). The experimental results for vf = 6.0 µm/s and simulation results for vf = 1.0 µm/s are shown in parenthesis. Indentation kcap,N/m F*, nN Z*, nm in vitro in silico 0.17±0.01(0.14±0.02) 0.71±0.08(0.72±0.07) 21.0±3.6(20.8±1.7) 0.11±0.01(0.11±0.02) 0.77±0.03(0.71±0.02) 24.7±2.1(25.5±0.9) Table II: Mechanical and thermodynamic properties of the CCMV capsid from indentation measurements in silico performed along the two-fold, three-fold, and five-fold symmetry axes (Fig. S1): critical force F*, indentation depth X*, spring constant kcap, and thermodynamic functions - Gibbs energy change ΔG, enthalpy change ΔH, and entropy change TΔS. Theoretical estimates of these quantities were obtained by averaging the results of 3 trajectories, using Rtip = 20 nm and vf = 1.0 µm/s (see also Fig. 4, and Figs. S2 and S3). The values of ΔG, ΔH, and TΔS correspond to the total change in these quantities observed at X = 20 nm (Z = 30 nm) indentation. The range of variation of kcap (from Figs. 4, S2, and S3) and the equilibrium estimates of ΔG, ΔH, and TΔS (from Umbrella Sampling calculations) are shown in parentheses. F*, nN X*, nm kcap, N/m Symmetry TΔSind, MJ/mol 0.71±0.02 9.1±1.0 0.11 (0.06-0.14) 4.5(6.9) 11.5(12.8) 7.0 (5.8) two-fold three-fold 0.68±0.02 11.9±0.5 0.10 (0.05-0.10) 5.1(6.8) 11.7(12.6) 6.6(5.8) 0.69±0.02 14.2±0.5 0.12 (0.04-0.12) 4.1(6.5) 12.5(11.5) 8.4(5.1) five-fold ΔGind, MJ/mol ΔHind, MJ/mol Table III: Mechanical and thermodynamic properties of the CCMV capsid from indentation measurements in silico performed at vf = 1.0 µm/s, along the two-fold symmetry axis (Fig. S1) - spring constant kcap, and Gibbs energy change ΔGind, enthalpy change ΔHind, and entropy change ΔSind - are compared for the spherical tips of different radius Rtip = 20 nm, 10 nm, and 5 nm. The estimates of kcap, ΔGind, ΔHind, and TΔSind are obtained from a single FX curve for each different Rtip and correspond to the total change in these quantities observed at X = 20 nm (Z = 30 nm) indentation. Simulation data for Rtip = 20 nm and 10 nm are shown in Fig. 4 and S4, respectively. Rtip, nm kcap, N/m ΔGind, MJ/mol ΔHind, MJ/mol TΔSind, MJ/mol 20 0.090 4.5 11.5 7.0 18 Kononova et al. Mechanical Properties of CCMV Capsid 10 5 0.075 0.069 3.9 1.8 9.6 4.9 5.7 2.2 19 Kononova et al. Mechanical Properties of CCMV Capsid Figure 1 (Kononova, Snijder, Brasch, Cornelissen, Dima, Marx, Wuite, Roos, Barsegov) 20 Kononova et al. Mechanical Properties of CCMV Capsid Figure 2 (Kononova, Snijder, Brasch, Cornelissen, Dima, Marx, Wuite, Roos, Barsegov) 21 Kononova et al. Mechanical Properties of CCMV Capsid Figure 3 (Kononova, Snijder, Brasch, Cornelissen, Dima, Marx, Wuite, Roos, Barsegov) 22 Kononova et al. Mechanical Properties of CCMV Capsid Figure 4 (Kononova, Snijder, Brasch, Cornelissen, Dima, Marx, Wuite, Roos, Barsegov) 23 Kononova et al. Mechanical Properties of CCMV Capsid Figure 5 (Kononova, Snijder, Brasch, Cornelissen, Dima, Marx, Wuite, Roos, Barsegov) 24 Kononova et al. Mechanical Properties of CCMV Capsid Structural Transitions and Energy Landscape for Cowpea Chlorotic Mottle Virus Capsid mechanics from nanomanipulation in vitro and in silico* Olga Kononova,†‡ Joost Snijder,§ Melanie Brasch, ¶ Jeroen Cornelissen, ¶ Ruxandra I. Dima, Kenneth A. Marx, † Gijs J. L. Wuite,§ Wouter H. Roos§* and Valeri Barsegov†‡* †Department of Chemistry, University of Massachusetts, Lowell, MA 01854; ‡Moscow Institute of Physics and Technology, Moscow Region, Russia 141700; §Natuur- en Sterrenkunde and LaserLab, Vrije Universiteit, 1081 HV Amsterdam, The Netherlands; ¶Biomoleculaire Nanotechnology, Universiteit Twente, 7500 AE Enschede, The Netherlands; Department of Chemistry, University of Cincinnati, Cincinnati, OH 45221. Supporting Material *Corresponding authors: e-mail: [email protected], tel: 978-934-3661, fax: 978-94-3013 e-mail: [email protected], tel: +31 20 59 83974, fax: +31 20 59 87991 Running Title: Mechanical properties of CCMV capsid 25 Kononova et al. Mechanical Properties of CCMV Capsid Self-Organized Polymer (SOP) Model Parametrization: Numerical values of εn were determined from equilibrium all-atom Molecular Dynamics (MD) simulations in implicit water for the CCMV shell at T=300 K (CHARMM19 force field) (1). We employed the Solvent Accessible Surface Area (SASA) model of implicit solvation (2). We used the CCMV structure from PDB (PDB code: 1CWP) to carry out a 5 ns equilibrium simulation run for the entire capsid. First, we calculated the number of native contacts, which stabilize the native state of the capsid, based on a standard cut-off distance RC=8 Å between the Cα-atoms. The native contacts were divided into the inter-chain contacts and intra-chain contacts. There were Nintra ≈ 20,554 intra-chain contacts in the capsid proteins, and Ninter ≈ 3,405 inter-chain contacts at their interfaces (due to the van-der-Waals and Coulomb interactions). Next, we calculated the total energy of non-covalent interactions for each contact group. We found Eintra = 25,898 kcal/mol for the intra-chain contacts and Einter = 3,746 kcal/mol for the inter-chain contacts. Finally, we divided the total energy by the number of contacts for each group of contacts. We obtained εintra = Eintra/Nintra = 1.26 kcal/mol for the intra-chain energy, and εinter = Einter/Ninter = 1.1 kcal/mol for the inter-chain energy. These parameters of the SOP force field were used in the simulations of mechanical indentation of CCMV. Langevin simulations of CCMV indentation: The indentation dynamics were obtained by integrating the Langevin equations for each particle position ri in the over-damped limit, ηdri/dt = -∂U/∂ri + gi(t). Here, U is the total potential energy, which accounts for the contribution from the capsid conformation USOP and for the interaction of the i-th particle with the spherical tip Utip(ri) (see Materials and Methods in the main text). Also, gi(t) is the Gaussian distributed random force and η is the friction coefficient. The Langevin equations were propagated with the time step Δt = 0.08τH = 20 ps, where τH = ζεhτL/kBT. Here, τL = (ma2/εh)1/2 = 3 ps, ζ = 50.0 is the dimensionless friction constant for a residue in water (η = ζm/ τL ), and m ≈ 3×10-22 g is the residue mass (3,4). Simulations of mechanical indentation were carried out at room temperature using the bulk water viscosity, which corresponds to the friction coefficient η = 7.0×105 pN ps/nm. Umbrella Sampling Method: We calculated the potential of mean force U as a function of indentation depth X for the process of mechanical compression of CCMV particle. In this approach, the potential energy of the system depends on X (reaction coordinate), which gradually at several values of X (<…> changes during the simulation, and one accumulates denotes the ensemble averaging). The potential of mean force can then be estimated as (5,6). Because the indentation depth X is the distance travelled by the cantilever corresponds to the indentation force (F), i.e. the force exerted on the tip by the virus tip, particle. Since F = κdX, where κ is the cantilever spring constant and dX is the tip displacement, the measurements of F can be used to estimate and to calculate U. We performed one simulation run to collect 280 data points. In each step, the cantilever (base) was moved by 0.1 nm; next, the system was equilibrated for 0.2 ms, and then the values of the tip position were . We used the cantilever spring constant κ= 1.4 sampled for 0.2 ms to calculate N/m. 26 XU/dλλU/XU/XU/XUF/ Kononova et al. Mechanical Properties of CCMV Capsid Normal Mode Analysis: In the NMA, the potential energy is expanded in a Taylor series in terms of the mass-weighted coordinates , where Δxi is the displacement of the i-th particle from the energy minimum and mi is its mass. The Hessian matrix of second derivatives H with information about (i,j=1,2,…,3N) carries the matrix elements equilibrium dynamics of the system in question. The NMA involves the following three steps: 1) minimization of the potential energy, 2) calculation of the Hessian matrix, and 3) diagonalization of the Hessian matrix (7). The energy minimization was performed using the steepest descent algorithm. Each residue was represented by its center of mass. The numerical calculation of the Hessian matrix was i.e. carried forces , out using atomic , where , is the unit vector in the direction of atom j, and h is the displacement along this direction. We used the transformation to obtain the Hessian matrix for the centers of mass HIJ from the Hessian matrix for atomic displacements Hij. Here, MI and MJ are the reduced mass of residue I and J, where I, J=1,2,…,N and N is the total number of residues. To build the Hessian matrix for the displacements of the centers of mass of amino acids, and to solve numerically the eigenvalue problem HR=ΛR, we used an algorithm implemented in the GROMACS package (8). We determined the I-th eigenvalue, λI, and the I-th eigenvector RI. Each eigenvector specifies a normal mode coordinate QI, i.e. , which oscillates with the characteristic frequency . The spectrum of eigenvalues {λI} was used to construct the spectrum of eigen-frequencies in the center-of-mass representation (c is the speed of light). Essential Dynamics: In the Essential Dynamics approach, one assumes that the most important displacements reside in a subspace of a few degrees of freedom, whereas the remaining degrees of freedom represent less important fluctuations (9). Dynamic correlations between particle positions at time t, X(t)={X1(t), X2(t),…, XN(t)}, and the position in the reference (equilibrium average) structure X0={X1(0), X2(0),…, XN(0)} can be expressed through the covariance matrix C(t) = <(X(t) – X0)(X(t) – X0)T>, where <…> denotes the ensemble averaging and the superscript T represents the transposed matrix. By construction, C is a symmetric matrix, which can be diagonalized by an orthogonal transformation T, C = TLTT, where L is the diagonal matrix of eigenvalues and T is the matrix of eigenvectors of C. In the three-dimensional space, there are 3N-6 eigenvectors with non-zero eigenvalues for a system of N particles (excluding translations and rotations). The eigenvalue lI in the center-of-mass representation (I =1,2,…,N) is the amplitude of displacement X(t) – X0 along the I-th eigenvector tI. The principal coordinates PI(t) are obtained by projecting X(t) – X0 onto each eigenvector PI(t) = tI(X(t) – X0). In the Cartesian basis, these projections are given by X(t) = PI(t) tI + X0 (4,6). . 27 iiiΔxm=qjiijqqV=H/2iiqV=f/2/)ehq(feh+qf=Hjijiij}{iqqjeijJIIJH)MM(=H/1IIJIqR=QIIλ=ωIIc)2/1( Kononova et al. Mechanical Properties of CCMV Capsid Figure S1. The CCMV shell: Forced indentation of CCMV along the two-fold, three-fold, and five-fold symmetry axis. The top structures show the top view of symmetry (filled black circle). Pentamers and hexamers are shown, respectively, in blue and red color. The grey structures show the tip-capsid contact area (in red) for different values of Z = 3 nm, 10 nm, 20 nm, and 25 nm and for different symmetry. 28 Kononova et al. Mechanical Properties of CCMV Capsid Figure S2. Forced indentation in silico of the CCMV capsid along the three-fold symmetry axis (see Fig. S1). Shown are the same quantities as in Fig. 4 in the main text, obtained using Rtip= 20 nm, vf = 1.0 µm/s (red and blue curves) and vf = 25 µm/s (dashed black curves), but for a different symmetry type (the top view of symmetry is shown in panel D). These results should be compared with the results of indentation along the two-fold symmetry axis (Fig. 4 in main text), and five-fold symmetry axis (Fig. S3). Also shown are the snapshots of CCMV shell (top view and profiles). The growing with indentation depth tip-capsid contact area is shown in red color (see also Fig. S1). 29 Kononova et al. Mechanical Properties of CCMV Capsid Figure S3. Forced indentation in silico of the CCMV capsid along the five-fold symmetry axis (see Fig. S1). Shown are the same quantities as in Fig. 4 in the main text, obtained using Rtip= 20 nm, vf = 1.0 µm/s (red and blue curves) and vf = 25 µm/s (dashed black curves), but for a different symmetry type (the top view of symmetry is shown in panel D). These results should be compared with the results of indentation along the two-fold symmetry axis (Fig. 4 in main text), and three-fold symmetry axis (Fig. S2). Also shown are the snapshots of CCMV shell (top view and profiles). The growing with indentation depth tip-capsid contact area is shown in red color (see also Fig. S1). 30 Kononova et al. Mechanical Properties of CCMV Capsid Figure S4. Forced indentation in silico of the CCMV capsid along the two-fold symmetry axis type (shown in panel D). Shown are the same quantities as in Fig. 4 in the main text but for a smaller tip of radius Rtip= 10 nm. The force peaks in the FX curves (panel A) are lower (~0.6 nN) compared to the results for Rtip= 20 nm, and the spring constant kcap now varies between 0.03 and 0.11 N/m (panel B). The X-dependence of kcap is still bimodal, as for Rtip= 20 nm, but the second peak at F ≈ 15 nm is now weaker. The structure overlap ξ shows that the CCMV particle in the collapsed state remains ~70-75% similar to the native (un-indented) state (panel C). The associated enthalpy change ΔHind 10 MJ/mol is slightly lower, and the entropy change ΔSind  6 MJ/mol is roughly the same as for the case of Rtip= 20 nm (panel D). Also shown are the snapshots of CCMV shell (top view and profiles). The growing with indentation depth tip- capsid contact area is shown in red color (see also Fig. S1). 31 Kononova et al. Mechanical Properties of CCMV Capsid Figure S5. Equilibrium modes of motion of CCMV particle. The spectrum of normal eigen- frequences (histogram) for isolated single hexamers (shown by gray bars) is compared with the spectrum of eigen-frequences for the entire CCMV shell (blue bars). Structures display the scale of normal displacements from the global modes, which include displacements of chains and entire capsomers (lefthand structure) to local modes of motion of secondary structure elements (α-helices and β-strands, middle structure), and to more localized modes (vibrations of amino acid residues, righthand structure). 32 Kononova et al. Mechanical Properties of CCMV Capsid Figure S6. Surface map of the potential energy (color scale for USOP is in the graph) for three representative structures of the CCMV shell (top view) observed in the course of deformation in silico at Z = 3 nm (panel A), 18 nm (panel B), and 27 nm (panel C). The direction of motion of the tip is perpendicular to the CCMV surface as indicated. The map shows a gradual increase in 33 Kononova et al. Mechanical Properties of CCMV Capsid the potential energy of proteins forming pentamers and hexamers as changes to the global structure occur. Movie S1. Mechanical nanoindentation in silico of CCMV shell along 2-fold symmetry axis. The light blue sphere represents cantilever tip of radius Rtip = 20 nm, moving along -z direction with constant velocity υf = 1.0 μm/s against a mica-surface (not shown). Supporting References: 1. MacKerell, A. D., D. Bashford, M. Bellott, R. L. Dunbrack, J. D. Evanseck, M. J. Field, S. Fischer, J. Gao, H. Guo, S. Ha, D. Joseph-McCarthy, D. T. Nguyen, B. Prodhom, W. E. Reiher, III, B. Roux, M. Schlenkrich, J. C. Smith, R. Stote, J. Straub, M. Watanabe, J. Wiorkiewicz-Kuczera, D. Yin, and M. Karplus. 1998. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B. 102:3586–3616. 2. Ferrara, P., J. Apostolakis, and A. Caflisch. 2002. Evaluation of a fast implicit solvent model for molecular dynamics simulations. Proteins 46:24-33. 3. Barsegov, V., D. Klimov, and D. Thirumalai. 2006. Mapping the energy landscape of biomolecules using single molecule force correlation spectroscopy (FCS): Theory and applications, Biophys. J. 90:3827-3841. 4. Kononova, O., L. Jones, and V. Barsegov. 2013. Order statistics inference for describing topological coupling and mechanical symmetry breaking in multidomain proteins. J. Chem. Phys. 139: 121913-121925. 5. Buch, I., S. Kashif Sadiq, and G. De Fabritiis. 2011. Optimized potential of mean force calculations for standard binding free energies. J. Chem. Theory Comput. 7:1765-1772. 6. Kumar, S., J. M. Rosenberg, D. Bouzida, R. H. Swendsen, and P. A. Kollman. 1992. The weighted histogram analysis method for free-energy calculations on biomolecules. I. The method. J. Comput. Chem. 13:1011–1021. 7. Hayward, S., and B. L. de Groot. 2008. Normal modes and Essential Dynamics. Methods Mol. Biol. 443:89-106. 8. Lindahl, E., B. Hess, and D. van der Spoel. 2001. GROMACS 3.0: a package for molecular simulation and trajectory analysis. J. Mol. Model. 7:306-317. 9. Amadei, A., A. B. M. Linssen, and H. J. C. Berendsen. 1993. Essential Dynamics of proteins. Proteins: Struct., Funct., Bioinf. 17:412-425. 34
1305.4705
1
1305
2013-05-21T04:03:11
Tribological Analysis of Ventral Scale Structure in a Python Regius in Relation to Laser Textured Surfaces
[ "physics.bio-ph" ]
Laser Texturing is one of the leading technologies applied to modify surface topography. To date, however, a standardized procedure to generate deterministic textures is virtually non-existent. In nature, especially in squamata, there are many examples of deterministic structured textures that allow species to control friction and condition their tribological response for efficient function. In this work, we draw a comparison between industrial surfaces and reptilian surfaces. We chose the python regius species as a bio-analogue with a deterministic surface. We first study the structural make up of the ventral scales of the snake (both construction and metrology). We further compare the metrological features of the ventral scales to experimentally recommended performance indicators of industrial surfaces extracted from open literature. The results indicate the feasibility of engineering a Laser Textured Surface based on the reptilian ornamentation constructs. It is shown that the metrological features, key to efficient function of a rubbing deterministic surface, are already optimized in the reptile. We further show that optimization in reptilian surfaces is based on synchronizing surface form, textures and aspects to condition the frictional response. Mimicking reptilian surfaces, we argue, may form a design methodology potentially capable of generating advanced deterministic surface constructs capable of efficient tribological function.
physics.bio-ph
physics
Tribological Analysis of Ventral Scale Structure in a Python Regius in Relation to Laser Textured Surfaces H. A. Abdel-Aal* Arts et Métiers ParisTech, LMPF-EA4106, Rue Saint Dominique BP 508, 51006, Châlons-en-Champagne, France *[email protected] M. El Mansori Ecole Nationale Supérieure d'Arts et Métiers, 151 Boulevard de l'Hôpital, 75013 PARIS, France ABSTRACT Laser Texturing is one of the leading technologies applied to modify surface topography. To date, however, a standardized procedure to generate deterministic textures is virtually non- existent. In nature, especially in squamata, there are many examples of deterministic structured textures that allow species to control friction and condition their tribological response for efficient function. In this work, we draw a comparison between industrial surfaces and reptilian surfaces. We chose the python regius species as a bio-analogue with a deterministic surface. We first study the structural make up of the ventral scales of the snake (both construction and metrology). We further compare the metrological features of the ventral scales to experimentally recommended performance indicators of industrial surfaces extracted from open literature. The results indicate the feasibility of engineering a Laser Textured Surface based on the reptilian ornamentation constructs. It is shown that the metrological features, key to efficient function of a rubbing deterministic surface, are already optimized in the reptile. We further show that optimization in reptilian surfaces is based on synchronizing surface form, textures and aspects to condition the frictional response. Mimicking reptilian surfaces, we argue, may form a design methodology potentially capable of generating advanced deterministic surface constructs capable of efficient tribological function. Nomenclature hp hs Ra Symbols λ λp λs Φ Φs Abbreviations AE-PE BDP-S COF FIEL FAR LL LTS Anterior Posterior Axis Bottom Dead Position Snake Coefficient of Friction Friction-Induced Energy Losses Fibril Aspect Ratio Left lateral direction Laser Textured Surface Fibril row intra spacing (μm) Intra Spacing on Laser Textured Surface Intra Spacing on Snake Skin Dimple Base diameter Dimple Base diameter for snake skin Dimple height Denticulation (fibril) height Average Roughness height   1  Meddle position-Snake Right lateral direction Ventral Scale Aspect Ratio Total Area ratio Top Dead Position Snake Surface Aspect ratio Dimple Slenderness Ratio Length to Circumference Ratio MP-S RL VSAR TAR TDP-S SAR DSR LCR Introduction Friction-Induced Energy Losses, FIEL, of a rubbing system has two contributions. The first is a result of friction between the micro-topography at the interface between the contacting bodies. The second is a consequence of the friction between the lubricants, if present, with the interface. The magnitude of the second component increases upon using a lubricant with high viscosity (which is necessary to support high frictional loads). Reduction of the frictional tractions allows using lubricants of lower viscosities and thereby it reduces the losses due to lubricant friction. Therefore, currently, many efforts address the possibility of engineering topographies in order to improve the quality of surface-interaction in rubbing assemblies. Successful engineering of surface topography, therefore, leads to reduction in the overall FIEL. Ideally, the target is to engineer surfaces that yield predetermined rubbing response, and are, in the same time, capable of self-adapting such response in accordance with changes in sliding conditions. Such surfaces, termed as “deterministic surfaces” comprise artificial textures embossed on the rubbing interface. The texture building block is a micron-sized 3-Diensional geometrical shape (cone, hemisphere, rounded apex, chevron etc.,) which repeats as an array over the desired area of the surface. There are several techniques to produce these textures (e.g., multistep honing, helical-slide honing, controlled thin layer deposition [1-4], and laser texturing [5-9] which is the most advanced and is considered by many as a promising enabling technology [10-13]. Although available since the seventies of the twentieth century application of Laser Texturing to frictional surfaces however, began early this century when it was initially applied to mechanical seals [5-6] then to piston rings and cylinder bores [7,8]. The process involves creation of an array of micro-dimples, either positive (protruding above) or negative (carved into) the target surface using a material ablation process with a pulsating laser beam. Theoretical analysis identified several dimensional groups that influence the tribological performance of a textured surface [15-26]. To date, however, there is no agreement on the optimal values a particular surface should acquire. More importantly, a well-defined methodology for the generation of textures for optimized surface designs is virtually non-existent because of the absence of a holistic surface-design methodology that merges function, form and topography to achieve lean performance. While Surface Design Optimization, in essence, has not matured, as of yet, within the realm of human engineering it is advanced in natural designs especially within the scaled reptiles (squamata). Squamate Reptiles present diverse examples where surface structure, texturing, and modifications through submicron and nano-scale features, achieve frictional regulation manifested in: reduction of adhesion [27], abrasion resistance [28], and frictional anisotropy [29].   2  Squamata comprises two large clades: Iguania and Scleroglossa. The later comprises 6,000 known species, 3100 of which are referred-to as “lizards,” and the remaining 2,900 species as “snakes” [30]. Snakes are found almost everywhere on earth. Their diverse habitat presents a broad range of tribological environments. This requires customized response that manifests itself in functional practices and surface design features. This, potentially, can inspire deterministic solutions for many technical problems. Many authors studied appearance and structure of snakeskin in relation to friction. Results point at the relation of surface topography in snakes to tribological performance [29-41]. The motion of a snake is a delicate balance between the propulsive forces generated by the muscles and the friction tractions due to contact with the substratum. In some cases, the snake makes use of friction to generate thrust. However, for economy of effort, the COF needs to be minimized (especially in rectilinear locomotion) since friction opposes motion. As such, a self- regulating mechanism to control frictional tractions should exist in the snake. The texture of the surface (i.e., the micron sized fibril structures or denticulations) are a major component of such a mechanism. The geometry and topology of the fibril structures allow the snake to modify the frictional profiles in response to changes in contact situations. The presence of the fibrils contributes to the dynamic control of the real area of contact between the skin and the substratum upon sliding [42]. The function of the denticulations (micron-sized fibrils) in this sense is similar to that of the deterministic textures used to regulate friction. Such a similarity raises a curious question. Namely, if we consider the denticulations present on the ventral side of a snakeskin as deterministic textures, would their descriptive parameters coincide with, or at least match, the range of values recommended by researchers based on laser texturing? Furthermore, in case of the validity of such a preposition, can the texturing of a snake inspire a new paradigm in deterministic surface texturing? There are several similarities between a LTS and the ventral side of a snake, both in metrology and topology. The main similarity is the dependence of the metrological features on the scale of observation. That is both of the snakeskin and the LTS belong to the so-called multi- scale surfaces [40]. In addition, the basic building block in each case is a textural element that repeats in an array. The snake’s ventral side is composed of identical micron-sized fibrils (denticulations) distributed over the skin area in a particular pattern. Spacing, length, orientation and shape of denticulation are, in general, common to a particular family of snakes. LTS, on the other hand, by their very definition, comprise an individual textural building block (cone, dimple, chevron etc.,) that also forms an array on the surface. Therefore, both types of surfaces share a common constructal origin. A snake, however, has to be self-sustaining over a wide spectrum of sliding terrains and conditions. The ventral texturing, therefore, has to be efficient over the spectrum of contact conditions that the species experiences. Such a feature offers an advantage to the natural surface over an engineered LTS where the dimples satisfy efficient performance over a narrow domain within possible contact conditions. The question, therefore, becomes how to optimize LTS in order to extend efficient performance, similar to that of a snake. In other words, how to capture the essence of texturing within the ventral side of a snakeskin and then incorporate it into a manmade surface to enhance tribological performance. The answer to this question is rather complex due to the many factors involved. A simple answer is to mimic, or replicate, the textural designs present on the reptile in a technological surface. However, the success of this approach depends on the existence of a one-to-one match between the function and perhaps, the material, of the target surface and that of a reptile. This match will limit the replication process to polymeric surfaces (because the skin is elastomeric).   3  However, bio-inspired surface texturing has a broader goal than merely replication. In essence, it seeks to extend the potential tribological benefits of reptilian surfaces to the domain of man- engineered surfaces. Transfer of design benefits requires developing multi-aspect compatibility metrics in order to evaluate the feasibility of the inspiration process itself (i.e., basing the design of a technical surface on that of a biological analogue). Two features are important in this context: the quality of frictional performance of the snake and matching of the deterministic surface parameters of the biological target to those recommended for the technological surface. Frictional performance of snakes was a subject of several investigations that confirmed the unique features of the ventral skin of snakes and their optimized tribological response with respect to energy losses and resistance to abrasion and wear [36, 42-44]. Other results [43, 45, 46] attribute optimization of tribological function to the geometrical patterns and metrology of the ventral micro-texture. These results, while promising, are to be complimented by assessing the deterministic features of the ventral textures along the same guidelines followed in examining LTS (and manmade surfaces in general). This study, therefore, presents a comparative study of the deterministic surface parameters of the ventral texturing in a snake (Python regius) and those parameters recommended from LTS. The choice of the python species for this study aims enhances the reliability of the comparison results due to the similarities between the mode of motion of this species and the kinematics of many rubbing surfaces. Pythons manifest some of the heaviest constrictor snakes. Their length and weight limits their locomotion to the rectilinear mode [47]. Due to this limitation, the snake depends on continuous frictional adaptation for propulsion. It also depends mainly on the ventral side for propulsion, which implies that the textural denticulations mostly encounter linear friction. This is similar to the mode of friction dominant within many rubbing contacts where linear relative motion, between complying surfaces, takes place. Moreover, the shape of the tips of the denticulations resembles a dimple (spherically capped asperity). Such a shape is the most studied within LTS literature. This work is organized in three parts. The first part details the general appearance and structure of the ventral skin of the Python. The emphasis is on the dimensional metrology of the micron-sized denticulation structures on the ventral side. Following this part, we present the essential deterministic metrology of the texture. The third part of the paper presents the detailed comparison between the parameters of the python and those recommended for LTS. In this part we further, evaluate the feasibility of the idea of building a tribological surface based on Python texturing and we identify several design lessons deduced from the texturing of the snake. 2. Laser Textured Surfaces 2.1 Description of surface Features Surfaces entail features of many different scales. Any surface engineered to meet a predetermined functional requirement, such as enhanced lubrication or reduced frictional response, includes surface features of different sizes and distribution. The distribution of the surface features, along with the type, will also determine the method of describing the performance metrics. Traditional surfaces (surfaces that result from conventional manufacturing processes such as grinding, turning etc.,) the organization of the surface features (height, intervals, size etc.,) is stochastic in essence. Description of surface metrics in such a case stems from signal processing techniques (e.g., spectral analysis, auto correlation functions, root mean square height etc.). The reason being that on a fundamental level, the stochastic elements resulting from the surface generation process dominate the texture. When, on the other hand, a surface contains dominant deterministic patterns (such as tessellations, axially symmetrical   4  patterns, rotationally symmetric patterns) a stochastic description of the surface features becomes meaningless [48]. The conventional approach for describing the compositional features of a surface invokes statistical techniques. Consequently, the distinctive property of any texture descriptor, for a given stochastic surface, is that this descriptor is a statistical variable (i.e., mean, maximum, root mean square, Skew etc.). The value of a characteristic texture feature in this frame is the expected value of the relevant statistical variable (s) or, in some cases, is the relationship between values of two statistical variables. By invoking statistics, one can account for individual variations between all textural features of the same kind present in the studied surface. For example, one can describe the roughness of the surface by accounting for the variation in heights of all the asperities present on the surface. The roughness feature in this case, Ra, will be the statistical average of the heights of all summits present on the surface. In the case of deterministic surfaces, however, the quasi-invariance in the size of surface features renders statistical analysis meaningless. This is because in the limit, the value of the statistical variables, conventionally used to express surface features, will converge to the size of the described feature. In other words, the statistical description of surface features of a deterministic surface will converge to the scales and dimensions of the building elements of the texture (which are deterministic to start with). 2.2 Descriptive metrics for dimpled surfaces For the simplest shape, the dimple, the height of the dimple (hp), dimple base diameter (Φp), and the center line-to-centerline spacing between dimples λ describe the texture of the surface (figure 1). a Di mple (building block of surfac e) Φ p h p λ LTS Textura l e lem en t λ p Figure 1 Definition of the primitive geometrical attributes for Laser Textured Surfaces (a) a sample laser textured (dimpled) surface The height of the dimple (h), Dimple base diameter (Φ), and the center line-to-centerline spacing between dimples λ. Many authors, [5-8, 49-52] use these descriptors to identify three ratios as key performance indicators for the functional quality of LTS. These are the Total Area Ratio (TAR), the Dimple Slenderness Ratio (DSR), and the Surface Aspect Ratio (SAR) (see table 1). Table 1 Definition and calculation formulae of the main parametric relations used to describe the deterministic features of a dimpled LST (all symbols defined in figure 1) Parameter Definition Formula   5  Total Area Ratio Total area of the surface occupied by the texturing element to the total area of the surface Dimple Slenderness Ratio Ratio of the height to the diameter of the (DSR) dimple Surface Aspect Ratio (SAR) Ratio of the centerline-to-centerline spacing between dimples to the height of the dimple Total area of dimples total surface area DSR = SAR = ph Φ p λ + Φ p h p p 3. The Python species Python regius is a non-venomous species native to West Africa. The build of the reptile, figure 2-a, is non-uniform (i.e., the ratio of the body length to the diameter of the body varies along the AE-PE axis. The head-neck region and the tail region are relatively thinner than the main portion of the body (trunk). The trunk comprises the principle load bearing region of the body (i.e., it is where most of the generation of frictional tractions takes place). Table-2 gives a summary of the taxonomic rank of the reptile as well as its’ major characteristics. Table-2 Summary of the taxonomy of the Python regius species and main geometrical features of the reptile Python regius Taxonomy Pythonidae Family Python Subfamily P. regius Genus Body Geometry and dimensions 150 Maximum Length (cm) 208 Number of ventral scales 10.2 Ratio of length to maximum diameter Mass (Kg) 1.3 The ventral side of the reptile comprises hexagonal scales that are elongated along the lateral axis of the reptile (see figure 2-a). The areas of individual scales vary along the AE-PE axis of the reptile. Individual ventral scales, although hexagonal in shape, are not straight edged. Rather their boundaries are arcs and not straight lines (see figure 2-b) and the arcs are curved toward the posterior end. The cross section of the body of the reptile parallel to the Dorso- Ventral axis is almost parabolic with the ventral side protruding outwards. This curved segment, comprising the ventral scale, curve A-A in figure 1-c, supports contact tractions and the weight of the reptile. The length of the curve A-A varies by location and therefore the ratio of the length to the circumference of the reptile varies along the AE-PE axis. The ventral side comprises two distinct regions scales and hinges. The scales are relatively rigid (almost like a membrane), whereas the hinge region is flexible. Scales contain micro fibril structures (also known as denticulations). The hinge region, on the other hand, contains a series of pores. Figure 3, depicts the composition of the scale region (figures 3-b, d, and e) and that of the hinge region (figures 3-a, c, and e) at different magnifications. The denticulations form wave-like rows separated by a distance, λ. The fibrils within the scale region point in the general direction of the posterior end (head-to- tail). Detailed description of the structural features of the ventral scales is given elsewhere [41, 42]. Here, however, we discuss those features pertaining to the current work.   6  Head a Major Longitudinal Bod y A x i s A B M L dorsal scales ventral scales Trunk Dorsa l b RL MLBA A A RL Ven tra l A A LL Tail LL   Figure 2: The ventral side of the snake: a- definition of the axis chosen to describe relative position of the ventral scales on the snake body, b- magnification of the ventral scales located within the mid-section of the reptile. The letter H denotes orientation of the head, c-schematic of a cross section of the body of the reptile along the dorso-ventral axis, and a plan view of a generalized ventral scale (AA is the lateral axis of the scale and BB is the longitudinal axis of the scale). The stocky build of the python regius species manifests distinctive change in the ratio of the snout-to-vent length to the circumference. In this work, we used this physical feature to determine the boundaries of the load bearing volumes on the body of the reptile. To this end, we define three regions. The first region, TDP-S, represents the top boundary of the load-bearing volume within the body. The highest LCR characterizes this region. The second, region, MP-S, represents the medium portion of the reptile body, and is the bulkiest of all regions. The third region, BDP-S, represents the lower boundary of the load-bearing volume of the body. Similar to the TDP-S this region also has a constricted circumference and a high LCR. Table 3 presents a summary of the location of the defined region along the AE-PE-axis and the defining geometrical rations. The geometry of the micron-sized fibrils within the three regions is similar. However, the spacing and density of the fibrils (and thereby the ratio of area occupied by fibrils to the rest of the area of a scale) will also differ in each region. Figure 4, presents two sets of SEM pictures depicting ventral scales located within three regions on the ventral side denoted as TDP-S, MP-S,   7  and BDP-S. The choice of these regions follows from the general build of the reptile. Table 3 presents a summary of the location of these regions on the AE-PE axis along with the LCR, which varies considerably at the beginning of each region. Variation in the LCR implies an analogous variation in bearing the capacity of friction-induced loads of each region. a b x=1000 c x=5000 e x=1000 10 mμ 10 mμ d λ 5 mμ x=5000 5 mμ f     x=10000 1 mμ x=10000 1 mμ   Figure 3 Details of the structure of the ventral scales of the Python regius species (a, c, and e) different magnification of the hinge region, (b, d, and f) details of the “scale” region. Table 3 Summary of non-dimensional location and geometrical features of the body regions shown in figure 4 on the AE-PE Circumference Lat. Diag. X/L Zone (cm) (cm) 0.01 TDP-S 1.5 0.67 MP-S 2.1 0.325 MP-S 1.95 0.775 BDP-S 1.6 6.75 10.5 9.75 7.845 Ratio of length to circumference 22.25 14.2857 15.385 19.125   8  4. Analogy between LTS and Snake skin Snakes use at least five unique modes of terrestrial locomotion (lateral undulation, side winding, concertina locomotion, rectilinear locomotion, and slide pushing). Each of these modes has unique energetic, as well as mechanical, requirements. Although there are distinct kinematic differences between the individual modes of motion, they all share their origin in muscle activity. Transfer of motion between the active muscle groups and the contacting substrate will thus depend on generation of sufficient tractions. The skin handles generation of tractions and accommodation of motion. The skin of a snake while transferring locomotion tractions also has to accommodate the energy consumed in resisting the motion. Muscular activity for locomotion comprises sequential waves of contraction and relaxation of appropriate muscle groups. The number, type, and sequence of muscular groups responsible for the initiation, and sustainment, of motion, and thus employed in propulsion vary according to the particular mode of motion. This implies that contact stiffness in dynamic as well as in static friction constantly varies. Generation of tractions for motion also depends on the habitat and the surrounding environment. This, in turn, affects the effort invested in initiation of motion and thereby affects the function of the different parts of the skin. In figure 4, left hand side pictures reveal fibril structures whereas, right hand side images provide details of the fibrils. The fibrils protrude over the background of the ventral scale and are arranged in waves rather than in orthogonal arrays (perpendicular rows and columns) as often practiced in man-manufactured surfaces. The spacing between fibrils appears to be non-uniform and the size of the individual fibrils seems to vary by location. Due to their relatively heavy weight, pythons use rectilinear locomotion (movement in a straight line). In this mode, the reptile slightly lifts the ventral scales from the ground, pulls them forward, and then downward and backward. However, because the scales "stick" against the ground, the body moves forward over them. Once the body has moved far enough forward to stretch the scales, the cycle repeats. This cycle occurs simultaneously at several points along the body. In rectilinear motion, the main force required is that to overcome external friction and for acceleration of the various regions of the body. An equal and opposite static reaction balances the forward propagation force under the stagnant segments fixed to the ground. Furthermore, the highly developed ventral coetaneous musculature, that generates a peristaltic wave along the snake body, facilitates locomotion. These contraction waves include modulation of both the area and pressure of contact between the skin and the ground [47] through interaction between the ventral skin and the substrate. In particular, the geometrical asymmetry of fibril tips induces precise adaptation of the frictional response [37-39, 41, 42, 53, 54]. AFM imaging of the fibrils, shown in figure 5 (a and b), reveals the elevation of the tips above the overall ventral cell plateau. In this sense, the tips resemble positive dimples raised above the general plane of a textured surface (similar to the micro-dimples created in a laser texturing process). A distinct difference however is the arrangement of the texturing in each case. For the shed skin, the distribution of the fibril tips (dimples) on the scale manifests wave arrangement (as opposed to the orthogonal array arrangement observed in manmade surfaces). The function of texturing is to modify the frictional behavior of the surface upon sliding. In this sense, both manmade and snake surface share the origin of tribological response modifiers: texturing. This observation provides the essence of the comparison between the topological features of the skin of a Python and an LTS.     9  TDP-S MP-S BDP-S Figure 4. SEM images depicting structural details of the ventral scales at three locations, TDP- S, MP-S, and BDP-S respectively. Images on the left hand side are taken at a magnification of X=250 (length marker 100 μm) and those on the right hand side are taken at a magnification X=10,000 (length marker 1μm). Note the arrangement of the fibrils in wave like rows, and the non-uniform distribution of the fibril dimensions. To compare the texture of the snakeskin to that of the LTS, there is a need to define texture descriptors for reptilian skin along the same lines used for LTS. The direct approach is to redefine the LTS descriptors in terms of the surface features present on the snakeskin. To this end, we use the subscript (s) to denote surface texture features present on the snakeskin, where as we use the subscript (p) to denote descriptor of LTS. Therefore, for the snakeskin the DSR is defined as the ratio of the fibril height (hs) to the base diameter of the fibril tip (Φs). The SAR meanwhile is redefined as the distance between two consecutive fibril waves (λs) averaged over the entire body and the protrusion distance. Finally, the TAR for the snake is redefined as the total area of the fibrils to the total surface area. Table 4 presents a summary comparison of the formulas used in computing the additional surface parameters for an LTS and a snake. The total fibril height of the snake, hs, manifests the protrusion of the fibril tip over the cell plateau. This is approximately equivalent to the average roughness of the surface, Ra, along the orientation of the fibrils (i.e., along the AE-PE-axis). According to the preceding, to compare the performance indicators of the shed skin to those of dimpled LTS one needs to calculate the four parametric ratios listed in table 4 for each surface.   10  Note that the descriptors of an LTS depend on the primitive geometry of the surface, those for a shed skin, however, are functions of metrological parameters (dimensional and textural). As such, for the shed skin, characterization of the essential metrological parameters should precede the calculation of the performance indicator ratios. Once the necessary parameters are determined, calculation of the four ratios of table 4 may take place.   a FR2 FR1 n ve ntra l cell plateau Fibril Tips A E-P E μM 4 3 b 2 1 Figure 5: Three dimensional AFM-images of the fibril structures within the ventral scales of the Python regius, a- general scan of a 45 μm x 45 μm area of a ventral scale. Yellow dots mark two fibril rows (FR1 and FR2). Note the protrusion of the fibrils above the general plane of the scale, b- AFM-image of a 5 μm x 5 μm area of the ventral scale. Note the orientation of the fibrils along the AE-PE Axis Table 4: Summary of formulas used to calculate parameters used in comparing LTS and Python Skin Parameter Python skin Python LTS LTS h p h s Total Area Ratio (TAR) Dimple Slenderness Ratio (DSR) Surface Aspect Ratio (SAR) p Φ λ p Total area of dimples total surface area λ Φ s s Total area of fibril tips total surface area DSR = p SAR = p h p Φ p λ + Φ p h p DSR = s R h t s ≈ Φ Φ s λ + Φ s h s s s p SAR = s 5. Materials and Methods All observations reported herein pertain to shed skin obtained from five male Ball pythons (Python regius). Skin shedding in snakes occurs naturally; as such no animals were injured in obtaining the examined skins. All the received shed skin was initially soaked in distilled water   11  kept at room temperature for two hours to unfold. Following soaking, the skin was dried using compressed air and stored in sealed plastic bags. To determine the parameters needed for comparison we identified thirty scales on the shed skin. The chosen scales cover the distance between the first ventral scale and the scale immediately preceding the anal opening of the reptile. The interspacing between the scales is 25 mm (centerline-to-centerline). For each of the chosen scales, we measured the lengths of the longitudinal chord (BB) and the horizontal chord (AA). The measurements then were used to calculate the aspect ratio and the area for each of the ventral scales and the fibrils. For each chosen scale a series of five SEM pictures (at a magnification of x=10,000) at different locations within the particular scale were recorded. Analysis of these pictures yielded fibril geometric information (counts, distance between fibrils, and length of individual fibrils) along the particular locations on the AE-PE axis. To extract data concerning the topography of the skin, we selected several swatches of skin from each of the thirty locations (1500 μm by 1500 μm) for examination using White Light Interferometery (WYKO 3300 3D automated optical profiler system). Analysis of all resulting White light Interferograms, to extract the surface parameters used two software packages: Vision ®v. 3.6 and Mountains® v 6.0. 6. Metrological features of Python skin 6.1 Dimensional metrology As mentioned in section 3, the ventral side of the reptile comprises non-uniform hexagonal unit cells (scales). The layout of these cells is such that the lateral diagonal (i.e., the diagonal in the direction of the lateral axis) is considerably longer than the diagonal in the AE-PE-direction. Moreover, the orientation of the scales is such that the lateral diagonal is perpendicular to the main direction of motion (along the AE-PE-axis for rectilinear locomotion). The variation in the length of the lateral and longitudinal diagonals of the ventral scales implies that the aspect ratio of the building block of the ventral side is greater than unity. It appears, moreover, that the aspect ratio for the ventral scales is not uniform along the AE-PE-axis. Therefore, as a point of entry to metrological characterization we determine the aspect ratio of the ventral scales and the micron-sized fibrils within. Such a step assumes importance in light of observations by other authors that increasing the aspect ratio of a hexagonal padding, above unity, contributes toward considerable reduction of friction [55]. W a b hinge fibr il asp ect ratio l/w l B A ve ntra l scale head   Figure 6 Definition of the geometric parameters used for scale characterization, (a) definition of the surface area of the ventral scale, Avs, and Ventral Scale Aspect Ratio (VSAR), (b) definition of the fibril aspect ratio. For a hexagonal configuration, the Ventral Scale Aspect Ratio (VSAR) may be calculated from:   12  = (1) FAR = VSAR L Lateral L AE PE ( ) − Where, L(AE-PE) is the length of the chord in the direction of the AE-PE axis (line BB in figure 6- a) and Llateral is the length of the chord in the direction of the lateral axis (line AA in figure 6-a). Similarly, the aspect ratio of an individual fibril may be calculated from: L fib W fib Where Lfib is the length of the individual fibril and Wfib is the width of the base of the fibril (see figure 6-b). Figure (7-a) depicts the variation in the aspect ratio of the scale VSAR (blue circles) and the aspect ratio of the fibrils FAR (red circles) along the AE-PE-Axis. Dashed lines accompanying the plots represent the statistical quadratic fit for the data points. In the figure, each data point represents the average of ten measurements from different hides. The raw data does not manifest significant variation. Rather the difference between the maximum and the minimum values is small (2.256 ≤ VS A R ≤ 2.606 (refer to table 5). (2) a Sca le Fib ril 3 2 1 o i t a r t c e p s A 1 .0 b ) x a m . c s A / c s A ( o i t a r a e r a l a n o i s n e m i d - n o n 0 .8 0 .6 0 .4 0 .2 0 .0 0 .0 1 .0 l/L to ta vs no rm_a rea_sca le X /L vs No rma lized_ f ib_ ra t io 0 .8 0 .6 0 .4 0 .2 n o r m a l i z e d f i b r i l a r e a r a t i o ( A f t o t / A s c . m a x ) 1 .0 0 .0 1 .0 0 0 .0 0 .2 0 .8 0 .6 0 .4 Non-dimens ional length of reptile x /L 0 .2 0 .8 0 .6 0 .4 Non-dimensional length of reptile x/L   Figure 7 Distribution of the critical geometrical attributes of the ventral scales along the AE-PE axis. Figure (7-a) distribution of the aspect ratio of the unit scale and the aspect ratio of the fibrils along the AE-PE – axis, Figure (7-b) distribution of the normalized area of the scales and the scale area ratio (area occupied by fibrils to area of scale) along the AE-PE-Axis Table 5 Summary variation of the key geometric attributes of the ventral scales within the leading and the trailing halves of the Python. Position Front Half TDP -S MP -S Trailing Half MP -S BDP -S Geometrical Parameter max min average max min average   13  Area of a ventral Scale mm2 127.25 66.82 104.75 120.447 86.282 103.59 2.61 Aspect ratio of a ventral Scale 2.25 2.47 2.487 2.61 2.549 4.94 2.93 3.63 0.783 6.04 3.46 Fibril density n/mm2 x 10-5 1.3 0.7 1.02 1.3 0.241 0.825 Length of individual fibril μm Fibril aspect ratio 2.6 1.4 2.03 0.482 2.6 1.65 Area of an individual fibril μm2 0.65 0.35 0.51 0.65 0.129 0.41 Area Ratio Fibrils/scale 0.24 0.129 0.18 0.31 0.049 0.14 Note that the distribution of the VSAR is asymmetrical with respect to the AE-PE axis. The values of the VSAR for scales located within the trailing half of the reptile are rather higher than the values pertaining to scales located within the leading half. The average values of the VSAR for each half, however, are almost equal (see table 2). The maximum value for each half, however, is practically invariable, whereas the minimum value for scales in the trailing half is slightly higher than that for the leading half. The maximum overall value of the VSAR belongs to scales located within the region in the middle of the trunk (VSAR ≥ 3) where the highest concentration of mass takes place. Unlike the VSAR, the FAR displays considerable variation along the AE-PE axis. In particular, it drops toward the trailing half of the body. The variation results from the change in the length of the fibrils rather than from the width of the fibril base. Referring to table 2, the maximum fibril length (1.3 μm) is the same within both halves of the body. However, the location of that maximum is different within each half. For the leading half of the body, the maximum length falls almost at the boundary of the thickest part of the trunk (which is approximately the end of the leading half). For the trailing half that maximum value falls at the beginning of that body region (almost at the MP-S section), past this location the length drops considerably. The width of fibrils, meanwhile, is almost invariable. The minimum FAR for the leading half of the reptile is almost three times the corresponding value within the trailing half. Similar to the VSAR, the distribution of the FAR is asymmetric. Fibrils located within the trailing half of the body are stout compared to those located within the leading half of the reptile. Figure 7-b depicts the variation of the area of individual ventral scales along the AE-PE axis. The values plotted in the figure (hexagonal markers) represent the ratio between the area of the particular scale and the area of the largest ventral scale. The later was located at the middle section of the reptile. The dashed line represents a quadratic best fit. Area of scales increases toward the middle section of the trunk. Ventra scales with the largest area are located at the thickest (stockiest) region of the trunk. Thus, the largest area accommodates the heaviest cross section of the reptile. Past the middle section, area of the scales decreases until they reach their minimum value just ahead of the anal opening. The maximum scale area within the leading half of the body is slightly larger than that within the trailing half. The minimum area, however, is smaller for the leading half compared to that within the trailing half (see table 5). Accordingly, ventral scales in the proximity of the tail region are larger than scales located at the proximity of the neck region. For complete characterization, there is a need to find the distribution of the ratio of the area occupied by the fibrils within the particular scale. This ratio is calculated by multiplying the number of fibrils present in a particular scale by the average area of one fibril, then dividing by the total area of the scale. Repeating this process for the chosen ventral scales yields the distribution of that ratio along the AE-PE axis. This distribution is represented by the second plot within figure 7-b (red hexagonal markers and right hand side y-axis). Similar to the   14  behaviour of the area distribution, the ratio of fibril area peaks at the middle section of the trunk. Again, this is where the mass is mostly concentrated. Beyond the middle-section, the ratio Afib/Ascale decreases. The values listed in table 5 indicate that while the average value within the leading and trailing halves of the body are invariable. The maximum and minimum values of Afib/Ascale are different with the area ratio being minimal toward the posterior end of the reptile. As mentioned elsewhere [41], fibrils of maximum length for the python regius were located within three regions: the top and the bottom boundaries of the trunk, and the mid-section. Their lengths were found to fall within the range 1.3 < l< 1.5 μm. The shortest fibrils meanwhile, were located within two regions: the head-neck region, and the trailing end of the load bearing volume. Their length was approximately 0.8 μm. Figure 8 presents a plot of the internal spacing, λ, (distance between fibril rows in μm), as a function of the non-dimensional distance x/l. Particular physical location on the snakeskin may be obtained by comparing the x/l values to entries in table 1. Data plotted in the figure represent the average of five separate measurements on different regions of the same SEM picture. 5 ) m μ ( λ g n i c a p s e l a c s l a n r e t n I 4 3 2 1 0 1 .0 0 .8 0 .6 0 .4 0 .2 0 .0 Non -Dimensiona l Loca tion on the AE -PE Axis (X /L )   Figure 8: Distribution of the inter-spacing between ventral fibril rows along the non- dimensional length of the reptile. (Regression line λ = -2.9149x2 + 2.3876x + 3.948, R2=0.0034) The distribution of the separation distance λ along the body is non-uniform. Internal spacing is larger within the trunk; the maximum is located roughly within mid-section (MP-S). The distance between fibril waves vary between 3.5µm <λ <.4.8µm. The shortest spacing (approximately 3.5µm) is roughly located within the non-load bearing portions of the body (i.e.; the head and tail sections). 6.2 Topographical Metrology of Skin Surface To determine the average profile roughness of the scales we utilized White Light Interferometery. Three scales within each of the skin pre-identified regions (i.e. TDP-S, MP-S, and BDP-S) were chosen at random. For each scale, five White Light Interferograms, WLI, were recorded. Each WLI covered a square area of about 150 μm by 150 μm. Since the size of fibril tips, the focus of the analysis, are of the order of magnitude as small roughness, statistics of roughness profiles would be more pertinent. To extract roughness   15  parameter information there is a need to filter form data out of the WLI. To this end, WLI for each spot on the scales were first treated to remove form data. Thereafter, we evaluated the average roughness, Ra, for small-scale profiles in the lateral and the AE-PE directions. Table 6 lists the extracted Ra values in both directions (lateral and AE-PE). Table 6 Values of the Profile Arithmetic Mean Deviation, Ra (μm), at selected regions within the skin Region AE-PE-Axis SD Lateral-Axis SD TDP-S ± 0.001 0.041 ± 0.01 0.126 MP-S ±0.0072 0.075 ±0.00813 0.06 BDP-S 0.039 ± 0.0085 0.14 ± 0.00624 Average 0.052 ± 0.0056 0.1087 ± 0.0081 The data show that the roughness in the AE-PE direction differs from that in the transverse direction. 7. Results This section presents the results of computing the performance indicators for the skin of the python. Table 7 presents a summary of the computed results. The table includes two data sets. The first pertains to the AE-PE axis. This data set resulted from using the Ra values for profiles along the AE-PE axis. The second set meanwhile resulted from using the Ra values of profiles along the lateral axis. It is noted that the TAR value for both directions is equal. This is a result of the definition of the TAR parameter, which in essence is the percentage of the area occupied by the fibrils within the area of the scale. This is not direction dependant. Values of the DSR and the SAR manifest some difference. For each of the computed quantities we also included the average value (entries in the last raw of table 7). Table: 7 values of the computed performance indicators along the two main axis of the skin the AE-PE and the lateral axis AE-PE-Axis Lateral Axis Region Ra TAR SAR DSR Ra DSR TAR SAR 0.041 TDP 55 0.21 0.1 0.126 0.13 43.65 0.21 0.076 MP 58.51 0.26 0.076 0.06 91.67 0.26 0.06 BDP 0.039 0.1 56.12 0.039 0.14 0.1 39.28571 0.14 Average 0.052 0.19 56.54 0.072 0.109 0.19 58.20 0.109 The slight difference between the computed values along the axes is rather deceptive. This is because of the sensitivity of the data and the original size of the surface features involved. To illustrate this point we introduce figure 10 (a and b). The figure is a plot of the effect of the textural anisotropy on the computed performance parameters. Values at the abscissa represent the textural anisotropy in Ra (defined as the ratio of Ra, in the lateral direction, to that in the AE-PE direction). The quantity represented in the ordinate differs according to the figure. In figure 9-a, the ordinate depicts data for the SAR anisotropy ratio (defined as SAR in the lateral direction to the SAR value in the AE-PE direction). For figure 9- b, however, the ordinate depicts data for the DSR anisotropy ratio (defined as DSR in the lateral direction to the DSR value in the AE-PE direction). In both plots, the dashed lines represent a quadratic regression function. The symbol labeled “average” in each plot marks the average value of the particular ratio of anisotropy, which was computed, based on average values.   16  The DSR anisotropy ratio manifests a trend that contrasts that of SAR anisotropy ratio. In particular, the DSR ratio increases with textural anisotropy whereas SAR anisotropy drops with textural anisotropy. A question that arises given the difference in directional values is: which of the data sets to be used upon comparing the textural construction of the skin to recommended optimal values for LTS? The answer depends on the principle direction of motion of the surface under consideration. Pythons are heavy snakes. Due to their weight, they move via rectilinear locomotion where the displacement is mainly along the AE-PE axis of the reptile. As such, the use of data along the AE-PE axis in all comparisons is appropriate. 5 1 .75 a b ) E P _ E A R A S / . t a L R A S ( y p o r t o s i n a 1 .50 1 .25 1 .00 0 .75 ave rage 4 3 2 1 ) E P _ E A R S D / . t a L R S D ( y p o r t o s i n a ave rage 0 .50 0 3 2 0 1 4 3 2 0 1 4 Ra Ra Ra Ra / / (AE- PE) (AE- PE)   (l ater al ) (l ater al ) Figure 9 Variation in the Ratio of Anisotropy for the SAR and DSR with textural anisotropy 8. Discussion 8.1 Comparison to Laser textured surfaces In this section, we draw a “performance prediction” comparison between the surface of the ventral side of the python and LTS. The scope of the comparison at this stage of the investigation is confined to comparing the numerical ranges of the performance indicators as computed from the geometry of the skin to those reported in tribology literature for LTS The comparison process entails constructing performance indicator maps (whenever possible) from available data. Thereafter, we superimpose the computed performance indicators for the skin on these maps. The performance indicators for the skin of the snake, as mentioned earlier, are functions of metrological parameters (both dimensional and topographical). The metrological parameters, on the other hand, differ by position and direction of profile extraction (refer to table 6). As such, performance indicators for the snakeskin are functions of the evaluated region on the body of the reptile. To generalize the comparison, we compute four variations of the skin performance indicators (and thereby we superimpose these on the constructed maps). As such, we evaluate the performance indicators for each of the pre-identified regions on the skin (TDP-S, MP-S, and BDP-S) using local data in addition to an average value for each of the comparison ratios. This average value is a function of the averaged metrological values. In this manner we identify which region, if any, is more suitable for in depth analysis that leads to surface-mimicry (i.e., simulations and experimental friction studies for surface replicas). The starting point of the devised procedure is to compare the shed skin to the experimental parameters of Kovalechenko et al [56] who performed extensive experiments to rate the tribological performance of LTS.   17  Kovalchenco and co-workers used a speed range of 0.15-0.75 m/s and nominal contact pressures that ranged from 0.16 to 1.6 MPa. They also used two lubrication oils with different viscosities (54.8 and 124.7 cSt at 40 o C). Table 3 presents a summary of the parameters describing the surfaces used in these and the equivalent parameters for the snakeskin. Table 8 Surface designation and parameters used in the experiments of Kovalchenko et al [56] Description Surface S-3 Standard LTS (SLTS) Higher Dimple Density (HDD) S-4 Standard unlapped (SU) S-5 Lower dimple density (LDD) S-6 P-AV Python skin based on averaged metrological quantities Parameter Python S-6 S-5 S-4 S-3 (average) LDD SU HDD SLTS 0.076 4 6.5 5 5.5 Depth of dimples h (μm) 0.05 0.09 0.07 0.06 0.03 Surface roughness between dimples Ra (μm) 1.0 58 80 58 78 Diameter of dimples Φ (μm) 4.5 200 200 80-100 200 Distance between dimples λ (μm) Dimple area density 0.1 0.7 0.12 0.15 0.12 Figure 10 presents a summary of the results obtained by Kovalchenko and co-workers. The figure depicts evolution of the coefficient of friction (COF) of the LTS used in the experiments. Data pertain to lubricated sliding in the presence of high viscosity motor oil. From the figure, we can distinguish the general trend of the COF. Depending on the sliding speed, two regimes take place upon sliding. The first is characterized by high friction associated with low load carrying capacity. The second regime meanwhile takes place at higher sliding speeds. Here the fluid film builds up to reach a thickness capable of establishing a hydrodynamic lubrication regime. Upon increasing the sliding speed, the thickness of the fluid film increases and the load carrying capacity increases consequently. This causes the friction to drop. 0.14 D isc 4 high density dimples D isc 5 dimpled unlapped D isc 6 low density dimples D isc 3 dimpled t n e i c i f f e o c n o i t c i r F 0.12 0.1 0.08 0.06 0.04 0.02 0 0.75 m/s 0.015 m/s 0.0375 m/s 0.075 m/s 0.0125 m/s 0.15 m/s 0.3 m/s 0.6 m/s 0.45 m/s 0 360 1080 1260 1440 1620 900 720 540 180 Sliding time (sec) and speed (m/s) Figure 10: Friction behavior of the surfaces used by Kovalchenko et al [56]   The value of the COF and the rate of transition to a full hydrodynamic lubrication regime constituted the criterion used to rate the tested surfaces. Accordingly, a better performing surface, thus, would exhibit low COF, and will transit to a full hydrodynamic regime at a sliding   18  speed lower than that of a poor performing surface. As shown in figure 10, the best performing surface is surface S-3 since it exhibits the lowest COF. Surface number S-6 exhibited the second lowest COF. As such, it was rated the second best performing surface and so on. In principle the ranking of the surfaces depends on the ability of the lubricating oil to establish a film of sufficient thickness capable of supporting the contact loads and whence separating the rubbing surfaces. Ability of the lubricant to build a suitable film depends on the optimization of texturing manifested in optimal values of performance indicators. Naturally, not all parameters are equally influential on lubrication quality and friction reduction. One may classify surface geometrical parameters in two categories: primary and auxiliary. Primary parameters directly influence the quality of lubrication; meanwhile auxiliary parameters play a role only when two surfaces share the same value of a primary parameter (or primary parameter set). For example, in table 6 surfaces S-3 and S-6 share the same distance between dimples, yet the COF exhibited by surface S-3 is less than that exhibited by surface S-6. This indicates that the better performance of surface S-3 relates to the difference in other parameters (dimple area density and dimple diameter for example). Similarly, surfaces S-3 and surface S-5 share the same dimple area density (0.12). Yet, surface S-3 produces a lower COF than surface S-5. So that the different values of the roughness parameter, Ra, and the diameter between the two surfaces for example should be the source of the contrast in performance. As such, complete mapping of the performance of tested surfaces in relation to their respective individual surface parameters (both primary and auxiliary) allows predicting the performance of untested surface. To illustrate this point, suppose that we construct a Ra-TAR map for which the respective parameters of the surfaces given in table 6 constitute the entries. Suppose, further, that we plot the Ra and TAR values of a surface of unknown performance on the same map. Then the ranking of the unknown surface will depend on its location with respect to surfaces S-3 and S-6. If the new surface falls between surfaces S-3 and S-6, then it is likely to perform better than surfaces S-4 and S-5. The overall ranking of the new surface also depends on how close it is located to surface S-3 on the Ra-TAR map. Repeating such a process with maps constructed by plotting the combination of parameter pairs of the surfaces in table 3, allows ranking of the unknown surface. In what follows we apply this procedure to rank the projected tribological performance of the skin. Tables 7, 8, and 9 contain the data necessary to rank the python skin with respect to the test surfaces used by Kovalechenko. Table 9 value of parameters used in comparing LTS and Python Skin Surface DSR TAR SAR 50.5 0.12 0.0705 S-3 31.6 0.15 0.0862 S-4 S-5 0.0812 0.12 43 64.5 0.07 0.0689 S-6 56.54 0.1753 0.076 Python (average) 55 0.214 0.1 TDP-S MP-S 0.07 0.26 58.51 BDP-S 56.122 0.1 0.039 Figure 11 (a-d) depicts four maps: Ra versus DSR (figure 11-a), DSR versus SAR (figure 11-b), Ra versus Total Area Ratio TAR (figure 11-c), and SAR versus TAR (figure 11-d). Figure 11-a shows that the Ra parameter is not a principal performance influence parameter. The DSR value is what primarily distinguishes between the better performing surfaces and those with lower   19  performance. Surfaces S-3 and surface S-6 fall approximately on the same DSR coordinate (the DSR values for both surfaces are so close-refer to table 8). Being almost equal, the Ra value would be the value that determines the quality of tribological performance. In this sense, the Ra value is an auxiliary influence parameter. The general trend exhibited by the data is that the performance is inversely proportional to the Ra value when the DSR values are equal. Consequently, if the snakeskin is to exhibit better performance it should be located almost on the same coordinate as S-3 and S-6. This is indeed the case (observe the location of the hexagonal symbol in the figure which denotes the average value for the python). Moreover, the python skin falls closer to S-3 than to S-6 (due to higher Ra than that of S-6) whence it is projected to perform better than S-6. To this end, a surface that mimics the python surface should fall second to S-3 under the same test conditions. Following the same criterion, the region, on the skin, that closely matches the better surface performance band is that for the mid-region MP-S. Compared to the average value, P-AV, however, the performance of a surface mimicked after the texture of the MP-S is predicted to be less efficient than S-3. Surfaces constructed after the textural features of the TDP-S region and the BDP-S regions fall out of the band of acceptable performance. The DSR in itself, however, may act as an initial filter for surfaces when mapped against the SAR, figure 11-b. Here, a narrow band of SAR values designates the best performing surfaces. Surfaces S-4 and S-5, for which the tribological performance is of lesser quality, fall at the higher end of the DSR axis. A linear fit of the data (the straight line shown in the figure) implies that the quality of performance is directly proportional to the SAR (the higher the SAR the better the performance). Interestingly, the snakeskin (pentagon symbol) is closer to S-3 and falls directly on the linear fit. Similar to the trend reflected in figure 11-a, among all regions within the skin texture within the MP-S region is predicted to offer the best performance. Upon considering the variation in the Total Area Ratio (TAR) with the Ra value (figure 11-c) it is noticed that a lower Ra value is a precondition for better performance for equal TAR values. Surfaces S-3 and S-5 share the same TAR value (0.12), yet S-3 exhibited better performance in the experiments. The skin of the Python, square symbol, falls close to S-3 (star symbol). Note that the Ra value for the Python is smaller than that of S-6. When, however, the SAR is plotted against the TAR (figure 13-d), most of the snake textures fall out of the performance band except for the BDP-S region, which falls almost mid-distance between S-3 and S-6. The BDP-S region, moreover, seems to be a good fit to the linear regression of the data. In all, in the four plots the python skin always fell between the two highly ranked surfaces (S-3 and S-6). The python surface (whether localized or average), additionally, was mostly located closer to S-3 than S-6. Very good agreement with the linear fit of the data presented in figure 11-b was observed. Such an agreement implies feasibility of more in-depth investigation of LST that mimics the construction and geometry of the python surface. 8.2 Comparison to recommended values Shinkarenko and co-investigators [57] provided optimal ranges for key performance indicator parameters. They deduced the range of an optimal DSR to fall within the interval 0.06<DSR< 0.08. In comparison, the equivalent value for the Python, see table 7, is 0.076. Further, they recommended a TAR value within the bounds 0.05 <TAR<0.5 with 0.2 being predicted as the optimal recommended value. The analogous value for the Python, again from table 7, is 0.175. Costa and Hutchings [49] extensively investigated the influence of surface topography on lubricant film thickness. They measured film thickness using a capacitance technique in   20  S-4 S -5 S-3 P-AV MP -S S-6 BDP -S 0 .02 0 .04 0 .06 0 .08 Average Roughness, Ra ( m)μ c MP -S TDP -S P -AV S-3 S -4 S-5 BDP -S S-6 ) R A S ( o i t a R t c e p s A e c a f r u S 60 50 40 30 20 10 0 .10 0 0 .000 70 60 50 40 30 20 10 ) R A S ( o i t a R t c e p s A e c a f r u S ) R S D ( o i t a R s s e n r e d n e l S e l p m i D 0 .10 0 .08 0 .06 0 .04 0 .02 0 .00 0 .00 0 .30 0 .25 0 .20 0 .15 0 .10 0 .05 ) R A T ( o i t a R a e r A l a t o T 0 .00 0 .000 reciprocating sliding of patterned plane steel surfaces. The study employed circular patterns (pockets), grooves, and chevrons. They also varied the sliding orientation relative to the texture. The results of Costa and Hutchings concluded that a TAR of 0.11 seems to achieve the maximum film thickness for circular pockets. It is interesting to note that the equivalent ratio for the Python (0.1) fits closely to that recommended value. The same authors also found that a sample with a TAR ratio of 0.06 produced maximum film thickness, when the DSR was about 0.07. Consequently, they rationalized that such a value is optimal for texturing. Again, in comparison, the average value for the Python (0.076) falls very close. 70 0 .12 a b S-6 TDP-S TDP -S S -3 MP -S P -AV S-5 BDP-S S-4 0 .075 0 .050 0 .025 D imp le Slenderness Ratio (DSR) 0 .100 d P-AV MP-S TDP -S S-6 BDP-S S -3 S-5 S -4 0 0 .00 0 .05 0 .100 0 .10 0 .15 0 .20 0 .025 0 .075 0 .050 Tota l Area Ratio (TAR) Average Roughness, Ra ( m)μ Figure 11 Comparison maps between snake skin and Laser Textured Surfaces, a- Ra versus DSR b-DSR versus SAR,c- Ra versus Total Area Ratio TAR, d-SAR versus TAR. Table 10 presents a collective comparison between the performance indicator values computed for the snakeskin and those recommended by several authors. To visualize the extent of agreement between the skin values and the recommended values data in table 9 we present figure 12. The figure depicts plots for three performance indicators: TAR, DSR, and SAR (note that 0 .25 0 .30   21  the figure plots the reciprocal of the later quantity for compatibility with literature values). The plot indicates that the values for the python skin are in good agreement with the recommended values. Table 10 Summary of recommended performance indicators extracted from open literature TAR DSR SAR Reference Shinkarenko et al [57] 0.05-0.5 0.06-0.08 Costa and Hutchins [49] 0.11 0.025-0.1 0.07 0.05 Dongsheng et al [50] Ronen et al [8] 0.05-0.2 Halperin et al [58] 0.02-0.25 0.25 Wang et al [59] 0.01-0.02 0.05 0.03-0.12 Yu and Zhou [60] Gonzalo et al [61] 0.6 Galda et al [62] 0.075-0.2 0.03-0.08 0.085 Hu, Hu [63] Yin et al [64] 0.03-0.1 0.2-0.4 Python (average) 0.017 0.076 0.1753 TDP-S 0.0182 0.1 0.214 0.26 0.07 0.017 MP-S 0.1 BDP-S 0.018 0.039 The extracted data were obtained under different conditions (different materials, different shapes of textural elements, different lubricants and lubrication conditions, etc.,). Despite these differences, it is remarkable that most of the recommended values converge to a relatively narrow band of optimal values. More remarkable is the fit of the computed snakeskin values within this narrow band, which points at the optimal construction of the topographical features of the reptile. Such a finding encourages additional in-depth studies of surface design of other snake species. 0 .6 0 .5 0 .4 0 .3 0 .2 0 .1 TAR Python DSR 1 - R A S n o h t y P n o h t y P 0 .020 0 .015 0 .010 0 .005 0 .0 0 .000   Performance Ind ica tor Parameter Figure 12 Summary plot to compare the performance indicators recommended in literature and those computed for the ventral skin of the Python.   22  Mimicry of reptilian surfaces does not suffice by itself for constructing a surface of optimal tribological quality. Rather it is the deduction of design rules, combined to observing reptilian dimensional proportionality that contributes to successful texturing. Observing the layout of dimples on many engineering surfaces reveals an array arrangement in which dimple arrangement is equispaced and is in patterns of parallel rows and columns. Equal spacing is hardly observed, if at all existent, in snake ventral scales (which are the principal frictional frontier in snakes). Rather, a wave arrangement is present in all sliding directions. An additional observation is the orientation of the fibrils along the body. The orientation of the fibrils along the AE-PE axis always points toward the posterior end. The fibrils are not necessarily parallel to the AE-PE axis. This orientation contrasts the orientation of the hexagonal ventral scales (in which the lateral diagonal is always perpendicular to the AE-PE axis). Combined to the waveform distribution of spacing, the orientation features of the fibrils enhance both the friction and wear performance of the skin. Yuan et al [71] performed exhaustive studies on the effect of textural orientation on friction and wear behavior of surfaces. The authors used several LTSs in lubricated sliding. They also varied the contact pressures and the pressure of lubricant supply. Yuan and coauthors concluded that the orientation of the micro- textural elements has a strong influence on friction performance of sliding surfaces. For relatively low contact pressure small sized grooves (order of few microns), that are oriented perpendicular to the sliding direction, reduce friction by about 40% compared to un-textured surfaces. Under high contact pressure, bigger grooves (order of 20 μm), but parallel to the sliding direction, yield better friction performance and reduced contact stresses. The best performance reported in the work of Yaun pertained to textural elements oriented at an angle to the principal direction of motion (regardless of size). This orientation resulted in reducing the friction by a factor of two. It is interesting to compare the findings of Yaun and coworkers to the texturing pattern observed on the ventral side of the Python (figures 3-b, d, and f and figure 4). The distribution of the fibrils, as revealed from the SEM photographs, is a wave of varying amplitude and period. More important, however, most of the fibrils make an angle with the AE- PE axis, which matches the findings of Yuan, and point at the possibility that ventral structure of the python is optimal for tribological function. Naturally, to generalize such a statement there is a need for extensive work to characterize the local friction response of multiple species and compare it to the ventral structure. Such an effort is currently undertaken in our laboratory. Conclusions In this work, we presented a comparison between the structure and geometrical metrology of the ventral skin of Python regius and industrial Laser Textured surfaces. Dimensional and metrological performance parameters, influential to efficient tribological function of textured surfaces, were found to be optimal in the case of the reptile. This, points at the feasibility of investigating more bio-analogues for developing a standard of performance that indexes the various Laser Textures currently in application. Several fundamental differences in both geometry and construction of the compared surfaces were noted. In particular, within the reptilian surface, arrangement of surface motifs is a- periodic, asymmetric and follows a wave pattern. This allows the surface to condition its tribological response upon sliding. Such an arrangement is not practiced in manufactured surfaces where dimples are positioned in matrices formed of perpendicular rows and columns. The ability of reptilian surfaces to self-adapt in response to changes in sliding conditions seems to originate from the holistic optimization that considers surface form, topography and   23  morphology. This consideration may form a foundation for a generation of deterministic surface textures in fabricated surfaces. References [1] M. Priest, C. M., Taylor, 2000 Automobile engine tribology-approaching the surface, WEAR 241, 193-203. [2] E. Willis, 1986 Surface finish in relation to cylinder liners, WEAR 109, 351–366 [3] N.W. Bolander and F. Sadeghi, 2007 Deterministic modeling of honed cylinder liner friction, Trib Trans 50, 248–256 [4] Santochi M., Vignale, M., 1982 A study on the functional properties of a honed surface. Ann. CIRP 31, 431–434. [5] Etsion I, Kligerman Y, Halperin G. 1999 Analytical and experimental investigation of laser-textured mechanical seal faces Trib. Trans 42, 511–6. [6] Etsion I, Halperin G. A 2002 laser surface textured hydrostatic mechanical seal. Trib. Trans. 45:430–4. [7] Ryk G, Kligerman Y, Etsion I. 2002 Experimental investigation of laser surface texturing for reciprocating automotive components. Trib. Trans 45, 444–449. [8] Ronen A, Etsion I, Kligerman Y. 2001 Friction-reducing surface texturing in reciprocating automotive components. Trib. Tran. 44, 359–66. [9] Golloch, R., Merker, G. P., Kessen, U., and Brinkmann, S., 2004 Benefits of Laser- Structured Cylinder Liners for Internal Combustion Engines, in Proceedings of the 14th Int. Colloquium Tribology, 321–328. [10] Dumitru, G., Romano, V., Weber, H. P., Gerbig, Y., Haefke, H., Bruneau, S., Hermann, J., and Sentis, M., 2004 Femtosecond Laser Ablation of Cemented Carbides: Properties and Tribological Applications, Appl. Phys. A: Mater. Sci. Process., 79, 629–632. [11] Hiroki Yamakiri, Shinya Sasaki, Tsuneo Kurita, Nagayoshi Kasashima, 2011 Effects of laser surface texturing on friction behavior of silicon nitride under lubrication with water, Trib Int. 44, 579-584 [12] N Hu, Qi Ding, Effective solution for the tribological problems of Ti-6Al-4V: 2012 Combination of laser surface texturing and solid lubricant film, Surface and Coatings Technology 206, 5060-5066. [13] Borghi, A., Gualtieri, E., Marchetto, D., Moretti, L, Valeri, S, 2008 Tribological effects of surface texturing on nitriding steel for high-performance engine applications, WEAR, 265, 1046-1051 [14] Qiu, Y., Khonsari, M.M., 2011 Experimental investigation of tribological performance of laser textured stainless steel rings, Trib. Int., 44, 635-644 [15] Brizmer, V. Kligerman, Y. 2012, A laser surface textured journal bearing, J of Trib., 134, .  http://dx.doi.org/10.1115/1.4006511 [16] Eichstät, J., Römer, G.R.B.E., Huis in't Veld, A.J., 2011, Towards friction control using laser-induced periodic Surface Structures, Physics Procedia 12, 7-15 [17] Kovalchenko, A., Ajayi, O., Erdemir, A. & Fenske, G. 2011, Friction and wear behavior of laser textured surface under lubricated initial point contact, WEAR, 271, 1719-1725. [18] Liu, H., Han, H., Xue, Y. Li, J. 2010, Influence of laser surface texturing distribution patterns on the hydrodynamic lubrication. Advanced Materials Research, 97-101, 1429 [19] Mann, M., Zum Gahr, K. 2012, Effect of a contact area texturing under unidirectional sliding load dependent on the viscosity of the liquid lubricant, Tribologie und Schmierungstechnik, 59, 35-40.   24  [20] Marian, V.G., Gabriel, D., Knoll, G., Filippone, S., 2011, Theoretical and experimental analysis of a laser textured thrust bearing, Trib. Lett. 44, 335-343. [21] Podgornik, B., Vilhena, L.M., Sedlaček, M., Rek, Z., Žun, I. 2012, Effectiveness and design of surface texturing for different lubrication regimes, Meccanica, 47, 7, 1613-1622. [22] Scaraggi, M. 2012, Textured Surface Hydrodynamic Lubrication: Discussion, Trib. Lett., doi: 10.1007/s11249-012-0025, 1-17. [23] Scaraggi, M., Mezzapesa, F.P., Carbone, G., Ancona, A., Tricarico, L. 2013, Friction Properties of Lubricated Laser-MicroTextured-Surfaces: An Experimental Study from Boundary- to Hydrodynamic-Lubrication, Trib. Lett. 49, 117-125. [24] Vilhena, L.M., Podgornik, B., Vižintin, J., Možina, J., 2011, Influence of texturing parameters and contact conditions on tribological behaviour of laser textured surfaces, Meccanica 46,. 567-575. [25] Yin, B., Li, X., Fu, Y., Yun, W. 2012, Effect of laser textured dimples on the lubrication performance of cylinder liner in diesel engine, Lubrication Science, 24, 293-312. [26] Zhan, J., Yang, M., 2012 Investigation on Dimples Distribution Angle in Laser Texturing of Cylinder-Piston Ring System, Trib. Tran. 55,693-697. [27] Arzt, E., Gorb, S., and Spolenak, R. 2003, From Micro to Nano Contacts in Biological Attachment Devices, Proceedings of National Academy of Sciences (PNAS), 100 19,10603. [28] Rechenberg, I. 2003 Tribological Characteristics of Sandfish, in Nature as Engineer and Teacher: Learning for Technology from Biological Systems, Shanghai, Oct. 8-11. [29] J. Hazel , M. Stone, Grace, M. S., Tsukruk, V. V, 1999 Nanoscale design of snake skin for reptation locomotions via friction anisotropy, J of Biomechanics, 32, 477-484. [30] Vitt, L.J., Pianka, E.R., Cooper Jr, W.E., Schwenk, K, 2003 History and the global ecology of squamate reptiles, Am. Nat. 162, 44–60. [31] Cheng Chang, Ping Wu, Baker, R, E., Maini, P, K., Alibardi, L, Cheng-Ming, Chiasson, R. B, Bentley D. L, Lowe C. H., 1989 Scale morphology in Agkistrodon and closely related crotaline genera. Herpetologica 45, 430–438 [32] Alibardi, L., Thompson, M. B. 2002. Keratinization and ultrastructure of the epidermis of late embryonic stages in the alligator (Alligator mississippiensis). J. Anat. 201, 71-84. [33] Ruibal R 1968 The ultrastructure of the surface of lizard scales Copeia 4, 698–703. [34] Jayne, B.C. 1988-a Mechanical behavior of snake skin. J. Zool., London 214, 125-140. [35] G. Rivera, A. H. Savitzky, J A. Hinkley, 2005 Mechanical properties of the integument of the common gartersnake, Thamnophis sirtalis (Serpentes: Colubridae), J of Exp. Biology 208, 2913-2922. [36] Berthé, R. A., Westhoff, G., Bleckmann, H., Gorb, S. N., 2009 Surface structure and frictional properties of the skin of the Amazon tree boa Corallus hortulanus (Squamata, Boidae) J Comp Physiol A Neuroethol Sens Neural Behav Physiol.;195, 311–318. [37] M. Saito, M. Fukaya, T. Iwasaki, 2000 Serpentine locomotion with robotics snakes, IEEE Control Syst. Mag. 20, 64–81 [38] M. Shafiei, A. T. Alpas, 2008 Fabrication of biotextured nanocrystalline nickel films for the reduction and control of friction, Materials Science and Engineering: C 28, 1340-1346. [39] M. Shafiei, A. T. Alpas, 2009 Nanocrystalline nickel films with lotus leaf texture for superhydrophobic and low friction surfaces, Applied Surface Science, 256, 3, 710-719.   25  [40] Abdel-Aal, H. A, El Mansori M, Mezghani S, .2010 Multi-Scale Investigation of Surface Topography of Ball Python (Python regius) Shed Skin in Comparison to Human Skin, Trib Lett., 37, 3, 517-528 [41] Abdel-Aal H. A., El Mansori M., 2011 Python Regius (Ball Python) Shed Skin: Biomimetic Analogue for Function-Targeted Design of Tribo-Surfaces, In: Biomimetics - Materials, Structures and Processes. Examples, Ideas and Case Studies, Eds: Bruckner D., Gruber P., Springer, ISBN: 978-3-642-11933-0, [42] Abdel-Aal, 2012 On Surface Structure and Friction Regulation in Reptilian Locomotion, J. Mech. Behav. Biomed. Mat. DOI 10.1016/j.jmbbm.2012.09.014 [43] H. A. Abdel-Aal, R. Vargiolu, H. Zahouani, M. El Mansori, Preliminary Investigation Of The Frictional Response Of Reptilian Shed Skin, WEAR 290–291 (2012) 51–60 doi./10.1016/j.wear.2012.05.015 [44] Benz, M.J., Kovalev, A. E., Gorb, S.N. 2012. Anisotropic frictional properties in snakes. Proc. of SPIE 8339, 83390X, doi: 10.1117/12.916972. [45] Stewart, G.R. and R.S Daniel. 1973. Scanning electron microscopy of scales from different body regions of three lizard species. J. Morph. 139, 377-388. [46] Christine V. Schmidt; Stanislav N. Gorb:Snake Scale Microstructure:, Phylogenetic Significance and Functional Adaptations, 2012 ISBN 978-3-510-55044-9. [47] Gray J..1946 The mechanics of locomotion in snakes. J. Exp. Biology 23: 101–120. [48] Jiang, X.J., Whitehouse, D.J., Technological shifts in surface metrology (2012) CIRP Annals - Manufacturing Technology, 61 (2), pp. 815-836. [49] Costa H. L., Hutchings I. M., 2007 Hydrodynamic lubrication of textured steel surfaces under reciprocating sliding conditions Trib. Int. 40, 1227–1238. [50] Yan, Dongsheng , Qu, Ningsong , Li, Hansong and Wang, Xiaolei(2010) 'Significance of Dimple Parameters on the Friction of Sliding Surfaces Investigated by Orthogonal Experiments', Tribology Transactions, 53: 5, 703 - 712 [51] Hoppermann, A., and Kordt, M., 2002 Tribological Optimisation Using Laser-Structured Contact Surfaces, Oelhydraulik und Pneumatik, Vereinigte Fachverlage Mainz, ISSN 0341-2660. [52] Huang, Z., Liu, H., Shen, Z., Li, P., Hu, Y., Liu, H., Du, D., Wang, X., 2012, Process parameters analysis on surface texturing under laser shock peening, Chinese J of Lasers, 39, 107-113 [53] Hazel, J., Stone, M., Grace, MS, Tsukruk, V. V, 1998 Tribological design of biomaterial surfaces for reptation motions. Polymer Preprints 39, 1187-1188. [54] Marvi H, Hu DL, Friction enhancement in concertina locomotion of snakes J. R. Soc. Interface rsif 20120132, doi:10.1098/rsif.2012.0132 1742-5662 [55] Varenberg, M., Gorb, S. N, 2009 Hexagonal surface micropattern for dry and wet friction, Advanced Materials 21, 483-486 [56] Kovalchenko A, Oyelayo A, Erdemir A, Fenske G, Etsion A, 2005 The effect of laser surface texturing on transitions in lubrication regimes during unidirectional sliding contact, Trib. Int 38: 219–225 [57] Shinkarenko A, Kligerman Y., Etsion I., 2009 The effect of surface texturing in soft elasto- hydrodynamic lubrication, Trib. Int. 42, 284–292. [58] Halperin G, Greenberg Y, Etsion I. 1997 Increasing mechanical seals life with laser- textured seal faces. 15th International Conference on Fluid Sealing, BHR, Maastricht: 3–11.   26  [59] Xiaolei Wang , WeiLiu, Fei Zhou, Di Zhu  2009 Preliminary investigation of the effect of dimple size on friction in line contacts, Trib. Int. 42,1118–1123 [60] Yu H.W., Wang, X.L., Zhou F., 2010 Geometric shape effects of surface texture on the generation of hydrodynamic pressure between conformal contacting surfaces Trib. Lett. 37, 123–130. [61] Gonzalo, B, G., Backhaus, K., Knoll, G., 2012 Numerical analysis of laser-textured piston- rings in the hydrodynamic lubrication regime, J. Trib. 134, Doi/10.1115/1.4007347 [62] L., Galda, Pawlus, P, Sep, J, 2009 Dimples shape and distribution effect on characteristics of Stribeck curve, Trib. Int. 42, 1505–1512 [63] Hu, T, Hu, L, 2012 The study of tribological properties of laser-textured surface of 2024 aluminium alloy under boundary lubrication, Lubrication Science 24, 84–93 [64] Yin, B., Li, X., Fu, Y. & Yun, W., 2012. Effect of laser textured dimples on the lubrication performance of cylinder liner in diesel engine. Lubrication Science, 24(7), pp. 293-312. [65] Pettersson U, Jacobson S. 2003 Influence of surface texture on boundary lubricated sliding contacts. Trib. Int. 36, 857–864. [66] Pettersson U, Jacobson S. 2004 Friction and wear properties of micro textured DLC coated surfaces in boundary lubricated sliding. Trib. Lett 17, 553–559. [67] Swift, J.A. and J.R. Smith, 2001 Microscopical investigations on the epicuticle of mammalian keratin fibres. J of Microscopy, 204, 203-211. [68] Baumgartner, W., Saxe, F , Weth, A., Hajas, D., Sigumonrong, D, Emmerlich, J., Singheiser, M., [69] Zhang, X., Shi, F., Niu, J., Jiang, Y.,Wang, Z., 2008 Super hydrophobic surfaces: from structural control to functional application, J Mater Chem. 18, 621–633. [70] Koch, K., Bhushan, B., Barthlott, W. 2009 Multifunctional surface structures of plants: An inspiration for biomimetics Prog. Mater. Sc.; 54, 137-178. [71] Sun, Yuan. S.H., 2011. Effect of micro-dimple shapes on tribological properties of specimen surfaces. Huanan Ligong Daxue Xuebao/Journal of South China University of Technology (Natural Science), 39(1), 106-110+123   27 
1411.5125
2
1411
2015-03-26T14:17:10
Bistable dynamics of control activation in human intermittent control
[ "physics.bio-ph", "eess.SY", "nlin.AO" ]
When facing a task of balancing a dynamic system near an unstable equilibrium, humans often adopt intermittent control strategy: instead of continuously controlling the system, they repeatedly switch the control on and off. Paradigmatic example of such a task is stick balancing. Despite the simplicity of the task itself, the complexity of human intermittent control dynamics in stick balancing still puzzles researchers in motor control. Here we attempt to model one of the key mechanisms of human intermittent control, control activation, using as an example the task of overdamped stick balancing. In so doing, we focus on the concept of noise-driven activation, a more general alternative to the conventional threshold-driven activation. We describe control activation as a random walk in an energy potential, which changes in response to the state of the controlled system. By way of numerical simulations, we show that the developed model captures the core properties of human control activation observed previously in the experiments on overdamped stick balancing. Our results demonstrate that the double-well potential model provides tractable mathematical description of human control activation at least in the considered task, and suggest that the adopted approach can potentially aid in understanding human intermittent control in more complex processes.
physics.bio-ph
physics
Bistable dynamics of control activation in human intermittent control Arkady Zgonnikov∗ and Ihor Lubashevsky† University of Aizu, Tsuruga, Ikki-machi, Aizuwakamatsu, Fukushima 965-8580, Japan When facing a task of balancing a dynamic system near an unstable equilibrium, humans often adopt intermittent control strategy: instead of continuously controlling the system, they repeatedly switch the control on and off. Paradigmatic example of such a task is stick balancing. Despite the simplicity of the task itself, the complexity of human intermittent con- trol dynamics in stick balancing still puzzles researchers in motor control. Here we attempt to model one of the key mechanisms of human intermittent control, control activation, using as an example the task of overdamped stick balancing. In so doing, we focus on the concept of noise-driven activation, a more general alternative to the conventional threshold-driven activation. We describe control activation as a random walk in an energy potential, which changes in response to the state of the controlled system. By way of numerical simulations, we show that the developed model captures the core properties of human control activation observed previously in the experiments on overdamped stick balancing. Our results demon- strate that the double-well potential model provides tractable mathematical description of human control activation at least in the considered task, and suggest that the adopted ap- proach can potentially aid in understanding human intermittent control in more complex processes. I. INTRODUCTION Humans face the task of balancing dynamic systems near an unstable equilibrium repeatedly throughout their lives, for example, when main- taining upright stance [1], carrying a cup of cof- fee [2], or driving a car on a highway [3]. The task of inverted pendulum (stick) balancing is an increasingly popular paradigm of studying human control behavior over unstable dynamic systems (see recent reviews in [4 -- 6]). Physics and control engineering have revealed a wide va- riety of continuous control strategies which can successfully stabilize upright position of inverted pendulum, including simple linear feedback [7], or, counterintuitively, high-frequency vibration applied to the pivot point of the stick [8]. How- ever, due to a number of physiological con- straints, continuous control strategies either can- not be realized or are ineffective when applied by humans [9]. For this reason, much research has been aimed at understanding the mechanisms of discontinuous, or intermittent control, which is often employed by humans in controlling unsta- ble systems. ∗ [email protected][email protected] Intermittent control is characterized by re- peated switching between periods of passive (control is off) and active (control is on) behavior of the controller. Previously it has been shown that intermittent control strategies are efficient and robust under presence of significant time de- lays and sensorimotor noise [9 -- 12]. Moreover, numerous signs of intermittency have been de- tected in experimental data on human postural sway [13 -- 15], stick balancing on fingertip [16], and, more recently, virtual stick balancing on the computer screen [17, 18]. Numerous attempts to model the dynamics observed in these exper- iments typically attribute the discontinuity of human control to the so-called "sensory dead- zone" [e.g. 6, 19, 20]. According to this hypoth- esis, the control remains switched off as long as the deviation of the controlled system from the desired state remains below certain thresh- old value. Whenever the deviation exceeds this threshold, the control is immediately switched on. Models which incorporate threshold-driven control activation can explain much dynamics observed, e.g., in balancing of inverted pendu- lum and quiet standing. Still, the question re- mains open: How accurately does the thresh- old model describe the process of control acti- 2 vation in humans? Recent investigations of vir- tual overdamped stick balancing have revealed that control triggering at large deviations oc- curs much more frequently then predicted by threshold-driven activation [18]. Besides, some of the most peculiar phenomena observed in hu- man control behavior still remain unexplained by the models based on threshold-driven con- trol activation. Importantly, such phenomena include high occurrence of large fluctuations in human-controlled systems, which result, e.g., in falls during stick balancing or quiet stand- ing [21]. All these considerations suggest that in controlling even the simplest unstable systems humans may employ a somewhat more complex control activation mechanisms than assumed by conventional threshold-based models. In a recent attempt to elucidate the proper- ties of control activation in humans, Zgonnikov et al. introduced the concept of intrinsically stochastic, noise-driven control activation [18]. The key hypothesis of this approach is that mathematical description of control activation should encompass stochasticity of human cogni- tive processes during decision when to start ac- tively controlling the unstable system. The plau- sibility of this hypothesis has been confirmed by the simple ad hoc model mimicking the effect of noise-driven activation. Yet, the adequate math- ematical formalism, which would allow one to model noise-driven control activation in differ- ent tasks, is still missing. The aim of the present work is to develop ex- plicit mathematical description of noise-driven control activation. The model we develop here describes competition between two cognitive states of the human operator, "wait" and "act", which are mapped onto the two states of a bistable dynamical system. The stochastic dy- namics of control activation is captured via the notion of random walk in a double-well poten- tial, which has proven its efficiency in modeling complex cognitive processes [e.g. 22 -- 24]. Using a simple example of overdamped stick balancing, we demonstrate that the model reproduces the behavior of human operators observed experi- mentally. Based on the results of numerical sim- ulations, we argue that the double-well approach can serve as a natural, easily adaptable frame- FIG. 1. cart. Inverted pendulum attached to a moving work for modeling mechanisms of control activa- tion in diverse human-controlled processes. II. HUMAN BALANCING OF OVERDAMPED STICK Although we hypothesize the model devel- oped here can be applied to virtually any human- controlled system, we exemplify our theoretical constructs using possibly the simplest unstable mechanical system, that is, overdamped stick balancing (Fig. 1). The linearized dynamics of the mechanical system in terms of the stick an- gle θ and cart velocity υ (which is a control vari- able in this case) are described by the differential equation (for derivation see [18]) τθ θ = sin θ − τ l υ cos θ, (1) where τθ > 0 is the intrinsic time scale of the stick motion, and l > 0 is the length of the stick. Experiments on human balancing of virtual overdamped stick on the computer screen re- vealed that the stick dynamics under human con- trol are universal across the subjects and within a range of kinetic parameters of the stick [18]. The task for the subjects was to keep the stick upwards, whereas the cart could be moved arbi- trarily within the limits of the computer screen. The experiments had shown that the subjects universally employed intermittent control strat- egy, with the cart velocity equal to zero for sig- nificant amount of time (Fig. 2). Importantly, the subjects often preferred to start correcting the stick position only when the deviation sig- 3 independent phase variable, so that its dynam- ics is defined by a separate differential equation, specifically υ = αlθξ − βυ, (2) where α and β are non-negative constants, and ξ is an order parameter describing the cognitive state of the operator in regards to the controlled system. The order parameter ξ switches inter- mittently between the states ξ = 0 and ξ = 1, which are mapped onto the operators' two cog- nitive states, "wait" and "act". Consequently, in either the "act" (ξ = 1) or "wait" (ξ = 0) states the cart dynamics are effectively linear; it is the nonlinear mechanism of switching between these states that encompasses stochasticity and complexity. In our model, the dynamics of ξ is driven jointly by the deterministic and random forces. The deterministic dynamics are governed by the double-well potential energy landscape, where the configuration of the landscape is determined by the state of the controlled system (in our case, the stick angle θ). The stochastic switching be- tween the two wells is caused by a random force. Such dynamics can be described by the Langevin equation (in the Ito interpretation) τξ ξ = − ∂H ∂ξ + √ εHζ, (3) where τξ > 0 is the constant parameter defining the time scale of the switching process, H(ξ, θ) is the Hamiltonian shaping the energy landscape of the system (Fig. 3), ζ is white noise, and ε > 0 is the parameter regulating the noise intensity. We wish to underline that the model utilizes multiplicative noise in Eq. (3) to reflect the as- sumption that the level of uncertainty of the op- erator's cognitive state is low in unambiguous situations (H ≈ 0), and that this uncertainty increases with ambiguity of the current system state (which is quantified by H). Specifically, we assume that in the situations when the de- viation θ is large, the operator's cognitive state is generally certain (that is, "act"). The square- root dependence of the noise intensity on H has been chosen for the sake of simplicity; in this case the form of the Langevin equation (3) does FIG. 2. Typical trajectory of overdamped inverted pendulum balancing by a human subject (based on the data reported in [18]). The trajectory reflects 15 seconds of balancing without stick falls. nificantly exceeded the scale of the angle sen- sory deadzone. The distribution of action points (stick deviations triggering reaction of the opera- tor) was found to be considerably non-Gaussian, which suggests that factors other than sensory threshold determine the process of control acti- vation. Without attempting to identify all such factors and model them in all their complex- ity, we will instead construct an integrated, phe- nomenological description of the control activa- tion process. III. DOUBLE-WELL DYNAMICS OF CONTROL ACTIVATION A. Model description In the case of overdamped inverted pendu- lum considered here, a model of the operator's behavior should specify how the control variable υ evolves in time given the state of the stick θ. Within the paradigm of intermittent control, such a model should describe how the control is activated and how it is executed once acti- vated. We have to emphasize, however, that the present model focuses mostly on the former issue. The latter problem, control execution, supposedly concerns the formalism of open-loop control, and, we believe, deserves detailed con- sideration elsewhere. Appealing to the phase space extension ap- proach [25], we consider the cart velocity υ an 0.500.250.000.250.50stick angle θ1.60.80.00.81.6cart velocity υ 4 FIG. 3. Double-well energy landscape H(ξ) depending on deviation of the stick from the desired position (as characterized by a(θ)). not depend on the stochastic process interpreta- tion. Namely, the difference between this equa- tion written in the Ito, Stratonovich, or Hanggi- Klimontovich forms is reduced to a minor renor- malization cofactor in the regular drift term (see e.g. [26]). The particular form of the Hamiltonian H is chosen in a way that where ∝ ξ(ξ − 1)(ξ − a(θ)), ∂H ∂ξ a(θ) = 1/(1 + (θ/η)2). (4) (5) Parameter η of the ansatz (5) characterizes the operator's perception, so that a ≈ 1 if θ (cid:28) η and a ≈ 0 when θ (cid:29) η. Thus, when the stick deviation is large (a ≈ 0), the energy landscape configuration is such that the "act" cognitive state (ξ ≈ 1) can be expected. On the contrary, the "wait" state (ξ ≈ 0) is most likely to be ob- served whenever the deviation is small (a ≈ 1). Intermediate values of θ lead to bistable dynam- ics, so both the states become possible (Fig. 3). We impose the following constraints on H Hξ=0, θ=0 = 0, Hξ=1, θ(cid:29)η = 0, Hξ=0, θ(cid:29)η = 1, Hξ=1, θ=0 = 1. (6) so that the random fluctuations diminish when the energy minimum H = 0 is achieved at the states ξ = 0 and ξ = 1 in case the state of the controlled system is unambiguous (θ = 0 and θ (cid:29) η, respectively). Conditions (4 -- 6) yield H(ξ, a(θ)) = 3ξ4− 4(1 + a)ξ3 + 6aξ2 + 1− a. (7) To simplify the model analysis, we linearize Eq. (1) around the upright position θ = 0 and rescale the time, stick angle, and cart velocity t → tτθ, θ → θη υ → υηl/τθ, so that Eqs. (1 -- 3) are transformed into θ = θ − υ, υ = γθξ − συ, √ τ ξ = − ∂H ∂ξ + Hζ, (8) where τ = τξ/τθ, γ = ατ 2 θ , σ = βτθ, and  = ε/τθ. In the rescaled variables the expression (5) takes form a(θ) = 1/(1 + θ2). (9) while the form of the energy function (7) is con- served. B. Model analysis The model (7 -- 9) has four parameters, τ , , γ, and σ. The feedback parameters γ and σ affect the system dynamics only when control is active (ξ ≈ 1). Under some weak assumptions (essentially, σ > 1, γ > σ) the particular values of these parameters have no substantial effect on the system dynamics, so in what follows we assume σ = 3.5, γ = σ2/2, the values which were previously shown to be physically plausible for overdamped stick balancing [18]. 010.00.51.01.5H(»)Zero deviationa(µ)=101»Intermediate deviationa(µ)=0:501Large deviationa(µ)=0:1 5 FIG. 4. Projections of the model-generated trajectories on the θ − υ plane. Each trajectory represents the system motion of duration 5000τ time units. The other two parameters, τ and , deserve close examination, because they define the very dynamics of the control activation process. The parameter τ denotes the time scale of the lo- cal dynamics of the operator's cognitive state ξ. More specifically, it defines how fast ξ responds to changes of the energy landscape H, i.e., varia- tions of the stick angle θ. Smaller values of τ lead to faster transient process, and the other way around. The parameter  affecting the noise in- tensity quantifies the overall level of uncertainty in the system. Specifically, for a given (fixed) configuration of the double-well potential H, the rate of intermittent switching between the two wells increases with . To illustrate the dynamics of the model de- pending on τ and , we have run numerical simulations using the stochastic Runge-Kutta scheme [27] for the Ito-type stochastic pro- cess (7 -- 9) with the discretization step ∆t = τ /10. We consider the values of τ such that τ (cid:28) 1: to enable the controller to keep the stick up- wards, the switching should apparently occur on time scales shorter than the time scale of the stick motion. The noise intensity parameter is also assumed to be small ( (cid:28) 1); otherwise, the noise term loses its physical meaning. To be specific, we illustrate the system dynamics for τ ∈ {0.01, 0.05, 0.2} and  ∈ {10−3, 10−2, 10−1}. 1011.50.01.5τ=0.012022.50.02.5τ=0.0510010201001020τ=0.22022.50.02.5cart velocity υ404505500508008040450515015stick angle θ2002060006008000800=0.001=0.01=0.1 The basic pattern of the system dynamics re- mains the same for all tested values of the sys- tem parameters (Fig. 4). The system initially perturbed by a small deviation of θ moves along the axis υ = 0, which represents the passive con- trol phase (ξ = 0). As the angle θ increases, the energy landscape changes so that the transition to the active phase (ξ = 1) becomes increasingly probable. This transition is induced by noise, so it occurs at probabilistically determined angle. Once the transition ξ : 0 → 1 occurs, the cor- rective feedback comes into action, and the stick is returned to the vicinity of the upright posi- tion. However, when the time scale τ is short and the noise intensity  is high (e.g., top left frame in Fig. 4), the control is always deacti- vated ξ : 1 → 0 before the stick reaches the vertical position (i.e., only undershooting is ob- served). In other tested cases, both the under- and overshooting corrections are implemented, which more resembles the actual human behav- ior (Fig. 2). It should be noted that the amplitude of the stick fluctuations increases with τ and decreases with . Indeed, the longer the time τ of the transient process ξ : 0 → 1, the larger the av- erage value reached by the stick angle when the control is finally switched on. On the opposite, increasing the level of noise  results in higher probability of switching even for small values of θ. These considerations prompt that long τ to- gether with small  reflect inability of the con- troller to switch the control on in response to the steadily increasing θ (hence large fluctuations in Fig. 4, bottom right frame; in the real balancing task the stick would just repeatedly fall in this case). The distribution of action points (AP; the values of θ matching the initiation of the tran- sition ξ : 0 → 1) highlights the diversity of the control activation dynamics captured by the model (Fig. 5). For short time scales (exem- plified by τ = 0.01), the log-scale AP distri- bution has parabolic shape similar to Gaussian, which suggests that in this regime the double- well activation works much like ordinary thresh- old. When τ is increased (τ = 0.05), small val- ues of  lead to substantial asymmetry in the AP distribution, which has approximately Laplace- 6 shaped tail part (pdf ∝ e−θ). Finally, if the time scale is further increased (τ = 0.2), the exponen- tial shape of the AP distribution tail transforms to the power-law shape with decreasing . Previously obtained experimental data on human overdamped stick balancing revealed distinct Laplace-like distribution of action points [18]. The proposed model captures well the tail part of the experimentally found dis- tribution (Fig. 6), although for small values of the stick angle the model-generated AP statistics differs from the experimental one. A threshold- based control activation yields the Gaussian dis- tribution of action points (cf. dashed line in Fig. 6), which practically eliminates any prob- ability of control activation at large values of θ. In this sense, noise-driven activation as captured in the double-well model provides a much more plausible explanation of the actual human con- trol strategy. IV. DISCUSSION In controlling unstable systems humans of- ten prefer to switch intermittently between the passive and active behavior instead of control- ling the system in a continuous manner. The present paper argues that some intricate prop- erties of intermittent control activation in hu- mans can be explained using the notion of ran- dom walk in a double-well potential. We focus on overdamped stick balancing as a representa- tive example of a human control task. In model- ing the control activation process, we extend the phase space of the physical stick under human control by an order parameter characterizing the operator's cognitive state. The dynamics of this order parameter is stochastic, and reflects inter- mittent switching between active and passive be- havior of the human operator. The switching process is defined in part by the double-well po- tential field (changing with the stick angle), and in part by the random force. We describe the latter using multiplicative noise, so that the un- ambiguous states of the controlled system evoke little noise compared to the uncertain ones. We demonstrate that the model has rich dynamics, and can explain a number of different control 7 FIG. 5. Probability distribution functions (pdf) of absolute values of action points (AP) exhibited by the model for various values of system parameters. The duration of numerical simulation needed to obtain stable distributions is of order 106τ . Values of θ are normalized so that the angle values of different scales can be compared. scopic level by just a few variables (order pa- rameters). This approach has been widely ap- plied for studying perceptual and cognitive dy- namics (see e.g. reviews in [29 -- 31]). At the same time, modern research on human control over unstable systems emphasizes mainly physi- ological and mechanical aspects of human behav- ior in diverse balancing tasks, while the under- lying cognitive mechanisms are rarely studied. In particular, control activation has been con- ventionally modeled as a threshold-driven pro- cess. However, a question had arisen recently as to whether threshold-based models can fully accommodate complex, often unpredictable dy- namics of human-controlled systems [4, 21, 32]. Models implementing noise-driven control ac- tivation can potentially serve as richer alterna- tives to the threshold-based models of intermit- tent motor control [18]. However, a physically tractable model of noise-driven control activa- tion has been missing up to now, which should necessarily incorporate explicit regulations gov- erning the activation dynamics. The present study develops such a model by adopting the concept of noise-driven switching widely used for describing bistable cognitive phenomena, e.g., as perceptual categorization [22, 33] and decision making under risk [24]. Noise has long been recognized as a key fac- tor in human intermittent control. Particularly, Milton, Cabrera et al. [6, 11, 16] have been devel- FIG. 6. Distributions of action points produced by the model (τ = 0.2,  = 0.02) and human sub- jects. Experimentally obtained distribution is aver- aged over ten subjects. Normal distribution trun- cated at zero is presented for reference. activation patterns. By comparing the model to the previously obtained data on human stick bal- ancing, we show that at least one of the modes of the model matches the actually observed hu- man behavior. Based on the presented results, we suggest that employing double-well potential approach to model control activation can aid in understanding complex properties of human in- termittent control. Human brain is a self-organizing system [28, 29]. Nonlinear interactions of large number of neurons and/or neuronal populations lead to emergence of complex dynamics, which, how- ever, can sometimes be described on the macro- 12345610-310-210-1100AP pdfτ=0.01123456normalized angle θ/std(θ)τ=0.05123456τ=0.2=0.1=0.01=0.001123456normalized angle θ/std(θ)10-310-210-1100AP pdfExperimentModelNormal dist. 8 oping the idea that state-dependent noise can aid in stabilizing (underdamped) inverted pendulum by forcing the feedback gain to oscillate back and forth across the stability boundary, which turns out to be beneficial for balance control un- der considerable response delay. Our approach is similar to theirs in that without noise the upright position of the stick is unstable. However, the principal distinction between these approaches is the constructive role of noise, which in our case is to induce switching of the operator's cognitive state between acting and waiting. Intrinsic stochasticity of the noise-driven switching allows the model to generate diverse distributions of action points (including Gaus- sian, Laplacian, and Pareto distributions). This suggests that the double-well model can capture control activation mechanisms of different na- ture. It remains for future experimental work to investigate whether and under what condi- tions these different mechanisms are employed by human operators. Moreover, one may spec- ulate that in the single control process human operators may utilize different control activa- tion mechanisms (for instance, for different spa- tial scales of the controlled system's motion). If this is true, there is need for further model- ing efforts. The future models may be based, for example, on the hypothesis that the charac- teristic time scale of control activation can be state-dependent: when the system approaches the desired position, the time scale may become shorter to enable fine tuning of the system state. This hypothesis may provide an explanation of the found mismatch between the double-well- model- and human-generated action point distri- butions in the vicinity of the zero angle (Fig. 6). We wish to underline that the presented model currently utilizes the double-well dynam- ics to capture not only control activation, but deactivation as well. However, the latter pro- cess is supposedly determined by the open-loop control execution mechanisms. Given the ex- perimentally observed variability of the correc- tive trajectories with respect to their end-points (Fig. 2), further development of mathematical description of open-loop control is needed to cap- ture the found control deactivation patterns in more detail. This would as well aid in represent- ing the whole spectrum of the possible corrective actions. We believe that the ideas of the present work may extend beyond the specific field of motor control to a more general class of cognitive pro- cesses. Complex, high-dimensional neural sys- tems can exhibit low-dimensional dynamics on the macrolevel [29]. An apt example is the phe- nomenon of bistable perception, which can be characterized by the double-well dynamics of switching, e.g., between two alternative inter- pretations of an ambiguous stimulus [23]. Ev- idence for double-well stochastic dynamics has also been found in other cognitive processes, for instance, perceptual categorization of speech patterns [22] and imagined actions [33]. Inspired by the results of the present study, we hypothe- size that further quantitative investigations may provide grounds for stronger link between such processes and the advanced concepts employed in describing dynamical systems in physics. Par- ticularly, treating the intensity of the neural noise as a state-dependent rather than constant quantity, on the one hand, is physiologically fea- sible [34, 35], and, on the other hand, may po- tentially enhance the explanatory power of the modern cognitive models of multistable phenom- ena. [1] Loram ID, Lakie M. Direct measurement of human ankle stiffness during quiet stand- ing: the intrinsic mechanical stiffness is insuf- ficient for stability. The Journal of Physiology. 2002;545(3):1041 -- 1053. [2] Mayer H, Krechetnikov R. Walking with cof- fee: Why does it spill? Physical Review E. 2012;85(4):046117. [3] Lubashevsky I, Wagner P, Mahnke R. Rational- driver approximation in car-following theory. Physical Review E. 2003;68(5):056109. [4] Asai Y, Tateyama S, Nomura T. Learning an Intermittent Control Strategy for Postural Bal- ancing Using an EMG-Based Human-Computer 9 Interface. PLoS One. 2013;8(5):e62956. ciety Interface. 2014;11(99):20140636. [5] Balasubramaniam R. On the Control of Unsta- ble Objects: The Dynamics of Human Stick Bal- ancing. In: Progress in Motor Control. Springer; 2013. p. 149 -- 168. [6] Milton JG. Intermittent Motor Control: The 'drift-and-act' Hypothesis. In: Progress in Mo- tor Control. Springer; 2013. p. 169 -- 193. [7] Kwakernaak H, Sivan R. Linear optimal control systems. Wiley-Interscience New York; 1972. [8] Kapitza P. Dynamic stability of a pendulum with an oscillating point of suspension. Jour- nal of Experimental and Theoretical Physics. 1951;21(5):588 -- 597. [9] Loram I, Gollee H, Lakie M, Gawthrop P. Hu- man control of an inverted pendulum: is contin- uous control necessary? Is intermittent control effective? Is intermittent control physiological? The Journal of Physiology. 2011;589(2):307 -- 324. [10] Milton J, Cabrera J, Ohira T. Unstable dynam- ical systems: Delays, noise and control. EPL (Europhysics Letters). 2008;83(4):48001. [11] Milton JG, Cabrera JL, Ohira T, Tajima S, Tonosaki Y, Eurich CW, et al. The time-delayed inverted pendulum: implications for human bal- ance control. Chaos. 2009;19(2):026110 -- 026110. [12] Milton JG. The delayed and noisy nervous sys- tem: implications for neural control. Journal of neural engineering. 2011;8(6):065005. [13] Loram ID, Maganaris CN, Lakie M. Active, non-spring-like muscle movements in human postural sway: how might paradoxical changes in muscle length be produced? The Journal of physiology. 2005;564(1):281 -- 293. [14] Loram ID, Maganaris CN, Lakie M. Human postural sway results from frequent, ballistic bias impulses by soleus and gastrocnemius. The Journal of Physiology. 2005;564(1):295 -- 311. [15] Bottaro A, Casadio M, Morasso PG, Sanguineti V. Body sway during quiet standing: Is it the residual chattering of an intermittent sta- bilization process? Human movement science. 2005;24(4):588 -- 615. [16] Cabrera JL, Milton JG. On-off intermittency in a human balancing task. Physical Review Letters. 2002;89(15):158702. [17] Bormann R, Cabrera JL, Milton JG, Eurich CW. Visuomotor tracking on a computer screen -- an experimental paradigm to study the dynamics of motor control. Neurocomputing. 2004;58:517 -- 523. [18] Zgonnikov A, Lubashevsky I, Kanemoto S, Miyazawa T, Suzuki T. To react or not to re- act? Intrinsic stochasticity of human control in virtual stick balancing. Journal of the Royal So- [19] Milton JG, Ohira T, Cabrera JL, Fraiser RM, Gyorffy JB, Ruiz FK, et al. Balancing with vi- bration: a prelude for "drift and act" balance control. PLoS One. 2009;4(10):e7427. [20] Gawthrop P, Loram I, Lakie M, Gollee H. Inter- mittent control: a computational theory of hu- man control. Biological cybernetics. 2011;104(1- 2):31 -- 51. [21] Cabrera JL, Milton JG. Stick balancing, falls and Dragon-Kings. The European Physical Journal Special Topics. 2012;205(1):231 -- 241. [22] Tuller B, Case P, Ding M, Kelso J. The non- linear dynamics of speech categorization. Jour- nal of Experimental Psychology: Human Per- ception and Performance. 1994;20(1):3. [23] Moreno-Bote R, Rinzel J, Rubin N. Noise- Induced Alternations in an Attractor Network Model of Perceptual Bistability. Journal of Neu- rophysiology. 2007;98(3):1125 -- 1139. [24] van Rooij MM, Favela LH, Malone M, Richard- son MJ. Modeling the Dynamics of Risky Choice. Ecological Psychology. 2013;25(3):293 -- 303. [25] Zgonnikov A, Lubashevsky I. Extended phase space description of human-controlled systems dynamics. Progress of Theoretical and Experi- mental Physics. 2014;2014(3):033J02. [26] Mahnke R, Kaupuzs J, Lubashevsky I. Physics of Stochastic Processes: How Randomness Acts in Time. Weinheim: WILEY-VCH Verlag GmbH & Co. KGaA; 2009. [27] Roessler A. Explicit order 1.5 schemes for the strong approximation of Ito stochastic differen- tial equations. Proceedings in Applied Mathe- matics and Mechanics. 2005;5(1):817 -- 818. [28] Haken H. Principles of brain functioning. Springer; 1996. [29] Kelso JS. Dynamic patterns: The self- organization of brain and behavior. MIT press; 1995. [30] Rolls ET, Deco G. The noisy brain: stochastic dynamics as a principle of brain function. Ox- ford university press Oxford; 2010. [31] Wagemans J, Feldman J, Gepshtein S, Kim- chi R, Pomerantz JR, van der Helm PA, et al. A century of Gestalt psychology in visual per- ception: II. Conceptual and theoretical founda- tions. Psychological bulletin. 2012;138(6):1218. [32] Bottaro A, Yasutake Y, Nomura T, Casadio M, Morasso P. Bounded stability of the quiet stand- ing posture: an intermittent control model. Hu- man movement science. 2008;27(3):473 -- 495. [33] van Rooij I, Bongers RM, Haselager W. A non- representational approach to imagined action. Cognitive Science. 2002;26(3):345 -- 375. [34] Lindner B, Schimansky-Geier L. Transmission of noise coded versus additive signals through a neuronal ensemble. Physical review letters. 2001;86(14):2934. [35] Silberberg G, Bethge M, Markram H, Pawelzik K, Tsodyks M. Dynamics of population rate codes in ensembles of neocortical neurons. Jour- nal of Neurophysiology. 2004;91(2):704 -- 709. 10
1706.02632
2
1706
2017-07-18T15:30:31
Acoustic resonance by fish schools. A proposal for the schooling mechanism
[ "physics.bio-ph", "physics.ao-ph" ]
The acoustic properties of a fish school have been modelled using a cloud of bubbles. Similar to how bubble clouds present a collective monopole mode, a fish school also shows a collective breathing mode in which the whole school resonates as a single body. Here, we conjecture that the underwater acoustic noise of the ocean amplified by the multiple scattering of swim bladders in the fish school might produce attractive acoustic forces that are strong enough to account for the schooling mechanism. Our model predicts the presence of attractive Bjerknes forces, as large as 30% of the fish weight for 20-metre large fish schools, and that the attractive force on every fish in the centre of the school is cancelled when the fish increase/decrease the volume/pressure of its gas bladder. To the best of our knowledge, this is the first example of a purely mechanical force that might contribute to fish being bound to or released from the school. The study may lead to new areas of research in many scientific fields beyond the nearest-neighbour interaction mechanism customarily used, and it would help in our understanding of collective processes of systems in ecology, physics, chemistry, and biology.
physics.bio-ph
physics
Acoustic resonance by fish schools: A proposal for the schooling mechanism F. Meseguer, and F. Ramiro-Manzano Instituto de Tecnología Química (CSIC-UPV). Universitat Politècnica de València, Av. Tarongers s/n 46022, Valencia, Spain Email: [email protected] Abstract The acoustic properties of a fish school have been modelled using a cloud of bubbles. Similar to how bubble clouds present a collective monopole mode, a fish school also shows a collective breathing mode in which the whole school resonates as a single body. Here, we conjecture that the underwater acoustic noise of the ocean amplified by the multiple scattering of swim bladders in the fish school might produce attractive acoustic forces that are strong enough to account for the schooling mechanism. Our model predicts the presence of attractive Bjerknes forces, as large as 30% of the fish weight for 20-metre large fish schools, and that the attractive force on every fish in the centre of the school is cancelled when the fish increase/decrease the volume/pressure of its gas bladder. To the best of our knowledge, this is the first example of a purely mechanical force that might contribute to fish being bound to or released from the school. The study may lead to new areas of research in many scientific fields beyond the nearest-neighbour interaction mechanism customarily used, and it would help in our understanding of collective processes of systems in ecology, physics, chemistry, and biology. Many scientists have been puzzled by the beautiful arrangements of fish usually called fish schools1. The pioneering work of Nicolis and Prigogine2 on self-organization was followed by many others trying to understand this collective behaviour in terms of the interactions between individuals forming the group3-5. Faucher et al.6 showed that the lateral line organ (LLO)7 plays an important role in the schooling mechanism. All reported fish schooling models have been based on the fish active response to an external stimulus3-7. Here, we show that a fish inside a school is passively bound to its position. The presence of the underwater acoustic noise of the ocean8, amplified by multiple scattering of the swim bladders in the school, might produce strong attractive acoustic Bjerknes9 forces comparable to the weight of the fish. The fish swim bladder is strongly affected by sound waves, and this bladder has been modelled by a gas bubble in water10 using the well-known Minnaert's11 monopole expression, (1) where R0 is the bubble radius, ρ is the water density, γ tis the specific heat ratio for the gas in the bubble, and p0 is the static pressure of the gas. There has being increasing interest in extending this bubble model to study the acoustical properties of fish schools12-17. In the pioneering paper of Feuillade et al.12, a fish school is emulated by a three-dimensional arrangement of bubbles, called a bubble cloud. This work uses the results of Otma14 and d'Agostino and Brennen15 for the acoustical properties of a bubble cloud. It was shown that the resonant mode with the lowest frequency value (monopole resonance) of a cloud is similar to the fundamental breathing mode of a single large bubble with a total volume equivalent to the volume of all bubbles forming the cloud. That is, the bubble resonates at a frequency value, ν1, given by the following expression (Figure 1): (2) where R1 and R0 are the radii of the cloud and the bubble, respectively, and α is the filling fraction of the bubble cloud. Although the expression (2) is a simplified model, it provides useful information regarding the low-frequency value of the fundamental collective breathing mode (see Supplementary Information). 1. Cloud bubble model with Figure fraction α, used for understanding the acoustical properties of a fish school. The building block c orresponds to a single bubble mimicking a single fish with a spherical swim bladder of radius R0. radius R1 and filling Many researchers have studied Bjerknes forces18-22. The secondary Bjerknes force, respectively, with a FSB(α, β), acts between two bubbles, α and β with radii distance between bubbles of and . When subjected to an acoustic field with a complex amplitude of pressure Am and an . The bubbles resonate at angular frequency values and angular frequency ω, the force can be written as18 where Here, ρ0 is the density of the fluid (water in this case), and are the damping losses of the α, β bubbles that depend on both the bubble size and the frequency value. loss δ has several contributions13,18, and in the case of fish and cm size bubbles (see Supplementary Information), we can write this loss as16,17 direction. In general, is a vector along the damping and the Figure 2. Bjerknes force and pressure. Left panel: Total pressure (black curve) and β factor (red curve) as a function of the bubble cloud radius at a sea depth h=10 m. Right panel: Total force per fish weight as a function of the number of fish in the fish school at a sea depth h=20 m (see text). all bubbles have the same size, R0, and the same gas pressure, p0. Thus, Secondary Bjerknes forces are attractive when , and repulsive when . As the fish school is formed by fish that are similar in size, we can assume ; . We assume all bubbles are located at different points on a Body Centred Cubic (BCC) lattice, with Cartesian coordinates α=(ijk) and β=(lmn). It is very difficult to calculate the Bjerknes force produced by the background noise of the ocean (BNO) between a single bubble and the rest of the bubble cloud, including multiple scattering effects, even for a modest number of bubbles. Therefore, we have made the following assumptions: A) The force on a single bubble (lmn) is the total contribution of the rest of the bubbles (ijk) integrated over the frequency region of interest. We assume a resonant frequency for all bubbles of radius R0. Then, as we have a single resonant frequency, all forces ; and become attractive. B) We also know that a cloud of bubbles behaves as an acoustic cavity, where the sound resonates with a lifetime value depending on the quality factor Q of the cavity, defined as . Therefore, the (lmn) bubble feels the scattering from the (ijk) bubble for a period of time τ=Q/ν1, which, in the case of large fish schools (see the Supplementary Table 1), can reach values of several seconds. Therefore, the Bjerknes force between factor, βijklmn, defined as each bubble pair should be , where c is the sound velocity in water and Lijklmn is the distance reinforced by a between the bubbles located at (ijk) and (lmn). The expression of the Bjerknes force for multiple scattering (MS) would be as follows (see Supplementary Information): where When mean field theory comes into play, the bubble cloud (and also the fish school) would 14,15 (on the order of magnitude of several behave as a large bubble of effective radius R01 tens of centimetres). It is important to emphasize that both the resonant frequency value ν1 and the damping constant δ1 decrease dramatically for increasing fish school sizes. This fact has two main consequences (see the Supplementary Information): 1. The resonant wavelength value is much larger than the fish school size regardless of the school size; i.e., all fish in the school would feel each other. 2. The amplification factor, βijklmn, can reach very large values; thus, it might dramatically amplify the Bjerknes forces. From the fish school example with R1=5 metres (see fourth row in the Supplementary Table 1), we obtain very long dwelling time values, about τ=2 seconds, which correspond to very large amplification factors, around β=104. The expression (5) is difficult to handle since it involves calculating the forces from all fish. The first task is to locate all neighbours (up to eight million) in a very large BCC lattice23. It also involves calculating the components of the force. However, as we are considering a spherical fish school arranged in a BCC lattice, we can safely assume an isotropic force on a single fish located in the centre of the fish school. Then, the pressure acting on the gas bladder can be written as and the total force That is, we assumed that the total pressure can be written as the sum of the module of all forces equally distributed over the surface of the gas bladder. We numerically calculated both the pressure for fish schools of different sizes at two different sea depths, h=10 m and h=20 m, as shown in the Supplementary Table 1. We considered the total force over the BNO in the frequency range 1 Hz < ν < 1000 Hz8 (see the Supplementary Figure 1). Figure 2 (left panel) shows the evolution of total pressure for different values of fish school radius. It also shows the evolution of the β factor (β=τc/d) as a function of the size of the fish school. In all cases, we used the following parameter values: a bubble radius R0=10-2 metres, and a filling fraction factor α=0.5%. This model represents fish schools of different sizes composed of fish of length L=29 cm, periodically distributed in BCC lattice, with a nearest-neighbour Figure 3. Total pressure on a bubble at the centre of the cloud as a function of the cloud radius, when either it has the same radius as the rest of the cloud (red curve) or the bubble radius is 20% larger than the rest of the bubbles (black curve). The bubble cloud is located at a depth h=10 metres. We used the same parameters as those in Figure 2 (see text). the fish 4% of pressure volume17. distance (NND) between fish of d=0,1 metres. We used a realistic value for the shear viscosity16, µ=50 N s/m2, which is associated with the swim bladder. We followed the Love model for calculating the fish volume, and we assumed that the gas bladder represents The β=τc/d factor is the maximum amplification factor due to multiple scattering effects. The acoustic the β factor increase nonlinearly with the fish school radius. The right panel in Figure 2 shows the evolution of the ratio of the total force to the fish weight as a function of the number of fish in the school at a sea depth of h=20 m. The total force increases quasi-linearly with the number of fish in the school, although it shows a certain tendency to saturation. This unexpected behaviour can be explained due to the very large values of both the multiple scattering factor β (Figure 2) and the large number of neighbours (see the Supplementary Figure 2). The calculated total forces can reach values of around 30% of the fish weight in 20-metre sized fish schools (see the Supplementary information). This result is consistent with some findings reporting large fish schools up to 1 Km24. and Finally, we would like to discuss the possibility that, through secondary Bjerknes forces, each fish can control at will its inclusion/exclusion to/from the school. Let us assume that a single fish from the school, located at the origin of coordinates (000), changes its gas bladder pressure from p0 to p'0. Then, it will resonate at a different frequency value that is completely detuned from the collective resonance of the school fish. Thus, the total Bjerknes force would be where The Bjerknes force will be attractive when the sound frequency ω is either ω > ω1, ω > ω0' or ω < ω1, ω < ω0', and repulsive when ω1 < ω < ω0'. Figure 3 shows the total pressure on a bubble located at the centre of the cloud as a function of the cloud radius, either when the bubble has the same radius as other bubbles in the cloud (red curve), or when the bubble radius is 20% larger than the rest of bubbles (black curve). In the second case, the attractive and repulsive forces cancel, releasing the fish from the school. Finally, this model also explains the short distance repulsive forces that avoid collapsing the whole school. Expression (3) predicts the repulsive forces between neighbouring school mates when they closely approach each other in a similar manner as it appears for the formation of stable bound bubble grapes18,21,25. If the BNO is the responsible for the binding forces in fish schools, the experimental evaluation disturbances (amplification/attenuation) when measured near a fish school, especially at frequency values near the collective resonance ν1 i.e., the whole fish school would act as a large acoustic antenna near its resonance frequency ν1, and the BNO would strongly be influenced by the multiple scattering effects of the whole fish school. PSD magnitude should show strong of the Acknowledgements. We would like to thank to Prof. C. Feuillade and to Dr. H. Estrada for the critical reading of the manuscript. This work has been partially supported by the Severo Ochoa Excellency SEV-2016-0683, the young investigator project TEC2015-74405-JIN and the PrometeoII/2014/026. Author Contributions. F.M. conceived the study and the model, performed the simulation and data analysis, and prepares the manuscript. F.R.-M. prepared the simulation code, contribute to the data analysis, and manuscript writing. References 1. Sumpter, D. J. T. The principles of collective animal behavior. Philos. Trans. R. Soc. Lond., B, Biol. Sci., 361, 5-22 (2006). 2. Nicolis, G. & Prigogine, I. Self-organization in nonequilibrium systems. Wiley, New York (1977). 3. Okubo, A. Dynamical aspects of animal grouping. Adv. Biophys. 22, 1–94 (1986). 4. Reynolds, C. W. Flocks, herds and schools: a distributed behavioural model. Comp. Graph., 21, 25–33 (1987). 5. Couzin, I. D., Krause, J., James, R., Ruxton, G. D. & Franks, N. R. Collective memory and spatial sorting in animal groups. J. Theor. Biol. 218, 1–11 (2002). 6. Faucher K., Parmentier E., Becco C., Vandewalle N., and Vandewalle P. Fish lateral system is required for accurate control of shoaling. Anim. Behav. 79, 679–687 (2010). 7. Blaxter J.H.S., Structure and development of the lateral line. Biol. Rev. Camb. Philos. Soc. 62, 471–514 (1987). 8. Dahl P. H., Miller J. J, Cato D.H., and Andrew R.K. Underwater ambient noise. Acoustics Today, January 2007. 9. Bjerknes V. F. K. Fields of Force. Columbia University Press, New York, (1906). 10. Love R. H. Resonant acoustic scattering by swimbladder-bearing fish. J. Acoust. Soc. Am. 64, 571–580 (1978). 11. Minnaert F. On musical air bubbles and the sounds of running water. Philos. Mag. 16, 235–248 (1933). 12. Feuillade C. Nero R. W., and Love R. H. A low-frequency acoustic scattering model for small schools of fish, J. Acoust. Soc. Am. 99, 196–208 (1996). 13. Leighton T. The Acoustic Bubble. Academic Press, (1994) ISBN9780124419209. 14. Otma R. Oscillations of a cloud of bubbles of small and not so small amplitude. J. Acous. Soc. Am., 82, 1018-1033. (1987). 15. D'Agostino L & Brennen C. Acoustical absorption and scattering cross section of spherical bubble clouds. J. Acous. Soc. Am., 84, 2126-2134 (1988). 16. Hahn T. R. Low frequency sound scattering from spherical assemblages of bubbles using effective medium theory. J. Acoust. Soc. Am. 122, 3252–3267 (2007). 17. Raveau M. and Feuillade C. Resonance scattering by fish schools: A comparison of two models. J. Acoust. Soc. Am. 139, 163–175 (2016). 18. Doinikov A. A. Bjerknes forces and translational bubble dynamics. Bubble and Particle Dynamics in Acoustic Fields. Modern Trends and Applications. Research Signpost, Trivandrum, Kerala, India, Editors: Doinikov A. A., pp.95-143. 19. Crum, L. A. Bjerknes forces on bubbles in a stationary sound field. J. Acoust. Soc. Am. 57, 1363-1370 (1975). 20. Lanoy M., Derec C., Tourin A., and Leroy V. Manipulating bubbles with secondary Bjerknes forces. Appl. Phys Lett. 107, 214101-4 (2015). 21. Rabaud D., Thibault P., Mathieu M., and Ph. Marmottant. Acoustically bound microfluidic bubble crystals. Phys. Rev. Lett. 106, 134501 (2011). 22. Ida M. Alternative interpretation of the sign reversal of secondary Bjerknes force acting between two pulsating gas bubbles. Phys. Rev. E67, 056617-4 (2003). 23. Wiley J.D. and Seman J.A. The enumeration of neighbors on Cubic and Hexagonal- based lattices. Bell System Technical Journal, 48, 366, (1970). 24. Radakov D. V. Schooling in the ecology of fish. J. Wiley, New York (1973) pp 173 25. Zabolotskaya E. A. Interaction of gas bubbles in a sound field, Sov. Phys. Acoust. 30, 365-368 (1984).
1512.06105
2
1512
2015-12-29T00:52:17
Rigidity of transmembrane proteins determines their cluster shape
[ "physics.bio-ph", "cond-mat.soft" ]
Protein aggregation in cell membrane is vital for the majority of biological functions. Recent experimental results suggest that transmembrane domains of proteins such as $\alpha$-helices and $\beta$-sheets have different structural rigidities. We use molecular dynamics simulation of a coarse-grained model of protein-embedded lipid membranes to investigate the mechanisms of protein clustering. For a variety of protein concentrations, our simulations under thermal equilibrium conditions reveal that the structural rigidity of transmembrane domains dramatically affects interactions and changes the shape of the cluster. We have observed stable large aggregates even in the absence of hydrophobic mismatch which has been previously proposed as the mechanism of protein aggregation. According to our results, semi-flexible proteins aggregate to form two-dimensional clusters while rigid proteins, by contrast, form one-dimensional string-like structures. By assuming two probable scenarios for the formation of a two-dimensional triangular structure, we calculate the lipid density around protein clusters and find that the difference in lipid distribution around rigid and semiflexible proteins determines the one- or two-dimensional nature of aggregates. It is found that lipids move faster around semiflexible proteins than rigid ones. The aggregation mechanism suggested in this paper can be tested by current state-of-the-art experimental facilities.
physics.bio-ph
physics
Rigidity of transmembrane proteins determines their cluster shape 1Department of Mechanical Engineering, Sharif University of Technology, P.O. Box: 11155-9567, Tehran, Iran Hamidreza Jafarinia1, Atefeh Khoshnood2, and Mir Abbas Jalali3∗ 2Reservoir Engineering Research Institute, Palo Alto, California 94301, USA 3Department of Astronomy, University of California, Berkeley, California 94720, USA Protein aggregation in cell membrane is vital for the majority of biological functions. Recent experimental results suggest that transmembrane domains of proteins such as α-helices and β-sheets have different structural rigidities. We use molecular dynamics simulation of a coarse-grained model of protein-embedded lipid membranes to investigate the mechanisms of protein clustering. For a variety of protein concentrations, our simulations under thermal equilibrium conditions reveal that the structural rigidity of transmembrane domains dramatically affects interactions and changes the shape of the cluster. We have observed stable large aggregates even in the absence of hydrophobic mismatch, which has been previously proposed as the mechanism of protein aggregation. According to our results, semi-flexible proteins aggregate to form two-dimensional clusters, while rigid proteins, by contrast, form one-dimensional string-like structures. By assuming two probable scenarios for the formation of a two-dimensional triangular structure, we calculate the lipid density around protein clusters and find that the difference in lipid distribution around rigid and semiflexible proteins determines the one- or two-dimensional nature of aggregates. It is found that lipids move faster around semiflexible proteins than rigid ones. The aggregation mechanism suggested in this paper can be tested by current state-of-the-art experimental facilities. PACS numbers: 87.15.kt,87.15.km,87.16.dt I. INTRODUCTION Transmembrane (TM) proteins are regulators of sev- eral cellular processes. In order to perform their func- tions, they often aggregate and distribute non-uniformly in the cell membrane [1]. The aggregation process of proteins sometimes becomes abnormal, and causes amy- loid diseases [2]. In recent years, there has been increas- ing interest in identifying various mechanisms that affect protein–membrane interactions [3, 4] and, consequently, the aggregation behavior of different proteins. The specific structure of TM proteins defines their physical properties and enables them to perform their biological functions [5]. TM proteins differ in size and physical properties [6], and may have single or multiple α-helical [7, 8] or β-structure [9] domains. Recent ex- perimental studies show that α-helical structures have softer domains than β-structures in both dry and hy- drated states, and β-barrels and β-sheets are more rigid structural units than α-helices [10]. The higher num- ber of hydrogen bonds per residue is also considered to be the probable cause of the more rigid structure of β- sheets compared to α-helices [11]. It has also been shown that the secondary structure of proteins affects the rigid- ity and dynamics of the protein. Proteins containing β- structures have a higher Young's modulus and higher fre- quency of the collective vibration [11]. Consequently, the effect of the class and structural rigidity of proteins on their aggregation behavior and biological function cannot be ignored. ∗ [email protected] Although lipid raft formation [12] and direct linking [13] of proteins cause the aggregation of membrane pro- teins, membrane curvature [14] and membrane-mediated interactions play important roles on the formation and fragmentation of protein clusters. Among mechanisms that generate lipid-mediated protein interactions, the hydrophobic mismatch interaction, which is due to the difference between the hydrophobic lengths of the inte- gral proteins and the hydrophobic thickness of their host membrane, has been widely studied by several groups [4, 15–17]. However, we know little about the effect of the structural properties of proteins on the cluster formation. It is known that the shapes and sizes of proteins deter- mine the distribution of their surrounding lipid molecules [17–20], which in turn affect the stability and function of proteins [21, 22]. What is poorly understood is how the structural rigidity of proteins integrates with other fac- tors to shape the patterns of aggregates. In this paper, we use coarse-grained molecular dynam- ics simulations to systematically investigate the effect of structural rigidity of TM proteins on the formation of clusters. We design specific model proteins to exclude the effect of hydrophobic mismatch and isolate the role of structural rigidity. We use two sets of proteins: semi- flexible and rigid. In §II, we present our model and sim- ulation method and setup. In §III, the results of molec- ular dynamics simulations are presented. We discuss our findings in §IV and show how they are comparable with experimental observations and previous theoretical mod- elings. Our simulations show that proteins form clusters even in the absence of hydrophobic mismatch and the fi- nal shapes of protein aggregates in lipid bilayers depend strongly on the rigidity of proteins. II. MODEL AND METHODS The model of lipid molecules is composed of one hy- drophilic head particle and a hydrophobic tail chain [23, 24], which contains four particles. Our TM pro- teins are modeled as hexagonal prisms with middle hy- drophobic particles and hydrophilic groups at both ends [16, 18]. The hydrophobic mismatch is tuned by chang- ing the length ∆r = rp − rl of the hydrophobic part of proteins, where rp is the length of the hydrophobic part of proteins, and rl is the average thickness of the hy- drophobic part of the bilayer. Models of lipid and protein molecules and their corresponding bonds are displayed in Fig. 1. Hydrophilic and hydrophobic particles are la- beled H and T, respectively. In this study, our length scale is σ = 1/3 nm and the energy unit is NAǫ = 2 kJ/mol, with NA being the Avogadro number. In each lipid or protein molecule, the adjacent ith and (i + 1)th particles interact through the harmonic bond potential Ub(ri,i+1) = kb(ri,i+1 − σeq)2, (1) where ri,i+1 is the distance between particles and σeq is the equilibrium length of the bonds. Three kinds bonds- planar, vertical, and oblique-are used to build two types of protein molecules-rigid and semiflexible. The planar and vertical bonds have identical spring con- stants of kb1 = 5000 ǫ/σ2, which is fixed in all simula- tions. For the oblique bonds of rigid proteins (RPs) and semiflexible proteins (SFPs) we set kb1 = 5000 ǫ/σ2 and kb2 = 35 ǫ/σ2, respectively. Both SFPs and RPs are more rigid than lipid molecules. We set σeq = σ for lipid bonds and planar and vertical bonds of proteins. For oblique protein bonds, we use σeq = √2 σ. SFPs are stiffer than lipid molecules and mostly maintain their hexagonal cross section when they are bent due to inter- actions with other proteins and lipids. The angle between consecutive bonds in a lipid molecule is controlled by U = ka(cos θ − cosθeq)2, (2) where ka = 1.85 ǫ and θeq = π. Interactions between the particles of different molecules are governed by soft-core and Lennard-Jones potentials, defined as , rij (cid:19)9 Usc(rij ) = 4ǫ(cid:18) σsc ULJ(rij ) = 4ǫ"(cid:18) σ rij(cid:19)12 −(cid:18) σ rij(cid:19)6# , (3) (4) where rij = ri− rj and σsc = 1.05 σ. In these equations, ri is the global position vector of the ith particle. The repulsive soft-core potential is used to model the interac- tion between hydrophobic and hydrophilic particles, and between solvent and hydrophobic particles. All other in- teractions are modeled by the Lennard-Jones potential. A cutoff radius of rc = 2.5 σ is applied to Usc and ULJ, which are then shifted in order to vanish at rij = 2.5 σ 2 Protein building blocks Lipid molecule Protein molecule FIG. 1. (Color online) Models of protein and lipid molecules and their bonds. In the three-dimensional perspective view of the protein, oblique bonds are shown by dashed lines. Hy- drophilic and the hydrophobic particles of the HT5H pro- tein are depicted as red and yellow spheres, respectively. Hy- drophilic and hydrophobic particles of lipid molecules are rep- resented by green and blue spheres, respectively. These col- oring schemes are also used in the snapshots of simulations. [23]. This guarantees the continuity of both potential fields. We carry out molecular dynamics simulations of N V T ensembles using the LAMMPS package. Periodic bound- ary conditions are imposed and the temperature is kept constant at T0 = 310 K (with kBT = 1.29 ǫ) utilizing the Nose-Hoover thermostat. Here kB is the Boltzmann constant and the lipid bilayer is in the liquid phase. All particles have the same mass, NAm = 36 gr/mol [23]. The integration time step is set to ∆t = 0.005 τ , where τ =pmσ2/ǫ is the intrinsic time scale. The actual value of ∆t is 7.07 fs. In all simulations, the number density of particles is n = 0.66/σ3 and the area per lipid (for a bilayer without proteins) is As = 2.09 σ2. As is the area of lipid bilayer divided by the number of lipids. We select the number of lipids such that bilayers with mini- mum surface tension and without permanent curvatures are obtained. This helps us to eradicate the effect of curvature-mediated interactions. Under this condition the surface tension of the bilayer is positive and approxi- mately equals 0.24 ǫσ−2, corresponding to 7 mN/m. This has led to stable bilayers in all of our simulations. It is remarked that the rupture surface tension of biological membranes varies from 1 to 30 mN/m, and it depends on the lipid composition of the bilayer [25, 26]. Simula- tions are performed for various concentrations of proteins embedded in the bilayer. We denote the concentration of proteins cp = Np/(Np + Nl), where Np and Nl are the numbers of protein and lipid particles, respectively. At the beginning of simulations, each protein molecule is placed in the bilayer by removing nine lipid molecules. RP SFP RP SFP 3 III. RESULTS We investigate the clustering process for low and high protein concentrations and for two protein types: RPs and SFPs. To quantify the flexibilities of proteins, we have compared the longitudinal flexibilities of our protein models with each other and, also, with a lipid patch that occupies the same area that proteins do. The torsional flexibilities of protein models have also been computed. We have carried out simulations for a single RP and SFP in vacuum as well as in a lipid bilayer and calculated the standard deviation of the length of the hydropho- bic part for each protein. A bilayer consisting of nine lipid molecules in a box of size 3.06σ × 3.06σ × 10σ was simulated. The area of this bilayer patch approximately equals the area of proteins, and it has the same area per lipid (As = 2.09σ2) as other bilayers in our simulations. Let us define the lengths of the hydrophobic parts of the RP and SFP by rRP and rSFP, respectively. Denoting the standard deviation SD(·), we find SD(rRP) = 0.025σ, SD(rSFP) = 0.067σ, (5) for proteins in vacuum, SD(rRP) = 0.025σ, SD(rSFP) = 0.055σ, (6) for proteins in the bilayer, and SD(rl) = 0.32σ for the lipid bilayer. It is seen that transmembrane RPs are stiffer than SFPs, and SFPs are stiffer than a bilayer that has the same surface area of proteins. Due to the lack of interactions with lipid and solvent molecules, SFPs are more flexible in vacuum. The weaker oblique bonds result in more torsional flex- ibility for SFPs. The torsional rigidity of proteins can be measured by the relative twist angle of their two hy- drophilic ends. Defining θRP and θSFP as the relative twist angles of the upper and lower hydrophobic parts, we obtain SD(θRP) = 0.11◦, SD(θSFP) = 0.28◦, (7) for proteins in vacuum and SD(θRP) = 0.11◦, SD(θSFP) = 0.22◦, (8) for proteins in the bilayer. These results show that RPs are torsionally stiffer than SFPs. Our findings are con- sistent with the higher conformational displacements of α-helices compared to β-sheets [11, 27]. In all of our simulations, the effect of hydrophobic mis- match is neutralized, ∆r = 0.01, by using the HT5H model proteins. Figure 2 shows snapshots of protein- embedded membranes with SFPs and RPs. We have used N = 3, 4, 6, and 8 protein molecules for both protein types. The size of the lipid bilayer is 16 σ×16 σ for N = 3, 23 σ × 23 σ for N = 4 and 6, and 35 σ × 35 σ for N = 8. The snapshots have been taken at t = 5 × 106∆t. The RP SFP RP SFP FIG. 2. (Color online) Shapes of clusters formed by SFPs and RPs. Top left: N = 3. Top right: N = 4. Bottom left: N = 6. Bottom right: N = 8. It is evident that RPs form string-shaped clusters. difference between SFP and RP clusters is prominent: RPs form string-like one-dimensional structures, while SFPs form two-dimensional clusters. These results dif- fer significantly from previous work [15], which suggests that there is only weak attraction between inclusions in the absence of mismatch. Our results, clearly, show that protein inclusions form stable clusters in the absence of hydrophobic mismatch. One of the distinct features of SFPs is their clustering in triangular structures. For the model with six and eight SFPs, two triangular structures can be seen in Fig. 2. RPs do not share this property. To quantify the discrepancies between one-dimensional and two-dimensional clusters, we measure the radius of gyration of protein structures as R2 g = 1 M Np Xi=1 mi ri − rc2 , (9) t1 t2−t1 R t2 where Np and M are the total number and the total mass of the head particles of proteins, respectively, and rc and ri define the position vectors of the ith par- ticle and the center of mass of protein heads, respec- tively. The mass of each particle is denoted mi. To com- pute Rg, we have used protein heads lying in one mono- layer. The variation of Rg over time has been plotted in Fig. 3 for a system with four and seven proteins. We have also plotted the time-averaged radius of gyration, hRgi = 1 Rg(t) dt, for several number of proteins in a cluster. We have used (t1, t2) = (30, 35 ns) for both RPs and SFPs. It is shown that for the same number of proteins, two-dimensional structures formed by SFPs always have a lower radius of gyration; the radius of gy- ration increases proportionally to the number of proteins in the cluster. Since the linear aggregates have more con- tact and interaction with surrounding lipids, these struc- tures are more likely to change their shape slightly, for example, from a completely straight linear structure to a curved line, which, in turn, alters the radius of gyration. As Fig. 2 shows, string-like clusters have a variety of configurations with large variations in their gyration radii (see Fig. 3). In simulations of SFPs we have observed de- 3.6 3.4 3.2 3 2.8 2.6 2.4 2.2 σ / g R Rigid Semi−flexible 5.5 5 4.5 4 3.5 σ / g R Rigid Semi−flexible 5 4.5 4 3.5 3 2.5 2 σ / > g R < 4 Rigid Semi−flexible 2 0 5 10 15 20 t(ns) 25 30 35 3 0 5 10 15 20 t(ns) 25 30 35 1.5 2 3 4 5 6 7 8 9 N FIG. 3. (Color online) Temporal evolution of the gyration radius of protein clusters for N = 4 (left), and N = 7 (middle). Black lines correspond to RPs; red lines to SFPs. Right panel : Time-averaged gyration radius versus number of proteins, N , in a cluster. The radius of gyration is consistently larger for RPs. formed cross sections of proteins (deviations from hexag- onal shapes), especially when they do not belong to a cluster. Deformations of proteins can occasionally stabi- lize them in the membrane, so that they do not partici- pate in cluster forming processes. So SFPs are stabilized by either contributing to a cluster or undergoing deforma- tion while they are singly dispersed in the membrane. It is remarked that we have repeated our simulations start- ing with different initial conditions and observed similar clustering trends for each protein class. Moreover, we continued our simulations over longer time scales, up to 200 ns, and obtained the same results as in our 35-ns-long simulations for both SFP and RP clusters. Figure 4 shows the temporal evolution of cluster for- mation from an initially random distribution of pro- teins. Interestingly, RPs immediately aggregate to one- dimensional clusters, whereas SFPs first make a cluster with several branches separated by lipid molecules, then evolve to a compact cluster as trapped lipids are released. t = 0 t = 1.2 ns t = 35 ns t = 0 t = 1.2 ns t = 35 ns FIG. 4. (Color online) Clustering process for SFPs (top) and RPs (bottom). Time increases from left to right. To eliminate possible boundary effects in simulation boxes with periodic boundary conditions, and also under- stand how the concentration of proteins affect the aggre- gation process, we have carried out simulations in a larger membrane, of size 90σ× 90σ with higher protein concen- trations, cp = 0.2, 0.27, 0.35, and 0.48. We have studied several snapshots of these simulations and our previous conclusions for smaller membranes are unaltered. A few more results for high-cp simulations are as follows: (i) SFPs rarely form one-dimensional clusters, though the lengths of such clusters are much shorter than the ag- gregates of RPs; (ii) one-dimensional structures of RPs may connect to each other to form longer or branched web-like structures; and (iii) on rare occasions RPs clus- ter as two-dimensional domains. Observations i and iii might be due to the longer relaxation time scales of large membranes with a high cp. The structure formation pro- cess by RPs and SFPs is better understood by computing the average number of neighboring proteins, Nav. For a given (subject) protein i, we find the number of neigh- boring proteins Ni within a distance of 6.1 σ, measured from the center of mass of the protein. We then cal- culate the average number Nav = 1 i=1 Ni. Figure 5 shows the variation of Nav over time for two protein types and several protein concentrations. It is shown that Nav,SFP is consistently larger than Nav,RP, and the dif- ference ∆Nav = Nav,SFP − Nav,RP increases versus time as the aggregate size and structure reach steady state. We note that the time for Nav to reach the steady-state value is shorter in systems with RPs and at higher pro- tein concentrations. N PN We explain the physical origin of different cluster- forming pathways of SFPs and RPs by investigating the distribution of lipids around proteins [17, 19]. Our nu- merical simulations show that the lipid head and tail den- sities around a subject protein depend on the rigidity of the protein: RPs induce order and structure in sur- rounding lipids, which have formed ring-like structures around proteins (Fig. 6). The lipid density around rigid inclusions is generally lower around SFPs because the 7 6 5 4 3 2 1 v a N (c p (c p (c p (c p =0.20) RP =0.20) SFP =0.27) RP =0.27) SFP =0.35) (c p RP =0.35) (c p SFP =0.48) (c p RP =0.48) (c p SFP 0 0 5 10 15 20 t (ns) 25 30 35 FIG. 5. (Color online) Evolution of the average number of neighboring proteins, Nav, for SFPs and RPs and four choices of protein concentrations. The difference ∆Nav = Nav Nav SFPs participate in the cluster formation process. RP between SFPs and RPs increases over time as more , SFP − , thermally vibrating flexible structure of relatively mas- sive proteins continuously kicks and scatters lighter lipid molecules. For a complex of two proteins, the ordered structure of lipids takes an oval shape and constrains the formation of a complex with three proteins. To investigate how the process works, we designed two experimental scenarios. In the first scenario, two pro- teins are kept close to each other almost at the center of the lipid membrane by means of a spring of constant 500 ǫ/σ2. We then fix the position of the third protein on the double-complex vertical symmetry line at d = 4 σ [Fig. 6(a)] with the same spring constant. The next step is to obtain the lipid distribution around the three- protein complex. When the triple-complex is composed of RPs, the high density of trapped lipids between the third protein and the double ones prohibits the forma- tion of a triple, triangular-shaped complex as Fig. 6(b) shows. For SFPs, however, lipids are weakly bound to the double-complex, easily diffuse out of the triple-contact area, and thus facilitate the formation of bigger two- dimensional clusters if the constraint is released [Fig. 6(c)]. In the second scenario, the third protein is placed close to the double-complex and along an oblique line with an angle of θ = 2π/3 [Fig. 6(d)], forming a string-like structure. In the case of three RPs, lipids are trapped between two proteins and have completely filled the re- gion between the two adjacent proteins. Therefore, these lipids resist the reduction of θ if the constraint is re- leased. This is how RPs maintain their one-dimensional structure [Fig. 6(e)]. Note that in Figs. 6(a) and 6(d), the arrows show the most probable path that the third protein chooses to follow in order to form a triangular 5 structure if the constraint is released. The distribution of lipids around SFPs is more homogeneous than that around RPs; compare Figs. 6(e) and 6(f). Our sim- ulations with three proteins show that lipid molecules more easily diffuse in the space between SEPs and re- sult in different arrangements of SEPs compared to RPs. More diffusive lipids around SFPs can also be identified by analyzing the mean squared displacement of lipids. For cp = 0.48, we have calculated the mean squared displacement for lipid molecules using the method in [28]. Let us define rk(t) as the component of the po- sition vector r(t) of lipid particles parallel to the bi- layer surface. The two-dimensional diffusion coefficients D = limt→∞hrk(t) − rk(0)2i/(4t) that we find are ap- proximately 43.1 × 10−7 and 56.2 × 10−7 cm2/s for the ensemble of lipid molecules around RPs and SFPs, re- spectively. The diffusion coefficients are different from experimental data because the model is coarse grained. Here the relative change in diffusion coefficients is im- portant. Our results show a 30% reduction in diffusion coefficient for a membrane with RPs in comparison with one hosting SEPs. This is consistent with our observa- tion of more restricted lipids surrounding RPs. We note that diffusion coefficients are measured from the part of the mean squared displacement profile that is linear in time. The initial anomalous region is not included in the calculations [28]. IV. DISCUSSION Using coarse-grained molecular dynamic simulations, we showed that the rigidity of proteins has a profound effect on the cluster shape of TM proteins in lipid mem- branes. For proteins with zero hydrophobic mismatch, regardless of the protein concentration, RPs aggregate to form one-dimensional clusters while SFPs form two- dimensional clusters. In contrast to previous studies [4, 15–17] where hydrophobic mismatch is considered to be essential for clustering, we showed that proteins form stable clusters even in the absence of mismatch. Lipid- induced depletion interactions have been suggested as one of the main contributing factors to the interactions of cylindrical inclusions in bilayers [4, 29, 30]. This type of attraction occurs at distances smaller than the diameter of one lipid molecule. It has also been suggested that the interaction between membrane proteins largely depends on indirect lipid-mediated interactions [4]. Therefore, in this study, the interactions between proteins at short dis- tances are most likely due to a strong depletion force of entropic origin explained by the Asakura-Oosawa model. This is a consequence of the high flexibility of the lipid chains. Direct protein–protein interactions can also play a role in the aggregation of proteins when protein parti- cles are in the range of the cut-off distance. Khoshnood et al. [31] report on the same effect of depletion force for aggregation of rigid inclusions com- pared with completely flexible ones while hydrophobic (a) d (d) (b) σ / Y (e) σ / Y 14 12 10 8 6 4 2 14 12 10 8 6 4 2 9 8 7 6 7 8 9 10 2 4 6 8 10 12 14 X /σ 11 10 9 8 6 7 8 9 9 8 7 6 7 8 9 10 2 4 6 8 10 12 14 X /σ 11 10 9 8 6 7 8 9 (c) σ / Y (f) σ / Y 14 12 10 8 6 4 2 14 12 10 8 6 4 2 2 4 6 8 10 12 14 2 4 6 8 10 12 14 X /σ X /σ 6 0.08 0.07 0.06 0.05 0.04 0.03 0.02 0.01 0 0.08 0.07 0.06 0.05 0.04 0.03 0.02 0.01 0 FIG. 6. (Color online) Understanding protein–protein interactions. (a) The third protein is placed at d = 4 σ on the vertical symmetry line of the double-protein complex. (b) Density of lipid heads around RPs. (c) Density of lipid heads around SFPs. It is shown that lipids are more scattered and less concentrated at the triple contact region in (c) compared with (b). (d) The third protein is placed along an oblique line with angle θ = 2π/3 relative to the double-protein complex. (e) Density of lipid heads around RPs. The high concentration of lipids near the contact zones prohibits θ from tending to small angles if the angle constraint is released. (f) Density of lipid heads around SFPs. Lipids are scattered enough to allow for the rotation of θ, and consequently, the formation of a triangular-shaped complex if the angle constraint is released. All insets: Closeups of contact regions. mismatch exists. A key factor in association of proteins is the distribution of lipid molecules in close proximity to the inclusions [20]. These lipid molecules lose their entropy due to interaction with the lateral surface of pro- teins and form patterns around the inclusion. These pat- terns are more structured around RPs than SFPs; com- pare Figs. 6(b) and 6(e) with Figs. 6(c) and 6(f), re- spectively. The strong induced lipid structures around RPs lead to a low mobility of lipids, and consequently they have lower diffusion coefficients compared to lipids in a system with SFPs. As a result, the formation of a two-dimensional cluster, is obstructed by the induced lipid structure surrounding RPs, and in such systems one-dimensional aggregates are dominant. We have observed fluctuation-induced attraction be- tween two RPs and SFPs by applying the same constraint described in [30]. This type of long-range interaction is caused by inclusions that affect the elastic properties of membranes and hence the fluctuation of lipid molecules. Long-range interactions play important roles in the clus- tering of both RPs and SFPs. However, the shape of the clusters is determined by a repulsive zone when pro- teins are close to each other and form groups of three or more proteins. The repulsive zone occurs when the dis- tances between proteins are slightly larger than the size of one lipid molecule (beyond the depletion zone) [29, 30]. When a dimer forms, the fluctuation-induced attraction between proteins overcomes the effect of the repulsive zone and RPs and SFPs are able to create string-like aggregates at the early stages of simulations. The ef- fect of the repulsive zone becomes more prominent when RPs want to create two-dimensional aggregates and can not be overcome by the fluctuation-induced attraction between inclusions. In the case of SFPs, the fluctuation- induced attraction overcomes the effect of the repulsive zone and pushes the third protein into the depletion zone. Both the fluctuation-induced attraction and the repulsive zone are classified as lipid-mediated interactions. Cell adhesion to extracellular matrices is regulated by the size and position of focal adhesions and, most im- portantly, how they are distributed [32]. The latter is controlled by integrin protein association. Integrin has both β and α subunits and the number of β and α sub- units varies depending on the type of integrin, which means a variety of structural rigidity. According to our results variation in integrin stiffness affects their aggre- 7 gation patterns and consequently may have an impact on the quality of cell adhesion. Cell receptor activation depends on receptors' cluster- ing and their conformational changes. The latter requires rearrangements of proteins in the cluster [33]. It means that the pattern in which receptors are attached to each other in a cluster plays a role in the activation process. Receptors may have subunits with different structural rigidities, and based on our findings this feature can af- fect their final aggregation pattern. Remodeling of the biomembrane is achievable by pro- teins [34] and nanoparticles [35], and it is a vital step in endocytosis and vesiculation [24]. Interestingly, in a system of spherical nano-particles on lipid membranes [35] different aggregation patterns have been induced by varying the membrane rigidity. The majority of living cell membranes are made up of phospholipids and the rigidity of the bilayer is known. Our results suggest that flexibility of the inclusion may work as a controlling pa- rameter for membrane remodeling when we can not al- ter the membrane rigidity. The new controlling param- eter can also help in the design of therapeutic peptides to tackle protein-aggregation diseases. This is obtain- able by protein engineering methods without perturbing significantly the overall stability or activity of the pro- tein [36]. Different mixtures of amino acids have differ- ent amounts of alpha-helical and beta-structural units and consequently have different rigidity [10]. Therefore, it is possible to design proteins with a specifically de- fined rigidity and certain aggregation behavior, which lead them to perform a specific biological function. The protein model in our study is a toy model with SFPs and RPs resembling α-helices and β-sheets, respec- tively, as their structural rigidities differ substantially [10]. The simulations by Parton et al. [37] with α-helical and β-barrel proteins on vesicles and flat membranes show that while α-helical proteins form two-dimensional clusters, β-barrel proteins constitute linear aggregates. In their experiments, a combination of several factors such as hydrophobic mismatch, membrane curvature, and the shape and class of proteins can be held respon- sible for the observed discrepancies in the aggregation patterns of proteins. Our results demonstrate that struc- tural rigidity as a sole factor determines the shape and pattern of protein aggregates. [1] J. J. Sieber, K. I. Willig, C. Kutzner, C. Gerding- Reimers, B. Harke, G. Donnert, B. Rammner, C. Eggeling, S. W. Hell, and H. Grubmuller, Science 317, 1072 (2007). Journal 88, 1778 (2005). [16] U. Schmidt, G. Guigas, and M. Weiss, Physical Review Letters 101, 128104 (2008). [17] F. J.-M. De Meyer, M. Venturoli, and B. Smit, Biophys- [2] F. Chiti and C. M. Dobson, Annu. Rev. Biochem. 75, ical Journal 95, 1851 (2008). 333 (2006). [18] U. Schmidt and M. Weiss, Biophysical Chemistry 151, [3] O. G. Mouritsen and M. Bloom, Biophysical Journal 46, 34 (2010). 141 (1984). [19] D. Morozova, M. Weiss, and G. Guigas, Soft Matter 8, [4] B. West, F. L. Brown, and F. Schmid, Biophysical Jour- 11905 (2012). nal 96, 101 (2009). [5] M. J. Buehler, Proceedings of the National Academy of Sciences 103, 12285 (2006). [20] J. Yoo and Q. Cui, Biophysical Journal 104, 128 (2013). [21] M. Ø. Jensen and O. G. Mouritsen, Biochimica et Bio- physica Acta (BBA)-Biomembranes 1666, 205 (2004). [6] T. Becker, M. Gebert, N. Pfanner, and M. van der Laan, [22] A. G. Lee, Biochimica et Biophysica Acta (BBA)- Current Opinion in Cell Biology 21, 484 (2009). Biomembranes 1666, 62 (2004). [7] E. V. Bocharov, P. E. Volynsky, K. V. Pavlov, R. G. Efremov, and A. S. Arseniev, Cell adhesion & migration 4, 284 (2010). [8] G. H. Westfield, S. G. Rasmussen, M. Su, S. Dutta, B. T. DeVree, K. Y. Chung, D. Calinski, G. Velez-Ruiz, A. N. Oleskie, E. Pardon, et al., Proceedings of the National Academy of Sciences 108, 16086 (2011). [23] R. Goetz and R. Lipowsky, The Journal of Chemical Physics 108, 7397 (1998). [24] B. J. Reynwar, G. Illya, V. A. Harmandaris, M. M. Muller, K. Kremer, and M. Deserno, Nature 447, 461 (2007). [25] E. Evans, V. Heinrich, F. Ludwig, and W. Rawicz, Bio- physical Journal 85, 2342 (2003). [9] C. Tang, Y. Zhao, R. Wang, C. H´elix-Nielsen, and [26] D. Needham and R. S. Nunn, Biophysical Journal 58, A. Fane, Desalination 308, 34 (2013). 997 (1990). [10] S. Perticaroli, J. D. Nickels, G. Ehlers, and A. P. Sokolov, Biophysical Journal 106, 2667 (2014). [11] S. Perticaroli, J. D. Nickels, G. Ehlers, H. O'Neill, Q. Zhang, and A. P. Sokolov, Soft Matter 9, 9548 (2013). [12] T. J. McIntosh and S. A. Simon, Annual Review of Bio- [27] A. M. Gaspar, M.-S. Appavou, S. Busch, T. Unruh, and W. Doster, European Biophysics Journal 37, 573 (2008). [28] A. Khoshnood and M. A. Jalali, Physical Review E 88, 032705 (2013). [29] P. Lague, M. J. Zuckermann, and B. Roux, Biophysical physics and Biomolecular Structure 35, 177 (2006). Journal 81, 276 (2001). [13] G. Feng, H. Tintrup, J. Kirsch, M. C. Nichol, J. Kuhse, [30] T. Sintes and A. Baumgartner, Biophysical Journal 73, H. Betz, and J. R. Sanes, Science 282, 1321 (1998). 2251 (1997). [14] A. H. Bahrami and M. A. Jalali, The Journal of Chemical [31] A. Khoshnood, H. Noguchi, and G. Gompper, The Jour- Physics 134, 085106 (2011). nal of Chemical Physics 132, 025101 (2010). [15] M. Venturoli, B. Smit, and M. M. Sperotto, Biophysical [32] K. K. Elineni and N. D. Gallant, Biophysical Journal 101, 2903 (2011). [35] A. Sari´c and A. Cacciuto, Physical Review Letters 108, [33] S. Minguet, M. Swamy, B. Alarc´on, I. F. Luescher, and 118101 (2012). W. W. Schamel, Immunity 26, 43 (2007). [34] M. Simunovic, A. Srivastava, and G. A. Voth, Proceed- ings of the National Academy of Sciences 110, 20396 (2013). [36] V. Villegas, J. Zurdo, V. V. Filimonov, F. X. Avil´es, C. M. Dobson, and L. Serrano, Protein Science 9, 1700 (2000). [37] D. L. Parton, J. W. Klingelhoefer, and M. S. Sansom, Biophysical Journal 101, 691 (2011). 8
1203.2673
2
1203
2012-03-14T19:30:49
Machines of life: catalogue, stochastic process modeling, probabilistic reverse engineering and the PIs- from Aristotle to Alberts
[ "physics.bio-ph", "cond-mat.stat-mech", "q-bio.SC" ]
Molecular machines consist of either a single protein or a macromolecular complex composed of protein and RNA molecules. Just like their macroscopic counterparts, each of these nano-machines has an engine that "transduces" input energy into an output form which is then utilized by its coupling to a transmission system for appropriate operations. The theory of heat engines, pioneered by Carnot, rests on the second law of equilibrium thermodynamics. However, the engines of molecular machines, operate under isothermal conditions far from thermodynamic equilibrium. Moreover, one of the possible mechanisms of energy transduction, popularized by Feynman and called Brownian ratchet, does not even have any macroscopic counterpart. But, {\it molecular machine is not synonymous with Brownian ratchet}; a large number of molecular machines actually execute a noisy power stroke, rather than operating as Brownian ratchet. The man-machine analogy, a topic of intense philosophical debate in which many leading philosophers like Aristotle and Descartes participated, was extended to similar analogies at the cellular and subcellular levels after the invention of optical microscope. The idea of molecular machine, pioneered by Marcelo Malpighi, has been pursued vigorously in the last fifty years. It has become a well established topic of current interdisciplinary research as evident from the publication of a very influential paper by Alberts towards the end of the twentieth century. Here we give a non-technical overview of the strategies for (a) stochastic modeling of mechano-chemical kinetic processes, and (b) model selection based on statistical inference drawn from analysis of experimental data. It is written for non-experts and from a broad perspective, showing overlapping concepts from several different branches of physics and from other areas of science and technology.
physics.bio-ph
physics
Machines of life: catalogue, stochastic process modeling, probabilistic reverse engineering and the PIs- from Aristotle to Alberts Debashish Chowdhury1, 2, ∗ 1Department of Physics, Indian Institute of Technology, Kanpur 208016, India 2Mathematical Biosciences Institute, The Ohio State University, Columbus, OH 43210, USA. (Dated: July 2, 2018) Molecular machines consist of either a single protein or a macromolecular complex composed of protein and RNA molecules. Just like their macroscopic counterparts, each of these nano-machines has an engine that "transduces" input energy into an output form which is then utilized by its cou- pling to a transmission system for appropriate operations. The theory of heat engines, pioneered by Carnot, rests on the second law of equilibrium thermodynamics. However, the engines of molecular machines, operate under isothermal conditions far from thermodynamic equilibrium. Moreover, one of the possible mechanisms of energy transduction, popularized by Feynman and called Brownian ratchet, does not even have any macroscopic counterpart. But, molecular machine is not synony- mous with Brownian ratchet; a large number of molecular machines actually execute a noisy power stroke, rather than operating as Brownian ratchet. The man-machine analogy, a topic of intense philosophical debate in which many leading philosophers like Aristotle and Descartes participated, was extended to similar analogies at the cellular and subcellular levels after the invention of opti- cal microscope. The idea of molecular machine, pioneered by Marcelo Malpighi, has been pursued vigorously in the last fifty years. It has become a well established topic of current interdisciplinary research as evident from the publication of a very influential paper by Alberts towards the end of the twentieth century. Here we give a non-technical overview of the strategies for (a) stochastic modeling of mechano-chemical kinetic processes, and (b) model selection based on statistical in- ference drawn from analysis of experimental data. It is written for non-experts and from a broad perspective, showing overlapping concepts from several different branches of physics and from other areas of science and technology. PACS numbers: 87.16.Ac 89.20.-a ∗Electronic address: [email protected] Table of contents: 1. Introduction II. A catalogue: typical examples of molecular machines and fuels II.A. Fuels for molecular machines II.B. Cytoskeletal motors II.B.1. Porters II.B.2. Machines for chipping filamentous tracks II.B.3. Contractility: motor-filament crossbridge and collective dynamics of sliders and rowers II.B.4. Push and pull of cytoskeletal filaments: nano-pistons and nano-springs II.C. Machines for synthesis, manipulation and degradation of macromolecules of life II.C.1. Membrane-associated machines for macromolecule translocation: exporters, importers and packers II.C.2. Machines for degrading macromolecules of life II.C.3. Machines for template-dictated polymerization II.C.4. Unwrappers and unzippers of packaged DNA: chromatin remodellers and helicases II.D. Rotary motors II.E. Processivity, duty ratio, stall force and dwell time II.F. Fundamental general questions III. Motoring in a viscous fluid: from Newton to Langevin III.A. Newton's equation: deterministic dynamics III.B. Langevin equation: stochastic dynamics III.B.1. Motoring in a ''sticky'' medium: Purcell's idea 2 III.B.2. Motoring under random impacts from surroundings: Brownian force IV. Energy transduction by molecular machines: from Carnot to Feynman IV.A. Molecular machines are run by isothermal engines IV.B. Inadequacy of equilibrium thermodynamics IV.C. Inadequacy of endo-reversible thermodynamics IV.D. Domain of non-equilibrium statistical mechanics IV.E. Defining efficiency: from Carnot to Stokes IV.F. Noisy Power stroke and Brownian ratchet IV.G. Physical realizations of Brownian ratchet in molecular motors V. Mathematical description of mechano-chemical kinetics: continuous landscape versus discrete networks V.A. Motor kinetics as wandering in a time-dependent mechano-chemical free-energy landscape V.B. Motor kinetics as wandering in the time-dependent real-space potential landscape V.C. Motor kinetics as a jump process in a fully discrete mechano-chemical network V.D. Balance conditions for mechano-chemical kinetics: cycles and flux VI. Solving forward problem by stochastic process modeling: from model to data VI.A. Average speed and load-velocity relation VI.B. Jamming on a crowded track: flux-density relation VI.C. Information processing machines: fidelity versus power and efficiency VI.D. Beyond average: dwell time distribution VI.D.1. DTD for a motor that never steps backward VI.D.2. Conditional DTD for motors with both forward and backward stepping VII. A summary of experimental techniques: ensemble versus single-machine VIII. Solving inverse problem by probabilistic reverse engineering: from data to model VIII.A. Frequentist versus Bayesian approach VIII.A.1. Maximum-likelihood estimate VIII.A.2. Bayesian estimate VIII.A.3. Hidden Markov models VIII.B. Extracting FP-based models from data? IX. Overlapping research areas IX.A. Symmetry braeking: directed motility and cell polarity IX.B. Self-organization and pattern formation: assembling machines and cellular morphogenesis IX.C. Dissipationless computation: polymerases as ''tape-copying Turing machines'' IX.D. Enzymatic processes: conformational fluctuations, static and dynamic disorder IX.E. Applications in biomimetics and nano-technology X. Concept of biological machines: from Aristotle to Alberts XI. Summary and outlook I. INTRODUCTION Cell is the structural and functional units of life. How "active" is the interior of a living cell? Imagine an un- der water "metro city" which is, however, only about 10µm long in each direction! In this city, there are "high- ways" and "railroad" tracks on which motorized "ve- hicles" transport cargo to various destinations. It has an elaborate mechanism of preserving the integrity of the chemically encoded blueprint of the construction and maintenance of the city. The "factories" not only sup- ply their products for the construction and repair works, but also manufacture the components of the machines. This eco-friendly city re-charges spent "chemical fuel" in uniquely designed "power plants". This city also uses a few "alternative energy" sources in some operations. Finally, it has special "waste-disposal plants" which de- grade waste into products that are recycled as raw mate- rials for fresh synthesis. All the automated processes in this high-tech micro-city are run by nano-machines [1]. This is not the plot of a science fiction, but a dramatized picture of the dynamic interior of a cell. A molecular machine is either a single protein or a macromolecular complex consisting of proteins and RNA molecules [2]. Just like their macroscopic counterparts, these nano-machines take an input (most often, chemical energy) and it transduces input energy into an output. If the output is mechanical work the machine is usually re- ferred to as a molecular motor [3 -- 5]. Similarly, a cell can also be regarded as a micron-size "energy transforming device". The concept of "machine" is not restricted only to subcellular or cellular levels of biological organization. In ancient times, philosophers proposed the concept of "living machine" to describe a whole animal, including a man; we'll discuss the evolution of this concept towards the end of this article. Not only animals, but even plants also move in re- sponse to external stimuli [6, 7] and movements take place in plants at all levels- from whole plant and plant cell to the subcellular level. I cannot resist the tempta- tion of quoting Thomas Huxley's poetic description of cy- toplasmic streaming [8], in particular, and of intracellular motor traffic, in general, in plants [9]: "..the wonderful noonday silence of a tropical forest is, after all, due only to the dulness of our hearing; and could our ears catch the murmur of these tiny maelstroms, as they whirl in the innumerable myriads of living cells which constitute each tree, we should be stunned, as with the roar of a great city". In this article, however, we focus attention almost ex- clusively on molecular machines that drive subcellular processes within living cells. The processes driven by molecular machines include not only intracellular mo- tor transport, but also manipulation, polymerization and degradation of the bio-molecules [10, 11]. Explaining the physical principles that govern these machine-driven pro- cesses will bring us closer to the ultimate goal of biologi- cal physics: understanding "what is life" in terms of the laws of physics (and chemistry). Biomolecular machines operate in a domain where the appropriate units of length, time, force and energy are, nano-meter, milli-second, pico-Newton and kBT , respec- tively (kB being the Boltzmann constant and T is the ab- solute temperature). Aren't the operational mechanism of molecular machines similar to their macroscopic coun- terparts except, perhaps, the difference of scale? NO. In spite of the striking similarities, it is the differences be- tween molecular machines and their macroscopic coun- terparts that makes the studies of these systems so inter- esting from the perspective of physicists. This article is a brief guide for a beginner embark- ing on an exploration of the exciting frontier of research on molecular machines. It begins with a catalog of the known molecular machines and a list of fundamen- tal questions on their operational mechanism that the explorer should try to address. Like any travel guide, this article also cautions the explorer about the counter- intuitive phenomena that await him/her in this new ter- ritory of his/her exploration where he/she may need new sophisticated technical tools that were not needed for handling macroscopic machines. For the explorer, this article also charts a tentative road map along with a sum- mary of the mathematical strategies that he/she might use to make progress in this frontier territory. Excursions to some of the listed border areas that overlap with areas of research in other branches of science and engineering could be enjoyable and intellectually rewarding for the explorer. For any explorer, it is good to know the experi- ence of the pioneers. The final section, on the evolution of the concept of biological machines, is likely to be en- joyed as a dessert by physicists, and as a fresh food for thought by historians and philosophers of science. 3 II. A CATALOGUE: TYPICAL EXAMPLES OF MOLECULAR MACHINES AND FUELS Based on the mode of operation, the biomolecular ma- chines can be divided broadly into two groups. Cyclic machines operate in repetitive cycles in a manner that is very similar to that of the cyclic engines which run our cars. In contrast, some other molecular machines are one-shot machines that exhaust an internal source of free energy in a single round. Force exerted by a compressed spring, upon its release, is a typical example of a one-shot machine. A. Fuels for molecular machines The most common way of supplying energy to a nat- ural nanomachine is to utilize the chemical energy (or, more appropriately, free energy) released by a chemi- cal reaction. Most of the machines use the so-called high-energy compounds- particularly, nucleoside triphos- phates (NTPs)- as an energy source to generate the me- chanical energy required for their directed movement. The most common chemical reaction is the hydrolysis of ATP (ADP): AT P → ADP +Pi. Some other high-energy compounds can also supply input energy; one typical ex- ample being the hydrolysis of Guanosine Triphosphate (GTP) to Guanosine Diphosphate (GDP). Under normal physiological conditions, hydrolysis of ATP is extremely slow. Most of the molecular motors fuelled by ATP func- tion also as ATPase enzyme (i.e., catalyzes hydrolysis of ATP) thereby speeding up the energy-supplying reaction by several orders of magnitude. If a cyclic machine runs on a specific chemical fuel then the spent fuel must be removed as waste products and fresh fuel must be supplied to the machine. Fortunately, normal cells have machineries for recycling waste prod- ucts to manufacture fresh fuel, e.g., synthesizing ATP from ADP. This raises an important question: since ATP is a higher-energy compound than ADP, how are the ATP-synthesizing machines driven to perform this ener- getically "uphill" task? Fortunately, chemical fuel is not the only means by which input energy can be supplied to intracellular molecular machines. A cell gets its energy from external sources. It has special machines to convert the input energy into some "energy currency". For example, chemical energy sup- plied by the food we consume is converted into an electro- chemical potential ∆µ that not only can be used to syn- thesize ATP, but can also directly run some other ma- chines. In plants similar proton-motive forces are gen- erated by machines which are driven by the input sun- light. Only hydrogen ion (H +), i.e., a proton, and sodium ion (N a+) are used in the kingdom of life to create the electro-chemical potentials, i.e., it used either a proton- motive force (PMF) or a sodium-motive force (SMF) as an energy currency. The advantages of using light, instead of chemical re- Motor superfamily Filamentous track Minimum step size Myosin Kinesin Dynein F-actin Microtubule Microtubule 36 nm 8 nm 8 nm TABLE I: Superfamilies of motor proteins and the corre- sponding tracks. action, as the input energy for a molecular motor are as follows: (i) light can be switched on and off easily and rapidly, (ii) usually, no waste product, which would re- quire disposal or recycling, is generated. Thus, study of molecular machines deals with two com- plementary aspects of bioenergetics: (a) conversion of energy input from the external sources into the energy currency of the cell, and (b) conversion of the energy currency to drive various active other active processes. B. Cytoskeletal motors The cytoskeleton of a cell is the analogue of the hu- man skeleton [3]. However, it not only provides mechan- ical strength to the cell, but its filamentous proteins also form the networks of "highways" (or, "tracks") on which cytoskeletal motor proteins [3, 4] can move. Filamen- tous actin (F-actin) and microtubules (MT) which serve as tracks are "polar" in the sense that the structure and kinetics of the two ends of each filament are dissimilar. The superfamilies of cytoskeletal motors and the cor- responding filamentous tracks are listed in table I. Every superfamily can be further divided into families. Mem- bers of every family move always in a particular direc- tion on its track; for example, kinesin-1 and cytoplasmic dynein move towards + and - end of MT, respectively. Similarly, myosin-V and myosin-VI move towards the + and - ends of F-actin, respectively. For their operation, each motor must have a track- binding site and another site that binds and hydrolyzes ATP (so-called ATPase site). Both these sites are lo- cated, for example, in the head domain of myosins and kinesins both of which walk on their heads! The motor- binding sites on the tracks are equispaced; the actual step size of a motor can be, in principle, an integral multiple of the minimum step size which is the separation between two neighboring motor-binding sites on the correspond- ing track. 1. Porters Some linear motors are cargo transporters [12]; how- ever, the size of the cargoes are usually much larger than 4 the motor itself! It is desirable that such a motor "walks" for a significant distance on its track carrying the cargo; for obvious reasons, such motors are referred to as porters [13]. Kinesin and dyneins attached simultaneously to the same "hard" cargo can get engaged in a tug-of-war lead- ing to a bidirectional movement of the cargo [14, 15]. In regions of overlap between MT and F-actin filaments, a large cargo may be hauled simultaneously by kinesins and myosins which, however, walk on their respective tracks. Alternatively, along its journey route, a cargo may be transferred from the MT-based transport net- work, which dominate at the cell center, to the F-actin based network that covers most of the cell periphery [16]. A "soft" cargo pulled by many kinesins can get elongated into a tube [17]. 2. Machines for chipping filamentous tracks A MT depolymerase is a kinesin motor that chips away its own track from one end [18]. Members of the kinesin- 13 family can reach either end of the MT diffusively (without ATP hydrolysis) and, then, start chipping the track from the end where it reaches. In contrast, mem- bers of the kinesin-8 family walk towards the plus end of the MT track hydrolyzing ATP and after reaching that end starts chipping it from there [19, 20]. Chipping by both families of depolymerase kinesins are energized by ATP hydrolysis. 3. Contractility: motor-filament crossbridge and collective dynamics of sliders and rowers Some motors are capable of sliding two different fila- ments with respect to each other by stepping simulate- nously on these two filaments [21]. Some sliders work in groups and each detaches from the filament after every single stroke; these are often referred to as rowers because of the analogy with rowing with oars [13]. Contractility, rather than motility, at the subcellular and cellular level are driven by the sliders and rowers [22, 23]. Some exam- ples of this category are listed in the table II; the details can be found in the cited review articles. Sliding filaments "Thin filaments" (F-actin) of muscle fibers "Stress fibers" (F-actin) of non-muscle cells Motor Myosin Myosin Myosin Cytokinetic "contractile ring" (F-actin) in eukaryotes Kinesin Dynein Dynein Interpolar microtubules in mitotic spindle Microtubules of megakaryocytes Microtubules of axoneme 5 Function (example) Muscle contraction [24] Cell contraction [25] Cell division [26] Mitosis [27, 28] Beating of eukaryotic flagella [29, 30] Blood platelet formation [31] TABLE II: Few example of cytoskeletal rowers and sliders as well as their biological functions. Polymer mode of force generation polymerizing pistion-like MT F-actin FtsZ MSP Type-IV pili MT spasmin Coiled actin polymerization polymerization polymerization polymerization de-polymerization spring-like spring-like Function (example) organizing cell interior [35] cell motility [36] bacterial cytokinesis [37] motility of nematode sperm cells [38] bacterial motility [39] Eukaryotic chromosome segregation [33] vorticellid spasmoneme [34] egg fertilization by sperm cell of the horse-shoe crab Limulus polyphemus [40] III: TABLE ing/depolymerizing, hooks and springs. Force generation by coiling/uncoiling filaments: polymeriz- pistons, 4. Push and pull of cytoskeletal filaments: nano-pistons and nano-springs Elongation of filamentous biopolymers that presses against a light object (e.g., a membrane) can result in a "push" [32]. Similarly, a depolymerizing tubular fila- ment can "pull" a light ring-like object by inserting its hook-like outwardly curled depolymerizing tip into the ring [33]. A flexible filament, upon compression by in- put energy, can store energy that can perform mechan- ical work when the filament springs back to its original relaxed shape [34]. Some typical examples are given in tableIII. C. Machines for synthesis, manipulation and degradation of macromolecules of life 1. Membrane-associated machines for macromolecule translocation: exporters, importers and packers In many situations, the motor remains immobile and pulls a macromolecule; the latter are often called translo- case. Some translocases export (or, import) either a pro- tein [41] or a nucleic acid strand [42, 43] across the plasma membrane of the cell or, in case of eukaryotes, across in- ternal membranes. A list is provided in table IV. The genome of many viruses are packaged into a pre- fabricated empty container, called viral capsid, by a pow- erful motor attached to the entrance of the capsid. As the capsid gets filled, The pressure inside the capsid in- creases which opposes further filling [44, 45]. The effec- tive force, which opposes packaging, gets contributions Membrane Nuclear envelope Polymer RNA/Protein [43, 47] Membrane of endoplasmic reticulum Protein [48] Membranes of mitochondria/chloroplasts Protein [49, 50] Membrane of peroxisome Protein [51] TABLE IV: Membrane-bound translocases. from three sources: (a) bending of stiff DNA molecule inside the capsid; (b) strong electrostatic repulsion be- tween the negatively charged strands of the DNA; (c) loss of entropy caused by the packaging [46]. The pack- aging motor has to be powerful enough to overcome such a high pressure. 2. Machines for degrading macromolecules of life Restriction-modification (RM) enzyme defend bacte- rial hosts against bacteriophage infection by cleaving the phage genome while the DNA of the host bacteria are not cleaved. Exosome and proteasome are machines that shred RNA and proteins into their basic subunits, namely, nucleotides and amino acids, respectively. Sim- ilarly, there are machines for degrading polysachharides, e.g., cellulosome (a cellulose degrading machine), starch degrading enzymes, chitin degrading enzyme (chitinase), etc. These machines are listed in table V. Polymer Examples of Machines DNA (polynucleotide) RNA (polynucleotide) Protein (polypeptide) Cellulose (polysachharide) RM enzyme [52] Exosome [53] Proteasome [54] Cellulosome [56] Starch (polysachharide) Starch degrad. enzyme [55] Chitin (polysachharide) Chitinase [57] TABLE V: Machines for degradation of macromolecules of life. a RNA secondary structure. During DNA replication, a helicase moves ahead of the polymerase, like a mine sweeper, unzipping the duplex DNA and dislodging other DNA-bound proteins. However, the transcriptional and translational machineries do not need assistance of any helicase because these are capable of unzipping DNA and unwinding RNA, respectively, on their own. A helicase can be monomeric, or dimeric or hexameric. 6 Function D. Rotary motors Machine Template Product DdDP DdRP RdDP RdRP DNA RNA DNA Reverse transcription [62] RNA Ribosome mRNA Protein DNA DNA RNA RNA DNA replication [58] Transcription [59 -- 61] RNA replication [63] Translation [64 -- 66] TABLE VI: Types of polymerizing machines, the templates they use and the corresponding product of polymerization. 3. Machines for template-dictated polymerization Two classes of biopolymers, namely, polynucleotides and polypeptides perform wide range of important func- tions in a living cell. DNA and RNA are examples of polynucleotides while proteins are polypeptides. Both polynucleotides and polypeptides are made from a lim- ited number of different species of monomeric building blocks, namely, nucleotides and amino acids,respectively. The sequence of the monomeric subunits to be used for synthesis of each of these are dictated by that of the corresponding template. These polymers are elongated, step-by-step, during their birth by successive addition of monomers, one at a time. The template itself also serves as the track for the polymerizer machine that takes chem- ical energy as input to polymerize the biopolymer as well as for its own forward movement. Therefore, these ma- chines are also referred to as motors. Depending on the nature of the template and prod- uct nucleic acid strands, polymerases can be classified as DNA-dependent DNA polymerase (DdDP), DNA- dependent RNA polymerase (DdRP), etc. as listed in the table VI. 4. Unwrappers and unzippers of packaged DNA: chromatin remodellers and Helicases In an eukaryotic cell DNA is packaged in a hierarchi- cal structure called chromatin. In order to use a sin- gle strand of the DNA as a template for transcription or replication, it has to be unpackaged either locally or globally. ATP-dependent chromatin remodelers [67] are motors that perform this unpackaging. However, only one of the strands of the unpackaged duplex DNA serves as a template; the duplex DNA is unzipped by a DNA helicase [68, 69]. Similarly, a RNA helicase unwinds Rotary molecular motors are, at least superficially, very similar to the motor of a hair dryer. Two rotary motors have been studied most extensively. (i) A rotary motor embedded in the membrane of bacteria drive the bacterial flagella [70, 71] which, the bacteria use for their swimming in aqueous media. (ii) A rotary motor, called ATP synthase [72, 73], is embedded on the membrane of mitochondria, the powerhouses of a cell. A synthase drives a chemical reaction, typically the synthesis of some product; the ATP synthase produces ATP, the "energy currency" of the cell, from ADP. E. Processivity, duty ratio, stall force and dwell time One can define processivity of a motor in three different ways: (i) Average number of chemical cycles in between attach- ment and the next detachment from the track; (ii) mean time of a single run, i.e., in between an attach- ment and the next detachment of the motor from the track; (iii) mean distance spanned by the motor on the track in a single run. Since the definitions (ii) and (iii) are directly accessible to experiments, these are more useful than the definition (i). Another related, but distinct, concept is that of duty ratio which is defined as the average fraction of the time that each head spends remaining attached to its track during one cycle. To translocate processively, a motor may utilize one of the three following strategies: Strategy I: the motor may have more than one track- binding domain (oligomeric structure can give rise to such a possibility quite naturally). Most of the cytoskele- tal motors like conventional two-headed kinesin use such a strategy [74]. One of the track-binding sites remains bound to the track while the other searches for its next binding site. Strategy II: A motor may possess non-motor extra do- mains or some accessory protein(s) bound to it which can interact with the track even when none of the motor domains of the motor is attached to the track. Dynein seems to exploit dynactin [75] to enhance its own proces- sivity. Strategy III: it can use a "clamp-like" device to remain attached to the track; opening of the clamp will be re- quired before the motor detaches from the track. For example, DdDP utilizes this strategy [76]. An external force directed opposite to the natural di- rection of walk of a motor is called a load force. Average velocity of a motor decreases with increasing load force. The magnitude of the load force at which the average velocity of the motor vanishes, is called the stall force. The force-velocity relation is one of the most fundamen- tal characteristic properties of a motor. Its status in bio- physics is comparable, for example, to that of the I-V characteristics of a device in semiconductor physics. Two motors with identical average velocities may ex- hibits widely different types of fluctuations. Therefore, as we'll show in detail, a deeper understanding of the op- erational mechanism of a motor can be gained from the distributions of their "dwells" at the successive spatial positions on the track. F. Fundamental general questions (i) Single-molecule mechanism [77]: How do the in- teractions among the component structural units of an individual motor, motor-track interactions, motor-ligand (fuel) interactions and the mechano-chemical kinetics of the system determine, for example, (a) the directional- ity, (b) processivity, (c) dwell-time distribution, (d) force- velocity relation, and (e) efficiency of transduction? Does a given dimeric motor walk hand-over-hand or crawl like an inchworm, and why? Do the ATPase domains of a hexameric motor "fire" (a) in series, or (b) in parallel, or, (c) in random sequence? (ii) Multi-motor coordination in a "workshop": The motors do not work in isolation in-vivo. What are the mechanisms and consequences of the spatio-temporal co- ordination of the motors? For example, a single mitotic spindle consists of many polymerizing and depolymeriz- ing MTs, several types of cytoskeletal motors, including depolymerases [28]. A replisome, the workshop for DNA replication, has to coordinate the operations of clamps and clamp loaders with those of primases, polymerases, etc. [58]. Similarly, a ribosome is a mobile platform on which the operations of several devices have to be coordi- nated properly during protein synthesis [64]. Well known co-directional as well as head-on collisions of polymerases and those of ribosomes can create traffic jam under some circumstances whereas a collision can restart stalled traf- fic in other circumstances [78]. III. MOTORING IN A VISCOUS FLUID: FROM NEWTON TO LANGEVIN Force is one of the most fundamental quantities in physics. As we'll argue in this section, some of the forces 7 which dominate the dynamics of molecular machines have negligible effect on macroscopic machines. A. Newton's equation: deterministic dynamics For simplicity, let us first consider a hypothetical sce- nario where neither the tracks nor the fuel molecules are present in the aqueous medium in which there is a free (i.e., not bound to any other molecule) motor protein. Suppose the medium consists of only N "particle-like" molecules and the center of mass of the motor protein is also represented by a "particle". Then, the exact tra- jectories of all these N + 1 "particles" can be obtained by solving the corresponding coupled Newton's equations that describe the dynamics of the N + 1 particles. In reality, this approach is impractical because, even with the largest and fastest computers available at present, we cannot get the trajectories over time intervals of the order of 1s which is relevant for majority of the motor proteins as long as N is large (N ∼ 1023). B. Langevin equation: stochastic dynamics A more pragmatic approach would be to solve only the equations of motion for the motor protein by treating the N particles of the medium as merely the constituents of a reservoir. In other words, one monitors the motion in a 6-dimensional subspace of the full 6(N + 1)-dimensional phase space of the system. The price one has to pay for this simplification is that instead of the original Newton's equation for the motor protein, one has to now solve a Langevin equation that describes the "Brownian" motion of the motor protein. Since we are soon going to extend the discussion in the presence of filamentous tracks which are essentially one-dimensional, we write the Langevin equation for the "Brownian" motion of the motor protein in one-dimension m(d2x/dt2) = Fd + ξ where m is the mass of the protein, Fd = −γv is the viscous drag and and ξ is the random Brownian force. Note that the Langevin equation is neither deterministic nor symmetric under time-reversal (t → −t). 1. Motoring in a "sticky" medium: Purcell's idea Using the Stokes formula γ = 6πηr for a globular pro- tein of radius r ∼ 10nm moving with an average ve- locity v ∼ 1m/sec (corresponding to 1nm/ns) in the aqueous medium of viscosity η ∼ 10−3P as we get an estimate Fd ∼ 200pN . Surprisingly, this viscous drag force is about 200 times larger than the elastic force it experiences when stretched by 1nm ! Consequently, the Reynold's number, which is the ratio of the inertial and viscous forces, is at most of the order O(10−2). The Reynold's number will be of the order of 10−2 if you ever try (not recommended) to swim in honey! Now let us supply the tracks and fuel molecules to the motor protein in the same medium. The Langevin equa- tion will now include additional terms Fext which rep- resents the net force arising from the interaction of the motor protein with its track and the ligand; the ligand could be the fuel molecule or the molecules produced by its "burning" (e.g., ATP or ADP). Besides, since the dy- namics of motor proteins is expected to be dominated by hydrodynamics at low Reynold's number [79], one gener- ally drops the inertial term. In this "overdamped" regime γv = Fext + ξ (1) 2. Motoring under random impacts from surroundings: Brownian force The energy released by the hydrolysis of a single ATP molecule is about 10−21J. Interestingly, the mean ther- mal energy kBT associated with a molecule at a temper- ature of the order of T ∼ 100K, is also kBT ∼ 10−21J. Moreover, equating this thermal energy with the work done by the thermal force Ft in causing a displacement of 1nm we get Ft ∼ 1pN . This is comparable to the elas- tic force experienced by a typical motor protein when stretched by 1nm. Thus, a motor protein that gets bom- barded from all sides by random thermal forces is similar to a tiny creature is a very strong hurricane! Therefore, in contrast to the deterministic dynamics of the macro- scopic machines, the dynamics of molecular motors is stochastic (i.e., probabilistic). Already in the first half of the twentieth century D'Arcy Thompson, father of modern bio-mechanics, real- ized the importance of viscous drag and Brownian forces at the cellular (and subcellular) level. He pointed out that [80] that in this microscopic world "gravitation is forgotten, and the viscosity of the liquid,..., the molecu- lar shocks of the Brownian movement, .... the electric charges of the ionized medium, make up the physical en- vironment ... predominant factors are no longer those of our scale". IV. ENERGY TRANSDUCTION BY MOLECULAR MACHINES: FROM CARNOT TO FEYNMAN Molecular motors generate force by transducing en- ergy. Interestingly, one of the two mechanisms of en- ergy transduction that I describe here, does not have any analogous macroscopic counterpart. I explain in con- siderable detail why thermodynamic formalisms, which have been successfully utilized to calculate the common performance characteristics of macroscopic machines, are inadequate for natural nano-machines. Molecular motors are made of soft matter whereas macroscopic motors are normally made of hard matter to withstand wear and tear. Nature seems to exploit the 8 high deformability of molecular motors for its biological function. But, this difference is minor compared to the others and will not be elaborated further here. A. Molecular machines are run by isothermal engines This is in sharp contrast to the macroscopic thermal engines which require at least two thermal reservoirs at different temperatures and which convert part of the in- put heat energy into mechanical work. Why can't a molecular machine work as a heat en- gine? In order to examine the viability of a intracellular heat engine, let us create a temperature gradient on a length scale ℓ ∼ 10nm. An elementary analysis of heat diffusion equation is adequate to show that a tempera- ture gradient on a length scale ℓ relaxes within a time interval τtemp ∼ chℓ2/κ where ch is the specific heat per unit volume and κ is the thermal conductivity. Using the typical characteristic values of ch and κ for water, one finds [81] τtemp ≃ 10−6 − 10−8s. Thus, temperature gradients cannot be maintained for the entire duration of even one single cycle of a cyclic molecular machine which is, typically, few orders of magnitude longer than τtemp. In other words, for all practical purposes, natural nano- machines operate isothermally. Because of the condition T = constant, the molecular machines operate as free energy transducers. B. Inadequacy of equilibrium thermodynamics A majority of the molecular motors are chemo- mechanical machines for which input and output are chemical energy and mechanical work, respectively. Re- call that for Carnot's thermo-mechanical machine, the thermodynamic efficiency is ηeq−th = 1 − (TL/TH) where TH and TL are the temperatures of the two reservoirs at high and low temperatures, respectively. Similarly, for an isothermal chemo-mechanical engine, the thermal reservoirs would be replaced by chemical reservoirs at fixed chemical potentials µH and µL (µH > µL). Con- sequently, the heat flow of the thermal engine would be replaced by particle flow in the chemical engine. Just as the difference of heat input and output in the heat engine is converted to mechanical work, the difference µH − µL would be converted into output work in the chemical engine. Obviously, the corresponding thermo- dynamic efficiency would be [82] ηeq−ch = 1 − (µL/µH ). Note that, because of the quasi-static nature of each step in the equilibrium thermodynamic theory of these cyclic engines, each cycle takes infinite time. Naturally, the power output Pout = 0 for both the thermal and chemi- cal Carnot engines. Can we apply the theory of such isothermal Carnot en- gines to a chemo-mechanical molecular machine by iden- tifying, for example, µH and µL (µH > µL) as the chem- ical potentials of ATP and ADP, respectively? The an- swer is: NO. The cycle time of the real molecular ma- chines is finite and their power output is non-zero. More- over, motors are examples of open systems [83] which continue to run in repetitive cycles as long as energy is pumped in. Such a system cannot be in thermodynamics equilibrium. C. Inadequacy of endo-reversible thermodynamics For macroscopic engines with finite cycle time, the characteristics of performance are often reliably calcu- lated within the theoretical framework of endo-reversible thermodynamics [84]. In this case, the ratio of the rates of output and input energies defines the efficiency of transduction. The rate of output work is called power output of the engine. The formalism of endo-reversible thermodyamics for heat engines is based on the assumption that the work- ing substance of the engine is coupled to the two thermal reservoirs by heat conductors of finite thermal conduc- tivity (non-zero thermal resistance). Entire dissipation is assumed to take place in the irreversible process of heat conduction through these thermal resistors whereas all the processes in the working material of the engine are assumed to be fully reversible (i.e., no entropy is gen- erated internally). In this scenario, the corresponding thermal conductances in the Carnot engine are infinite. Efficiency at maximum power output η(Pmax), rather than maximum efficiency itself, is the most appropriate quantitative measure of the performance of such engines. The upper-bound of η(Pmax) is given by the Curzon- Ahlborn expression [85] ηCA(Pmax) = 1 − (TL/TH)1/2. Similarly, within the framework of the endoreversible thermodynamics, a phenomenological theory for chem- ical engines can also be formulated if the cycle time is finite [86]. Can't we use this approach for molecular ma- chines which are also chemical engines with finite cycle times? Recall that endo-reversible thermodynamics is based on the assumption that dissipation takes place only in the process of transfer of matter between reservoirs and the working substance whereas processes that the work- ing substance goes through in each cycle are perfectly reversible (and, therefore, non-dissipative). This assump- tion is valid if the relaxations in the working substance of the engine are very rapid compared to the processes involving the coupling of the working substance with the reservoirs. The validity of this assumption requires clear separation of time scales of relaxation in the working substance and in the working substance-reservoir cou- pling. But, such separation of time scales does not hold in molecular machines which consist of macromolecules. 9 D. Domain of non-equilibrium statistical mechanics In general, molecular motors operate far beyond the linear response regime and, therefore, the formalism of non-equilibrium thermodynamics [87] for coupled mechano-chemical processes is not applicable. More- over, thermodynamic formalisms developed for macro- scopically large systems do not account for the sponta- neous fluctuations. On the other hand, because of the small size of a molecular motor and because of the low concentrations of the molecules involved in its operation, fluctuations of positions, conformations as well as the cycle times are intrinsic features of their kinetics. There- fore, one has to use the more sophisticated toolbox of stochastic processes and non-equilibrium statistical me- chanics for theoretical treatment of molecular machines. Interestingly, as we argue below, noise need not be a nui- sance for a motor; instead, a motor can move forward by gainfully exploiting this noise! E. Defining efficiency: from Carnot to Stokes The performance of macroscopic motors are charac- terized by a combination of its efficiency, power output, maximum force or torque that it can generate. Just like the performance of their macroscopic counterparts with finite cycle time, that of molecular motors [88] have also been characterized in terms of efficiency at maximum power, rather than maximum efficiency. However, the efficiency of molecular motors can be defined in several different ways [89]. The efficiency of a motor, with finite cycle time, is generally defined by η = Pout/Pin (2) where Pin and Pout are the input and output powers, respectively. The usual definition of thermodynamic effi- ciency ηT is based on the assumption that, like its macro- scopic counterpart, a molecular motor has an output power [90] Pout = −FextV. (3) where Fext is the externally applied opposing (load) force. Although this definition is unambiguous, it is unsatisfac- tory for practical use in characterizing the performance of molecular motors. As explained earlier, a molecular motor has to work against the omnipresent viscous drag in the intracelluar medium even when no other force op- poses its movement (i.e., even if Fext = 0). A generalized efficiency ηG is also represented by the same expression (2) where, instead of (3), the output power is assumed to be [91] Pout = FextV + γV 2. (4) This definition treats the load force and viscous drag on equal footing. In contrast, the "Stokes efficiency" ηS for a molecular motor driven by a chemical reaction is defined as [92] ηS = γV 2 (∆G)hri + FextV (5) where hri is the average rate of the chemical reaction and ∆G is the chemical free energy consumed in each reaction cycle. This efficiency is named after Stokes because the viscous drag is calculated from Stokes law. As we show in the next subsection, the directional movement of some motors arises from the rectification of random thermal noise. For such motors, an alternative measure of performance is the rectification efficiency [93]. Thus, different definitions of the efficiency of a molecular motor may arise from different interpretations of their tasks or may characterize distinct aspects of their oper- ation. Nevertheless, for any definition of efficiency, it is essential to ensure that its allowed values lie between 0 and 1. F. Noisy power stroke and Brownian ratchet If the input energy directly causes a conformational change of the protein machinery which manifests itself as a mechanical stroke of the motor, the operation of the motor is said to be driven by a "power stroke" mecha- nism. This is also the mechanism used by all man made macroscopic machines. However, in case of molecular motors, the power stroke is always "noisy" because of the Brownian forces acting on it. Let us contrast this with the following alternative sce- nario: suppose, the machine exhibits both "forward" and "backward" movements because of spontaneous thermal fluctuations. If now energy input is utilized to prevent "backward" movements, but allow the "forward" move- ments, the system will exhibit directed, albeit noisy, movement in the "forward" direction. Note that the for- ward movements in this case are caused directly by the spontaneous thermal fluctuations, the input energy rec- tifies the "backward" movements. This alternative sce- nario is called the Brownian ratchet mechanism [94, 95]. The concept of Brownian ratchet [96] was popularized by Richard Feynman with his ratchet-and-pawl device [97]. However, in the context of molecular motors, the Brownin ratchet mechanism was proposed by several groups in the 1990s [94, 98]. Thus, in principle, there are two idealized scenarios for a transition from a conformation A to a conformation B- one is by a pure power stroke and the other by a purely Brownian ratchet mechanism [99]. However, for a real molecular motor, it is difficult to unambiguously distin- guish between a power stroke and a Brownian ratchet [100]; the actual mechanism may be a combination of the two idealized extremes. •Are Brownian motors Maxwell's demon? Brownian ratchet mechanism transduces random ther- mal energy of the reservoir into mechanical work. Does 10 it imply that these motors are perpetual motors of the second kind that violate the second law of thermody- namics? In that case, does it have any similarity with the Maxwell's demon [101]? A short answer is: No, a Brownian ratchet is not a Maxwell's demon because it is an open system far from the state of equilibrium whereas the second law of thermodynamics is strictly valid only for systems in stable thermodynamic equilibrium. G. Physical realizations of Brownian ratchet in molecular motors Is there any real biomolecular motor which can be re- garded as a true physical realization of Brownian ratch- ets? The answer is an emphatic "yes" and we list a few typical examples here. KIF1A, a single-headed kinesin, is an example of cy- toskeletal motor whose operational mechanism can be in- terpreted as a Brownian ratchet [102 -- 104]. The original acto-myosin crossbridge model suggested by Huxley [105] can be interpreted as Brownian ratchet [106] although Huxley himself did not use this terminology. Brownian ratchet can also account for the force generation by a polymerizing cytoskeletal filament [107]. The crossing of a membrane during the export/import of a protein of length L (amino acid monomers) takes place at a faster rate by the Brownian ratchet mechanism compared to that by pure translational diffusion [108]. It has been claimed that the translocation of a mRNA molecule across the nuclear membrane of an eukaryotic cell takes place also by a similar Brownian ratchet mech- anism through the nuclear pore complex [109]. The elon- gation of a mRNA during transcription by a RNA poly- merase can also be a physical realization of the Brownian ratchet mechanism [110, 111]. Translocation, one of the crucial steps of the mechano-chemical cycle of a ribosome is a physical realization of Brownian ratchet [65]. V. MATHEMATICAL DESCRIPTION OF MECHANO-CHEMICAL KINETICS: CONTINUOUS LANDSCAPES VS. DISCRETE NETWORKS We combine the fundamental principles of (stochas- tic) nano-mechanics and (stochastic) chemical kinetics to formulate the general theoretical framework for a quan- titative description of the mechano-chemistry or chemo- mechanics of molecular motors. We mention a few alter- native formalisms. A. Motor kinetics as wandering in a time-independent mechano-chemical free-energy landscape This approach is useful for an intuitive physical ex- planation of the coupled mechano-chemical kinetics of molecular motors. At least one of the independent coor- dinate axes of this landscape represents the position of the motor in real space (i.e., its "mechanical coordinate") while at least another represents its chemical state (.e., "chemical coordinate"). Therefore, the minimum dimen- sion of this "land" is 2 and the free energy can be rep- resented by the hight at each point on the "land". Both the position and chemical state variables are assumed to be continuous. The profiles of the free energy along both the position and chemical coordinates, obtained by taking appropriate cross sections of the free energy land- scape, are periodic with the same periodicity in both the directions. But, the profile is tiled forward along the chemical direction so that the bottom of the minima are located deeper in the forward direction along the "chem- ical" coordinate; this tilt accounts for the lowering of free energy caused, for example, by ATP hydrolysis. B. Motor kinetics as wandering in the time-dependent real-space potential landscape Consider those special situations where the chemical states of the motor are long lived and change in fast dis- crete jumps so that the position of the motor can con- tinue to change without alteration in its chemical state, except during chemical transitions when position remains frozen. Thus, no mixed mechano-chemical transition is allowed in this scenario. In such situations, we can as- sume that the potential landscape in real space remains unchanged for a while, during which the wanderings of the motor in this landscape (i.e., the positional dynamics of the motor) is governed by the "frozen" spatial profile of the potential. The profile of the potential switches se- quentially from one form to another and the sequence of m profiles is repeated in each cycle although the switch- ing times are random. Different profiles correspond to different motor-track interactions which is dependent on the nature of the ligand (if any) bound to the motor. In this formulation we assume that the allowed po- sitions of the motor on its track form a continuum x. The discrete index µ = 1, 2, ..., M denote the M dis- tinct spatial profiles of the potential Vµ(x) that are pos- tulated apriori. At any instant of time t, the state of the motor is given by (x, µ). In the overdamped regime, the time-evolution of the position x(t) of the mo- tor obeys the Langevin equation (1) with [112] Fext = −dVµ(x)/dx + Fload. The time-dependence of the profile 11 of the potential Vµ(x) is governed by the master equation ∂Pµ(x, t) ∂t =Xµ′ Pµ′ (x, t)Wµ′→µ(x) −Xµ′ Pµ(x, t)Wµ→µ′ (x) (6) where Pµ(x, t) is the probability that the motor protein "sees" the landscape Vµ(x) at time t and Wµ→µ′ (x) is the transition probability per unit time for a transition from the chemical state µ to the chemical state µ′ at position x. Instead of the Langevin equation, an equivalent Fokker-Planck equation can be used to describe the wan- dering of the motor protein in the potential energy land- scape Vµ(x). The advantage of this alternative formu- lation is that the Fokker-Planck equation can be com- bined naturally with the master equation that describes the time-evolution of the potential energy landscape. In this hybrid formulation, Pµ(x, t) represents the probabil- ity that, at time t, the motor is located at x, while "see- ing" the potential energy landscape Vµ(x). The equation governing the time dependence of Pµ(x, t) ∂Pµ(x, t) ∂t = 1 η ∂ µ(x) − F }Pµ(x, t)(cid:21) ∂x(cid:20){V ′ η (cid:19) ∂2Pµ(x, t) ∂x2 Pµ′ (x, t)Wµ′→µ(x) + (cid:18) kBT + Xµ′ − Xµ′ Pµ(x, t)Wµ→µ′ (x) (7) where η is a phenomenological coefficient that captures the effect of viscous drag. Note that there is no term in this equation which would correspond to a mixed mechano-chemical transition. C. Motor kinetics as a jump process in a fully discrete mechano-chemical network In this formulation, both the positions and "internal" (or "chemical") states of the motors are assumed to be discrete. Let Pµ(i, t) be the probability of finding the motor at the discrete position labelled by i and in the "chemical" state µ at time t. Then, the master equation for Pµ(i, t) is given by ∂Pµ(i, t) ∂t = [Xj6=i + [Xµ′ + [Xj6=iXµ′ − Xj6=iXµ′ Pµ(j, t)kµ(j → i) −Xj6=i Pµ′ (i, t)Wµ′→µ(i) −Xµ′ Pµ′ (j, t)ωµ′→µ(j → i) Pµ(i, t)kµ(i → j)] Pµ(i, t)Wµ→µ′ (i)] Pµ(i, t)ωµ→µ′ (i → j)] (8) where the terms enclosed by the three different brackets [.] correspond to the purely mechanical, purely chemical and mechano-chemical transitions, respectively. As a concrete example, which will be used also for on several other occasions later in this review, consider a 2- state motor that, at any site j, can exist in one of the only two possible chemical states labelled by the symbols 1j and 2j. We assume the mechano-chemical cycle of this motor to be 1j ω1 ⇋ ω−1 2j ω2→ 1j+1 (9) where the rates of the allowed transitions are shown ex- plicitly above or below the corresponding arrow. Note that the transition 1j ⇋ 2j is purely chemical whereas the transition 2j → 1j+1 is a mixed mechano-chemical transition. The corresponding master equations are given by dP1(i) dt = ω2P2(i − 1) + ω−1P2(i) − ω1P1(i) dP2(i) dt = ω1P1(i) − ω−1P2(i) − ω2P2(i) (10) Suppose a load force Fext opposes the meacho-chemical transition 2j → 1j+1. The effect of the load force can be incorporated by assuming the load-dependence of the corresponding rate constant to be of the form ω2(Fext) = ω2(0)exp[−Fext/(kBT )] where ω2(0) is the value of the rate constant in the absence of any external force. We'll see some implications of these equations in several specific contexts later in this article. D. Balance conditions for mechano-chemical kinetics: cycles, and flux On a discrete mechano-chemical state space, each state is denoted by a vertex and the direct transition from one state (denoted by, say, the vertex i) to another (denoted by, say, the vertex j) is represented by a directed edge ij >. The opposite transition from j to i is denoted by the directed edge ji >. A transition flux can be defined along any edge of this diagram. The forward transition flux Jij> from the vertex i to the vertex j is given by WjiPi while the reverse flux, i.e., transition flux Jji> from j to i is given by Wij Pj. Therefore, the net transition flux in the direction from the vertex i to the vertex j is given by Jji = WjiPi − Wij Pj. A cycle in the kinetic diagram consists of at least three vertices. Each cycle Cµ can be decomposed into two di- rected cycles (or, dicycles) [113] Cd µ where d = ± corre- sponds the clockwise and counter-clockwise cycles. The net cycle flux J(Cµ) in the cycle Cµ, in the CW direction, is given by J(Cµ) = J(C+ µ ) − J(C− µ ). For each arbitrary dicycle Cd µ, let us define the dicycle ratio R(Cd µ) = Π<ij>ǫC +d µ 12 Wji/Π<ij>ǫC−d µ Wij = Πµ,d <ij>(Wji/Wij ). (11) where the superscript µ, d on the product sign denote a product evaluated over the directed edges of the cycle. So, by definition, R(C− µ ) = 1/R(C+ µ ). It has been proved rigorously [114] that, for detailed balance to hold, the necessary and sufficient condition to be satisfied by the transition probabilities is R(Cd µ) = 1 for all dicycles Cd µ (12) For a non-equilibrium steady state (NESS), one can µ) which is the µ completed per unit time in the define [113] the dicycle frequency Ωss(Cd number of dicycles Cd NESS of the system. Then, Ωss(C+ µ )/Ωss(C− µ ) = Πµ,d <ij>(Wji/Wij ) = R(Cd µ) (13) Clearly, R(Cd µ) 6= 1, in general, for any NESS. Detailed balance is believed to be a property of systems in equilibrium whereas the conditions under which molec- ular machines operate are far from equilibrium. Does it imply that detailed balance breaks down for molecular machines? The correct answer this subtle question needs a careful analysis [115 -- 117]. If one naively assumes that the entire system returns to its original initial state at the end of each cycle one would get the erroneous result that the detailed balance breaks down. But, strictly speaking, the free energy of the full system gets lowered by δG (e.g., because of the hydrol- ysis of ATP) in each cycle although the cyclic machine itself comes back to the same state. When the latter fact is incorporated correctly in the analysis [115 -- 117], one finds that detailed balance is not violated by molecular machines. This should not sound surprising- the transi- tion rates "do not know" whether or not the system has been driven out of equilibrium by pumping energy into it. VI. SOLVING FORWARD PROBLEM BY STOCHASTIC PROCESS MODELING: FROM MODEL TO DATA A. Average speed and load-velocity relation For simplicity, let us consider the kinetic scheme shown in fig. (1(a)). In terms of the Fourier transform ¯Pµ(k, t) = ∞ Xj=−∞ Pµ(xj , t)e−ikxj (14) of Pµ(xj , t), the master equations can be written as a matrix equation ∂ ¯P(k, t) ∂t = W(k) ¯P(k, t) (15) (a) (b) (c) FIG. 1: Three examples of different types of networks of dis- crete mechano-chemical states. The bullets represent the dis- tinct states and the arrows denote the allowed transitions be- tween the two corresponding states. The scheme is (a) is unbrached whereas that in (b) has branched pathways con- necting the same pair of states. The mechanical step size is unique in both (a) and (b) whereas steps of more than one size are possible in (c). (Adapted from fig.1 of ref.[120]). Taking derivatives of both sides of (16) with respect to q we get [118, 119] 13 i(cid:18) ∂ ¯P (k, t) ∂k (cid:19)k=0 ∂k(cid:19)2 +(cid:18) ∂P k=0 −(cid:18) ∂2 ¯P ∂k2(cid:19)k=0 = < x(t) > = < x2(t) > − < x(t) >2 (17) Evaluating ¯P (k, t), in principle, the stationary drift ve- locity (i.e., asymptotic mean velocity) V and the corre- sponsding diffusion constant D can be obtained from V = limt→∞i D = limt→∞ ∂ ∂t(cid:20)(cid:18) ∂ ¯P (k, t) ∂t(cid:20)(cid:18)− ∂k (cid:19)k=0(cid:21) ∂k2 (cid:19)k=0 ∂2 ¯P (k, t) ∂ 1 2 ∂k (cid:19)2 +(cid:18) ∂ ¯P (k, t) k=0(cid:21) (18) It may be tempting to attempt a direct utilization of the general form ¯P (k, t) =Xµ Bµeλµ(k)t (19) where the coefficients Bµ are fixed by the initial condi- tions and λµ(k) are the eigenvalues of W(k). However, for the practical implementation of this method analyti- cally the main hurdle would be to get all the eigenvalues of W(k). Fortunately, only the smallest eigenvalue λmin, which dominates ¯P (k) in the limit t → ∞, is required for evaluating V and D [118, 119]: ∂k V = i(cid:18) ∂λmin(k, t) (cid:19)k=0 2(cid:18) ∂2λmin(k, t) D = − ∂k2 1 (cid:19)k=0 (20) Even the forms (20) are not convenient for evaluating V and D. Most convenient approach is based on the characteristic polynomial Q(k) associated with the ma- trix W(k), i.e., Q(k; λ) = det[λI − W(k)]. Therefore, λmin(k) is a root of the polynomial Q(k; λ), i.e., solution of the equation where ¯P is a column vector of M components (µ = 1, ...M ) and W(k) is the transition matrix in the k-space (i.e., Fourier space). Summing over the hidden chemical states, we get Q(k; λ) = M Xµ=0 Hence [118, 119] Cµ(k)[λ(k)]µ = 0. (21) ¯P (k, t) = M Xµ=1 ¯Pµ(k, t) = M ∞ Xµ=1 Xj=−∞ Pµ(xj , t)e−ikxj . (16) C′ 0 C1(0) 0 − 2iC′ C′′ V = −i D = 1(0)V − 2C2(0)V 2 2C1(0) (22) where C′ µ = [∂Cµ(k)/∂k]k=0. For a postulated kinetic scheme, W is given. Then the expressions (22) are adequate for analytical deriva- tion of V and D for the given model. However, in order to calculate the distributions of the dwell times of the motors, it is more convenient to work with the Fourier- Laplace transform, rather than the Fourier transform, of the probability densities. Therefore, we now derive al- ternative expressions for V and D in terms of the full Fourier-Laplace transform of the probability density. Taking Laplace transform of (14) with respect to time Pµ(k, s) =Z ∞ 0 ¯Pµ(k, t)e−st, (23) the matrix form of the master equation reduces to P(k, s) = R(k, s)−1P(0) where R(k, s) = sI − W(k) and P(0) is the column vector of initial probabilities. Now the characteristic polynomial is the determinant of R(k, s) which can be expressed as [120] wj. The average velocity V of the motor is given by [5] 14 V = 1 RM(cid:20)1 − M−1 Yj=0(cid:18) wj uj(cid:19)(cid:21) RM = M−1 Xj=0 rj where with rj =(cid:18) 1 uj(cid:19)(cid:20)1 + M−1 Xk=1 k Yi=1(cid:18) wj+i uj+i(cid:19)(cid:21) while D is given by D =(cid:20) (V SM + dUM ) R2 M − (M + 2)V 2 (cid:21) d M (26) (27) (28) (29) R(k, s) = M Xµ=0 cµ(k)sµ = 0. (24) Equation (24) is formally similar to (21). As expected, we get [120] where and V = −i c′ 0 c1(0) 0 − 2ic′ c′′ D = 1(0)V − 2c2(0)V 2 2c1(0) (25) while, M−1 M−1 SM = sj Xj=0 Xk=0 (k + 1)rk+j+1 (30) UM = M−1 Xj=0 ujrjsj (31) •Example: chemical cycle an M-step unbranched mechano- As an illustrative example, let us consider the un- branched mechano-chemical cycle [5] with M = 4, as shown in fig.2. This special value of M is motivated by the typical example of a kinesin motor for which the four essential steps in each cycle are as follows: (i) a substrate- binding step (e.g., binding of an ATP molecule), (ii) a chemical reaction step (e.g., hydrolysis of ATP), (iii) a product-release step (e.g., release of ADP), and (iv) a mechanical step (e.g., power stroke). sj = 1 uj(cid:18)1 + M−1 k Xk=1 Yi=1 wj+1−i uj−i (cid:19). (32) For various extensions of this scheme see ref.[5]. In the simpler case shown in (9), where M = 2, and the second step is purely irreversible, using the step size ℓ explicitly (to make the dimensions of the expressions explicitly clear), we get V = ℓ(cid:20) ℓ2 D = ω1ω2 ω1 + ω−1 + ω2(cid:21) 2(cid:20) (ω1ω2) − 2(V /ℓ)2 ω1 + ω−1 + ω2 (cid:21) (33) FIG. 2: An unbranched mechano-chemical cycle of the molec- ular motor with M = 4. The horizontal dashed line shows the lattice which represents the track; j and j + 1 represent two successive binding sites of the motor. The circles labelled by integers denote different "chemical" states in between j and j + 1. (Adapted from fig.7 of ref.[121]). Suppose, The forward transitions take place at rates uj whereas the backward transitions occur with the rates Note that if, in addition, ω−1 vanishes, i.e., if both the steps are fully irreversible, then d/V = ω−1 2 , i.e., the average time taken to move forward by one site is the sum of the mean residence time in the two steps of the cycle. •Effects of branched pathways and inhomo- geneities of tracks 1 + ω−1 Microtubule-associated proteins and actin-related pro- teins can give rise to inhomogeneities of the tracks for cytoskeletal motors. The intrinsic sequence inhomogene- ity of DNA and RNA strands can nontrivial influences on the motors that use nucleic acid strands as their tracks [122]. B. Jamming on a crowded track: flux-density Let is consider the kinetic scheme shown below relation Mj + Pn + S ⇋ I1 → I2 ↓ → Mj+1 + Pn+1 15 (34) In the preceeding subsection we considered lone motor moving along a filamentous track. Now we consider a track with heavy motor traffic. In principle, unless con- trolled by some regulatory signals, formation of "traffic jams" [123 -- 125] on crowded tracks cannot be ruled out [126 -- 129]. These phenomena are formally similar to sys- tems of sterically interacting self-propelled particles. One of the simplest theoretical models for such systems is the totally asymmetric simple exclusion process (TASEP). In fact, TASEP was originally inspired by traffic of ri- bosomes on a mRNA track during translation [130 -- 133]. However, in recent years it has found application also in the context of traffic-like collective movements of ribo- somes in more realistic models [134, 135] as well as many other types of molecular motors, e.g., kinesins [103, 104], RNA polymerase [136, 137], growth of fungal hyphae [138], growth of bacterial flagella [139], crowding of de- polymerizing kinesins at the tip of a microtube [140], etc. C. Information processing machines: fidelity versus power and efficiency Power and efficiency are most appropriate quantities to characterize the performance of porters, rowers, sliders, etc. But, not all machines fall in these categories. For information processing machines, which we consider in this subsection, fidelity of information transfer is more important than power and efficiency. The sequence of the monomeric subunits of a polynu- cleotide or that of a polypeptide is dictated by that of the corresponding template strand. During the polymer- ization process, after preliminary tentative selection, the selected monomeric subunit is subjected to several levels of checks by the quality control system of the polymeriz- ing machinery. Only after a selected subunit is screened by such a stringent test, it is covalently bonded to the elongating biopolymer. Discrimination of monomers based merely on the dif- ferences of free-energy gains cannot account for the ob- served high fidelity of these processes. For example, "ki- netic proofreading" [141, 142] has been invoked to explain the kinetic processes that contribute to the high level of translational fidelity. Following three features are essen- tial for kinetic proofreading [143]: (i) formation of an initial enzyme-reactant complex, (ii) a strongly forward driven step that results in a high- energy intermediate complex, and (iii) one or more branched pathways along which dissocia- tion of the enzyme-reactant complex, and rejection of the reactant, is possible before the complex gets an opportu- nity to make the final transition to yield the product. where Mj denotes the polymerase motor located at the discrete position j on its template, Pn represents the elongating biopolymer consisting of n subunits and S is a single subunit. Note that the scheme shown in (34) is, at least, formally an extension of (9) where an extra intermediate state I2 and a branched path from I2 have been added. This scheme is one of the simplest possible implementations of kinetic proofreading. Kinetic proofreading leads to futile cycles in which at least part of the input free energy is dissipated with- out elongating the nascent polypeptide by an amino acid monomer. The effects of the consequent loose mechano- chemical coupling [144] of the translational machinery on the rate of polypeptide synthesis has been investigated [135, 145, 146]. Occasionally wrong monomers escape detection in spite of elaborate selection procedure. In case of DNA replication, the polymerase normally detects the error immediately and corrects it before elongating it further. For this purpose, the nascent DNA is transferred to an- other specific site on the same polymerase where the wrong monomer is cleaved before transferring the nascent DNA back to the site of elongation activity. The inter- play of the two contradictory activities of elongation and shortening of a nascent DNA exposes leads to the cou- pling of their corresponding rates in the overall rate of DNA polymerization by a DdDP [147]. D. Beyond average: dwell time distribution (DTD) Two motors with identical average velocities may ex- hibits widely different types of fluctuations. Suppose the successive mechanical steps are taken by the motor at times t1, t2, ..., tn−1, tn, tn+1, .... Then, the time of dwell before the k-th step is defined by τk = tk − tk−1. In between successive steps, the motor may visit several "chemical" states and each state may be visited more than once. But, the purely chemical transitions would not be visible in a mechanical experimental set up that records only its position. The number of visits to a given state and the duration of stay in a state in a given visit are random quantities. In order to appreciate the origin of the fluctions in the dwell times, let us consider the simple N -step kinetics: M1 ⇋ M2 ⇋ M3... ⇋ Mj ⇋ ... ⇋ MN (35) Suppose tµ,ν is the duration of stay of the motor in the µ-th state during its ν-th visit to this state. If τ is the dwell time, then τ = N nµ Xµ=1 Xν=1 tµ,ν (36) where nµ is the number of visits to the µ-th state. It is straightforward to check that < τ > = N Xµ=1 < nµ >< tµ > (37) where < nµ > is the avarege number of visitits to the µ-th state and < tµ > is the avarege time of dwell in the µ-th state is a single visit to it. More interestingly [148], N < τ 2 > − < τ >2= N (< t2 µ > − < tµ >2) < nµ > Xµ=1 (< n2 µ > − < nµ >2) < tµ >2 (< nµnν > − < nµ >< nν >) < tµ >< tν > (38) + Xµ=1 + 2Xµ>ν The first and second terms on the right hand side of (38) capture, respectively, the fluctuations in the lifetimes of the individual states and that in the number of visits to a kinetic state. Note that the number of visits to a particular state depends on the number of visits to the neighboring states on the kinetic diagram; this interstate correlation is captured by the third term on the right hand side of (38). Several different analytical and numerical techniques have been developed for calculation of the dwell time distribution [120, 149 -- 151]. Since the dwell time is essen- tially a first passage time [152], an absorbing boundary method [150] has been used. 1. DTD for a motor that never steps backward As an example, we consider again the simple scheme (9). In this case, the DTD is f (t) =(cid:18) ω1ω2 ω− − ω+(cid:19)(e−ω+t − e−ω−t) (39) where ω± = ω1 + ω−1 + ω2 2 ±(cid:20)r (ω1 + ω−1 + ω2)2 4 − ω1ω2(cid:21) (40) In the special case ω−1 = 0, ω+ = ω1 and ω− = ω2 and, hence, 16 It is possible to establish on general grounds that, for a motor with N mechano-chemical kinetic states like (35), the DTD is a sum of n exponentials of the form [148] f (t) = N Xj=1 Cje−ωjt (42) where N − 1 of the N coefficients Cj (1 ≤ j ≤ N ) are independent of each other because of the constraint im- posed by the normalization condition for the distribution f (t). Also note that the prefactors Cj can be both posi- tive or negative while ωj > 0 for all j. If we consider an even more special case of the scheme (9) where ω−1 = 0, ω1 = ω2 = ω, i.e., all the steps are completely irreversible and take place at the same rate ω, the DTD becomes the Gamma-distribution f (t) = ωN tN −1e−ωt Γ(N ) (43) In- where Γ(N ) is the gamma function with N = 2. terestingly, for the Gamma-distribution, the randomness parameter [154] (also called the Fano factor [155]) r = (< τ 2 > − < τ >2)1/2/ < τ > (44) is exactly given by r = 1/N . Can the experimentally measured DTD be used to determine the number of states N ? Unfortunately, for real motors, (i) not each step of a cycle is fully irreversible, (ii) the rate constants for dif- ferent steps are not necessarily identical, (iii) branched pathways are quite common. Consequently, 1/r may pro- vide just a bound on the rough estimate of N . Can one use the general form (42) of DTD to extract all the rate constants for the kinetic model by fitting it with the experimentally measured DTD? The answer is: NO. First, even if a good estimate of N is available, the number of parameters that can be extracted by fitting the experimental DTD data to (42) is 2N − 1 (n values of ωj and N − 1 values of Cj). On the other hand, the number of possible rate constants may be much higher [148]. For example, if transitions from every kinetic state to every other kinetic state is allowed, the total number of rate constants would be N (N − 1). In other words, in general, the kinetic rate constants are underspecified by the DTD. Second, as the expression (39) for the DTD of the example (9) shows expplicitly, the ω's that appear in the exponentials may be combinations of the rate con- stants for the distinct transitions in the kinetic model. It is practically impossible to extract the individual rate constants from the estimated ω's unless any explicit rela- tion like (40) between the estimated ω's and actual rate constants is apriori available. f (t) =(cid:18) ω1ω2 ω2 − ω1(cid:19)(e−ω1t − e−ω2t) (41) 2. Conditional DTD for motors with both forward and backward stepping Similar sum of exponentials for DTD have been derived also for machines with more complex mechano-chemical kinetics (see, for example, refs.[121, 146, 153]). For motors which can step both forward and backward, more relevant information on the kinetics of a motor are contained in the conditional dwell time distribution [156]. If n+ and n− are the numbers of forward and backward steps, respectively, then for large n = n+ + n−, the cor- responding step splitting probabilities are Π+ = n+/n and Π− = n−/n. The dwell times before a forward step and before a backward step can be measured separately. Hence, the prior dwell times τ+ and τ− can be obtained by restricting the summations in τ ← ± = 1 n± ± X τk (45) to just forward (+) or just backward (-) steps, respec- tively. In terms of splitting probabilities and prior dwell times, the mean dwell time hτ i can be expressed as hτ i = Π+τ ← + + Π−τ ← − (46) Compared to the prior dwell times, more detailed in- formation on the stepping statistics is contained in the four conditional dwell times, which are defined as follows: τ++ = dwell time between a + step followed by a + step τ+− = dwell time between a + step followed by a − step τ−+ = dwell time between a − step followed by a + step τ−− = dwell time between a − step followed by a − step (47) It is helpful to introduce pairwise step probabilities Π++, Π+−, Π−+, Π−−. Note that Π++ + Π+− = 1, and Π−+ + Π−− = 1. Neglecting finite time corrections, Π++ = n++/(n++ + n+−), Π+− = n+−/(n++ + n+−) Π−+ = n−+/(n−+ + n−−), Π−− = n−−/(n−+ + n−−) Hence τ ← + = Π++τ++ + Π+−τ−+ τ ← − = Π−+τ+− + Π−−τ−− (48) (49) Defining the post dwell times τ → that used for defining the prior dwell times, we get ± in a fashion similar to and τ → + = Π++τ++ + Π+−τ+− τ → − = Π−+τ−+ + Π−−τ−− hτ i = Π+τ → + + Π−τ → − (50) (51) This formalism has been applied, for example, single- headedkinesin KIF1A [157]. VII. A SUMMARY OF EXPERIMENTAL TECHNIQUES: ENSEMBLE VERSUS SINGLE-MACHINE The experimental techniques for probing the struc- ture and dynamics of molecular machines can be di- vided broadly into two groups: (i) ensemble-averaged, (ii) single-machine. 17 X-ray diffraction and electron microscopy, two of the most powerful biophysical techniques, yield the structure of molecular machines averaged over an ensemble. Elec- tron microscopy is a powerful alternative for the deter- mination of structures of those macromolecules whose crystals are not available. Cryo-electron microscopy [1, 158] combines the technique of electron microscopy with cryogenics-based sample preparation. The single-molecule techniques [159 -- 164] can be fur- ther classified into two groups: (i) methods of imaging, and (ii) methods of manipulation. These are not mutu- ally exclusive and can be probed by a single experimental set up. The most direct single-molecule imaging is car- ried out by fluorescence-based optical microscopy. Tra- ditional diffraction-limited optical microscopy provides a "hazy" image of the machine. However, modern opti- cal nanoscopy has broken the resolution limit imposed by diffraction [165]. Fluorescence microscopy provides a glimpse (howsoever hazy) of single molecules. Imaging a fluorescently labelled molecular machine in real time enables us to study its dynamics just as ecologists use "radio collars" to track individual animals. The techniques developed for the mechanical manipu- lation of a single machine can be classified on the basis of tranducers used; the table below provides a summary. Mechanical manipulation of a single biomolecule ւ ց Mechanical transducers Field-based t ransducers ւ ց SFM Micro-needle ւ ց EM-field ւ ց Flow-field Electric field Ma gnetic field VIII. SOLVING INVERSE PROBLEM BY PROBABILISTIC REVERSE ENGINEERING: FROM DATA TO MODEL A discrete kinetic model of a molecular motor can be regarded as a network where each node represents a dis- tinct mechano-chemical state. The directed links denote the allowed transitions. Therefore, such a model is unam- biguously specified in terms of the following parameters: (i) the total number N of the nodes, (ii) the N × N ma- trix whose elements are the rates of the transitions among these states; a vanishing rate indicates a forbidden direct transition. In the preceeding sections we handled the "forward problem" by starting with a model that is formulated on the basis of apriori hypotheses which are, essen- tially, educated guess as to the mechano-chemical kinet- ics of the motor. Standard theoretical treatments of the model yields data on various aspects on the modelled motor; this approach is expressed below schematically. Theoretical model → Experimental data Consistency between theoretical prediction and experme- ntal data validates the model. However, any inconsis- tency between the two indicates a need to modify the model. In most real situations the numerical values of the rate constants of the kinetic model are not known apriori. In principle, these can be extracted by analyzing the exper- imental data in the light of the model. However, not only the rate constants but also the number of states and the overall architecture of the mechano-chemical network as well as the kinetic scheme postulated by the model may be uncertain. In that case, the experimental data should be utilized to "select" the most appropriate model from among the plausible ones. In fact, more than one model, based on different hypotheses, may appear to be consis- tent with the same set of experimental data within a level of accuracy. The experimental data can be exploited at least to "rank" the models in the order of their success in accounting for the same data set. The "inverse problem" of inferring the model from the observed experimental data has to be based on the the- ory of probability. Such "statistical inference" [166] can be drawn by following methods developed by statisti- cians over the last one century. Inferring the complete network of mechano-chemical states and kinetic scheme of a molecular machine from its observed properties is reminiscent of inferring the operational mechanism of a given functioning macroscopic machine by "reverse en- gineering". This inverse problem, which is the main aim of this section, is expressed below schematically. Theoretical model ← Experimental data It is worth emphasizing that both the directions of in- vestigations, i.e. the forward problem and the inverse problem are equally important and complementary to each other [167]. A. Frequentist versus Bayesian approach Suppose, ~m be a column vector whose M compo- nents are the M parameters of the model, i.e., the trans- pose of ~m is ~mT = (m1, m2, ..., mM ) Let the data ob- tained in N observations of this model are represented by the N -component column vector ~d whose transpose is ~dT = (d1, d2, ..., dN ). Our "inverse problem" is to in- fer information on ~m from the observed ~d. The philoso- phy underlying the frequentist approach, i.e., approaches based on maximum-likelihood (ML) estimation and the Bayesian approach for extracting these information are different in spirit, as we explain in the next two subsub- sections [168]. For simplicity, let us assume that a device has only two possible distinct states denoted by E1 and E2. E1 kf ⇋ kr E2 (52) Let us imagine that we are given the actual sequence of the states, over the time interval 0 ≤ t ≤ T , generated 18 by the Markovian kinetics of the device. But, the mag- nitudes of the rate constants kf and kr are not supplied. We'll now formulate a method, based on ML analysis [169] to extract the numerical values of the parameters kf and kr. and t(2) Suppose t(1) j j denote the time interval of the j- th residence of the device in states E1 and E2, respectively. Moreover, suppose that the device makes N1 and N2 vis- its to the states E1 and E2, respectively, and N = N1 +N2 is the total number of states in the sequence. Therefore, j=1 t(1) j j=1 t(2) total time of dwell in the two states are T1 = PN1 and T2 =PN2 Since the dwell times are exponentially distributed for a Poisson process, the likelihood of any state trajectory S is the conditional probability j where T1 + T2 = T . j (cid:19) P (Skf , kr) = (cid:18)ΠN1 j=1kf e−kf t(1) j (cid:19) (cid:18)ΠN2 j=1kre−krt(2) r e−krT2(cid:19) (53) f e−kf T1(cid:19)(cid:18)kN2 = (cid:18)kN1 1. Maximum-likelihood estimate ML approcah [170] is based on finding the estimates of the set of model parameters that corresponds to the maximum of the likelihood P (~d ~m) for a fixed set of data ~d. For the kinetic scheme shown in equation (52), the the ML estimates of kf and kr are obtained by using (53) in d[lnP (Skf , kr)]/dkf = 0 = d[lnP (Skf , kr)]/dkr. It is straightforward to see [169] that these estimates are kf = N1/T1 and kr = N2/T2. 2. Bayesian estimate For drawing statistical inference regarding a kinetic model, the Bayesian approach has gained increasing pop- ularity in recent years [171 -- 176]. The areas of research where this has been applied successfully include various biological processes in, for example, genetics [177, 178], biochemistry [179], cognitive sciences [180], ecology [181], etc. In the Bayesian method there is no logical distinc- tion between the model parameters and the experimental data; in fact, both are regarded as random. The only dis- tinction between these two types of random variables is that the data are observed variables whereas the model parameters are unobserved. The problem is to estimate the probability distribution of the model parameters from the distributions of the observed data. The Bayes theorem states that P ( ~m~d) = P (~d ~m)P ( ~m) P (~d) where P (~d) can be expressed as P (~d) =Z P (~d ~m)P ( ~m)d ~m (54) (55) The likelihood P (~d ~m) is the conditional probability for the observed data, given a set of particular values of the model parameters, that is predicted by the kinetic model. However, implementation of this scheme also re- quires P ( ~m) as input. In Bayesian terminology P ( ~m) is called the prior because this probability is assumed apriori by the analyzer before the outcomes of the exper- iments have been analyzed. In contrast, the left hand side of equation (54) gives the posterior probability, i.e., after analyzing the data. Thus, an experimenter learns from the Bayesian anal- ysis of the data. Such a learning begins with an input in the form of a prior probability; the choice of the prior can be based on physical intuition, or general arguments based, for example, on symmetries. Prior choice can be- come simple if some experience have been gained from previous measurements. Often an uniform distribution of the model parameter(s) is assumed over its allowed range if no additional information is available to bias its choice. To summarize, Bayesian analysis needs not just the likelihood P (~d ~m) but also the prior P ( ~m). For the kinetic scheme shown in equation (52), the Bayes' theorem (54) takes the form P (kf , krS) = P (Skf , kr)P (kf , kr) P (S) 19 on the recording of the position of the motor in a single motor experiment. This problem is similar to an older problem in cell biology: ion-channel kinetics [182, 183]. Current passes through the channel only when it is in the "open" state. However, if the channel has more than one distinct closed states, the recordings of the current reveals neither the actual closed state in which the chan- nel was nor the transitions between those closed states when no current was recorded. Hidden Markov Model (HMM) [184 -- 187] has been applied to analyze FRET trajectories [188, 189], step- ping recordings of molecular motors [190, 191], and acto- myosin contractile system [192] to extract kinetic infor- mation. For a pedagogical presentation of the main ideas be- hind HMM, we start with the kinetic scheme shown in (54) and a simple, albeit unrealistic, situation and then by gradually adding more and more realistic features, ex- plain the main concepts in a transparent manner [169]. First, suppose that the actual sequence of states (trajec- tory) itself is visible; this case can be analyzed either by the ML-analysis of by Bayesian approach both of which we have presented above. We now relax the strong as- sumption about the trajectory and proceed as below. •If state trajectory is hidden and visible trajec- tory is noise-free The sequence of states of the device is, as before, gen- erated by a Markov processs which is hidden. Suppose the device emits photons from time to time that are de- tected by appropriate photo-detectors. For simplicity, we assume just two detection channels labelled by 1 and 2. For the time being, we also assume a perfect one-to-one correspondence between the state of the light emitting device and the channel that detects the photon. If the channel 1 (2) clicks then the light emitting device was in the state E1 (E2) at the time of emission. The inter- val ∆tj = tj+1 − tj between the arrival of the j-th and j + 1-th photons (1 ≤ j ≤ N ) is random. Thus, from the photo-detectors we get a visible se- quence of the channel index (a sequence made of a bi- nary alphabet) which we call "noiseless photon trajec- tory" [169]. The sequence of states in the noiseless photo trajectory is also another Markov chain that is conven- tionally referred to as the "random telegraph process". Note that the photon detected by channel 1 (or, channel 2) can take place at any instant during the dwell of the device in state 1 (or, state 2). Therefore, the sequence of states in the noiseless photon trajectory does not re- veal the actual instants of transition from one state of the device to another. Since the noiseless phton trajectory corresponds to a random telegraph process, the transition probabilities for P (Skf , kr)P (kf , kr) r)P (k′ P (Sk′ f , k′ f ,k′ r f , k′ r) (56) = Pk′ We now assume a uniform prior, i.e., constant for positive kf and kr, but zero otherwise. Then, P (kf , krS) is pro- portional to the likelihood function P (Skf , kr) (within a normalization factor). Normalizing, we get P (kf , krS) =(cid:20) T N1+1 Γ(N1 + 1) 1 kN1 f e−kf T1(cid:21)(cid:20) T N2+1 Γ(N2 + 1) 2 (57) The mean of kf is N1 + 1)/T1 whereas the most-likely es- timate is N1/T1. Similarly, the mean and most probable estimates of kr are obtained by replacing the subscripts 1 by subscripts 2. Moreover, the variance of kf and kr are (N1 + 1)/T 2 1 and (N2 + 1)/T 2 2 , respectively. kN2 r e−krT2(cid:21) 3. Hidden Markov Models The actual sequence of states of the motor, generated by the underlying Markovian kinetics, is not directly visi- ble. For example, a sequence of states that differ "chem- ically" but not mechanically do not appear as distinct this process are P (E1E1; kf , kr, ∆tj) = kr kf + kr + kf kf + kr e−(kf +kr )∆tj P (E1E2; kf , kr, ∆tj) = P (E2E1; kf , kr, ∆tj) = P (E2E2; kf , kr, ∆tj) = kr kf + kr kf kf + kr kf kf + kr [1 − e−(kf +kr )∆tj ] [1 − e−(kf +kr )∆tj ] + kr kf + kr e−(kf +kr )∆tj (58) where P (EµEν; kf , kr, ∆tj) is the conditional probability that state of the device is Eµ given that it was in the state Eν at a time ∆t earlier. The likelihood of the visible data sequence {V } is now given by P ({V }kf , kr) = P (V1kf , kr)ΠN −1 j=1 P (Vj+1Vj ; kf , kr, ∆tj) (59) where the first factor on the right hand side is the ini- tial probability (usually assumed to be the equilibrium probability). The transition probabilities on the right hand side of equation (59) are the conditional probabili- ties given in equation (58). Unlike the previous simpler case, where the state sequence itself was visible, no ana- lytical closed-form solution is possible in this case. Nev- ertheless, analysis can be carried out numerically. •If state trajectory is hidden and visible trajec- tory is noisy In the preceeding case of a noise-free photon trajectory, we assumed that from the channel index we could get perfect knowledge of the state of the emitting device. More precisely, the conditional probabilities were P (1E1) = 1 P (1E2) = 0 P (2E1) = 0 P (2E2) = 1 (60) However, in reality, background noise is unavoidable. Therefore, if a photon is detected by the channel 1, it could have been emitted by the device in its state E1 (i.e., it is, indeed, a signal photon) or it could have come from the background (i.e., it is a noise photon). Suppose ps is the probability that the detected photon is really a signal that has come from the emitting device. Suppose pb1 is the probability of arrival of a background photon in the channel 1. The probability that a background photon arrives in channel 2 is 1 − pb1. Then [169], E(1E1) = ps + (1 − ps)pb1 E(2E1) = 1 − P (1E1) = (1 − ps)(1 − pb1) E(1E2) = (1 − ps)pb1 E(2E2) = 1 − P (1E2) = ps + (1 − ps)(1 − pb) (61) 20 Thus, in this case, the relation between the states of the hidden and visible states is not one-to-one, but one-to- many.Therefore, given a hidden state of the device, a set of "emission probabilities" determine the probability of each possible observable state; these are listed in equa- tions (61) for the device (52). •HMM: formulation for a generic model of molec- ular motor On the basis of the simple example of a 2-state sys- tem presented above, we conclude that, for data analysis based on a HMM four key ingredients have to be speci- fied: (i) the alphabet of the "visible" sequence {µ} (1 ≤ µ ≤ N ), i.e., N possible distinct visible states; (ii) the alpha- bet of the "hidden" Markov sequence {j} (1 ≤ j ≤ M ), i.e., M allowed distinct hidden states, (iii) the hidden- to-hidden transition probabilities W (j → k), and (iv) hidden-to-visible emission probabilities E(j → µ). In ad- dition to the transition probabilities and emission proba- bilities, which are the parameters of the model, the HMM also needs the initial state of the hidden variable(s) as in- put parameters. ... Vt ⇑ V1 ⇑ Visible: V0 ⇑ VT ⇑ Hidden: H0 → H1 → ...→ Ht → HT Suppose P ({V }HM M, {λ}) denotes the probability that an HMM with parameters {λ} generates a visible sequence {V }. Then, P ({V }HM M, {λ}) = X{H} P ({V }{H}; {λ})P ({H}{λ}) (62) where P ({H}{λ}) is the conditional probability that the HMM generates a hidden sequence {H} for the given parameters {λ} and P ({V }{H}; {λ}) is the conditional probability that, given the hidden sequence {H} (for pa- rameters {λ}) the visible sequence {V } would be ob- tained. Once P ({V }HM M, {λ}) is computed, the parame- ter set {λ} are varied to maximize P ({V }HM M, {λ}) (for the convenience of numerical computation, often lnP ({V }HM M, {λ}) is maximized. The total number of possible hidden sequences of length T is T MN . In order to carry out the summation in equation (62) one has to enumerate all possible hidden sequences and the corresponding probabilities of occurrences. A successful implementation of the HMM requires use of an efficienct numerical algorithm; the Viterbi algorithm is one such candidate. In case of a molecular motor, the "chemical states" are not visible in a single molecule experiment. More- over, even its mechanical state that is "visible" in the recordings may not be its true position because of (a) measurement noise, and (b) steps missed by the detector. Let is denote the "visible" sequence by the recorded posi- tions {Y } whereas the hidden sequence is the composite mechano-chemical states {X, C} where X and C denote the true position and chemical state, respectively. The transition probabilities are denoted by W (Xt−1, Ct−1 → Xt, Ct) One possible choice for the emission probabilities E is [190] E(Xt → Yt) =s 1 (2πσ2 t ) exp[− (Yt − Xt)2 (2σ2 t ) ] (63) where 21 In this case, P ({Y }HM M, {λ}) = X{X,C} P ({Y }{X, C}; {λ})P ({X, C}{λ}) (64) P ({X, C}{λ}) = PX0,C0W (X0, C0 → X1, C1)W (X1, C1 → X2, C2).....W (XT −1, CT −1 → XT , CT ) (65) and P ({Y }{X, C}; {λ}) = E(X0 → Y0)E(X1 → Y1)...E(Xt → Yt)...E(XT → YT ) (66) The usual strategy [188, 190] consists of the following steps: Step I: Initialization of the parameter values for it- eration. Step II: Iterative re-estimation of parameters for maximum likelihood: the parameters {W (Xt−1, Ct−1 → Xt, Ct)}, {E(Xt → Yt)} and PX0,C0 are re-estimated iter- atively till P ({Y }HM M, {λ}) saturates to a maximum. Step III: Construction of "idealized" trace: using the final estimation of the model parameters, the position of the motor as a function of time is reconstructed; natu- rally, this trace is noise-free. Step IV: Extraction of the distributions of step sizes and dwell times: from the ideal- ized trace, the distributions of the steps sizes dwell times are obtained. These distributions can be compared with the corresponding theoretical predictions. B. Extracting FP-based model from data? In this section so far we have discussed methods for ex- tracting the master-equation based models that describe the kinetics of motors in terms of discrete jumps on a fully discrete mechano-chemical state space. However, as we summarized in section V, kinetics of molecular motors can be formulated also in terms of FP equations which treat space as a continuous variable. Can one extract the parameters of such FP-based models from the data collected from single-molecule experiments? One of the key ingredients of the FP-based approach is the profile of the potential Vµ(x) felt by the motor in the "chemical" state µ. To my knowledge, it has not been possible to extract it from any experimental data. Now, let us define dφ(x)/dx = 1 Pµ Pµ(x)Xµ Pµ(x)[dVµ(x)/dx] (67) to be the profile of the potential averaged over all the chemical states. It can be shown [193] that φ(x) kBT = F xkBT − ln[Xµ Pµ(x)] − J D Z x 1 dy 0 Pµ Pµ(y) (68) Based on this observation, a prescription has been sug- gested [193] to extract φ(x) by analyzing the time se- ries of motor positions obtained from single-motor ex- periments. IX. OVERLAPPING RESEARCH AREAS A. Symmetry breaking: directed motility and cell polarity Energy is a scalar quantity whereas velocity is a vector. How does consumption of energy give rise to a non-zero average velocity of a molecular motor? Moreover, a di- rected movement that a motor exhibits on the average, requires breaking the forward-backward symmetry on its track. What are the possible cause and effects of this broken symmetry at the molecular level? As far as the cause of this asymmetry is concerned, the asymmetry of the tracks alone cannot explain the "di- rected" movement of the motors, because on the same track members of different families of motors can, on the average, move in opposite directions. Obviously, the structural design of the motors and their interactions with the respective tracks also play crucial roles in deter- mining their direction of motion along a track. Further- more, this broken symmetry at the molecular level, e.g., the "directed" movement of molecular motors, has im- portant effects on various biological phenomena, partic- ularly "vectorial" processes, at the sub-cellular and cel- lular levels. In general, a cell is not isotropic. Motors are essential in breaking the cellular symmetry [194]. Can we establish a unique set of basic principle underlying the symmetry breaking [194, 195]? Therefore, the cause and effects of broken symmetry of molecular motors can be examined in the broader con- text of the fundamental principles of symmetry breaking in physics and biology [196] (see also other articles in the special "Perspectives on Symmetry Breaking in Biol- ogy" [197]). For macroscopic systems in thermodynamic equilibrium, symmetry breaking is explained in terms of the form of the free energy. However, since living cells are far from thermodynamic equilibrium, the theory of symmetry breaking in those systems cannot be based on thermodynamic free energy. Kinetics cannot be ignored in the study of symmetry breaking in living systems. B. Self-organization and pattern formation: assembling machines and cellular morphogenesis The interior of a living bacterial cell is far from ho- mogeneous; the intracellular space of eukaryotes are di- vided into separate compartments. Scaling is one of the interesting properties of many physical quantities that gives rise to some well known universalities. The scal- ing properties of a cell and the subcellular compartments [198, 199]. often depend on the machineries which as- semble them. The size, shape, location and number of intracellu- lar compartments as well as modular intracellular ma- chineries are self-organized, rather than self-assembled. Dissipation takes place in "self-organization" and distin- guishes it from "self-assembly"; the latter corresponds to the minimum of thermodynamic free energy whereas self- organized system does not attain thermodynamic equi- librium [195, 200 -- 202]. Molecular motors and their fila- mentous tracks play important roles in the intracellular self-organization process [195, 200] and even in the cellu- lar morphogenesis which may be regarded as a problem of pattern formation far from equilibrium. C. Dissipationless computation: polymerases as "tape-copying Turing machines" The concept of information in the context of biolog- ical systems has been discussed at length in the past [203]. The operation of molecular machines involved in genetic processes can be analyzed in terms of storage and transmission of information. In fact, a particular class of machines carry out what may be viewed as data trans- mission whereas that of others may be viewed as digital- to-analog conversion [204] . For these obvious reasons, a broad class of molecular machines are interesting also from the perspective of information theory, electronic en- gineering and computer science. Computation can be viewed as a transition from one state to another. However, in a conventional digital com- puter an elementary operation is logically irreversible. To understand the meaning of this term, consider the two summations 3 + 1 = 4 and 2 + 2 = 4. If the computer 22 retains only the output, i.e., 4 and erases the input num- bers (i.e., 3 and 1, or 2 and 2, as the case may be) after summing the two input numbers, then, given only the output (i.e., 4) it is impssible to figure out whether the input were 3, 1 or 2, 2. Note that erasing every bit leads to loss of information which may also be interpreted as an increase of entropy. However, strategies for logically reversible computation have been developed [205, 206]. For example, the simplest strategy for logically reversible computation is based on the prescription that neither the initial input nor the data in any intermediate step should be erased; instead, these should be retained in an auxil- iary register. In practice, computation is carried out with a device that is governed strictly by the laws of physics that in- cludes thermodynamics. Since entropy increases in ev- ery irreversible physical process, it would be tempting to associate the logical irreversibility of computation with a physical irreversibility of the computational device. Since each bit of a classical digital computer has only two pos- sible states, at first sight, one would expect dissipation of kBT ln2 energy for every bit erased [207]. But, the pos- sibility of logically reversible computation raises a funda- mental question: is it possible for a physical computing device to carry out physically reversible (and, hence non- dissipative) computation? Operation of a polymerase (and that of a ribosome) can be regarded as computation. More precisely, such a computing machine can be viewed as a "tape-copying Turing machine" that polymerizes its output tape, in- stead of merely writing on a pre-synthesized tape [205] Dissipationless operation of these machines is possible only if every step is error-free which, in turn, is achiev- able in the vanishingly snall speed, i.e., reversible limit [208]. D. Enzymatic processes: conformational fluctuations, static and dynamic disorder For a motor that doesn't step backwards, the position advances in the forward direction one step at a time; how- ever, the time gap between the successive steps, i.e., dwell time, is a random variable. Similarly, in single-molecule enzymology, the population of the product molecules in- creases by one in each enzymatic cycle, the time gap be- tween the release of the successive product molecules is random. The distribution of this time [209, 210] is analo- gous to that of the dwell times of molecular motors [120]. Therefore, the research fields of single-motor mechanics and single-enzyme reactions can enrich each other by ex- change of concepts and techniques [148]. As a concrete example, consider the chemical reaction E + S ω1 ⇋ ω−1 ω2→ E + P I1 (69) catalyzed by the enzyme E where S and P denote the substrate and product, respectively. The enzyme and the substrate form, at a rate ω1, an intermediate com- plex I1 that can either dissociate into the free enzyme and the substrate at a rate ω−1, or get converted irre- versibly into the product and free enzyme at a rate ω2. In a bulk sample, where a large number of enzymes con- vert many substrates into products by catalyzing this re- action simultaneously, the measured rate of the reaction is actually an average over the ensemble. This rate was calculated by Michaelis and Menten about 100 years ago and is given by the celebrated Michaelis-Menten equation [211]. Formally, the scheme (69) is very similar to the scheme (9) that we presented earlier as a very simple example of the mechano-chemical cycle of a molecular motor. The time taken by the chemical reaction (69) fluctuates from one enzymatic cycle to another. One of the fundamen- tal questions is: what are the conditions under which the average rate would still satisfy the Michaelis-Menten equation [209, 210]? Another topic that overlaps research on molecular mo- tor kinetics and chemical kinetics is allosterism [212]. A motor protein has separate sites for binding the fuel molecule and the track. How do these two sites commu- nicate? How does the binding of ligand at one site affect the binding affinity of the other? The mechanochemical cycle of a motor can be analyzed [213] from the perspec- tive of allosterism which is one of the key mechanisms of cooperativity in protein kinetics [214]. E. Applications in biomimetics and nano-technology Initially, technology was synonymous with macro- technology. The first tools applied by primitive humans were, perhaps, wooden sticks and stone blades. Later, as early civilizations started using levers, pulleys and wheels for erecting enormous structures like pyramids. Until nineteenth century, watch makers were, perhaps, the only people working with small machines. Using magnifying glasses, they worked with machines as small as 0.1mm. Micro-technology, dealing with machines at the length scale of micrometers, was driven, in the second half of the twentieth century, largely by the computer miniatur- ization. In 1959, Richard Feynman delivered a talk [215] at a meeting of the American Physical Society. In this talk, entitled "There's Plenty of Room at the Bottom", Feyn- man drew attention of the scientific community to the unlimited possibilities of manupilating and controlling things on the scale of nano-meters. This famous talk is now regarded by the majority of physicists as the defin- ing moment of nano-technology [216]. In the same talk, in his characteristic style, Feynman noted that "many of the cells are very tiny, but they are very active, they manufacture various substances, they walk around, they wiggle, and they do all kinds of wonderful things- all on a very small scale". 23 From the perspective of applied research, the natural molecular machines opened up a new frontier of nano- technology [217 -- 222]. The miniaturization of compo- nents for the fabrication of useful devices, which are es- sential for modern technology, is currently being pursued by engineers following mostly a top-down (from larger to smaller) approach. On the other hand, an alternative approach, pursued mostly by chemists, is a bottom-up (from smaller to larger) approach. Unlike man-made machines these are products of Na- ture's evolutionary design over billions of years. In fact, cell has been compared to an "archeological excavation site" [223], the oldest modules of functional devices are the analogs of the most ancient layer of the exposed site of excavation. We can benefit from Nature's billion year experience in nano-technolgy. The term biomimetics has already become a popular buzzword [220, 221]; this field deals with the design of artificial systems utilizing the principles of natural biological systems. X. CONCEPT OF BIOLOGICAL MACHINES: FROM ARISTOTLE TO ALBERTS The concept of "living machine" evolved over many centuries. Some of the greatest thinkers in human history made significant contributions to this concept. It started with the man-machine (and, more generally, animal- machine) analogy. Aristotle [224] distinguished between the "body" and the "soul" of an organism. However, after more than one and half millenia, an intellectu- ally provocative idea was put forward by Rene Descartes when he argued that animals were "living machines". This idea took its extreme form in Julien Offray de La Mettrie's book [225] L'homme Machine ('man a ma- chine'). Leibnitz [226] wrote that the body of a living being is a kind of "divine machine or natural automa- ton" and it "surpasses all artificial automata", because not each part of a man-made machine is itself a machine. In contrast, he argued, living bodies are "machines in their smallest parts ad infinitum". Is this statement to be interpreted as Leibnitz's speculation for the existence of machines within machines in a living organism? The debate over this interpretation continues [227, 228]. All the great thinkers from Aristotle to Descartes and Leibnitz compared the whole organism with a machine, the organs being the coordinated parts of that machine. Cell was unknown; even micro-organisms became visible only after the invention of the optical microscope in the seventeenth century. Marcelo Malpighi, father of micro- scopic anatomy, speculated in the 17th century about the existence molecular machines in living systems. He wrote (as quoted in english by Marco Piccolino [229]) that the organized bodies of animals and plants been constructed with " very large number of machines". He went on to characterize these as "extremely minute parts so shaped and situated, such as to form a marvelous organ". Un- fortunately, the molecular machines were invisible not only to the naked eye, but even under the optical micro- scopes available in his time. In fact, individual molec- ular machines could be "caught in the act" only in the last quarter of the 20th century. We highlight here the progress made during the last three centuries when grad- ually the analogy with machine got extended to cover all levels of biological organization- from the topmost level of organisms down to cellular and subcellular levels. Not surprisingly, muscles seem to have attracted maximum attention in the context of machinery of life. Thomas Henry Huxley, dubbed as "Darwin's bulldog" for his strong support for Darwinian ideas of evolution, delivered his famous lecture, titled "On the Physical Ba- sis of Life", on 8th November, 1868 (see ref.[9] for the published version). Among the provocating statements and insightful comments in this article, which received lot of hostile criticism at that time, I quote only a few that are directly relevant from the perspective of molec- ular machines. Huxley had the foresight to see that [9] "speech, gesture, and every other form of human action are, in the long run, resolvable into mascular contrac- tion, and muscular contraction is but a transitory change in the relative positions of the parts of a muscle"- a re- markably insightful comment on contractility and motil- ity because the mechanism of muscle contraction was dis- covered almost 90 years later, one of the discoverers being his grandson! David Ferrier, a pioneering neurologist and a younger contemporary of Thomas Huxley, in his lecture [230] de- livered at the Middlesex Hospital Medical School, on October 5th, 1870, not only echoed similar ideas but stressed the role of physical energy, rather than any hy- pothetical vital action, for sustaining life of an organism. A few years later, Robert Henry Thurston, the first pres- ident of the ASME (American Society of Mechanical En- gineers) and a pioneer of modern engineering education, investigated what he called a "vital machine" (or "prime motor") [231] from an engineer's perspective. In the initial stages, most of the visionaries compared animals with a machine. Although movements of plants received much less attention, results of pioneering sys- tematic studies of these phenomena were reported al- ready in the nineteenth century by Charles Darwin in a classic book [232] which was co-authored by his son. Later, in the preface of his report on the classic investi- gations on the mechanical response of plants to stimuli, Jagadis Chandra Bose [233] wrote: "From the point of view of its movements a plant may be regarded ...simply as a machine, transforming the energy supplied to it, in ways more or less capable of mechanical explanation". Interestingly, the first chapter was titled "The plant as a machine" [233]. Based on the well known fact that all animals exhibit irritability and contractility [234], Thomas Henry Huxley had already speculated in ref.[9], that "it is more than probable, that when the vegetable world is thoroughly explored, we shall find all plants in possession of the same powers..". In support of this possibility he described a 24 phenomenon that is now known as cytoplasmic streaming [8]. His speculation that "the cause of these currents lie in contraction" was established a century later when cytoplasmic streaming was shown to be caused by an acto-myosin system. "The story of the living machine" [235] narrated by Herbert William Conn, one of the founding members and the third president of the Society of Americam Bacteriol- ogists (renamed, in 1960, American Society for Microbi- ology), is a critical overview of the understanding of these machines at the end of the nineteenth century. Because of his deep understanding of the basic principles of phys- ical sciences as well as evolutionary biology, his narrative on the fundamental principles of living machines remains as contemporary today as it was at the time of its publi- cation. Conn asked whether the operations of individual organisms, i.e., the living machines, could be "reduced to the action of still smaller machines" [235]. Conn ar- gued that one can look upon each constituent cell of an organism also "a little engine with admirably adapted parts" [235]. His summary [235] that a living organism is a "series of machines one within the other" sounds very similar to Leibnitz's philosophical idea. Conn went even further: "As a whole it is machine, and its parts are sep- arate machines. Each part is further made up of still smaller machines until we reach the realm of the micro- scope. Here still we find the same story. Even the parts formerly called units, prove to be machines". He spec- ulated [235] "we may find still further machines within" cells. He ended his summary with the statement "And thus vital activity is reduced to a complex of machines, all acting in harmony with each other to produce together the one result- life" [235]. Conn [235] not only discussed "the running of the living machine", but also "the origin of the living ma- chine". In the latter context, while pointing out the dif- ferences in the principles of engineering design of man- made machines and evolutionary principles of nature's nano-machines, he wrote [235]: "It is something as if steam engine of Watt should be slowly changed by adding piece after piece until there was finally produced the mod- ern quadruple expansion engine, but all this change being made upon the original engine without once stopping its motion." Jaques Loeb, a famous embryologist, delivered a se- ries of eight lectures at the Columbia university in 1902 (see ref.[236] for a more complete published version). In the first lecture, Loeb started by saying "In these lec- tures we shall consider living organisms as chemical ma- chines, consisting essentially of colloidal material, which possess the peculiarities of automatically developing, pre- serving, and reproducing themselves" [236]. He empha- sized the crucial differences between natural and artificial machines by the following statement: "The fact that the machines which can be created by man do not possess the power of automatic development, self-preservation, and reproduction constitutes for the pesent a fundamen- tal difference between living machines and artificial ma- chines". However, just like some of his other visionary predecessors, he also admitted that "nothing contradicts the possibility that the artificial production of living mat- ter may one day be accomplished" [236]. The concept of "living machine" as a "transformer of energy", from the level of a single cell to the level of a multi-cellular organism as complex as a man, found men- tion in many lectures and books of the leading biologists in the late nineteenth and early twentieh centuries (see, for example, [237]). Research on molecular machines was focussed almost exclusively on the mechanism of muscle contraction during the first half of the 20th century and it was dominated by Archibald Hill and Otto Meyerhof [238, 239] who shared the Nobel prize in Physiology (or medicine) in 1922. By mid-twentieth century, physical sciences already witnessed spectacular progress and life sciences was just embarking on its golden period. In his Guthrie lecture (also titled "The Physical Basis of Life"), delivered on 21st November, 1947 (see ref.[240] for the full text), J. D. Bernal drew attention of the audience to the structure and dynamics of machines. Ten years later, in a review article, the title of which again had the words "physi- cal basis of life", Schmitt [241] made even more concrete references to molecular machines covering almost all the types of machines that we have sketched in this article. The concept of molecular machines [242] was men- tioned on many occasions in the late twentieth century by leading molecular biologists who made outstanding con- tributions in elucitaing their mechanisms of operation. The strongest impact was made, however, by the influ- ential paper of Bruce Alberts [243], then the president of the National Academy of Sciencs (USA). He wrote that "the entire cell can be viewed as a factory that contains an elaborate network of interlocking assembly lines, each of which is composed of a set of large protein machines" [243]. Why did I start my story with Aristotle? In D'Arcy Thompson's words [244] "We know that the history of biology harks back to Aristotle by a road that is straight and clear, but that beyond him the road is broken and the lights are dim". Moreover, a section of the mod- ern enterprise in animal sciences is pursuing important investigations on the efficiency of energy utilization by animals [245] treating an entire animal as an a "combus- tion engine", a modern and scientifically correct version of the original philosophical idea developed by Aristotle and propagated by his followers. Why am I closing my story with Alberts? Because of a fortutious coincidence, Alberts' vision for training the next generation of biologists for studying natural nano- machines coincided with the explosive beginning of nano- technology that includes research programs on artificial nano-machines. Alberts' article [243] was appeared at a time when the statistical physics of Brownian ratchets was getting lot of attention [94]. The concept of molec- ular machine has matured fully into an area of inter- disciplinary research in the twenty-first century. A new exciting era of research has just begun! 25 XI. SUMMARY AND OUTLOOK So far as the current status of our understanding is concerned, I would say that till the middle of the 20th century we had never seen or manipulated a single molec- ular machine although machine-like operation of a cell or an entire organism was fairly well established. In the last 60 years this area has seen explosive growth in activity. The structures as well as the mechanisms of many ma- chines no longer appear any more mysterious than those of macroscopic machines. It is practically impossible to predict new ideas in any field of research; molecular machines are no exception. The reason, as Peter Medawar [246] admitted in his pres- idential address at the Cambridge meeting of the British Association for the Advancement of Science, is as follows: "to predict an idea is to have an idea, and if we have an idea it can no longer be the subject of prediction". Therefore, in this section, I do not propose any new idea but merely mention a few systems and phenomena which need new ideas for their studies and understanding. Let me begin with solving the forward problem with process modeling. Here we need new ideas to make progress in two opposite directions: (a) in-silico modeling of single individual machines in terms of its coordinat- ing parts in an aqueous medium- aquatic nano-robotics; (b) integration of nano-machines and machine-assemblies into a micro-factory, the living cell. So far as the point (a) is concerned, serious efforts have been made by de- veloping coarse-grained computational models [247 -- 249] that can be regarded as the substitutes for full molecular dynamics based models. However, a technical (or algo- rithmic) breakthrough is needed to achieve the ultimate goal of "seeing" the operation of a machine in its natu- ral aqueous environment by carrying out an experiments in-silico. Next let me explain the aim of developing integrated models. It has been strongly argued at the dawn of this millenium [250] that the machinery of life is modular. The machineries of transcription, translation, replication, chromosome segregation, etc. are all examples of mod- ules which perform specific tasks. The components of some of these modules are parts of a single machine, e.g., all devices participating in translation function on a sin- gle platform provided by a ribosome. Signal transduction machinery is an extreme example of the other types of functional modules which are spatially distributed over a significantly large region of intracellular space without need for direct physical contacts among the parts. The machines that we have considered here are all more or less spatially-confined functional modules. Nor- mally, functional modules are not completely isolated and must communicate and coordinate their function with other modules. For example, DNA replication and chro- mosome segregation machineries require proper coordi- nation. To my knowledge, no attempt has been made so far for quantitative stochastic process modeling incor- porating more than one functional module in a seamless fashion. Such an enterprise may look like what is now the happening in systems biology: to integrate the op- erations of machineries for different functional modules within a single theoretical framework. Finally, let me point out some of the standard prac- tices of statistical inference that, to my knowledge, have not been followed so far while reverse-engineering molec- ular machines. The experimental data for molecular ma- chines have been analyzed by several groups to extract an underlying kinetic model. However, most often the analysis carried out during such reverse engineering is based on a single working hypothesis. It would be desir- able to follow Platt's [251] principle of "strong inference" which is an extension of Chamberlin's [252] "method of multiple working hypothesis". By multiple model I do not mean models hierarchically nested such that each one is a special case of the model at the next level. By the term multiple model, I mean truly competing models that may, however, overlap partially. The relative scores of the competing models (and the corresponding underly- ing hypothese) would be a true reflection of their merits. For example, in case of a closely related systems in cell bi- ology and systems biology, experimental data have been analyzed recently within the framework of this method [253]. In the case of molecular machines, a method for model selection has been developed [254]; it extracts the best model by optimizing the maximum evidence, rather than maximum likelihood, that treats, for example, the 26 number of discrete states of the system as a variational parameter. This is a step in the right direction. However, I am not aware of any work that ranks alternative models according to relative scores computed on the basis of the principle of strong inference. The molecular mechanism of muscle myosin and the structure of double-stranded DNA were both discovered in the 1950s. These discoveries set the stage for the re- search on cytoskeletal motors as well as on machines that make, break and manipulate nucleic acids and proteins. Last 50 years has seen enormous progress. I wanted to tell you about the PIs and their contributions during this glorious period. But, I have run out of space. So, I'll nar- rate this fascinating story in detail elsewhere [255]. Acknowledgements: I apologize to all those re- searchers whose work, although no less important, could not be included in this article because of space limita- tions. I thank all my students and colleagues with whom I have collaborated on molecular machines. I am also in- debted to many others (too many to be listed by name) for enlightening discussions and correspondences over the last few years. I thank Ashok Garai for a critical read- ing of the manuscript. Active participation in the MBI annual program on "Stochastics in Biological Systems" during 2011-12 motivated me to look at molecular ma- chines from the perspectives of statisticians. I thank Marty Golubitsky, Director of MBI, for the hospitality at OSU. This work has been supported at IIT Kanpur by the Dr. Jag Mohan Garg Chair professorship, and at OSU by the MBI and the National Science Foundation under grant DMS 0931642. [1] J. Frank, (ed.) Molecular machines in biology: workshop of the cell (Cambridge University Press, 2011). [2] D. Chowdhury, Resource Letter PBM-1: Physics of biomolecular machines, Amer. J. Phys. 77, 538-594 (2009). [3] J. Howard, Mechanics of motor proteins and the cy- toskeleton, (Sinauer Associates, Sunderland, 2001) . [4] M. Schliwa, (ed.) Molecular Motors, (Wiley-VCH, 2003). [5] A.B. Kolomeisky and M.E. Fisher, Molecular motors: a theorist's perspective, Annu. Rev. Phys. Chem. 58, 675-695 (2007). [6] J.M. Skotheim and L. Mahadevan, Physical limits and design principles for plant and fungal movements, Sci- ence 308, 1308-1310 (2005). [7] P.T. Martone, M. Boller, I. Burgert, J. Dumais, J. Ed- wards, K. Mach, N. Rowe, M. Rueggberg, R. Seidel and T. Speck, Mechanics without muscle: biomechanical in- spiration from the plant world, Integrative and Compar- ative Biology 50, 888-907 (2010). [8] T. Shimmen The sliding theory of cytoplasmic stream- ing: fifty years of progress, J. Plant Res. 120, 31-43 (2007). [9] T.H. Huxley, On the physical basis of life, Fortnightly Review, Feb. (1869). [10] N. Cozzarelli, G.J. Cost, M. Nollmann, T. Viard and J.E. Stray, Giant proteins that move DNA: bullies of the genomic playground, Nat. Rev. Mol. Cell Biol. 7, 580-588 (2006). [11] W.M. Stark, B.F. Luisi and R.P. Bowater, Machines on genes: enzymes that make, break and move DNA and RNA, Biochem. Soc. Trans. 38, 381-383 (2010). [12] R. Mallik and S.P. Gross, Molecular motors as cargo transporters in the cell- the good, the bad and the ugly, Physica A 372, 65-69 (2006). [13] S. Leibler and D.A. Huse, Porters versus rowers: a uni- fied stochastic model of motor proteins, J. Cell Biol. 121, 1357-1368 (1993). [14] S. P. Gross, Hither and yon: a review of bi-directional microtubule-based transport, Phys. Biol. 1, R1-R11 (2004). [15] M.A. Welte, Bidirectional transport along microtubules, Curr. Biol. 14, R525-R537 (2004). [16] M. Maniak, Organelle transport: a park-and-ride system for melanosomes, Curr. Biol. 13, R917-R919 (2003). [17] C. Leduc, O. Campas, J.F. Joanny, J. Prost and P. Bassereau, Mechanism of membrane nanotube forma- tion by molecular motors, Biochim. Biophys. Acta 1798, 1418-1426 (2010). [18] J. Howard and A.A. Hyman, Microtubule polymerases and depolymerases, Curr. Opin. Cell Biol. 19, 31-35 (2007). [19] G. Klein, K. Kruse, G. Cuniberti and F. Julicher, Fila- ment depolymerization by motor molecules, Phys. Rev. Lett. 94, 108102 (2005). [20] B.S. Govindan, M. Gopalakrishnan and D. Chowdhury, Europhys. Lett. 83, 40006 (2008). [21] A. Zemel and A, Mogilner, Motor-induced sliding of mi- crotubule and actin bundles, Phys. Chem. Chem. Phys. 11, 4821-4833 (2009). [22] J.R. Sellers, Fifty years of contractility research post sliding filament hypothesis, J. Muscle Res. Cell Motil- ity 25, 475-482 (2004). [23] J.M. Squire and D.A.D. Parry, Fibrous proteins: muscle and molecular motors, (Elsevier 2005). [24] J.D. Jontes, Theories of muscle contraction, J. Struct. Biol. 115, 119-143 (1995). [25] S. Pellegrin and H. Mellor, Actin stress fibres, J. Cell Sci. 120, 3491-3499 (2007). [26] A. Zumdieck, K. Kruse, H. Bringmann, A.A. Hyman and F. Julicher, Stress generation and filament turnover during actin ring constriction, PLoS ONE 2, e696 (2007). [27] E. Karsenti and I. Vernos, The mitotic spindle: a self- made machine, Science 294, 543-547 (2001). [28] G. C. Scholey and J.M. Scholey, Mitotic force generators and chromosome segregation, Cell. Mol. Life Sci. 67, 2231-2250 (2010). [29] P. Satir, L.B. Pedersen and S.T. Christensen, The pri- mary cilium at a glance, J. Cell Sci. 123, 499-503 (2010). [30] C.B. Lindemann and K.A. Lesich, Flagellar and ciliary beating: the proven and the possible, J. Cell Sci. 123, 519-528 (2010). [31] J.E. Italiano Jr., S. Patel-Hett and J.H. Hartwig, Me- chanics of proplatelet elaboration, J. Thrombosis and Haemostasis, 5(Supp.1), 18-23 (2007). [32] A. Mogilner and G. Oster, Polymer motors: pushing out the front and pulling up the back, Curr. Biol. 13, R721-R733 (2003). [33] J.R. McIntosh, V. Volkov, F.I. Ataullakhanov and E.L. Grischuk, Tubulin depolymerization may be an ancient biological motor, J. Cell Sci. 123, 3425-3434 (2010). [34] L. Mahadevan and P. Matsudaira, Motility powered by supramolecular springs and ratchets, Science 288, 95-99 (2000). [35] I.M. Tolic-Norrelykke, Push-me-pull-you: how micro- tubules organize the cell interior, Eur. Biophys. J. 37, 1271-1278 (2008). [36] M.F. Carlier (ed.), Actin-based motility: cellular, molec- ular and physical aspects, (Springer, 2010). [37] H.P. Erickson, D.E. Anderson and M. Osawa, FtsZ in bacterial cytokinesis: cytoskeleton and force generator all in one, Microbiol. Mol. Biol. Rev. 74, 504-528 (2010). [38] C.W. Wolgemuth, L. Miao, O. Vanderlinde, T. Roberts and G. Oster, MSP dynamics drives nematode sperm locomotion, Biophys. J. 88, 2462-2471 (2005). [39] E. Nudleman and D. Kaiser, Pulling together with type IV pili, J. Mol. Microbiol. and Biotechnol. 7, 52-62 (2004). [40] L. mahadevan, C.S. Riera and J.H. Shin, Structural dy- namics of an actin spring, Biophys. J. 100, 839-844 (2011). [41] G. Schatz and B. Dobberstein, Common principles of protein translocation across membranes, Science 271, 1519-1526 (1996). 27 transport machines, Cold Spring Harb. Perspect. Biol. 2, a000406 (2010). [43] M. Stewart, Nuclear export of mRNA, Trends in Biochem. Sci. 35, 609-617 (2010). [44] P. Guo and T. J. Lee, Viral nanomotors for packaging of dsDNA and dsRNA, Molec. Microbiol. 64, 886-903 (2007). [45] J. Yu, J. Moffitt, C.L. Hethrington, C. Bustamante and G. Oster, Mechanochemistry of a viral DNA packaging motor, J. Mol. Biol. 400, 186-203 (2010). [46] A.S. Petrov and S.C. Harvey, Structural and thermody- namic principles of viral packaging, Structure 15, 21-27 (2007). [47] L.C. Tu and S.M. Musser, Single molecule studies of nucleoplasmic transport, Biochim. Biophys. Acta 1813, 1607-1618 (2011). [48] T.A. Rapoport, Protein transport across the endoplas- mic reticulum membrane, FEBS J. 275, 4471-4478 (2008). [49] W. Neupert and J.M. Herrmann, Translocation of pro- teins into mitochondria, Annu. Rev. Biochem. 76, 723- 749 (2007). [50] H. Aronson and P. Jarvis, The chloroplast protein im- port apparatus, its components, and their roles, in: Plant Cell Monogr. (Springer 2008). [51] S.J. Gould and C.S. Collins, Peroxisomal- protein im- port: is it really that complex?, Nat. Rev. Mol. Cell Biol. 3, 382-389 (2002). [52] A. Pingoud, M. Fuxreiter, V. Pingoud and W. Wende, Type II restriction endonucleases: structure and mech- anism, Cell. Mol. Life Sci. 62, 685-707 (2005). [53] K. Buttner, K. Wenig and K.P. Hopfner, The exosome: a macromolecular cage for controlled RNA degradation, Mol. Microbiol. 61, 1372-1379 (2006). [54] E. Lorentzen and E. Conti, The exosome and the pro- teasome: nano-compartments for degradation, Cell 125, 651-654 (2006). [55] A.M. Smith, S.C. Zeeman and S.M. Smith, Starch degra- dation, Annu. Rev. Plant Biol. 56, 73-98 (2005). [56] E.A. Bayer, J.P. Belaich, Y. Shoham and R. Lamed, The cellulosomes: multienzyme machines for degradation of plant cell wall polysaccharides, Annu. Rev. Microbiol. 58, 521-554 (2004). [57] L. Duo-Chuan, Review of fungal chitinases, Myco- pathologia 161, 345-360 (2006). [58] A. Kornberg and T. Baker, DNA replication, 2nd edn. (W.H. Freeman and Co., New York, 1992). [59] L. Bai, T.J. Santangelo and M.D. Wang, Single-molecule analysis of RNA polymerase transcription, Annu. Rev. Biophys. Biomol. Struct. 35, 343-360 (2006). [60] K.M. Herbert, W.J. Greenleaf and S.M. Block, Single- molecule studies of RNA polymerase: motoring along, Annu. Rev. Biochem. 77, 149-176 (2008). [61] R. Kornberg, The molecular basis of eukaryotic tran- scription, (Nobel lectures by R. Kornberg), Angewandte Chemie Int. ed. 46, 6957-6965 (2007). [62] A. Herschhorn and A. Hizi, Retroviral reverse transcrip- tase, Cell. Mol. Life Sci. 67, 2717-2747 (2010). [63] J.N. Barr and R. Fearns How RNA viruses maintain their genome integrity, J. Gen. Virol. 91, 1373-1387 (2010). [64] A.S. Spirin, Ribosomes, (Kluwer Academic/Plenum, 1999). [42] B. Burton and D. Dubnau, Membrane-associated DNA [65] J. Frank and R.L. Gonzalez Jr., Structure and dynam- ics of a processive Brownian motor: the translating ri- bosome, Annu. Rev. Biochem. 79, 381-412 (2010). [66] See the nobel lectures by A. Yonath, V. Ramakrishnan and T.A. Steitz in Angewandte Chemie International Ed. 49, 4340-4398 (2010). [67] C.R. Clapier and B.R. Cairns, Annu. Rev. Biochem. 78, 273 (2009). [68] A.M. Pyle, Translocation and unwinding mechanisms of RNA and DNA helicases, Annu. Rev. Biophys. 37, 317- 336 (2008). [69] A. Garai, D. Chowdhury and M.D. Betterton, Two-state model for helicase translocation and unwinding of nu- cleic acids, Phys. Rev. E 77, 061910 (2008). [70] H.C. Berg, E. coli in motion, (Springer, 2004). [71] Y. Sowa and R.M. Berry, Bacterial flagellar motor, Quart. Rev. Biophys. 41, 103-132 (2008). [72] G. Oster and H. Wang, How protein motors convert chemical energy into mechanical work, in: M. Schliwa, (ed.) Molecular Motors, (Wiley-VCH, 2003). [73] C. von Ballmoos, G.M. Cook and P. Dimroth, Unique rotary ATP synthase and its biological diversity, Annu. Rev. Biophys. 37, 43-64 (2008). [74] M.T. Valentine and S.P. Gilbert, To step or not to step? How biochemistry and mechanics influence processivity in kinesin and Eg5, Curr. Opin. Cell Biol. 19, 75-81 (2007). [75] J.R. Kardon and R.D. Vale, Regulators of the cytoplas- mic dynein motor, Nat. Rev. Mol. Cell Biol. 10, 854-865 (2009). [76] C. Indiani and M. O'Donnell, The replication clamp- loading machine at work in the three domains of life, Nat. Rev. Mol. Cell Biol. 7, 751-761 (2006). [77] S.M. Block, Kinesin motor mechanics: binding, step- ping, tracking, gating, and limping, Biophys. J. 92, 2986-2995 (2007). [78] R.T. Pomerantz and M. O'Donnell, What happens when replication and transcription complexes collide?, Cell Cycle 9, 2537-2543 (2010). [79] E.M. Purcell, Life at low Reynolds number, Am. J. Phys. 45, 3-11 (1977). [80] D'Arcy Thompson, On Growth and Form, reprinted 2nd edition (Cambridge University Press, 1963). [81] F. Julicher, Statistical physics of active processes in cells, Physica A 369, 185-200 (2006). [82] K. Sasaki, Molecular chemical engines: pseudo-static processes and the mechanism of energy transduction, J. Phys. Soc. Jap. 74, 2973-2980 (2005). [83] L. von Bertalanffy, The theory of open systems in physics and biology, Science 111, 23-29 (1950). [84] R.S. Berry, V.A. Kazanov, S. Sieniutycz, Z. Szwast and A.M. Tsirlin, Thermodynamic optimization of finite- time processes, (Wiley, 2000). [85] F.L. Curzon and B. Ahlborn, Efficiency of a Carnot engine at maximum power output, Amer. J. Phys. 43, 22-24 (1975). [86] J.M. Gordon and V.N. Orlov, Performance characteris- tics of endoreversible chemical engines, J. Appl. Phys. 74, 5303-5309 (1993). [87] A. Katchalsky and P.F. Curran, Nonequilibrium ther- modynamics in biphysics, (Harvard university press, 1967). [88] T. Schmiedl and U. Seifert, Efficiency at maximum power: an analytically solvable model for stochastic heat engines, EPL 81, 20003 (2008). 28 [89] H. Linke, M.T. Downton and M.J. Zuckermann, Perfor- mance characteristics of Brownian motors, Chaos 15, 026111 (2005). [90] A. Parmeggiani, F. Julicher, A. Ajdari and J. Prost, En- ergy transduction of isothermal ratchets: generic aspects and specific examples close to and far from equilibrium, Phys. Rev. E 60, 2127-2140 (1999). [91] I. Derenyi, M. Bier and R.D. Astumian, Generalized effi- ciency and its application to microscopic engines, Phys. Rev. Lett. 83, 903-906 (1999). [92] H. Wang and G. Oster, The Stokes efficiency for molec- ular motors and its applications, Europhys. Lett. 57, 134-140 (2002). [93] D. Suzuki and T. Munakata, Rectification efficiency of a Brownian motor, Phys. Rev. E 68, 021906 (2003). [94] F. Julicher, A. Ajdari and J. Prost, Modeling molecular motors, Rev. Mod. Phys. 69, 1269-1281 (1997). [95] R.D. Astumian, Design principles of Brownian molecu- lar machines: how to swim in molasses and walk in a hurricane, Phys.Chem.Phys. 7, 5067-5083 (2007). [96] P. Reimann, Brownian motors: noisy transport far from equilibrium, Phys. Rep. 361, 57-265 (2002). [97] R.P. Feynman, R.B. Leighton and M. Sands, The Feyn- man lectures on physics, vol.I (Basic Books, 2011). [98] R.D. Vale and F. Oosawa, Protein motors and Maxwell's demons: does mechanochemical transduction involve a thermal ratchet?, Adv. Biophys. 26, 97-134 (1990). [99] J. Howard, Protein power strokes, Curr. Biol. 16, R517- R519 (2006). [100] H. Wang and G. Oster, Ratchets, power strokes and molecular motors, Appl. Phys. A 75, 315-323 (2002). [101] H. Leff and A.F. Rex, (eds.) Maxwell's demon 2: en- tropy, classic and quantum information, computing, (Taylor and Francis, 2002). [102] Y. Okada and N. Hirokawa, Mechanism of single-headed processivity: diffusional anchoring between the K-loop of kinesin and the C terminus of tubulin, PNAS 97, 640- 645 (2000). [103] K. Nishinari, Y. Okada, A. Schadschneider and D. Chowdhury, Intracellular transport of single-headed molecular motors KIF1A, Phys. Rev. Lett. 95, 118101 (2005). [104] P. Greulich, A. Garai, K. Nishinari, A. Schadschnei- der and D. Chowdhury, Intra-cellular transport by single-headed kinesin KIF1A: effects of single-motor mechanochemistry and steric interactions, Phys. Rev. E, 75, 041905 (2007). [105] A.F. Huxley, Muscle structure and theories of contrac- tion, in: Prog. Biophys. Biophys. Chemistry, 7, eds. J.A.V. Butler and B. Katz, (Pergamon Press, 1957). [106] N.J. Cordova, B. Ermentrout and G.F. Oster, Dynamics the thermal ratchet model, of single-motor molecules: PNAS 89, 339-343 (1992). [107] C.S. Peskin, G.M. Odell and G.F. Oster, Cellular mo- the Brownian ratchet, tions and thermal fluctuations: Biophys. J. 65, 316-324 (1993). [108] S.M. Simon, C.S. Peskin and G.F. Oster, What drives the translocation of proteins?, PNAS 89, 3770-3774 (1992). [109] M. Stewart, Ratcheting mRNA out of the nucleus, Mol. Cell 25, 327-330 (2007). [110] G. Bar-Nahum, V. Epshtein, A.E. Ruckenstein, R. Rafikov, A. Mustaev and E. Nudler, A ratchet mecha- nism of transcription elongation and control, Cell 120, 183-193 (2005). [111] Q. Guo and R. Sousa, Translocation by T7 RNA poly- merase: a sensitively poised Brownian ratchet, J. Mol. Biol. 358, 241-254 (2006). [112] H. Wang and T.C. Elston, Mathematical and computa- tional methods for studying energy transduction in pro- tein motors, J. Stat. Phys. 128, 35-76 (2007). [113] R. Lipowsky and S. Liepelt, Chemomechanical coupling of molecular motors: thermodynamics, network repre- sentations, and balance conditions, J. Stat. Phys. 130, 39-67 (2008). [114] N. G. Van Kampen, Stochastic processes in physics and chemistry, (Elsevier, 2007). [115] H. Wang, Chemical and mechanical efficiencies of molecular motors and implications for motor mech- anism, J. Phys. Condens. Matter 17, S3997-S4014 (2005). [116] R. D. Astumian, Biasing the random walk of a molecu- lar motor, J. Phys. Condens. Matter, 17, S3753-S3766 (2005). [117] N. Thomas, Y. Imafuku and K. Tawada, Molcular mo- tors: thermodynamics and the random walk, Proc. Roy. Soc. Lond. B 268, 2113-2122 (2001). [118] Z. Koza, General technique of calculating the drift ve- locity and diffusion coefficient in arbitrary periodic sys- tems, J. Phys. A 32, 7637-7651 (1999). [119] Z. Koza, Diffusion coefficient and drift velocity in peri- odic media, Physica A 285, 176-186 (2000). [120] Y.R. Chemla, J.R. Moffitt and C. Bustamante, J. Phys. Chem. B 112, 6025 (2008). [121] A. Garai, D. Chowdhury, D. Chowdhury and T.V. Ra- makrishnan, Stochastic kinetics of ribosomes: Single motor properties and collective behavior, Phys. Rev. E 80, 011908 (2009). [122] Y. Kafri, D.K. Lubensky and D.R. Nelson, Dynamics of molecular motors and polymer translocation with se- quence heterogeneity, Biophys. J. 86, 3373 (2004). [123] D. Chowdhury, L. Santen and A. Schadschneider, Sta- tistical physics of vehicular traffic and some related sys- tems Phys. Rep. 329 199-329 (2000). [124] D. Chowdhury, A. Schadschneider and K. Nishinari, Physics of Transport and Traffic Phenomena in Biology: from molecular motors and cells to organisms, Phys. of Life Rev. 2 318-352 (2005). [125] A. Schadschneider, D. Chowdhury and K. Nishinari, Stochastic transport in complex systems: from molecules to vehicles (Elsevier, 2010). [126] R. Lipowsky, Y. Chai, S. Klumpp, S. Liepelt and M.J.I. Muller, Physica A 372, 34 (2006). [127] A. Parmeggiani, T. Franosch and E. Frey, Totally asym- metric simple exclusion process with Langmuir kinetics, Phys. Rev. E 70, 046101 (2004). 29 ological transport, Rep. Prog. Phys. 74, 116601 (2011). [133] T. von der Haar, Mathematical and computational mod- eling of ribosomal movement and protein synthesis: an overview, Comp. Struct. Biotechnol. J. 1, e201204002 (2012). [134] A. Basu and D. Chowdhury, Traffic of interacting ribo- somes: effects of single-machine mechano-chemistry on protein synthesis, Phys. Rev. E 75, 021902 (2007). [135] A.K. Sharma and D. Chowdhury, Stochastic theory of protein synthesis and polysome: Ribosome profile on a single mRNA transcript, J. Theor. Biol. 289, 36 (2011). [136] T. Tripathi and D. Chowdhury, Interacting RNA poly- merase motors on a DNA track: effects of traffic conges- tion and intrinsic noise on RNA synthesis, Phys. Rev. E 77, 011921 (2008). [137] S. Klumpp, Pausing and backtracking in transcription under dense traffic conditions, J. Stat. Phys. 142, 1252- 1267 (2011). [138] K.E.P. Sugden, M.R. Evans, W.C.K. Poon and N.D. Read, Model of hyphal tip growth involving microtubule- based transport, Phys. Rev. E 75, 031909 (2007). [139] M. Schmitt and H. Stark, Modelling bacterial flagellar growth, EPL 96, 28001 (2011). [140] L. Reese, A. Melbinger and E. Frey, Crowding of molec- ular motors determines microtubule depolymerization, Biophys. J. 101, 2190-2200 (2011). [141] J.J. Hopfield, Kinetic proofreading: a new mechanism for reducing errors in biosynthesis process requiring high specificity, PNAS 71, 4135-4139 (1974). [142] J. Ninio, Kinetic amplification of enzyme discrimina- tion, biochimie 57, 587-595 (1975). [143] M. Yarus, Proofreading, NTPases and translation: con- straints on accurate biochemistry, Trends in Biochem. Sci. 17, 130-133 (1992). [144] F. Oosawa, The loose coupling mechanism in molecular machines of living cells, Genes to Cells 5, 9-16 (2000). [145] A.K. Sharma and D. Chowdhury , Quality control by a mobile molecular workshop: quality versus quantity, Phys. Rev. E 82, 031912 (2010). [146] A.K. Sharma and D. Chowdhury, Distribution of dwell times of a ribosome: effects of infidelity, kinetic proof- reading and ribosome crowding, Phys. Biol. 8, 026005 (2011). [147] A. K. Sharma and D. Chowdhury, Error correction during DNA replication: DNAP as Dr. Jekyll-and-Mr. Hyde, (2012) arXiv:1112.3276 [148] J. R. Moffitt, Y.R. Chemla and C. Bustamante, Methods in statistical kinetics, Methods in Enzymology 475, 221- 257 (2010). [149] J. W. Shaevitz, S. M. Block and M. J. Schnitzer, Statis- tical kinetics of macromolecular dynamics, Biophys. J. 89, 2277-2285 (2005). [128] V. Popkov, A. Rakos, R.D. Willmann, A.B. Kolomeisky [150] J.C. Liao, J.A. Spudich, D. Parker, S.L. Delp, PNAS and G.M. Schutz, Phys. Rev. E 67, 066117 (2003). 104, 3171 (2007). [129] M.R. Evans, R. Juhasz and L. Santen, Phys. Rev. E 68, 026117 (2003). [130] C. MacDonald, J. Gibbs and A. Pipkin, Kinetics of biopolymerization on nucleic acid templates, Biopoly- mers, 6, 1-25 (1968). [131] C. MacDonald and J. Gibbs, Concerning the kinetics of polypeptide synthesis on polyribosomes, Biopolymers, 7, 707-725 (1969). [132] T. Chou, K. Mallick and R.K.P. Zia, Non-equilibrium statistical mechanics: from a paradigmatic model to bi- [151] M. Linden and M. Wallin, Biophys. J. 92, 3804 (2007). [152] S. Redner, A guide to first-passage processes, (Cam- bridge University Press, 2001). [153] T. Tripathi, G. M. Schutz and D. Chowdhury , RNA polymerase motors: dwell time distribution, velocity and dynamical phases, J. Stat. Mech.: Theory and Experi- ment, P08018 (2009). [154] M.J. Schnitzer and S.M. Block, Statistical kinetics of processive enzymes, Cold Spring Harbor Symp. Quanti- tative Biol. LX, 793-802 (1995). 30 [155] U. Fano, Ionization yield of radiations.II: the fluctua- tions of the number of ions, Phys. Rev. 72, 26-29 (1947). [156] D. Tsygankov, M. Linden, M.E. Fisher, Back-stepping, hidden substeps, and conditional dwell times in molecu- lar motors, Phys Rev E. i75, 021909 (2007). [157] A. Garai and D. Chowdhury, Stochastic kinetics of a single-headed motor protein: dwell time distribution of KIF1A, EPL 93, 58004 (2011). [158] R.M. Glaeser, Cryo-electron microscopy of biological [178] M.A. Beaumont and B. Rannala, The Bayesian revolu- tion in genetics, Nat. Rev. Genet. 5, 251-261 (2004). [179] A. Golightly and D.J. Wilkinson, Bayesian parameter inference for stochastic biochemical network models us- ing particle Markov chain Monte Carlo, JRS Interface Focus 1, 807-820 (2011). [180] J.K. Kruschke, What to believe: Bayesian methods for data analysis, Trends in Cognitive Sci. 14, 293-300 (2010). nanostructures, Phys. Today January 48-54 (2008). [181] A.M. Ellison, Bayesian inference in ecology, Ecology [159] J. Zlatanova and K. van Holde, Single-molecule biology: what is it and how does it work?, Mol. Cell 24, 317-329 (2006). [160] F. Ritort, Single-molecule experiments in biological physics: methods and aplications, J. Phys. Cond. Matt. 18, R531-R583 (2006). [161] P.V. Cornish and T. Ha, A survey of single-molecule techniques in chemical biology, ACS chemical biology, 2, 53-61 (2007). [162] A. N. Kapanidis and T. Strick, Biology, one molecule at a time, Trends in Biochem. Sci. 34, 234-243 (2009). [163] J.R. Moffitt, Y.R. Chemla, S.B. Smith and C. Busta- mante, Recent advances in optical tweezers, Annu. Rev. Biochem. 77, 205-228 (2008). [164] W. J. Greenleaf, M.T. Woodside and S.M. Block, High- resolution, single-molecule measurements of biomolecu- lar motion, Annu. Rev. Biophys. Biomol. Str. 36, 171- 190 (2007). [165] L. Schermelleh, R. Heintzmann and H. Leonhardt, A guide to super-resolution fluorescence microscopy, J. Cell Biol. 190, 165-175 (2010). [166] D. Balding, Inference in complex systems, JRS Interface Focus 1, 805-806 (2011); see also the other articles in the same special issue of JRS Interface Focus on statistical inference. [167] D.B. kell and S.G. Oliver, Here is the evidence, now what is the hypothesis? The complementary roles of in- ductive and hypoesis-driven science in the post-genomic era, Bioessays 26, 99-105 (2003). [168] G. Cowan, Data analysis: frequently Bayesian, Phys. Today 60, 82-83 (2007). [169] M. Andrec, R.M. Levy and D.S. Talaga, Direct determi- nation of kinetic rates from single-molecule photon ar- rival trajectories using hidden Markov models, J. Phys. Chem. 107, 7454-7464 (2003). [170] I.J. Myung, Tutorial on maximum likelihood estimation, J. Math. Psychol. 47, 90-100 (2003). [171] D.E. Raeside, A physicist's introduction to Bayesian statistics, Amer. J. Phys. 40, 688-694 (1972). [172] D.E. Raeside, A physicist's introduction to Bayesian statistics. II, Amer. J. Phys. 40, 1130-1133 (1972). [173] T.J. Ulrych, M.D. Sacchi and A. Woodbury, A Bayes tour of inversion: a tutorial, Geophys. 66, 55-69 (2001). [174] J.A. Scales and L. Tenorio, Prior information and un- certainty in inverse problems, Geophys. 66, 389-397 (2001). Lett. 7, 509-520 (2004). [182] F.G. Ball and J.A. Rice, Stochastic models of ion chan- nels: introduction and bibliography, Mathematical Bio- sciences 112, 189-206 (1992). [183] A.G. Hawkes, Stochastic modelling of single ion chan- nels, in: Computational neuroscience: a comprehensive approach, (CRC press, 2004). [184] S.R. Eddy, What is a hidden markov model?, Nat. Biotechnol. 22, 1315-1316 (2004). [185] L. R. Rabiner, A tutorial on hidden markov models and selected applications in speech recognition, Proc. IEEE 77, 257-286 (1989). [186] D.S. Talaga, Markov processes in single molecule flu- orescence, Curr. Opin. Colloids and Interface Sci. 12, 285-296 (2007). [187] C. Vogl and A. Futschik, Hidden Markov models in bi- ology, in:Data mining techniques for the life sciences, (Humana Press, 2010). [188] S.A. McKinney, C. Joo and T. Ha, Analysis of single- molecule FRET trajectories using hidden Markov mod- eling, Biophys. J. 91, 1941-1951 (2006). [189] T. H. Lee, Extracting kinetics information from single- molecule Fluorescence Resonance Energy Transfer data using Hidden Markov Models, J. Phys. Chem. 113, 11535-11542 (2009). [190] F.E. Mullner, S. Syed, P.R. Selvin and F.J. Sigworth, Improved hidden markov models for molecular motors, Part 1: basic theory, Biophys. J. 99, 3684-3695 (2010). [191] S. Syed, F.E. Mullner, P.R. Selvin and F.J. Sigworth, Improved hidden markov models for molecular motors, Part 2: extensions and application to experimental data, Biophys. J. 99, 3684-3695 (2010). [192] D.A. Smith, W. Steffen, R.M. Simmons and J. Sleep, Hidden Markov methods for the analysis of single- molecule actomyosin displacement data: the variance- hidden-Markov method, Biophys. J. 81, 2795-2816 (2001). [193] H. Wang, Motor potential profile and a robust method for extracting it from time series of motor positions, J. Theor. Biol. 242, 908-921 (2006). [194] R. D. Mullins, Cytoskeletal mechanisms for breaking cellular symmetry, Cold Spring Harb. Persp. Biol. 2, a003392 (2009). [195] M. Kirschner, J. Gerhart and T. Mitchison, Molecular vitalism, Cell 100, 79-88 (2000). [196] R. Li and B. Bowerman, Symmetry breaking in biology, [175] S.R. Eddy, What is Bayesian statistics, Nat. Biotechnol. Cold Spring Harb. Perspect. Biol. 2: a003475 (2010). 22, 1177-1178 (2004). [176] S.C. Kou, X.S. Xie and J.S. Liu, Bayesian analysis of single-molecule experimental data, Apl. Statistics 54, 469-506 (2005). [177] J.S. Shoemaker, I.S. Painter and B.S. Weir, Bayesian statistics in genetics: a guide for the uninitiated, Trends in Genet. 15, 354-358 (1999). [197] Special "Perspectives on Symmetry Breaking in Biol- ogy", Cold Spring Harb. Perspect. Biol. (2010). [198] W. F. Marshall, Cellular length control systems, Annu. Rev. Cell Dev. Biol. 20, 677-693 (2004). [199] W.F. Marshall, Origins of cellular geometry, BMC Biol. 9, 57 (2011). [200] T. J. Mitchison, Self-organization of polymer-motor sys- 31 tems in the cytoskeleton, Phil. Trans. Roy. Soc. Lond. B 336, 99-106 (1992). 1907). [225] J. O. de la Mettrie, L'homme Machine ('man a ma- [201] T. Misteli, The concept of self-organization in cellular chine'), (Open Court, La Salle, Illinois, 1912). architecture, J. Cell Biol. 155, 181-185 (2001). [202] E. Karsenti, Self-organization in cell biology: a brief his- tory, Nat. Rev. Mol. Cell Biol. 9, 255-262 (2008). [203] P. Godfrey-Smith, Information in biology, in: The Cam- bridge Companion to the Philosophy of Biology, eds. D. Hull and M. Ruse (Cambridge University Press, 2006). [204] H. Noll, The digital origin of human language- a syn- thesis, Bioessays 25, 489-500 (2003). [205] C. H. Bennett, The thermodynamics of computation- a review, Int. J. Theor. Phys. 21, 905-940 (1982). [206] C.H. Bennett, Notes on Landauer's principle, reversible computation, and Maxwell's Demon, Stud. Hist. Phil. Mod. Phys. 34, 501-510 (2003). [207] M.B. Plenio and V. Vitteli, The physics of forgetting: Landauer's erasure principle and information theory, Contemp. Phys. 42, 25-60 (2001). [208] C. H. Bennett, Dissipation-error tradeoff in proofread- ing, Biosystems 11, 85-91 (1979). [209] S.C. Kou, B.J. Cherayil, B.P. English and X.S. Xie, Single-molecule Michaelis-Menten equations, J. Phys. Chem. B 109, 19068-19081 (2005). [210] S. C. Kou, Stochatic networks in nanoscale biophysics: modeling enzymatic reaction of a single protein, J. Amer. Stat. Assoc. 103, 961-975 (2008). [211] K.A. Johnson and R.S. Goody, The original Michaelis constant: translation of the 1913 Michaelis-Menten pa- per, Biochem. 50, 8264-8269 (2011) (see the supporting information for the english translation of the original 1913 MM paper). [212] D. Bray and T. Duke, Conformational spread: the prop- agation of allosteric states in large multiprotein com- plexes, Annu. Rev. Biophys. Biomol. Struct. 33, 53-73 (2004). [213] A. Vologodskii, Energy transformation in biological molecular motors, Phys. of Life Rev. 3, 119-132 (2006). [214] A. Whitty, Cooperativity and biological complexity, Nat. [226] G.W. Leibnitz, The Philosophical works of Leibnitz, (Tanslated from original latin and french), with notes by G. M. Duncan (Tuttle, Morehouse, & Taylor, pub- lishers, 1890). [227] J.E.H. Smith, Divine machines: Leibnitz and the science of life (Princeton univ. press, 2011). [228] J.E.H. Smith and O. Nachtomy (eds.), Machines of nature and corporeal substances in Leibnitz (Springer, 2011). [229] M. Piccolino, Biological machines: from mills to molecules, Nature Rev. Mol. Cell Biol. 1, 149-153 (2000). [230] D. Ferrier, Life and vital energy considered in relation to physiology and medicine, The British Medical Journal, Oct. 22, 429-432 (1870). [231] R.H. Thurston, The animal as a machine and a prime motor, and the laws of energetics, (John Wiley & Sons, 1894). [232] C. Darwin and F. Darwin, The power of movement in plants (D. Appleton and Co., New York, 1897). [233] J. C. Bose, Plant response as a means of physiological investigation, (Longmans, Green, and Co., 1906). [234] M. Verworn, Irritability: a physiological analysis of the general effect of stimuli in living substance, (Yale Univ. Press, 1913). [235] H. W. Conn, The story of life's mechanism, (George Newnes, London, 1899). [236] J. Loeb, The Dynamics of Living Matter, (Columbia University Press, 1906). [237] E.B. Wilson, The Physical Basis of Life (1925). [238] A.V. Hill, Living machinery (Harcourt, Brace and Co., New York, 1927). [239] D.M. Needham, Machina Carnis: the biochemistry of muscular contraction in its historical development, (Cambridge University Press, 1971). [240] J.D. Bernal, The Physical Basis of Life, Proc. Phys. Chem. Biol. 4, 435-439 (2008). Soc. B 62, 597-618 (1949). [215] R. P. Feynman 1959, included also in: The pleasure of finding things out, (Perseus Books, Cambridge, Mas- sachusetts, 1999), chapter 5. [216] A. Junk and F. Riess, From an idea to a vision: there's plenty of room at the bottom, Am. J. Phys. 74, 825-830 (2006). [217] D.S. Goodsell, Bionanotechnolgy: lessons from Nature, (Wiley, 2004). [218] M.G. van der Heuvel and C. Dekker, Motor proteins at work for nanotechnology, Science 317, 333-336 (2007). [219] V. Balzani, M. Venturi and A. Credi, Molecular devices and machines: a journey into the nano-world (Wiley- VCH, 2003). [220] Y. Bar-Cohen, ed. Biomimetics: biologically inspired technologies (Taylor and Francies, 2005). [221] M. Sarikaya, C. Tamerler, A. K. -Y. Jen, K. Schulten and F. Baneyx, Molecular biomimetics: nanotechnology through biology, Nat. Materials, 2 577-585 (2003). [222] W.R. Browne and B.L. Feringa, Making molecular ma- chines work, Nat. Nanotechnol. 1, 25-35 (2006). [223] A. Szent-Gyorgyi, The development of bioenergetics, Bioenergetics 3, 1-4 (1972). [241] F. O. Smitt, Molecular biology and the physical basis of life processes, Rev. Mod. Phys. 31, 5-10 (1959). [242] C. Mavroidis, A. Dubey and M.L. Yarmush, Molecular Machines, in: Annual Rev. Biomed. Engg., 6, 363-395 (2004). [243] B. Alberts, The cell as a collection of protein machines: preparing the next generation of molecular biologists, Cell 92(3), 291-294 (1998). [244] D'Arcy W. Thompson, On Aristotle as a Biologist, (Clarendon Press, Oxford, 1913). [245] D.E. Johnson, C.L. Ferrell and T.G. Jenkins, The his- tory of energetic efficiency research: where have we been and where are we going?, J. Anim. Sci. 81, E27-E38 (2003). [246] P. Medawar, A biological retrospect, Nature 207, 1327- 1330 (1965). [247] F. Tama and C.L. Brooks III, Symmetry, form, and shape: guiding principles for robustness in macromolec- ular machines, Annu. Rev. Biophys. Biomol. Struct. 35, 115-133 (2006) (see also chapter 4 in: J. Frank, (ed.) Molecular machines in biology: workshop of the cell (Cambridge University Press, 2011)). [224] Aristotle, De Anima, (with translation, introduction and notes) by R.D. Hicks (Cambridge University Press, [248] J. Ma, Usefulness and limitations of normal mode anal- ysis in modeling dynamics of biomolecular complexes, Structure 13, 373-380 (2005). 148, 754-759 (1965). [249] E.C. Dykeman and O.F. Sanky, Normal mode analysis and applications in biological physics, J. Phys. Condens. Matter 22, 43202 (2010). [250] L.H. Hartwell, J.J. Hopfield, S. Leibler and A.W. Mur- ray, From molecular to modular cell biology, Nature 402, C47-C52 (1999). [251] J.R. Platt, Strong inference, Science 146, 347-353 (1964). [252] T.C. Chamberlin, The method of multiple working hy- potheses, Science 15, 92-96 (1890); reprinted in Science [253] D.A. Beard and M.J. Kushmerick, Strong inference for systems biology, PLoS Comp. Biol. 5, e1000459 (2009). [254] J. Bronson, J. Fei, J.M. Hofman, R.L. Gonzales Jr., C.H. Wiggins, Learning rates and states from biophysi- cal time series: a Bayesian approach to model selection and single-molecule FRET data, Biophys. J. 97, 3196- 3205 (2009). [255] D. Chowdhury, in preparation (to be published). 32
1704.01655
1
1704
2017-04-05T21:33:42
Quantifying protein densities on cell membranes using super-resolution optical fluctuation imaging
[ "physics.bio-ph" ]
Surface molecules, distributed in diverse patterns and clusters on cell membranes, influence vital functions of living cells. It is therefore important to understand their molecular surface organisation under different physiological and pathological conditions. Here, we present a model-free, quantitative method to determine the distribution of cell surface molecules based on TIRF illumination and super-resolution optical fluctuation imaging (SOFI). This SOFI-based approach is robust towards single emitter multiple-blinking events, high labelling densities and high blinking rates. In SOFI, the molecular density is not based on counting events, but results as an intrinsic property due to the correlation of the intensity fluctuations. The effectiveness and robustness of the method was validated using simulated data, as well as experimental data investigating the impact of palmitoylation on CD4 protein nanoscale distribution in the plasma membrane of resting T cells.
physics.bio-ph
physics
1 Quantifying protein densities on cell membranes using super-resolution optical fluctuation imaging Tomáš Lukeš1,2 Daniela Glatzová3,4, Zuzana Kvíčalová3, Florian Levet5,6, Aleš Benda3,7,Tomáš Brdička4, Theo Lasser1* & Marek Cebecauer3* 1Laboratoire d'Optique Biomédicale, Ecole Polytechnique Fédérale de Lausanne, Lausanne, Switzerland 2Department of Radioelectronics, FEE, Czech Technical University in Prague, Prague, Czech Republic 3J. Heyrovsky Institute of Physical Chemistry, Czech Academy of Sciences, Prague, Czech Republic 4Institute of Molecular Genetics, Czech Academy of Sciences, Prague, Czech Republic 5Interdisciplinary Institute for Neuroscience, UMR 5297 CNRS, Université de Bordeaux, Bordeaux, France 6Bordeaux Imaging Center, UMS 3420 CNRS, Université de Bordeaux, US4 INSERM, Bordeaux, France 7Imaging methods core facility, BIOCEV, Vestec u Prahy, Czech Republic *Corresponding authors Abstract Surface molecules, distributed in diverse patterns and clusters on cell membranes, influence vital functions of living cells. It is therefore important to understand their molecular surface organisation under different physiological and pathological conditions. Here, we present a model-free, quantitative method to determine the distribution of cell surface molecules based on TIRF illumination and super-resolution optical fluctuation imaging (SOFI). This SOFI-based approach is robust towards single emitter multiple-blinking events, high labelling densities and high blinking rates. In SOFI, the molecular density is not based on counting events, but results as an intrinsic property due to the correlation of the intensity fluctuations. The effectiveness and robustness of the method was validated using simulated data, as well as experimental data investigating the impact of palmitoylation on CD4 protein nanoscale distribution in the plasma membrane of resting T cells. 1. Introduction Numerous cellular functions are controlled by molecules at the cell surface among which proteins form the largest pool. An growing body of evidence supports the hypothesis that plasma membrane proteins are not distributed homogeneously but rather in complexes, clusters and other higher order patterns1. It has been experimentally demonstrated that these protein clusters are involved in the regulation of signal transduction and other vital cell processes2. This has been the driving motivation to develop a robust method for investigation of molecular organization at the plasma membrane under various conditions. The size of protein assemblies varies and is frequently smaller than 200 nm in diameter which is below the resolution limit of classical fluorescence microscopy. During the last two decades, super-resolution techniques have been developed which overcome the diffraction limit3,4 and provide a detailed view of structures smaller than 200 nm. Single molecule localization microscopy (SMLM) has been frequently used to characterize membrane protein assemblies5 -- 9. SMLM techniques such as photoactivated localization microscopy (PALM)10 and stochastic optical reconstruction microscopy (STORM)11 rely on temporal discrimination of otherwise spatially overlapping fluorophore images. In sequences of at least several thousand images, the position of fluorescent markers is determined by fitting a model function to the imaged point spread functions (PSFs). In high density samples, this fitting procedure may meet its limit leading to under-counting errors with significant localization errors for overlapping molecules. The stochastic blinking behaviour of of fluorophores may result in multiple localizations from single molecules12. High photoswitching rates in combination with high emitter densities can give rise to the appearance of artificial clusters13. These limitations may compromise the quantification of densely packed proteins. Characterization of protein clusters becomes a challenge because current methods for cluster analysis6 -- 9,13 -- 16 rely both on difficult-to-model photophysical properties and on acquisition parameters of the SMLM data. In this work, we readdress these problems with a novel approach based on SOFI and present an innovative and general method 2 to study molecular distribution on cell membranes which overcomes the aforementioned limitations. SOFI is an optical super-resolution technique which exploits the spatio-temporal photon traces created by stochastically blinking fluorophores. SOFI disentangles the overlapping PSFs by employing higher order statistics. The strong temporal cross-correlation over several neighbouring pixels is the underlying cause of SOFI super-resolution17,18. The achieved resolution improvement results from the properties of spatio-temporal cross-cumulants calculated from the entire image sequence of 2D17 or 3D images19. SOFI can be used to analyse SMLM data, but tolerates much higher emitter densities20,21. Balanced SOFI (bSOFI) combines the information content of several cumulant orders in a system of linear equations allowing one to extract physically meaningful parameters such as brightness, emitter density and the on-time ratio of the blinking emitters22. Therefore molecular density appears as a calculated parameter based on the full image sequence. Multiple blinking of individual fluorophores improves the bSOFI signal and therefore the accuracy of these statistically estimated parameters. In addition, bSOFI suppresses uncorrelated noise22, which leads to improved image contrast. We exploited these features of bSOFI and present here a novel SOFI-based quantitative assessment of protein distributions, resulting in protein density maps. In particular, we investigated the impact of palmitoylation on CD4 nanoscale organization at the surface of resting T cells. 2. Results For quantifying the protein distribution in the plasma membrane of T-cells, we acquired image sequences with a total internal reflection fluorescence (TIRF) microscope equipped with an EMCCD camera to detect the fluorescence originating from individual fluorescent emitters (see Methods). The proteins of interest were labelled with adequate blinking fluorophores i.e. emitters cycling between dark/bright states. 2.1 Molecular density analysis The algorithm work flow is shown in Fig. 1. All acquired image sequences are first drift corrected with sub-pixel precision. Using ThunderSTORM23, we measured lateral drift using fluorescent beads (fiducial markers) present in the images. These drift corrected image sequences were then processed by our bSOFI algorithm using 2nd, 3rd and 4th order cumulant analysis (see Methods). We extracted molecular density maps by combining the cumulant images in a system of linear equations, (see Methods). Fig. 2 shows a data processing example (molecular density analysis) for a single cell. As shown previously21, the accuracy of the density calculation is mainly determined by the size of the input image sequence. We acquired image sequences of 5000 frames for each dataset, choosing the number of frames by analysing the signal to noise ratio (SNR)21. We further evaluated the density maps generated by bSOFI by systematically increasing the density threshold (see Methods). Starting with a low threshold, large regions with a low average density are segmented. Increasing the threshold step by step allows precise density quantification (see Supplementary Fig. 1). We analysed each region of interest (ROI) by calculating, for each density threshold, the average number and area of high density regions (HDRs), as well as the relative area occupied by the HDRs. The averaged data across all cells for each protein variant over the range of density thresholds is shown in (Fig. 3a-c). This analysis provides an overview of HDR parameters in relation to the density threshold, unravelling the overall clustering behaviour of the cell samples under study. Inset images in Fig. 3a and Supplementary Fig. 1 indicate how the density threshold affects the detection of HDRs. Detailed statistics of the quantitative molecular density data can be further presented for the optimal density threshold (Fig. 3d; the threshold determined from simulations, see Supplementary Fig. 2) or any other threshold selected, for example, based on biological reasoning. When calculating the 4th order SOFI image, the pixel size of the resulting SOFI density map is 26.25 nm. According to the Shannon-Nyquist sampling theorem, the smallest detectable HDR would have a diameter of 52.5 nm. The density analysis can distinguish differences in HDR diameters in increments of 26.25 nm. Higher resolution could be possible with higher order SOFI images at the expense of more input images, i.e. longer acquisition times. The simulations indicate good performance of the analysis across a broad range of HDR densities (500 -- 3000 mol/µm) and HDR to background ratios (20 -- 100). Accuracy of HDR detection increases with increasing HDR to background ratio (see Methods and Supplementary Fig. 3). 3 2.2 Protein nanoscale organization We expressed four different mutant variants of mEOS2-labelled CD4 and analysed individual protein distributions on the plasma membrane of resting T cells immobilised on poly-L-lysine coated glass coverslips. Using TIRF microscopy, we imaged 20 cells for each CD4 variant (i.e. 80 in total) acquiring 5000 frames per cell. Tested mutants were native CD4 protein (WT), palmitoylation mutant (CS1) and truncated variants lacking the extracellular (dD1D4) and cytoplasmic (dCT) domains (Supplementary Fig. 4). Segmentation of SMLM data acquired for CD4-mEos2(WT) indicated the accumulation of native CD4 in HDRs with irregular shape, frequently forming networks (Fig. 2 and Supplementary Fig. 5). SMLM-based cluster analysis of these localisation data would be a challenge13. We therefore based our analysis on bSOFI imaging and extracted the corresponding density maps (see Methods). To minimize cell-size dependency and aiming for a true comparative protein density analysis among different CD4 variants, we selected a 3 x 3 µm region of interest (ROI) in each cell. Fig. 3 summarizes the quantitative data on CD4 membrane organization and indicates significant differences between the tested protein variants at the cell surface of resting T cells. As shown in Fig. 3d, native (WT) CD4 are organized in HDRs covering a large part of the plasma membrane as indicated in Roh et al.24. Such arrangements depend on the intact extracellular domain and palmitoylation of CD4 since mutants lacking these structures exhibit more random distribution with rare accumulation in a rather small HDRs (Fig. 3d). Truncation of the cytoplasmic domain had only a minor effect. The results presented in Fig. 3 point to the ability of our new method to identify HDRs with irregular shape and varying densities. The imaged cells exhibited a high level of intercellular variability, especially in case of the intermediate phenotype (CD4-dCT), and heterogeneity between HDRs identified within ROIs (see Supplementary Fig. 6). 3. Discussion In this work, we introduced a novel method for the characterisation of molecular organisation on cellular surfaces. Our quantitative analysis is based on SOFI, which provides several distinct advantages as applicable to densely populated regions (overlapping fluorescence emitters), no need for multiple blinking corrections, and inherent access to molecular density without a priori assumptions about the clustered molecules. Our approach does not require molecular localisation coordinates to calculate clustering properties of proteins (or other molecules) on the cell surface. Our algorithm provides quantitative molecular density analysis of membrane protein distributions independent of any user-defined parameters. We demonstrated the applicability of the proposed method by analysing the surface distribution of CD4 glycoprotein which forms large, dense, and interconnected regions on human T cells. Our molecular density analysis indicates the importance of the extracellular domain and of receptor palmitoylation for the organisation of CD4 on the plasma membrane. Our new method has the potential to be extended for various molecular density studies of surface molecules accessible for fluorescent labelling under physiological, pathological or pharmacological conditions. 4 4. Methods 4.1 Microscope setup A customized setup built on an inverted optical microscope (IX71, Olympus) was used for cell imaging. A 150 mW, 561 nm laser (Sapphire, Coherent) and a 100 mW 405 nm laser (Cube, Coherent) provided the excitation and activation, respectively. An acousto-optic tuneable filter (AOTFnC-400.650-TN, AA Optoelectronics) provided fast switching of the laser sources. Both lasers were combined and focused into the back focal plane of an objective (UApoN 100x, NA=1.49, Olympus). Total internal reflection was achieved with a commercial TIRF module (IX2-RFAEVA-2, Olympus) and the fluorescence emission was detected by an EMCCD camera (iXon DU-897, Andor). 1.1 Sample preparation Jurkat T cells in RPMI-1640 media (Sigma-Aldrich), complemented with glutamine and 10% foetal calf serum (Life Technologies) were grown in an incubator under controlled conditions of 37°C, 5% CO2, and 100% humidity. The cells were transiently transfected using the Neon® transfection system (Life Technologies). 1 µg of vector DNA per shot (3 pulses of 1325 V lasting for 10 ms) per 200,000 cells was used (see manufacturer's instructions). 25 mm diameter microscope coverslips were cleaned by incubation with 2% Hellmanex (Sigma-Aldrich) overnight at 42C and subsequently washed with MiliQ water. Prior to use, the coverslips were coated with poly-L-lysine (Sigma-Aldrich). Twenty hours after transfection, the cells were washed with PBS, resuspended in phenol red-free RPMI-1640 media (Sigma-Aldrich), supplemented with 10% foetal calf serum (Life Technologies), seeded on the poly-L-lysine coated coverslips, and incubated for 5 min at 37°C under 5% CO2. After a quick PBS wash the cells were fixed using 4% paraformaldehyde in PBS at 37 °C for 10 minutes under 5% CO2. After removal of excess liquid, the fixation was stopped with 0.1 M NH4Cl in PBS and the cells were washed with PBS. Finally, the coverslip was placed into a ChamLide holder for imaging. 1.2 Imaging Fixed cells were imaged in a 0.9% NaCl solution at room temperature. For monitoring drift, 200 nm gold beads (BBI international) were added to the sample. The mEos2 fluorophore was excited at 561 nm with power of ~30 mW and activated by a 405 nm laser with power of ~ 3 mW (both measured at the sample plane ). Cells were imaged with an EMCCD camera using an EM gain of 300 and an exposure time of 32 ms. 4.2 SOFI molecular density analysis The SOFI molecular density analysis can be subdivided into three distinct steps: Drift correction SOFI needs the sample to be immobile during image acquisition, and imaging beyond the diffraction limit demands drift correction. Tracking the positions of gold nanoparticles provides translational motion vectors in between consecutive frames. Registering consecutive frames with sub-pixel precision using bilinear interpolation was used for drift correction. Bleaching correction The drift corrected image sequence was sub-divided into sub-sequences of 500 frames each. These sub-sequences were chosen sufficiently short in order to minimize the influence of photobleaching19,21. In each subsequence, SOFI images of 2nd, 3rd and 4th order were calculated. These SOFI images were then averaged across all sub-sequences. Molecular density analysis The SOFI based molecular density analysis was programmed in MATLAB taking into account a linearization procedure as described in21. Combining SOFI images of different orders allows one to extract density maps (see Supplementary Note). Molecular density (i.e. number of emitters per pixel area) at pixel position r is given as 222222111221112212221212()(),3322332232(1)2()(32)2(32)gNKKKKKKKKKKKKKKKK=−−−−−−−−−−rrwhere 23132()()gKgµµ=rr, 24242()()gKgµµ=rr, 234,,gggrepresents cumulant images of 2nd, 3rd and 4th order, respectively. {()}nnVUµ=r, where {()}nVUr is the expectation value of the PSF (()nUr) of the nth order cumulant image. For more details, see Supplementary Note. Areas containing only background are removed using the bSOFI image as a mask. The threshold filtering procedure is described in more detail in Supplementary Figure 1. The algorithm loop through a 5 whole range of density levels presented in the sample starting with a threshold equal to zero and increasing the threshold step by step in each iteration. For each threshold, the data are further processed to acquire quantitative parameters describing regions with local protein density above the threshold i.e. high density regions (HDRs). The algorithm calculates the area, equivalent diameter, number of HDRs, and the proportion of the total area of the ROI covered with HDRs. As a result, we obtain a graph which describes the dependence of HDR parameters on the molecular density and reveals the overall clustering behaviour of the sample under study. The reliability of the algorithm was investigated under a broad range of simulations (see Supplementary Figure 3). The absolute values of SOFI density map depend on expression of the fluorescent markers and parameters of the microscope (particularly excitation intensity and a camera gain). Therefore, we use relative densities and investigate relative changes of local density. For simulations, the molecular density maps were normalized by the mean density calculated over all ROIs. For the experimental data, samples were split into four groups according to four CD4 variants. In each group, a group mean density was calculated across ROIs of twenty samples. ROI of each sample was first normalized to the same mean within the group and then by the maximum of all group means. This normalization procedure largely removes the expression dependence in between the experiments while preserving the differences in relative density in between the CD4 variants. Therefore, normalized SOFI density values can be compared across experiments, as long as the parameters of image acquisition remain the same. 4.3 Simulations The simulation assumed photokinetics as known for fluorescent proteins in PALM experiments21. A photon time-trace for each fluorophore was simulated providing the number of emitted photons over time. The pixel intensity at a given time point corresponds to the integration over the brightness originating from fluorophores in the conjugated object localizations. The number of converted photo-electrons was estimated by a Poisson distributed random distribution. The average value was taken as a pixel value multiplied by the detection efficiency. Additive noise corresponding to thermal noise, read out noise and gain variations was added as a Gaussian noise contribution. Optical system and camera parameters are matched to the microscope system settings (NA, wavelength, magnification, pixel size etc.). The ground truth object is composed of 10 HDRs randomly distributed over an area of 3 x 3 µm. The diameter of generated HDRs varies over the range 60 -- 180 nm, whereas the molecular density in HDRs varies over the range 500 -- 3000 µm. In between the HDRs, individual molecules were randomly distributed such that the HDR to background ratio was {20, 50, 100}. For each test scenario, we simulated a random distribution of labelled molecules, including no clusters as a control. In total, we generated and analysed 720 simulated image sequences (see Supplementary Figure 3). The simulation proves that the algorithm performs well under a broad range of conditions. Acknowledgements We would like to thank Guy Hagen for his helpful remarks and comments. We acknowledge Peter Kapusta for technical assistance and professional advice. This work was funded by Czech Science Foundation (M.C.: 15-06989S). T.La. acknowledge the partly funding of the Horizon 2020 project AD-gut (SEFRI 16.0047) and the Swiss National Science Foundation (SNSF, http://www.snf.ch/) under grants 200020-159945 and 205321-138305. T.Lu. acknowledges a SCIEX scholarship (13.183) and a CTU student grant (SGS16/167/OHK3/2T/13). Author contributions T.Lu., T.La. and M.C. conceived the study. T.Lu., T.La. developed the molecular density analysis. D.G. and Z.K. prepared the samples. D.G, T.Lu. performed the experiments, T.Lu., F.L., and Z.K. analyzed the data. A.B. made the microscope setup. A.B., T.B. provided research advice. T.Lu., T.La. and M.C. wrote the paper. All authors reviewed and approved the manuscript. 6 References 1. Cebecauer, M., Spitaler, M., Sergé, A. & Magee, A. I. Signalling complexes and clusters: functional advantages and methodological hurdles. J. Cell Sci. 123, 309 -- 20 (2010). 2. Prior, I. A., Muncke, C., Parton, R. G. & Hancock, J. F. Direct visualization of ras proteins in spatially distinct cell surface microdomains. J. Cell Biol. 160, 165 -- 170 (2003). 3. Huang, B., Bates, M. & Zhuang, X. Super-resolution fluorescence microscopy. Annu. Rev. Biochem. 78, 993 -- 1016 (2009). 4. Vandenberg, W., Leutenegger, M., Lasser, T., Hofkens, J. & Dedecker, P. Diffraction-unlimited imaging: from pretty pictures to hard numbers. Cell Tissue Res. 360, 151 -- 178 (2015). 5. Baumgart, F. et al. Varying label density allows artifact-free analysis of membrane-protein nanoclusters. Nat. Methods 13, 661 -- 4 (2016). 6. Spahn, C., Herrmannsdörfer, F., Kuner, T. & Heilemann, M. Temporal accumulation analysis provides simplified artifact-free analysis of membrane-protein nanoclusters. Nat. Methods 13, 963 -- 964 (2016). 7. Rubin-Delanchy, P. et al. Bayesian cluster identification in single-molecule localization microscopy data. Nat. Methods 12, 1072 -- 1076 (2015). 8. Sengupta, P. et al. Probing protein heterogeneity in the plasma membrane using PALM and pair correlation analysis. Nat. Methods 8, 969 -- 975 (2011). 9. Levet, F. et al. SR-Tesseler: a method to segment and quantify localization-based super-resolution microscopy data. Nat. Methods 12, 1065 -- 71 (2015). 10. Betzig, E. et al. Imaging intracellular fluorescent proteins at nanometer resolution. Science 313, 1642 -- 5 (2006). 11. Rust, M. J., Bates, M. & Zhuang, X. Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM). Nat. Methods 3, 793 -- 795 (2006). 12. Annibale, P., Vanni, S., Scarselli, M., Rothlisberger, U. & Radenovic, A. Identification of clustering artifacts in photoactivated localization microscopy. Nat. Methods 8, 527 -- 528 (2011). 13. Burgert, A., Letschert, S., Doose, S. & Sauer, M. Artifacts in single-molecule localization microscopy. Histochem. Cell Biol. 144, 123 -- 131 (2015). 14. Owen, D. M. et al. PALM imaging and cluster analysis of protein heterogeneity at the cell surface. J. Biophotonics 3, 446 -- 454 (2010). 15. Ester, M., Kriegel, H.-P., Sander, J., Xu, X. A. A density-based algorithm for discovering clusters in large spatial databases with noise. KDD 96, 226 -- 231 (1996). 16. Ankerst, M., Breunig, M., Kriegel, H. & Sander, J. OPTICS: Ordering Points To Identify the Clustering Structure. ACM SIGMOD Int. Conf. Manag. data 28, 49 -- 60 (1999). 17. Dertinger, T. & Colyer, R. Fast, background-free, 3D super-resolution optical fluctuation imaging (SOFI). Proc. … 106, (2009). 18. Dertinger, T., Colyer, R., Vogel, R., Enderlein, J. & Weiss, S. Achieving increased resolution and more pixels with Superresolution Optical Fluctuation Imaging (SOFI). Opt. Express 18, 18875 -- 85 (2010). 19. Geissbuehler, S. et al. Live-cell multiplane three-dimensional super-resolution optical fluctuation imaging. Nat. Commun. 5, (2014). 20. Geissbuehler, S., Dellagiacoma, C. & Lasser, T. Comparison between SOFI and STORM. Biomed. Opt. Express 2, 408 -- 420 (2011). 21. Deschout, H. et al. Complementarity of PALM and SOFI for super-resolution live-cell imaging of focal adhesions. Nat. Commun. 7, 13693 (2016). 22. Geissbuehler, S. et al. Mapping molecular statistics with balanced super-resolution optical fluctuation imaging (bSOFI). Opt. Nanoscopy 1, (2012). 23. Ovesný, M., Křížek, P., Borkovec, J., Švindrych, Z. & Hagen, G. M. ThunderSTORM: A comprehensive ImageJ plug-in for PALM and STORM data analysis and super-resolution imaging. Bioinformatics 30, 2389 -- 2390 (2014). 24. Roh, K.-H., Lillemeier, B. F., Wang, F. & Davis, M. M. The coreceptor CD4 is expressed in distinct nanoclusters and does not colocalize with T-cell receptor and active protein tyrosine kinase p56lck. Proc. Natl. Acad. Sci. U. S. A. 112, E1604 -- 13 (2015). 7 Figure 1: The workflow of SOFI-based molecular density analysis. SOFI images of different cumulant orders were calculated and used to extract molecular densities. The background was removed using the bSOFI image as a transparency mask. High density regions (HDRs) were segmented by varying the threshold parameter over the whole range of available density levels. For each threshold, the area, equivalent diameter, and number of HDRs were extracted and plotted as a function of the density threshold. (see Fig. 3; given as multiples of mean density in tested ROIs). This procedure is then repeated for each sample and ROI. 8 Figure 2: Example of data processing for a single cell expressing CD4(WT)-mEos2 fusion protein. (a) bSOFI image. (b) Molecular density map. (c and d) Segmentation of the 3 x 3 µm region of interest indicated in b by the red square for a relative density threshold equal to 2.2 times the mean density. (e) Histogram of equivalent diameters (i.e. diameter of a circle of the same area as the non-circular region). (f) Histogram of measured area (in px2) of high density regions (HDRs) in the ROI shown in d. 9 Figure 3: SOFI analysis of four CD4 protein variants in resting T cells immobilized on poly-L-lysine coated coverslips. Native CD4 (WT), palmitoylation mutant (CS1) and variants lacking the extracellular (dD1D4) and cytosolic parts (dCT) were tested (n=20 per variant). (a) Number of high density regions (HDRs) averaged over all samples for each CD4 variant. Density thresholds are related to the mean density calculated over the 3 x 3 µm ROIs of all samples. The inset images show examples of the segmented HDRs for various thresholds indicated above the image. (b) HDR area averaged over all samples for each CD4 variant in px2, where pixel size is 25 nm. (c) Relative area occupied by HDRs related to the total area of the ROI. (d) Box-plots of the properties of HDRs for a threshold equal to 2.6δavg (marked by the vertical dash-dot line). The chosen threshold is the value, where Gaussian function of a random distribution (marked in a by the dashed line) falls below 1 (see Supplementary Fig. 2). Semi-transparent colour areas in a,b,c represent standard deviation. In each box-plot in d, the box represents the interquartile range (IQR), the central mark is the median, and the whiskers extend to the most extreme data points. Supplementary Material for the article: Quantifying protein densities on cell membranes using super- resolution optical fluctuation imaging Tomáš Lukeš1,2, Daniela Glatzová3,4, Zuzana Kvíčalová3, Florian Levet5,6, Aleš Benda3,7,Tomáš Brdička4, Theo Lasser1 & Marek Cebecauer3 1Laboratoire d'Optique Biomédicale, Ecole Polytechnique Fédérale de Lausanne, Lausanne, Switzerland 2Department of Radioelectronics, Faculty of Electrical Engineering, Czech Technical University in Prague, Prague, Czech Republic 3J. Heyrovsky Institute of Physical Chemistry, Academy of Sciences, Prague, Czech Republic 4Institute of Molecular Genetics, Czech Academy of Sciences, Prague, Czech Republic 5Interdisciplinary Institute for Neuroscience, UMR 5297 CNRS, Université de Bordeaux, Bordeaux, France 6Bordeaux Imaging Center, UMS 3420 CNRS, Université de Bordeaux, US4 INSERM, Bordeaux, France 7Imaging methods core facility, BIOCEV, Vestec u Prahy, Czech Republic 10 Supplementary Figure 1: SOFI-based molecular density analysis - threshold filtering. PN l=1 PN PK KLN k=1 Threshold filtering of simulated datasets containing (a) high density clusters, (b) randomly dis- tributed emitters. First a mean density over the wall region of interests (ROIs) is calculated as δavg = 1 n=1 d(k, l, n), where d(k, l, n) is a molecular density per pixel located in the k-th row and l-th column of n-th ROI, N is the total number of ROIs, k,l runs through all rows and columns of the ROI, respectively. The threshold parameter is given as a multiple of the mean density taken over the selected ROI. i.e. threshold = 2 corresponds to 2δavg. For each threshold setting, densities above the threshold determine the boundary of the density dependent area providing number of segments, area size and equivalent diameter. (a.1 - a.4) shows an example for threshold values: 0.1, 1, 2, and 4. Repeating this procedure step by step for the whole range of thesholds, we obtain charts that show number of HDRs as a function of the density threshold (a.5,b.5; blue line) for the case with HDRs (a.5) and with randomly distributed molecules (b.5). For the first case, 6 HDRs are detected at the density threshold 4 (a.5; green dashed line). For the second case, 0 HDRs is detected at this density threshold (b.5; green dashed line). 11 a.2a.3a.4a.5a.1b.2b.3b.4b.5b.1ab Supplementary Figure 2: SOFI-based molecular density analysis - threshold detection. Number of high density regions (HDRs) as a function of density threshold for a simulated sam- ple which (a) contains high density regions, or (b) contains randomly distributed molecules. Fitting a sum of two Gaussian functions reveals a component which corresponds to the random patterns (red dashed line) and a component corresponding to non random HDRs (yellow dashed line). (c) Number of HDRs as a function of density threshold averaged over all cell samples (i.e. 80 samples). Averaging across all samples allows us to obtain one density threshold for all samples and thus compare HDRs size of different CD4 variants at the same density level (Fig. 3). A sum of two Gaussian functions fitted to the data ("gauss2"). Vertical dash-dot line indicates the detected threshold where the value of the Gaussian function (red dashed line), which corresponds to randomly distributed molecules, falls below 1. 12 Number of high density regions05101520253035High density regionsCell samplesdatagauss2 agauss2 bgauss2Number of high density regions05101520253035Randomly distributed moleculesdatagauss2 agauss2 bgauss1acb1 2 3 4 5 6 7 8 9 10Number of high density regions020406080100120data stddata meangauss2gauss2 agauss2 bdetected threshold1 2 3 4 5 6 7 8 9 101 2 3 4 5 6 7 8 9 10Relative density threshold (# δavg)Relative density threshold (# δavg)Relative density threshold (# δavg) Supplementary Figure 3: Simulation of SOFI-based molecular density analysis under controlled conditions. Estimation of number of high density regions (HDRs): 10 simulated HDR per ROI as ground truth. The HDR radius was in the range {30, 60, 90} nm, the molecular density per cluster was in the range {500, 1000, 2000, 3000} molecules per µm2. In between the HDRs, molecules were randomly distributed such that HDR/background ratio was equal to {100, 50, 20}. Each test scenario was repeated 10 times. In total, 360 datasets were generated and evaluated (120 datasets for each HDR/background case). The number of photons per emitter per frame was set to 100, which corresponds well to the experimental conditions. The number of frames of each image sequence was 5000. Dashed blue line in (a),(b),(c) marks the ground truth. Overall, the simulation validates the algorithm and estimation under a broad range of conditions. (d) HDR detection efficiency score is a probability that the estimated number of HDRs is in the range (7-13). The accuracy of the estimation increases with increasing HDR/background ratio and increasing HDR density. The simulations were calculated for the grid points. The dashed line marks the conditions where the signal to noise ratio (SNR) and signal to background ratio (SBR) of simulated data correspond to SNR and SBR of the real cell datasets in our experiments. (e) Estimation of HDR radius for grid points marked in d by the circles. 13 Number of HDRs051015202530102030405060708090100HDR / background ratio = 100Number of HDRs051015202530102030405060708090100HDR / background ratio= 50Number of HDRs051015202530102030405060708090100HDR / background ratio = 20Density # µm-1500 100020003000HDR / background ratioHDR detection efficiency score (%)HDR radius (nm)20 50 100707580859095100r = 30 nnr = 60 nmr = 90 nm020406080100120abced Supplementary Figure 4: Schematic representation of wild-type (WT) CD4 and its variants: palmi- toylation mutant (CS1), mutant missing the extracellular domain (dD1D4) and mutant missing the intracellular domain (dCT). TM = transmembrane domain, Ig = four immunoglobulin type domains D1- D4, Palm = palmitoylations sites C419 and C422, Non-palm = non-palmitoylated variant with mutations C419,422S. Proportions are not in scale. 14 TMPalmitoylIgcTMIgTMPalmitoylccTMNon-palmitIgc*WTCS1dD1D4dCT Supplementary Figure 5: SMLM analysis of the plasma membrane organisation of CD4 variants. Plasma membrane organisation of the CD4 protein variants in resting T cells characterized using photo-activation localisation microscopy (PALM) followed by a Voronoï-based segmentation algorithm [1]. High density regions (HDRs) of irregular shape frequently forming networks of connected areas are identified. Yet, since the acquisition exhibited a high density of molecules with high blinking rates, the quantification of these HDRs can be affected by localization errors and under- or overcounting artifacts as described by Burgert et al. [2]. (a) Original dataset composed of 1,747,681 localizations, scale bar 2 µm. (b) Magnification of the central area of the cell. Segmented HDRs are displayed in red, scale bar 1 µm. (c) Zoom on a HDR composed of 65 localizations, scale bar 50 nm. As shown by its time trace, the localizations in this region are originating from a single fluorophore, making its blinking correction simple. (d-e) Zoom on a denser HDR (d, 384 localizations, scale bar 50 nm) and interconnected HDRs (e, 12,886 localizations, scale bar 200 nm) with their time trace. Even for a small HDR, the presence of multiple emitters complicates the blinking correction. The problem becomes even more apparent with interconnected HDRs because of their high-density of molecules. (f-g) Three frames of the original image acquisition, pixel size 105 nm. (g) Corresponding zoom to the region covered by one HDR. The high-density of molecules makes it difficult to properly separate each emitter, resulting in localization errors as well as under-counting. 15 afg bcde # photons# photons# photons0246825003750500002468300040005000 02468300037504500Distance (px)0300060009000120000200040006000 030006000900012000025005000750010000 0300060009000120000100002000030000 # photons# photons# photonsFrameFrameFrame Supplementary Figure 6: Each row represents four selected ROIs of one CD4 variant. Colorbar represents relative density (#δavg). Scale bar 500 nm. 16 012345678910012345678910012345678910012345678910012345678910012345678910012345678910012345678910012345678910012345678910012345678910012345678910012345678910012345678910012345678910012345678910WTCS1dCTd1d4 Supplementary Note: SOFI density estimation The technical requirements for SOFI are a classical widefield microscope merged with a fast high sensitivity digital camera. SOFI image processing is based on higher order statistics and exploits the temporal sequence of blinking fluorescent emitters [3, 4]. Calculating spatio-temporal cross- cumulants allows SOFI to obtain a super-resolved, background-free and noise-reduced images. Higher-order cumulants contain information about the photo-physics of the emitters. Com- bining SOFI images of different cumulant orders, allows one to extract physical parameters like molecular density [5], which we applied to investigate plasma membrane distribution of proteins. 1.1 SOFI principle and theory As stated by Dertinger et al. [3], the fluctuating emitters should switch between at least two optically distinguishable states (e.g. a dark and a bright state) repeatedly and independently in a stochastic manner. Images of stochastically blinking emitters are recorded such that the point-spread function (PSF) extends over several camera pixels. Acquiring a sequence of images results in a time dependent intensity trace for each pixel. Assuming N independently fluctuating emitters, the detected intensity is given as I(r, t) = kU(r − rk)sk(t) + b(r) + n(r, t), (1) NX k=1 NX k=1 where k is the molecular brightness, U(r − rk) is the PSF at the position rk, sk(t) denotes a switching function (normalized fluctuation sequence, sk(t) ∈ {0, 1}), b(r) is a constant back- ground, and n(r, t) represents an additive noise contribution. For each pixel, an nth order cumulant is calculated for disentangling emitters inside the PSF. By applying the nth order cumulant to Eq. (1), we obtain κn{I(r, t)}(τ) = κn ) kU(r − rk)sk(t) + b(r) + n(r, t) ( MX k=1 (τ). (2) Using additivity and semi-invariance properties of cumulants [6], the nth order cumulant with zero time lag can be written as κn{I(r, t)} = k U n(r − rk)κn{sk(t)} + κn{b(r)} + κn{n(r, t)}. n (3) √ For (n ≥ 2), the Gaussian noise (κ{n(r, t)})) and stationary background (κ{b(r)} ) terms are eliminated by the cumulant analysis as an intrinsic property of cumulants. For an nth order cumulant, the PSF is raised to the nth power (see Eq. 3). As a consequence, the PSF is narrowed and the spatial resolution is improved by a factor of n [3]. Therefore, increasing the cumulant order yields an image with an enhanced spatial resolution. Since a multiplication in the spatial domain corresponds to a convolution in the frequency domain, the cut-off frequency of the spectrum U n (k) is n-times higher than that of U (k). By applying a deconvolution and a subsequent rescaling, the nth order cumulant image exhibits an up to n-fold resolution improvement [4]. As shown in [4], virtual pixels can be calculated in between the physical pixels using cross-cumulants and followed by a flattening operation i.e. assigning proper weights to these virtual pixels [4, 7, 8]. SOFI assumes a blinking model where the fluorophores reversibly switch between a bright and a dark state. In Deschout et al. [9], SOFI was applied to the PALM photo-physical model. In the PALM photo-physical model, the emitter activation is assumed as non-reversible, however, once the emitter is activated, it exhibits several fast blinking events prior to the final bleaching 17 event [10]. The emitter fluctuates between two different states (an on-state Son and a dark state Soff), which is expressed by the on-time ratio as ρ = τon τon + τoff , (4) where τon and τoff are the characteristic lifetimes of the Son and Soff states. The nth order cumulant κn{sk(t)} is in this model described by a Bernoulli distribution with probability ρon [5] and approximated by an nth order polynomial function for the on-time ratio as fn(ρon) = ρon(1 − ρon) ∂fn−1 ∂ρon . Under these conditions, the nth order cumulant can be approximated as [5] κn{I(r, t)} ≈ nfn(ρon) U n(r − rk). NX k=1 (5) (6) 1.2 Estimation of density maps Geissbuehler et al. [5] used three cumulant images (2nd, 3rd, and 4th order) to estimate molecular parameters: on-time ratio, brightness and molecular density. Here, we generalize this concept to any three cumulant images of distinct orders. If we assume spatially varying but locally constant on-time ratios and molecular brightness, the cumulants (for the cumulant order n > 1) can be approximated by [5] gn(r) ≈ n(r)fn(ρon)N(r)EV {U n(r)}. (7) where EV {U n(r)} is the expectation value of U n(r), N(r) is the number of emitters inside a detection volume V. Approximating the PSF near the interface in a total internal reflection (TIR) configuration by a lateral 2D Gaussian profile combined with an axial exponential profile, we obtain EV {U nTIR(r)} = c(σx,y, σz, dz) n2 , (8) where dz represents the exponential decay of the TIR illumination [11]. Using 3 consecutive cumulant images of orders n, (n − 1), (n − 2), we obtain for the ratios K1 = gn−1 gn−2 K2 = gn gn−2 K3 = gn gn−1 = fn−1(ρon(r)) fn−2(ρon(r)) (r) µn−1 µn−2 = fn(ρon(r)) fn−2(ρon(r)) 2(r) µn µn−2 = fn(ρon(r)) fn−1(ρon(r)) (r) µn µn−1 , where µn = EV {U n(r)}. Substitution of Eq. (11) into Eq. (9) leads to (cid:26) gngn−2 g2 n−1 f3(ρon(r)) fn(ρon(r)) (r) = gn gn−1 = fn(ρon(r))fn−2(ρon(r)) . f 2 n−1(ρon(r)) µn−1 µn , N(r) = gn(r) n(r)fn(ρon)µn (9) (10) (11) (12) (13) (cid:27) . Building up the ratios K1 and K2 from the Eq. (14) for cumulants of 2nd, 3rd, and 4th order, we obtain 18 K1(r) = µ2g3 µ3g2 K2(r) = µ2g4 µ4g2 (r) = (r)(1 − 2ρon(r)) (r) = 2(r)(1 − 6ρon(r) + 6ρ2 on(r)), (14) (15) Solving for molecular brightness , on-time ratio ρon, we obtain two solutions for the on-time ratio ρon, and molecular brightness  1 ± K1 3K2 ρon(r) = q 1 − 2K2 − 2K2 3K2 1 − 2K2) 2(3K2 , (r) = ∓q 3K2 1 − 2K2  , where the first solution corresponds to a negative brightness and will be discarded. The molecular density (number of molecules N per pixel area) is (16) (17) (18) (19) (20) , (21) N(r) = g2(r) 2(r)ρon(r)(1 − ρon(r)) . For cumulants of 3rd, 4th, and 5th order, the ratios K1 and K2 become K1(r) = (r)(1 − 6ρon + 6ρ2 on) (1 − 2ρon(r)) K2(r) = 2(r)(12ρ2 on − 12ρon + 1) which ends in four solutions. Two correspond to positive molecular brightness K2 1(4K2 1 − 3K2) 1(4K2 1 − 3K2) − 3K2 − 9K1K2 q 3 12K3 1 ± √ q 4K2 1 ∓ 2 ρon(r)1,2 = r (r)1,2 = 1 − 3K2) − 3K2 . K2 1(4K2 √ 3 r q 1 ∓ 2 K2 1 − 3K2) 6K1(4K2 4K2 Using a combination of higher order cumulants for molecular parameters can theoretically provide higher spatial resolution of the molecular parameter maps assuming high enough SNR of the cumulant images used. For the combination of 4th, 5th, 6th order cumulant, it is also possible to find a solution in a closed form, but due to its complexity, a numerical approach might be preferred. Therefore SOFI extracts density without counting individual events in the image. Density simply results from a correlation/cumulant analysis of intensity time traces. 19 Supplementary References [1] I. Chamma, F. Levet, J.-B. Sibarita, M. Sainlos, and O. Thoumine, "Nanoscale organization of synaptic adhesion proteins revealed by single-molecule localization microscopy," Neurophotonics, vol. 3, p. 041810, 2016. [2] A. Burgert, S. Letschert, S. Doose, and M. Sauer, "Artifacts in single-molecule localization microscopy," Histochemistry and Cell Biology, vol. 144, pp. 123 -- 131, 2015. [3] T. Dertinger and R. Colyer, "Fast, background-free, 3D super-resolution optical fluctuation imaging (SOFI)," Proceedings of the . . . , vol. 106, no. 52, 2009. [4] T. Dertinger, R. Colyer, R. Vogel, J. Enderlein, and S. Weiss, "Achieving increased resolution and more pixels with Superresolution Optical Fluctuation Imaging (SOFI).," Optics express, vol. 18, pp. 18875 -- 85, Aug. 2010. [5] S. Geissbuehler, N. L. Bocchio, C. Dellagiacoma, C. Berclaz, M. Leutenegger, and T. Lasser, "Mapping molecular statistics with balanced super-resolution optical fluctuation imaging (bSOFI)," Optical Nanoscopy, vol. 1, no. 1, 2012. [6] J. M. Mendel, "Tutorial on Higher-Order Statistics (Spectra) in Signal Processing and System Theory: Theoretical Results and Some Applications," Proceedings of the IEEE, vol. 79, pp. 278 -- 305, 1991. [7] S. C. Stein, A. Huss, D. Hähnel, I. Gregor, and J. Enderlein, "Fourier interpolation stochastic optical fluctuation imaging.," Optics express, vol. 23, pp. 16154 -- 63, 2015. [8] W. Vandenberg, S. Duwé, M. Leutenegger, B. Krajnik, T. Lasser, and P. Dedecker, "Model-free uncertainty estimation in Stochastical Optical Fluctuation Imaging ( SOFI ) leads to a doubled temporal resolution," vol. 2402, pp. 1347 -- 1355, 2015. [9] H. Deschout, T. Lukes, A. Sharipov, D. Szlag, L. Feletti, W. Vandenberg, P. Dedecker, J. Hofkens, M. Leutenegger, T. Lasser, and A. Radenovic, "Complementarity of PALM and SOFI for super-resolution live-cell imaging of focal adhesions," Nature Communications, vol. 7, p. 13693, 2016. [10] N. Durisic, L. Laparra-Cuervo, A. Sandoval-Álvarez, J. S. Borbely, and M. Lakadamyali, "Single-molecule evaluation of fluorescent protein photoactivation efficiency using an in vivo nanotemplate.," Nature methods, vol. 11, pp. 156 -- 62, 2014. [11] K. Hassler, T. Anhut, R. Rigler, M. Gösch, and T. Lasser, "High Count Rates with Total Internal Reflection Fluorescence Correlation Spectroscopy," Biophysical journal, vol. 88, pp. L01 -- L03, 2005. 20
1806.02258
1
1806
2018-06-06T15:47:19
Partitioning of 1-alkanols in composite raft-like lipid membrane
[ "physics.bio-ph", "cond-mat.soft" ]
Though 1-alkanols are well known to have anesthetic properties, the mode of operation remains enigmatic. We perform extensive atomistic molecular dynamics simulation to study the partitioning of 1- alkanols of different chain-lengths in raft-like model membrane made up of mixture of unsaturated dioleoyl-phosphatidylcholine (DOPC) and saturated dipalmitoyl-phosphatidylcholine (DPPC) and choles- terol (Chol) exhibiting phase coexistence of liquid-ordered (lo) - liquid disordered (ld) phase domains. Our simulation shows that the effect of 1-alkanols on the mechanical properties of the membrane has been pronounced with the chain-length of it. In particular, the 1-alkanols prefer to partition in the ld phase domain. The penetration of the 1-alkanols in the membrane increases significantly for the long-chain alkanol. We also have found the dependency of the order of chains of the lipids and rigidity of the membrane on the 1-alkanols.
physics.bio-ph
physics
Partitioning of 1-alkanols in composite raft-like lipid membrane Anirban Polley1 1Department of Chemical Engineering, Columbia University, New York City, New York-10027, USA Abstract Though 1-alkanols are well known to have anesthetic properties, the mode of operation remains enig- matic. We perform extensive atomistic molecular dynamics simulation to study the partitioning of 1- alkanols of different chain-lengths in 'raft-like' model membrane made up of mixture of unsaturated dioleoyl-phosphatidylcholine (DOPC) and saturated dipalmitoyl-phosphatidylcholine (DPPC) and choles- terol (Chol) exhibiting phase coexistence of liquid-ordered (lo) - liquid disordered (ld) phase domains. Our simulation shows that the effect of 1-alkanols on the mechanical properties of the membrane has been pronounced with the chain-length of it. In particular, the 1-alkanols prefer to partition in the ld phase domain. The penetration of the 1-alkanols in the membrane increases significantly for the long-chain alkanol. We also have found the dependency of the order of chains of the lipids and rigidity of the membrane on the 1-alkanols. PACS numbers: 61.41.+e, 64.70.qd, 82.37.Rs, 45.20.da I. INTRODUCTION cific proteins [2, 3] and block the protein func- The phenomena of general anesthesia [1] is known from a long time and the use of anes- thetics in the hospitals are very common for all painless surgical operation. However, the molec- ular level of understanding in the mechanism of the general anesthesia is not yet understood. It is still debatable whether the general anesthesia tion by changing conformation or it is an indirect lipid-mediated mechanism [4 -- 6] while the anes- thetics alter the membrane properties such as total volume of the membrane, the volume oc- cupied by the anesthetics within the membrane, the phase transition temperature, the lipid chain order, membrane thickness or the lateral pres- sure profile of the membrane. is caused by the binding of anesthetics to spe- Apart from anesthesia, the alkanol has been 1 arXiv:1806.02258v1 [physics.bio-ph] 6 Jun 2018 well known as a penetration enhancer for trans- the symmetric 'raft-like model membrane', com- dermal drug delivery. However, the understand- prising DOPC, DPPC and Chol, which exhibits ing of the mechanism of the action of penetra- phase coexistence of lo-ld domains. 'Membrane- tion enhancer remains opaque. It is interesting raft' exhibiting lipid-based compositional het- to know that 1-alkanol as anesthetics or as pen- erogeneity has been thought to facilitate several etration enhancer exhibits 'cutoff-effect'. The physiological activities like protein sorting, sig- potency of 1-alkanols as anesthetics is increased naling processes [28 -- 36]. Thus it is important up to a certain cutoff chain length (dodecanol) to study the effect of 1-alkanols into the model [7] and 1-alkanols with a chain length above cut- raft-like membrane. off length show no anesthetic potency [7]. Sim- The article is arranged as follows: we first ilarly, the penetration enhancement property discuss the details of the atomistic molecular dy- of 1-alkanols increase with increasing the chain namics simulations of the multicomponent bi- length up to decanol and decrease again for 1- layer membrane. Next, we present our main alkanol with longer than cut-off chain length. results of the spatial heterogeneity of the com- The potency of 1-alkanols as anesthetics and as ponents, effects of 1-alkanols on the thickness, penetration enhancer both is decreased with the penetration, partitioning, order parameter and branching of the carbon chains. bending rigidity of the membrane. We end by The several experimental studies have been summarizing our findings. conducted to explore the effect of the anesthet- ics in the property of the lipid-membrane such as NMR spectroscopic studies that show ethanol II. METHODS molecules having disordering effect on the lipids Model membrane : We have studied symmetric [8 -- 13]; X-Ray studies that reveals ethanol hav- 3-component bilayer membrane embedded in ing potential to change in the lipid membrane an aqueous medium by atomistic molecular above its main phase transition temperature[14 -- dynamics simulations (MD) using GROMACS- 16]. Few molecular dynamics studies also have 5.1 [37]. We prepare the bilayer membrane been performed to study the influence of the at 23◦C at the relative concentration, 33.3% anesthetics in the lipid membrane [11, 17 -- 27]. of DOPC, DPPC, and Chol, respectively. To In the present work, we investigate the effects the symmetric ternary bilayer membrane, we of 1-alkanols: ethanol, pentanol, and octanol on have introduced 25% of 1-alkanols (ethanol, 2 pentanol, and octanol) to water layer from and are hydrated with 32768 water molecules. both sides of the symmetric ternary bilayer To the symmetric composite membrane, we membrane, respectively. All 4 multicomponent introduce 256 1-alkanols (ethanol, pentanol, bilayer membranes comprise 512 lipids in each and octanol) uniformly in both sides of the leaflet (with a total 1024 lipids) and 32768 water layers of the bilayer membrane using water molecules (such that the ratio of water PACKMOL [48] as shown in Fig. S1 in the to lipid is 32 : 1) to hydrate the bilayer mem- supplementary information (SI). brane. In our simulation study, we choose the compositions of the membrane such that we Choice of ensembles and equilibration : We obtain the bilayer membrane exhibiting lo-ld simulate the symmetric bilayers for 50 ps in phase coexistence [38]. the NVT ensemble using a Langevin thermo- stat to avoid bad contacts arising from steric Force fields : The force field parameters of constraints and then for 500 ns in the NPT en- DOPC, DPPC, and Chol are taken from the semble (T = 296 K (23◦C), P = 1 atm). To get previously validated united-atom description enough data, we repeat the simulations 4 times [39 -- 43]. We have taken the same previously (total 2µs) to study each system. We run the used force-field parameters for the ethanol, simulations in the NPT ensemble for the first pentanol, and octanol [41, 44 -- 47]. In our simu- 50 ns using Berendsen thermostat and barostat, lation study, we use improved extended simple then using Nose-Hoover thermostat and the point charge (SPC/E) model to simulate water Parrinello-Rahman barostat to produce the molecules, with an extra average polarization correction to the potential energy function. correct ensemble using a semi-isotropic pressure coupling with the compressibility 4.5 × 10−5 bar−1 for the simulations in the NPT ensemble, Initial configurations : We generate the initial long-range electrostatic interactions by the configurations of the symmetric multicompo- reaction-field method with cut-off rc = 2 nm, nent bilayer membrane using PACKMOL [48], and the Lennard-Jones interactions using a where all the components are homogeneously cut-off of 1 nm [40, 43, 49]. We use last 200 ns mixed. All the symmetric bilayer membrane of the trajectories of the four independent comprises 342 DOPC, 340 DPPC, and 342 repeated simulations for the data analysis. Chol with spatially uniformly randomly placed, 3 Computation of bending rigidity : To calcu- values attend its asymptotic behavior with a late bending rigidity, we have employed spectral very small fluctuation, we can be satisfied as we method, where we need to calculate structure obtain equilibrated bilayer membrane [40, 50]. factor of the fluctuation of the lipid head groups. We start our simulation from the initial con- In this method, we need to use a large mem- figuration, where lipid components are homo- brane, otherwise, we would get very noisy data. geneously mixed and the 1-alkanols are in the Therefore, we have carried out the simulation of 9216 lipid membrane size ≈ 48×48 nm using the already equilibrated last configuration of 1024 aqueous layers of both sides of the bilayer mem- brane shown Fig. S1 in SI. After the long sim- ulation, we get the equilibrated patch of bilayer lipid patch of the 500 ns simulation. Here, we membrane and the last frame of the simulation translate the 1024-lipid membrane patch with 9 positions like a 3 × 3 matrix-form using vmd, such that we generate the initial configuration of a large bilayer membrane of 9 × 1024 lipids for the further simulation. We run large bi- layer membrane for 100 ns in NPT ensemble us- ing Nose-Hoover thermostat and the Parrinello- Rahman barostat to produce the correct ensem- ble using a semi-isotropic pressure coupling with the compressibility 4.5×10−5 bar−1 as described above. After the simulation, we collect the po- have been shown in Fig. 1, where all the com- ponents, especially 1-alkanols are highlighted to show their preferable arrangement in the com- posite bilayer membrane. A. Transverse heterogeneity of the com- ponents in membrane The electron density profile of components in the membrane without and with 1-alkanols has been shown in Fig. 2. Here, the electron density sition of the center of mass of the head group of of the head groups and tails of DPPC, DOPC lipids and use to calculate the structure factor. lipids and Chol and 1-alkanols, respectively in III. RESULTS AND DISCUSSION the z-axis are plotted (In our simulation, z-axis is representing the normal to the surface of the membrane lying in the xy-plane). To confirm that we attend thermally and We find the peaks of the head-group of the chemically equilibrated bilayer membrane, we lipids (DPPC and DOPC) are pronounced av- calculate the time evolution of the total energy eraged at 2.058 nm, 1.915 nm, 1.887 nm and of the system and area per lipid and when the 1.859 nm respectively for without and with 4 FIG. 1. The snapshot of symmetric composite bilayer membrane comprising DOPC (purple), DPPC (orange), Chol (green), water (cyan) (A) without 1-alkanol, (B) with Ethanol (ETH) (red), (C) Pentanol (PCT) (blue) and (D) Octanol (OCT) (gray) are shown respectively. 1-alkanols (ethanol, pentanol, and octanols). respectively as shown in Fig. 3 (a). (This is the averaged value of the thickness of Again, we find from the electron density pro- the one leaflet of the membranes.) file of the 1-alkanols highlighted in brown shown From the two peaks of the head-group the in Fig. 2, that the peaks of the electron den- lipids (DPPC and DOPC), we can measure the sity of 1-alkanols are closest (average in two average thickness of the membranes with and leaflets 1.1509 nm) to 0 (center of the membrane, without 1-alkanols. Here, we find that the z = 0nm) for octanol and farthest for ethanol average thickness of the lipid membrane has (average in two leaflets 1.732 nm), and in be- been decreased with the number of carbon chain tween value (average in two leaflets 1.527 nm) for length in 1-alkanol as the carbon chain length of pentanol, which is related to the depth of pen- ethanol, pentanol, and octanol are 2, 5 and 8, etration of 1-alkanols at equilibrium. To illus- 5 (A) (C) (B) (D) trate the depth penetration into the membrane, nent of the bilayers with ethanol, pentanol, and we plot the mean depth of penetration for 1- octanol are shown in Figure 4 (A), (B) and (C), alkanols with increasing chain length as shown respectively. We find that all the symmetric in Fig. 3 (b), which shows that the depth of pen- membrane exhibits phase coexistence of DPPC- etration of 1-alkanols in the membrane increases rich lo domain and DOPC-rich ld domains, while with the increase of carbon chain length of the Chol is also accumulated in DPPC-rich domains. 1-alkanols. Again, we find that ethanol, pentanol and oc- tanol are partitioning in the DOPC-rich do- B. Lateral heterogeneity of the compo- mains shown in Fig. 4. nents in membrane In Fig. 5, the joint probability distribution (JPD) shows that a distinct peak along the di- To elucidate the spatial heterogeneity of the agonal for the JPD of DOPC to all three 1- components, especially 1-alkanols, we calculate alkanols, while there is no diagonal peak in JPD the spatial number density of each component for any 1-alkanols to the DPPC-rich domain. of the membrane. We collect the last 200 ns tra- jectories of the simulations of each system and To extend our the investigation on the calculate the center of mass of each component spatial partitioning of the anesthetics, we define and project it on the xy-plane and calculate the the 'correlation coefficient' from the normalized number density by binning the data with bin- cross-correlation size 0.4 nm. between the 1-alkanols solutes to the The spatial number density of each compo- DOPC/DPPC-rich domain [31, 50] as C(ρDOP C/DP P C(r), ρ1alkanol(r)) = hρDOP C/DP P C (r)ρ1−alkanol(r)i−hρDOP C/DP P C (r)ihρ1alkanol(r)i √hρDOP C/DP P C (r)2i−hρ1alkanol(r)i2√hρDOP C/DP P C (r)2i−hρ1alkanol(r)i2 averaged over space (denoted by Cuu), while ρDOP C/DP P C(r) and ρ1−alkanol(r) are the spatial We compute Cuu for DOPC to ethanol, pen- tanol, and octanol and PSM to ethanol, pen- number densities of the DOPC/DPPC and 1- tanol, and octanol, respectively bilayer mem- alkanol, respectively with the same positional branes shown in Figure 6 which indicates that coordinate r(x, y). the values of Cuu are high for DOPC-rich do- 6 FIG. 2. The electron density of each component vs. z-axis (normal to membrane) of the symmetric composite bilayer membrane (A) without 1-alkanol, (B) with Ethanol (ETH), (C) Pentanol (PCT) and (D) Octanol (OCT) are shown in respectively. main to ethanol/ pentanols/ octanols-rich do- domain. mains while that values are low for DPPC to all 1-alkanols. This suggests that all three 1- alkanols (ethanol, pentanol, and octanol) prefer C. Spatial heterogeneity of thickness to partition into the DOPC-rich domain, while Here, we compute the spatial heterogeneity 1-alkanols do not partition in DPPC-Chol rich of thickness for all membranes. To compute the 7 (A) (C) (B) (D) FIG. 3. (A) The average thickness of the membrane without 1-alkanol and ethanol (chain length =2), pentanol (chain length =5) and octanol (chain length =8) and (B) the penetration of ethanol, pentanol and octanol are shown in respectively. local thickness of the membrane, we first col- of the membrane. lect the position of P-atom of both DPPC and Figure 7 shows that the phase coexistence of DOPC lipids of both upper and lower leaflets. thicker lo and thinner ld phase domains are ex- Now, we divide the xy-plane in both leaflets into hibited in all bilayer membranes. many small boxes with grid size 0.4 nm to cal- culate the mean coordinate of each grid in each leaflet from the collected data of positions of P- atoms. The difference between the z-coordinate corresponding to two layers with same (x,y) - coordinate gives us the thicknss of the mem- brane, d at that local coordinate (x,y). Thus, we calculate the spatial thickness heterogeneity 8 D. Effects of 1-alkanols on the order pa- rameter We calculate the deuterium order parameter, S of the acyl chains of the DOPC and DPPC lipids in the membrane respectively, which is de- fined as S = h 3 2i where θ is the an- 2(cos2θ) − 1 (A) (B) FIG. 4. The spatial number density of each component in the membrane with ethanol (ETH), pentanol (PCT) and octanol (OCT) are shown in (A), (B) and (C), respectively[40, 50]. gle between the carbon atom - hydrogen (deu- To illustrate the effect of the 1-alkanols on S, terium) atom and the bilayer normal shown in Fig. 8 (a) and (b). Here, we find two impor- tant things: one is that the value of the S for DPPC is more than that of DOPC, which sug- we define, δS = SA−S0 S0 where SA and S0 are the order pa- rameter of the lipid chains of the symmetric mul- ticomponent bilayer membrane with and without gests that DPPC-rich domain is more ordered ethanol, pentanol and octanol, respectively and (lo) phase, while DOPC-rich domain is less or- we find that change of the order parameter of the dered (ld) phase, where acyl chains of the lipids lipids is highest in presence of octanol, while the are more floppy. Secondly, we find that the value change of that is lowest in ethanol. of S is affected most in octanol, the value of that is changed lowest in ethanol. 9 (A) (i) DOPC (ii) DPPC (iii) CHOL (iv) ETH (B) (i) DOPC (ii) DPPC (iii) CHOL (iv) PCT (C) (i) DOPC (ii) DPPC (iii) CHOL (iv) OCT FIG. 5. The joint probability distribution between (A1) ethanol (ETH) and DOPC, (A2) ETH and DPPC, (B1) pentanol (PCT) and DOPC, (B2) PCT and DPPC, (C1) octanol (OCT) and DOPC and (C2) OCT and DPPC are shown respectively [40]. E. Orientation of O-H and C1-Cn bonds surface of the membrane. Fig. 9 shows the orientational order param- We characterize the orientation of the vector eter, Iβ for the ethanol, pentanol, and octanol, (OH) between O and H -atoms and that another vector (C1Cn) between the coordinate of first C-atom (C1) and last C-atom (Cn), where the respectively. We find that the value of Iβ close to ±1 near to the head groups of the lipid mem- brane and the value of it are zero and switch value of n = 2, 5 and 8 for ethanol, pentanol its value across the center of the membrane. and octanol, respectively. To quantify the ori- The value of Iβ of the OH-vector for ethanol entation of the two vectors: OH and C1Cn, we is greater than that of C1Cn vector, while the define an orientational order parameter, Iβ as Iβ = hCos(β)i, where β is the angle between the vector OH/C1Cn and the normal (z-axis) of the value of OH-vector is less than that of C1Cn- vector for octanol. Therefore, Fig. 9 suggests that the value of 10 .5 .5 2 11 00 012345 0.5 1 1.5 ρDOPC (nm−2) 0.5 1 1.5 ρDPPC (nm−2) 1.5 1 0.5 0 0 1.5 1 0.5 0 0 (C1) ρOCT (nm−2) (C2) ρOCT (nm−2) .5 .5 2 11 00 012345 0.5 1 1.5 ρDOPC (nm−2) 0.5 1 1.5 ρDPPC (nm−2) 1.5 1 0.5 0 0 1.5 1 0.5 0 0 (B1) ρPCT (nm−2) ρPCT (nm−2) (B2) .5 .5 2 11 00 012345 0.5 1 1.5 ρDOPC (nm−2) 0.5 1 1.5 ρDPPC (nm−2) 1.5 1 0.5 0 0 ρETH (nm−2) (A1) (A2) 1.5 1 0.5 0 0 ρETH (nm−2) FIG. 6. The cross-correlation between ethanol (ETH) and DOPC (A-X), ETH and DPPC (A-Y), pentanol (PCT) and DOPC (B-X), PCT and DPPC (B-Y), octanol (OCT) and DOPC (C-X), and OCT and DPPC (C-Y) are shown respectively[31, 50]. (1) Z (κ∇2u(x, y)2 +γ∇u(x, y)2)dxdy 1 2 H[u(x, y)] = Here, the first term in the Hamiltonian is the energy cost due to the bending of the membrane with a constant κ, the bending rigidity whereas orientation order parameter for C1Cn increases with the acyl-chain length as the acyl chain of the 1-alkanol arranged in more ordered fashion with the increase in chain length. F. Effects of 1-alkanols on bending mod- the second term corresponds to the area fluctu- ulus and elasticity ation of the surface with constant γ, the surface tension of the membrane. If we assume our bilayer as a flat surface u(x, y) with fluctuation over minimum surface energy at u(x, y) = u0 and the energy costs due to the thermal fluctuation, then we can write by u(~r) = P~q If we transform the Eqn.1 into Fourier Space u(~q)ei~q·~r where, ~r ≡ (x, y) in two dimensional real space and ~q ≡ (qx, qy) in two dimensional Fourier space, then we get by the the Hamiltonian of the membrane as use of equipartition theorem, 11 ����� ����� ����� ������ ������ ��� ��� ��� ��� ��� ��� FIG. 7. The spatial thickness of composite bilayer membrane (A) without 1-alkanol, (B) with ethanol (ETH), (C) with pentanol (PCT) and (D) with octanol (OCT) are shown respectively [40]. Eqn.2 turns into S(q) = KBT AL(κq4 + γq2) (2) S(q) = KBT ALκq4 (3) However, the Eqn.2 tells us that the nonzero where, S(q) ≡ 1 2N < u(q)2 > is the struc- ture factor of the membrane surface u(x, y) with projected area per lipid Ah = AL/N of N num- surface tension γ affects the fluctuation spectra κ). S(q) significantly (S(q) ∝ q−2 for q (cid:28)p γ Fig. 10 shows the structure factor S(q) vs. q, ber of lipids in each leaflet with total area AL. which is fitted to the straight line for small wave As the surface tension γ for the lipid bilayer membrane is approximately zero (γ ≈ 0), the 12 vector, q (See Section: Methods). From the slope of the Fig. 10, we calculate (A) (C) (B) (D) FIG. 8. (A) the deuterium order parameter (S) vs. carbon number of the acyl chain of the lipids of the multicomponent symmetric membrane without (black) and with ethanol (red), chloroform (blue), and methanol(green), respectively. (B) the change of deuterium order parameter (δs) vs. carbon number of the acyl chain of the lipids in the membrane with anesthetics compared to without anesthetic lipid membrane. the bending rigidity of the membrane without membrane decrease with the acyl-chain length and with ethanol, pentanol, and octanol, re- of 1-alkanols. spectively using equation 3 and are tabulated in Table-I [51]. It suggests that the rigidity of the 13 ��� ���� ��� ���� ��� ��� FIG. 9. Orientation of O-H and C1-Cn vectors of the ethanol, pentanol, and octanol are shown in Fig. 9, respectively. System κ (kBT ) DOPC,DPPC,CHOL 50.3857 ± 0.9590 DOPC,DPPC,CHOL and ethanol 33.5272 ± 0.6002 DOPC,DPPC,CHOL and pentanol 29.3434 ± 0.5157 DOPC,DPPC,CHOL and octanol 22.9672 ± 0.7419 TABLE I. The value of the κ of the multicomponent bilayer membrane without and with 1-alkanols are given in Table-I. IV. CONCLUSION decreases in presence of 1-alkanol and it's value We present the effect of 1-alkanols with dif- ferent chain-length in the symmetric 'raft mem- brane' comprising DOPC, DPPC and Chol us- ing atomistic molecular dynamics simulation. Here, we explore the consequence of 1-alkanols by computing the physical properties of the membrane. Our main findings in the present study are: (i) the thickness of the membrane 14 decrease with the increase of the chain length of the 1-alkanols. (ii) the penetration of the 1- alkanols increases with length of chain length of it. (iii) 1-alkanols prefer to partition in the DOPC-rich ld domain in the 'raft-membrane'. (iv) The deuterium order parameter of the acyl chain of the lipids are decreased in presence of 1- alkanols and its value decreases with the chain length of 1-alkanols. (v) With the increase of ��� ��� ��� FIG. 10. The fluctuation spectra S(q) versus wave vector q for the membrane (A) without 1-alkanol, (B) with ethanol, (C) with pentanol and (D) with octanol, respectively are shown. Here, the bending rigidity is calculated using the long wavelength undulation spectrum. The fitting is carried out using the theoretical q−4 line and is denoted as the dashed line. chain-length in the 1-alkanols, the orientation in presence of 1-alkanols. order parameter of the acyl chain of 1-alkanols increases. (vi) The value of orientation order pa- rameter for acyl chain (C1Cn) increases with the chain length of acyl chain of the 1-alkanol. (vii) The bending rigidity of the membrane decreases Our study shows that the mechanical prop- erties of the membrane change significantly in presence of 1-alkanols, which partition in the ld domain and the effect on the membrane is pronounced with chain-length of 1-alkanol 15 q (nm-1) q (nm-1) q (nm-1) S(q) (B) S(q) (D) q (nm-1) S(q) (A) (C) S(q) molecules. Our work would be useful for the V. ACKNOWLEDGEMENT study of 1-alkanols and its anesthetic effect. AP gratefully acknowledges the hospitality of generous the computing facilities of 'RRIHPC' clusters at RRI, Bangalore, India and Columbia University, New York. The author also thanks the support of TCSC - Tampere Center for Sci- entific Computing resources, Finland. [1] P. Seeman, Pharmacol Rev. 24, 583 (1972). [11] M. Patra, E. Salonen, E. Terama, I. Vat- [2] N. P. Franks and W. R. Lieb, Nature 367, 607 tulainen, R. Faller, B. W. Lee, J. Holopainen, (1997). and M. Karttunen, Biophys. J. 90, 1121 [3] N. P. Franks and W. R. Lieb, Nat. Med. 3, 377 (2006.). (1997). [12] J. T. Marques, A. S. Viana, and F. M. Ro- [4] S. Turkyilmaz, W. H. Chen, H. Mitomo, and drigo de Almeida, Biochim. Biophys. Acta. S. L. J. Regen, J. Am. Chem. Soc. 131, 5068 1808, 405 (2011). (2009). [13] I. Vorobyov, W. F. Drew Bennett, P. D. Tiele- [5] I. Ueda and A. Suzuki, Biophys. J. 75, 1052 man, T. W. Allen, and S. Noskov, J. Chem. (1998). Theory. Comput. 8, 307 (2012). [6] I. Ueda and A. Suzuki, Keio J.Med. 50, 20 [14] R. S. Cantor, Biochem. 36, 2339 (1997). (2001). [15] R. S. Cantor, J. Phys. Chem. B 101, 1723 [7] J. Alifimoff, L. Firestone, and K. W. Miller, (1997). Br J Pharmacol. 96, 9 (1989). [16] R. S. Cantor, Toxicol. Lett. 100, 451 (1998). [8] S. E. Feller, C. A. Brown, D. T. Nizza, and [17] J. Chanda and S. Bandyopadhyay, Chem. K. Gawrisch, Biophys. J. 82, 15 (2002). Phys. Lett. 392, 249 (2004). [9] L. Holte and K. Gawrisch, Biochem. 36, 4669 [18] J. Chanda and S. Bandyopadhyay, Langmuir (1997). 22, 3775 (2006). [10] J. A. Barry and K. Gawrisch, Biochem. 34, [19] M. Kranenburg and B. Smit, FEBS Lett. 568, 8852 (1995). 15 (2004). 16 [20] E. Terama, O. H. Ollila, E. Salonen, A. C. M. Rao, and S. Mayor, Cell 161 (3), 581 Rowat, C. Trandum, P. Westh, M. Patra, (2015). M. Karttunen, and I. Vattulainen, J. Phys. [32] S. Mayor and M. Rao, Traffic 5(4), 231 (2004). Chem. B. 112, 4131 (2008). [33] J. F. Hancock, Nature Reviews Molecular Cell [21] B. Griepernau, S. Leis, M. F. Schneider, Biology 7, 456 (2006). M. Sikor, D. Steppich, and R. A. Bockmann, [34] P. Sharma, R. Varma, R. Sarasij, Ira, K. Gous- Biochim. Biophys. Acta. 1768, 2899 (2007). set, G. Krishnamoorthy, M. Rao, and [22] R. Dickey, A. N.and Faller, Biophys. J. 92, S. Mayor, Cell 116(4), 577 (2004). 2366 (2007). [35] D. Goswami, K. Gowrishankar, S. Bilgrami, [23] U. Vierl, L. Lobbecke, N. Nagel, and G. Cevc, S. Ghosh, R. Raghupathy, R. Chadda, R. Vish- Biophys. J. 67, 1067 (1994). wakarma, M. Rao, and S. Mayor, Cell 135, [24] T. Heimburg and A. D. Jackson, Biophys. J. 1085 (2008). 92, 31593165 (2007). [36] K. Gowrishankar, S. Ghosh, S. Saha, C. Ru- [25] J. A. Barry and K. Gawrisch, Biochem. 34, mamol, S. Mayor, and M. Rao, Cell 149, 8852 (1995). 13531367 (2012). [26] C. W. Hollars and R. C. Dunn, Biophys. J. 75, [37] E. Lindahl, B. Hess, and D. van der Spoel, 342 (1998). Journal of Molecular Modeling 7, 306 (2001). [27] S. A. Simon and T. J. Mcintosh, Biochim. Bio- [38] D. Scherfeld, N. Kahya, and P. Schwille, Bio- phys. Acta. 773, 169 (1984). phys J. 85, 3758 (2003). [28] K. Simons and E. Ikonen, Nature 387, 569 [39] J. de Joannis, P. S. Coppock, F. Yin, M. Mori, (1997). A. Zamorano, and J. T. Kindt, J Am Chem [29] D. Lingwood and K. Simons, Science 327, 46 Soc. 133, 3625 (2011). (2010). [40] A. Polley, S. Vemparala, and M. Rao, J Phys [30] K. Simons and D. Toomre, Nat Rev Mol Cell Chem B. 116(45), 13403 (2012). Biol 1, 31 (2000). [41] A. Polley and S. Vemparala, Chem Phys [31] R. Raghupathy, A. A. Anilkumar, A. Pol- Lipids. 166, 1 (2013). ley, P. P. Singh, M. Yadav, C. Johnson, [42] D. P. Tieleman and H. J. Berendsen, Biophys S. Suryawanshi, V. Saikam, S. D. Sawant, J. 74(6), 2786 (1998). A. Panda, Z. Guo, R. A. Vishwakarma, 17 [43] P. S. Niemela, S. Ollila, M. T. Hyvnen, [48] L. Martnez, R. Andrade, E. Birgin, and M. Karttunen, and I. Vattulainen, PLoS Com- J. Martnez, J Comput Chem. 30(13), 2157 put Biol. 3(2), e34 (2007). (2009). [44] B. Griepernau and R. A. B´'ockmann, Biophys [49] M. Patra and M. Karttunen, J. Phys. Chem. J. 95, 5766 (2008). B 108 (14), 4485 (2004). [45] M. Patra, E. Salonen, E. Terama, I. Vat- [50] A. Polley, S. Mayor, and M. Rao, J. chem. tulainen, R. Faller, B. W. Lee, J. Holopainen, physics. 141, 064903 (2014). and M. Karttunen, Biophys. J. 90, 1121 [51] G. Khelashvili, B. Kollmitzer, P. Heftberger, (2006). Pabst, G., and D. Harries, J Chem Theory [46] R. Reigada, J. Phys. Chem. B. 115, 2527 Comput. 9, 38663871 (2013). (2011). [47] R. Reigada, PLoS Comput Biol. 8, e52631 (2013). 18 Supplementary Information Partitioning of 1-alkanols in composite raft-like lipid membrane Anirban Polley Fig. S1 Snapshot of initial configuration of the symmetric bilayer membranes comprising DOPC (purple), DPPC (orange), Chol (green), water (cyan) (A) without 1-alkanol, (B) with Ethanol (ETH) (red), (C) Pentanol (PCT) (blue) and (D) Octanol (OCT) (gray) are shown respectively. arXiv:1806.02258v1 [physics.bio-ph] 6 Jun 2018 S1 (A) (B) (C)
1505.06108
2
1505
2015-07-17T07:45:48
General formulation of Luria-Delbr{\"u}ck distribution of the number of mutants
[ "physics.bio-ph", "physics.data-an", "q-bio.PE" ]
The Luria-Delbr{\"u}ck experiment is a cornerstone of evolutionary theory, demonstrating the randomness of mutations before selection. The distribution of the number of mutants in this experiment has been the subject of intense investigation during the last 70 years. Despite this considerable effort, most of the results have been obtained under the assumption of constant growth rate, which is far from the experimental condition. We derive here the properties of this distribution for arbitrary growth function, for both the deterministic and stochastic growth of the mutants. The derivation we propose uses the number of wild type bacteria as the independent variable instead of time. The derivation is surprisingly simple and versatile, allowing many generalizations to be taken easily into account.
physics.bio-ph
physics
General formulation of Luria-Delbrück distribution of the number of mutants. Bahram Houchmandzadeh CNRS, LIPHY, F-38000 Grenoble, France Univ. Grenoble Alpes, LIPHY, F-38000 Grenoble, France The Luria-Delbrück experiment is a cornerstone of evolutionary theory, demonstrating the ran- domness of mutations before selection. The distribution of the number of mutants in this experiment has been the subject of intense investigation during the last 70 years. Despite this considerable ef- fort, most of the results have been obtained under the assumption of constant growth rate, which is far from the experimental condition. We derive here the properties of this distribution for arbitrary growth function, for both the deterministic and stochastic growth of the mutants. The derivation we propose uses the number of wild type bacteria as the independent variable instead of time. The derivation is surprisingly simple and versatile, allowing many generalizations to be taken easily into account. I. INTRODUCTION. The Darwin-Wallace theory of evolution rests upon mutation of living organisms and their selection. In their seminal article [1], Luria and Delbrück (LD) described an experiment demonstrating the randomness of mutational events before the selection process. The experiment con- sists of growing C cultures of bacteria in parallel in iden- tical environments, beginning with a small number N0 (typically 103) in each batch. After a sufficient growth period, the cultures saturate and the number of wild type bacteria reaches N (typically 109 − 1010). Each culture is then tested against an antibacterial agent, a phage virus in the LD case, and the number of surviving bacteria aris- ing from mutation in the cultures is counted by a plating method. If C is large, the probability P (m) of having m mutants can be experimentally determined; in practice, C cannot be large and therefore only statistical quanti- ties such as the mean and the variance of the number of mutants can be estimated. The LD experiment has spurred a large interest and many authors have developed increasingly refined mod- els to estimate statistical properties of the random vari- able X of the number of mutants, such as its cumu- lant/probability/moment generating functions (cgf, pgf, mgf), from which the mutation rate or probability can be estimated. The pioneering authors were Lea and Coulson[2], Armitage [3], Bartlett[4], Crump and Hoel[5], Mandelbrot[6], Sarkar, Ma and Sandri[7] who set the LD distribution on a solid mathematical ground, generalized the model to take into account stochastic growth of the mutant and the wild type, and developed algorithms to estimate the mutation rate. The works of these and other authors have been reviewed in an elegant article by Zheng[8] which also contains original results and cor- rections of some of the errors contained in the previous works. These investigations have been extended during the last 15 years by authors such as Angerer[9], Dewanji et al[10], Ycart[11], Kessler and Levine[12, 13]. A descrip- tion of some of these more recent works will be given in the following sections. The fundamental LD experiment is now currently used to estimate mutation rates in various setups such as an- tibiotic resistance or experimental investigation of the evolutionary process[14 -- 16]. However, nearly all of the existing computations have focused on the exponential growth (either deterministic or stochastic) of the wild type and mutant bacteria, al- though Dewanji et al.[10] have extended these results to Gompertz growth. The reason behind this choice is that in these formulations, an explicit expression for the num- ber of wild type (WT) cells as a function of time is needed in order to compute the statistical properties of the num- ber of mutants. The assumption of constant growth rate is however too restrictive. Experimentally, the growth is never exponen- tial but follows a Monod curve[17] : the growth rate is not constant, but begins with a value close to zero (called the lag phase), increases gradually to a maximum value and then decreases as the number of bacteria increases, to finally reach zero when the culture is saturated. Var- ious functions (logistic, Gompertz, Richards, Stannard, ...) are used in the literature to model the growth curve and their relevance has been studied in depth by Zwi- eteting et al. [18]. The real time however is not the relevant independent variable in terms of which the system may be described. The WT population grows from an initial number of cells N0 to reach a final value N . Each time a WT cell divides, there is a small probability ν that a mutant having the desired trait (phage or antibiotic resistance) appears. It does not matter how much time the system spends be- tween WT population size n and n + 1, but only the fact that once a division has taken place, a mutation may ap- pear. For the mutants growing in the same environment as the WT, their growth curve will be similar (but not necessarily equal) to that of the WT. The only quanti- ties that are indeed measured in an experiment are the number of mutants m, the initial WT population N0 and the final population N + m. Even though the growth curve can theoretically be measured, its determination is cumbersome and, as we will see, not relevant. In this paper, we shall use the WT population size n as the independent variable. It appears that this formu- lation of the LD distribution is surprisingly simple and applies to any growth curve for the bacteria, including of course exponential growth. As in the case of classical derivation of the LD distribution, this derivation is valid in the limit of large final population size, which is typi- cal of experiments involving bacteria. The formulation is versatile and can take into account many generalizations, some of which are considered in this article. A similar approach has been taken by Kessler and Levine[12, 13] when solving directly for the Master equation governing the dynamics of the mutant population. This article is organized as follows: in the next section, we present the basic concept for the simple case of deter- ministic growth of both WT and mutant bacteria, where only mutations apparitions are random. We present some generalizations, such as different growth rates for WT and mutants and non-constant mutation probability. In the following section, we generalize the model to the case of stochastic growth of the mutant, where we consider (i) a linear birth process for the mutant and (ii) a random relative growth rate for the mutant. We stress that in the following computations, the growth rate is not con- stant, but can have an arbitrary form. The last section is devoted to a discussion of possible extensions of this work and conclusions. An appendix, containing straight- forward mathematical derivations is included in order to make this article self sufficient. II. DETERMINISTIC MODEL. A. Equal growth of WT and mutant. Consider a culture of WT bacteria growing from size N0 to size N . The growth curve can be as general as pos- sible assuming that no death event takes place. Let Xn be the random variable describing the occurrence of a mu- tant when the WT population increases from n to n + 1. Throughout this paper, we use the term mutant to des- ignate an individual which acquire a trait (e.g. phage or antibody resistance) which will be tested once the growth is stopped. Denoting the mutation probability by ν, Pr(Xn = 0) = 1 − ν ; Pr(Xn = 1) = ν (1) This may seem to be an approximate description, be- cause if during a cell division, a mutation has occurred, obviously the number of WT cells has not increased from n to n + 1. A more precise formulation would be Pr(Xn = k) = (1 − ν)νk, i.e., the number of mutants when the WT population increases by one unit is geo- metrically distributed. However, as ν ≪ 1 (ν is usually of the order of 10−8), we will use the relation (1) to de- scribe the random variable Xn. The generalization to the geometric distribution of Xn is straightforward (see appendix 1). Note that most formulations of LD distri- bution use the above approximation. 2 Figure 1: The number of mutants X at WT population size N is the sum of the contributions of mutants lineages appearing at WT population size ni. The size of the lineage of mutant i appearing at WT population size ni is mi(n) = n/ni for n ≥ ni. The i−th mutant, appearing at size 1 at WT population ni will be present at size N/ni in the final population. We assume in this subsection that both mutant and the WT population follow a deterministic, equal growth. We do not assume the growth rate to be a constant. As the mutants are similar to the WTs, a mutant appearing in one copy at WT population size n will contribute N/n to the number of mutants when the WT population reaches size N (figure 1). In other words, the proportion of the number of this mutant to the number of WT population will remain constant. Let Y N n = (N/n)Xn be the contribution of this mu- tant to the final number of mutants X, when the WT population reaches size N . Then X = N Xn=N0 Y N n As the mutant occurrences are independent random vari- ables, the moment (mgf) and cumulant generating func- tions (cgf) are N φ(s) = (cid:10)esX(cid:11) = ψ(s) = log(cid:0)(cid:10)esX(cid:11)(cid:1) = n E Yn=N0DesY N Xn=N0 N log(cid:16)DesY N n E(cid:17) On the other hand, by its very definition, DesY N n E = 1 − ν + νesN/n which gives the cgf as N ψ(s) = Xn=N0 = N 1 x0 log(cid:16)1 − ν + νesN/n(cid:17) log(cid:16)1 − ν + νes/x(cid:17) dx (2) 3 where in the second expression, we have used the con- tinuous approximation for the sum, x = n/N and x0 = N0/N . The relative error in using the continuous approx- imation is at most O (1/N0 log(x0)). Note that this derivation is analogous to the filtered Poisson process derivation used when the problem is for- mulated in real time ([8],eq. 14). However, because the problem is formulated in terms of WT population size, the propagator is simply a straight line regardless of the growth rate function (figure 1). Expanding the expression (2) to the first order in ν and restricting the domain of definition to s . −x0 log ν, ψ(s) = −θφ + θ 1 x0 es/xdx + O(ν2) (3) Where θ = N ν and φ = 1 − x0. For the particular case of exponential growth, the expression (3) for the cgf has been obtained by Crump and Hoel [5] and in closed form by Zheng ([8],eq.14). Note that for Zheng, N = exp(βt), i.e. the initial number of WT bacteria is 1. The first two cumulant coefficients κp = ψ(p)(0) are then Figure 2: Contribution of mutants with different growth rate. The size of the lineage of mutant i appearing at WT popula- tion size ni is mi(n) = (n/ni)c for n ≥ ni. the population size is n will contribute (N/n)c to the final number of mutants (figure 2). The computation of the preceding section can be repeated and leads to κ1 = −θ log x0 κ2 = θφ x0 (4) (5) ψ(s) = N 1 x0 log(cid:16)1 − ν + νes/xc(cid:17) dx These expressions for the average (κ1 ) and the variance (κ2 ) have been obtained originally by LD for the ex- ponential growth of bacteria [1]. As we see here, the hypothesis of exponential growth is superfluous and the expressions (4-5) are valid for arbitrary growth curves. To the leading order in ν and x0 , the general expression for cumulant coefficients is Keeping only the leading term in ν and x0, for c 6= 1/p, the p−th cumulant coefficient is given by κp = θ cp − 1 (x1−cp 0 − 1) The case c = 1/p can be recovered from the above for- mula by taking the limiting value for c → 1/p and reads κp+1 = θ pxp 0 p > 0 (6) κp = −θ log x0 These expressions are known for the special case of ex- ponential growth ([8],eq.9 for equal growth rate). For the particular case of exponential growth (α(n, t) = α ), this is the expression given by Zheng ([8],eq.9). B. Different growth of WT and mutant. C. Variable mutation probability. Let us now consider the case where WT (n) and mu- tants (m) (once they have appeared) have similar but different growth rates: dn dt = α(n, t)n ; dm dt = cα(n, t)m where c is a constant. We do not specify any particular form for the growth rate, but we suppose that the mu- tant follows the same law as the WT, within a constant multiplicative factor. This is the case for example where the resources are depleted by the growth of the bacteria, and the mutant is inferior to the WT for its duplication. Time can be eliminated between the above equations: dm/dn = c(m/n). A mutant appearing at one copy when In most models the mutation probability ν is consid- ered to be a constant and independent of time, i.e., pop- ulation size. This is a sound hypothesis when the muta- tion involves only point-mutations on the chromosome. However, traits such as virus or antibiotic resistance may involve many point mutations before the trait is func- tional. Bacteria at the end of the growth process, having achieved more divisions, may be more prone to mutate to the given trait than bacteria at the early stage of the growth. A crude approximation of the above phenomena will be a mutation rate that depends on the population size ν = ν(n). The formulation for the number of mu- tants of the previous section does not suppose a constant rate of mutation and the relation 2 for ψ(s) remains valid. tion size N is ψ(s) = N Xn=N0 log(cid:16)1 − ν + νDesKN n E(cid:17) . 4 (8) Knowing the statistical properties of the propagator gives access directly to the statistical properties of the total number of mutants. Before applying this concept to spe- cific cases, let us compute the first two cumulant coeffi- cients: κ1 = ν κ2 = ν N Xn=N0(cid:10)K N n (cid:11) n (cid:1)2E − ν2 Xn=N0D(cid:0)K N N N n (cid:11)2 Xn=N0(cid:10)K N (9) (10) The mean κ1 is what we already had in the deterministic n (cid:11) = N/n. Let us express the second n as a function of its mean and variance case, where (cid:10)K N moment of K N V N n D(cid:0)K N n (cid:1)2E = V N n (cid:11)2 n +(cid:10)K N Then κ2 = ν(1 − ν) N n (cid:11)2 Xn=N0(cid:10)K N + ν N Xn=N0 V N n The first term on the RHS of the above relation is what we already had in the case of deterministic growth. The second term is the contribution of the stochasticity of the propagator to the variance of the number of mutants at population size N . B. Linear birth process. Consider the case where the growth of the WT is de- terministic and continuous dn dt = α(n, t)n while the mutant, once it has appeared, follows a stochas- tic growth with transition probability density W (m → m + 1) = α(n, t)m where W (m → m + 1)dt is the probability that this lin- eage of the mutant has increased its size by one unit in the interval [t, t+dt]. This model was first introduced by Lea and Coulson[2] for exponential growth case α(n, t) = α. A note of caution should be made here. Although widely used, the linear birth model may not be very re- alistic, as bacterial division times are not exponentially distributed. Indeed, after a division, a bacterium needs to elongate again to its original size before being able to Figure 3: The stochastic propagator K N pearing at WT population size n. n of the mutant ap- The first two cumulants are given by κ1 = N 1 κ2 = N 1 x0 x0 ν(x) x dx ν(x) − ν2(x) x2 dx III. STOCHASTIC GROWTH OF MUTANTS. A. General discussion. Until now, we have considered the deterministic prop- agation of the mutant from its appearance at population size n to its final value at population size N . We will denote the propagator as K N n . The contribution of the mutant appearing at WT population size n to the final population N was expressed as Y N n = XnK N n For the deterministic case, K N Y N n was simply n = N/n and the mgf of DesY N n E = 1 − ν + νesN/n We will now consider for the mutant a stochastic prop- n (figure 3). Because Xn takes only the values agator K N 0 or 1, DesY N n E = DesXnKN n E = 1 − ν + νDesKN n E (7) (see appendix 2). Therefore, all the discussion of the pre- ceding section naturally generalizes to stochastic propa- gation and the cgf for the number of mutants at popula- divide again, so that the next division time cannot be smaller than a finite time τ . In fact, the distribution of division times around the time τ is fairly narrow and the division process is much less random than a linear birth process. The phenomenon has been experimentally in- vestigated by a microfluidic device by Wang et al [19] ; the overestimation of the mutation rate by a linear birth model, for the exponential growth case, has been inves- tigated by Ycart [11]. Let us now come back to the linear birth model. As in the previous section, the real time is not the best choice of independent variable and we can write the stochastic growth of the mutant in terms of WT population size: noting W (m → m + 1)dn the probability that a mutant has divided when the WT population size ∈ [n, n + dn], we have W (m → m + 1) = W (m → m + 1) dt dn = m n (11) Note that the equation for the mean of this lineage is d hmi dn = hmi n which conserves the ratio between the size of this lineage and the WT population, in agreement with the deter- ministic case investigated above (subsection II A). The master equation governing the growth of the mu- tant is ∂P (m; n) ∂n = W (m − 1 → m)P (m − 1; n) − W (m → m + 1)P (m; n) (12) and the mgf of the propagator K N pendix 3): n is given by (see ap- 5 growth case can be recovered from the above expression and is equal to the expressions given by Lea and Coulson and Zheng ([8], eq.52-53). Other cumulant coefficients can be readily recovered by multiple derivation of expression (14). Restricting the computations to the leading order of ν and x0, the ex- pression for the cumulant coefficients is (see appendix 4): κp = θ p! (p − 1)xp−1 0 p ≥ 2 (15) Note that even in the case of exponential growth, no general expression for the cumulant coefficients could be obtained by classical methods ([8]). Comparing the above expression with the deterministic case where κp = N ν/(p − 1)xp−1 , we see that indeed the linear birth pro- cess induces large amplification of the p−th cumulant coefficient by a factor of p!. 0 The probabilities. To compute the probabilities, it is more advantageous to use the probability generating function (pgf) G(z) = (cid:10)zX(cid:11) = eψ(z) where ψ(z) is defined from (14) by setting z = es, i.e.: ψ(z) = N 1 x0 log(cid:18)1 − ν + νzx z(x − 1) + 1(cid:19) dx (16) Approximating the above expression to the first order in ν, we have ψ(z) = θ( 1 z − 1) log (1 − φz) + O(ν2) (17) where θ = N ν and φ = 1 − x0. We therefore obtain a simple expression for the probability generating function DesKN n E = nes (n − N )es + N (13) G(z) = (1 − φz)θ(1/z−1) (18) Using now the expression (8) and the continuous variable x = n/N , we obtain the cumulant generating function of the number of mutants ψ(s) = N 1 x0 log(cid:18)1 − ν + νesx es(x − 1) + 1(cid:19) dx (14) Note that the above integral can be expressed in an an- alytical, albeit cumbersome, form. The expressions for the two first cumulant coefficients are κ1 = −θ log x0 κ2 = (2 − ν) + θ log x0 θφ x0 θφ x0 ≈ 2 + θ log x0 For the case of exponential growth, the above expres- sion (without the φ factor) was first discovered by Lea and Coulson[2]; omitting the φ factor however results in divergent moments. The correct expression can be seen in the Zheng review ([8], eq 65). We stress again that relation (18) is very general and does not depend on the assumption of exponential growth. Note that in the case of exponential growth, θ ∼ exp(βt) and all the mo- ments diverge as t → ∞. This divergence, discussed by Bartlett and later by Zheng [8], cannot be cured within the framework of the exponential growth model. No such divergence exists in the present formulation, as the WT population size, following any realistic growth curve, will remain finite. In order to evaluate the probabilities, we have first to compute ψ(p)(0). Expanding expression (17) in powers of z we have Where as before, θ = N ν and φ = 1 − x0. The variance of the number of the mutants is now approximately twice what we had for the deterministic case. The exponential ψ(z) = −θφ + θ ∞ Xn=1 (cid:18) φn n − φn+1 n + 1(cid:19) zn Keeping only the first leading terms in θ and x0 we have 6 ψ(z = 0) = −θφ θ ψ(p)(z = 0) = p(p + 1) − θpx2 0 p + 1 (19) (20) 1 p! where the second term in x2 0 can be neglected for p ≪ 1/x0 and θφ ≈ θ. We can now use the Faà Di Bruno Formula [20] to compute the k−th derivative of G(z) at z = 0 and obtain the probabilities : P (k) = 1 k! G(k)(0) k = e−θ X{mj } Yj=1 1 (mj)! (cid:18) θ j(j + 1)(cid:19)mj Pk where the sum is taken over all k−tuples {mj} such that j=1 jmj = k. The above formula can be easily pro- grammed to compute numerically the probabilities. The only linear term in θ in the above formula is for mk = 1, mi<k = 0. Therefore, for θ ≪ 1, the probabilities take the simple form of P (0) = e−θ P (k) = e−θ θ k(k + 1) k > 0 The explicit expression for the probabilities in the linear birth model has been intensely investigated by many au- thors, and reviewed by Zheng ([8], 5.3) and the above ex- pressions are known for the constant linear birth model. An explicit expression in terms of Landau distribution has been found recently by Kessler and Levine [12, 13] for both the deterministic and stochastic growth model. Experimentally, obtaining the probabilities implies a very large number of parallel cultures and the above ex- pressions may not be of great practical use. Different growth rate. The discussion of the preced- ing subsection can be extended to take into account a different relative growth rate for the mutant compared to the WT: dn dt = α(n, t)n W (m → m + 1) = cα(n, t)m Repeating the discussion of the preceding sections, the cumulant generating function in this case is (see appendix 3) ψ(s) = N 1 x0 log(cid:18)1 − ν + ν xces (xc − 1)es + 1(cid:19) (21) from which the cumulant coefficients can be deduced as before. Figure 4: Deterministic growth, random relative growth rate. The size of the lineage of mutant i appearing at WT popula- tion size ni is mi(n) = (n/ni)c for n ≥ ni. This time however, c is a random variable. C. random relative growth rate. The traditional formulation of LD distribution assumes that all mutants are similar in their growth function. As many different mutations can bring a bacteria to the same phage resistance, this assumption may seem too restric- tive and can be easily relaxed. For example, a compre- hensive study of this phenomenon has recently been pub- lished [21] where the growth rate of all mutants in the gene TEM-1, conferring resistance to the antibiotic ce- fotaxime, where measured and shown to be variable in some conditions. Consider the case where the growth of both mutants and WT are deterministic as in subsection II B dn dt = α(n, t)n ; dm dt = cα(n, t)m but the relative growth rate c is now a random variable (figure 4): when it appears, a mutant picks a relative growth rate c from a given distribution, which is trans- mitted to its progeny. Following the discussion of subsection III A, the prop- agator this time is K N n = (N/n)c where c is now a random variable. Let us denote ρ(s) its moment generating function ρ(s) = hesci Then according to relations (9-10), the first two cumu- lants are now, to the first order in ν: κ1 = θ − log x0 κ2 = θ − log x0 0 0 ρ(z)e−zdz ρ(2z)e−zdz For example, if c follows a Normal distribution c = N (µ, σ), then ρ(s) = exp(µs + (1/2)σ2s2) and these ex- pressions can be evaluated in terms of the error function. For the specific case of µ = 1 and σ ≪ 1, restricting the computation to the leading orders of ν and x0, we have κ1 = −θ log x0(cid:18)1 + σ2 6 (log x0)2(cid:19) κ2 = θ x0 (cid:0)1 + 2σ2(cid:0)1 + (1 + log x0)2(cid:1)(cid:1) IV. DISCUSSION AND CONCLUSION. In the previous sections we have given the general so- lution of the LD distribution through the derivation of its cumulant generating function. The key point to this derivation is to change the independent variable from time to WT population size, which is indeed the relevant variable : whatever the growth function, a mutation can occur only when a WT cell divides and WT population size changes from n to n+1. This consideration consider- ably simplifies the solution of the problem and allows us to extend the solution to arbitrary growth curves for the WT population. The mathematical formulation applies equally well to the case of deterministic and stochastic growth. This mathematical formulation is sufficiently simple to allow for many generalizations, some of which have been considered in this article: for example variable mutation probability (subsection II C) or random relative growth rate (subsection III C) have been investigated. Other generalizations that we have not developed can be considered. For example, for the stochastic growth case, we have only considered the linear birth process. Other, more realistic cases can be envisaged where the distribution of the division times are not exponential, along the lines developed by Ycart[11]. One can also con- sider the case where both mutants and WT grow stochas- tically. These generalizations would be straightforward if the moment generating function of the propagator K N n can be derived in explicit form. Another possible gener- alization would be to take into account the experimen- tal uncertainty on the initial and final value of the WT population, and its influence on the estimation of mu- tation probability, as has been considered by Ycart and Veziris[22]. To summarize, we have developed in this article a ver- satile method for investigating the Luria Delbrück distri- bution with an arbitrary growth function. The method uses only very few measurable parameters, namely the initial and final number of the WT population. We be- lieve that the method we propose here can be used as a simple basis for further investigations of the LD distribu- tion. 7 Appendix: Mathematical details. 1. Geometric distribution of mutants appearing at WT population size n. We have noted ν the probability of apparition of a mutant when a WT cell divides, and Xn the random variable tracking the number of mutants appearing when WT population changes its size from n to n + 1. The probability of producing k mutant during this change is geometrically distributed: Pr(Xn = k) = (1 − ν)νk As ν ≪ 1, we have approximated Xn by a binary process (Xn = 0, 1) in the article. This constraint can be relaxed. Following notations of subsection II A, DesY N n E = 1 − ν 1 − νesN/n and N ψ(s) = n E(cid:17) log(cid:16)DesY N log(cid:18) 1 − ν Xn=N0 = N 1 = −θφ − N 1 x0 1 − νes/x(cid:19) dx log(cid:16)1 − νes/x(cid:17) dx x0 where θ = −N log(1 − ν) and φ = 1 − x0. The above expression is equal to expression (2) to the first order in ν. 2. Moment generating function of Y N n in the stochastic case. Consider the random variable Y N n where Xn is a boolean variable P (Xn = 0) = 1 − ν and P (Xn = 1) = ν (subsection III A); For simplicity, we suppose that K N n is a positive discrete random variable. Then, n = XnK N n = 0) = 1 − ν + νP (K N P (Y N n = k 6= 0) = νP (K N n = k) P (Y N n = 0) The mgf is then (cid:10)esYn(cid:11) = ∞ Xk=0 P (Yn = k)esk = 1 − ν + νP (K N n = 0) + ∞ Xk=1 νP (K N n = k)esk n E = 1 − ν + νDesKN which is the expression (7). 3. Moment generating function of the propagator K N n . This equation can be transformed into equation (A.1) by a simple scaling n → nc and the mgf is therefore given by 8 A mutant appearing in one copy when the WT popu- lation size is n0 will reach size K n n0 when the WT popu- lation reaches size n. The forward master equation gov- erning the probabilities of the propagator, Pr(K n n0 = m) is given by equation (12). The mgf of the propagator, φ(s, n) therefore obeys the following equation : n ∂φ ∂n + (1 − es) ∂φ ∂s = 0 (A.1) with the boundary conditions φ(s, n = n0) = es φ(s = 0, n) = 1 Equation (A.1) is a linear first order partial differential equation that can be solved by the method of character- istics: φ(s, n) = (n0/n)es (n0/n − 1)es + 1 Changing now the notation to denote n as the WT popu- lation size of the mutant occurrence, and N the final size of the WT population, we obtain the expression (13). If the relative growth rate of the mutant is not 1 but c, where c is an arbitrary constant, the transition proba- bility for the mutant once it has appeared is W (m → m + 1) = c m n and φ(s, n) obeys the following equation: 1 c n ∂φ ∂n + (1 − es) ∂φ ∂s = 0 φ(s, n) = (n0/n)ces ((n0/n)c − 1) es + 1 from which the cumulant generating function of the num- ber of mutants can be deduced. 4. Cumulant coefficients for the stochastic growth. To the first order in ν, the cgf for the linear birth model (subsection III B) is given by 1 θ ψ(s) = (e−s − 1) log (1 − es + x0es) Expanding the above function into powers of (1 − e−s), we have ψ(s) = (s + log x0)(e−s − 1) + 1 (1 − e−s)n n − 1 xn−1 0 Evaluating the p − th derivative at s = 0, we have ∞ Xn=2 ψp(0) = p! (p − 1)xp−1 0 + O( 1 xp−2 0 ) which is the expression given in equation (15). Acknowledgements. I sincerely thank Marcel Vallade, Olivier Rivoire and Erik Geissler for the critical reading of the manuscript. (A.2) [1] S E Luria and M Delbrück. Mutations of Bacteria from Virus Sensitivity to Virus Resistance. Genetics, 28(6):491 -- 511, 1943. [2] D. E. Lea and C. A. Coulson. The distribution of the numbers of mutants in bacterial populations. Journal of Genetics, 49(3):264 -- 285, 1949. [3] P. Armitage. The statistical theory of bacterial popula- tions subject to mutations. J. Roy. Stat.Soc. B, 14:1 -- 33, 1952. [4] M. S. Bartlett. An introduction to stochastic processes. Cambridge Univ Press, London, 1977. [5] Kenny S. Crump and David G. Hoel. Mathematical models for estimating mutation rates in cell populations. Biometrika, 61(2):237 -- 252, 1974. [6] B. Mandelbrot. A population birth and mutation process I. J. Applied Probability, 1974:437, 1974. [7] S. Sarkar, W. T. Ma, and G. v. H. Sandri. On fluctu- ation analysis: a new, simple and efficient method for computing the expected number of mutants. Genetica, 85(2):173 -- 179, 1992. [8] Qi Zheng. Progress of a half century in the study of the Luria-Delbrück distribution. Mathematical Biosciences, 162(1-2):1 -- 32, 1999. [9] Wolfgang P. Angerer. An explicit representation of the Luria-Delbrück distribution. Journal of Mathematical Biology, 42(2):145 -- 174, 2001. [10] A Dewanji, E G Luebeck, and S H Moolgavkar. A gener- alized Luria-Delbrück model. Mathematical biosciences, 197(2):140 -- 52, 2005. [11] Bernard Ycart. Fluctuation analysis: can estimates be trusted? PloS one, 8(12):e80958, 2013. [12] David A Kessler and Herbert Levine. Large popula- tion solution of the stochastic Luria Delbruck evolution model. Proceedings of the National Academy of Sciences of the United States of America, 110(29):11682 -- 7, 2013. [13] David A. Kessler and Herbert Levine. Scaling Solution in the Large Population Limit of the General Asymmetric Stochastic Luria Delbruck Evolution Process. Journal of 9 Statistical Physics, 158(4):783 -- 805, 2014. 1990. [14] W A Rosche and P L Foster. Determining mutation rates in bacterial populations. Methods (San Diego, Calif.), 20(1):4 -- 17, 2000. [15] P D Sniegowski, P J Gerrish, and R E Lenski. Evolution of high mutation rates in experimental populations of E. coli. Nature, 387(6634):703 -- 5, 1997. [16] Csaba Pal, María D Maciá, Antonio Oliver, Ira Schachar, and Angus Buckling. Coevolution with viruses drives the evolution of bacterial mutation rates. Nature, 450(7172):1079 -- 81, 2007. [17] Jacques Monod. The Growth of Bacterial Cultures. An- nual review of microbiology, 3:371 -- 394, 1949. [18] M H Zwietering, I Jongenburger, F M Rombouts, and K van 't Riet. Modeling of the bacterial growth curve. Applied and environmental microbiology, 56(6):1875 -- 81, [19] Ping Wang, Lydia Robert, James Pelletier, Wei Lien Dang, Francois Taddei, Andrew Wright, and Suckjoon Jun. Robust growth of Escherichia coli. Current biology : CB, 20(12):1099 -- 103, 2010. [20] Frank W J Olver, Daniel W Lozier, Ronald F Boisvert, and Charles W Clark. NIST handbook of mathematical functions. Cambridge University Press, 2010. [21] Michael A. Stiffler, Doeke R. Hekstra, and Rama Ran- ganathan. Evolvability as a Function of Purifying Selec- tion in TEM-1 β-Lactamase. Cell, 160(5):882 -- 892, 2015. [22] Bernard Ycart and Nicolas Veziris. Unbiased estimation of mutation rates under fluctuating final counts. PloS one, 9(7):e101434, 2014.
1609.03837
2
1609
2017-03-09T07:50:01
Nonlinear amplitude dynamics in flagellar beating
[ "physics.bio-ph" ]
The physical basis of flagellar and ciliary beating is a major problem in biology which is still far from completely understood. The fundamental cytoskeleton structure of cilia and flagella is the axoneme, a cylindrical array of microtubule doublets connected by passive crosslinkers and dynein motor proteins. The complex interplay of these elements leads to the generation of self-organized bending waves. Although many mathematical models have been proposed to understand this process, few attempts have been made to assess the role of dyneins on the nonlinear nature of the axoneme. Here, we investigate the nonlinear dynamics of flagella by considering an axonemal sliding control mechanism for dynein activity. This approach unveils the nonlinear selection of the oscillation amplitudes, which are typically either missed or prescribed in mathematical models. The explicit set of nonlinear equations are derived and solved numerically. Our analysis reveals the spatiotemporal dynamics of dynein populations and flagellum shape for different regimes of motor activity, medium viscosity and flagellum elasticity. Unstable modes saturate via the coupling of dynein kinetics and flagellum shape without the need of invoking a nonlinear axonemal response. Hence, our work reveals a novel mechanism for the saturation of unstable modes in axonemal beating.
physics.bio-ph
physics
Nonlinear amplitude dynamics in flagellar beating David Oriola,1,∗ Hermes Gadelha,2,3 Jaume Casademunt 1 1Departament de F´ısica de la Mat`eria Condensada Facultat de F´ısica, Universitat de Barcelona Avinguda Diagonal 647, E-08028 Barcelona, Spain. 2Department of Mathematics, University of York, YO10 5DD, UK. 3Wolfson Centre for Mathematical Biology, Mathematical Institute University of Oxford, Oxford OX2 6GG, UK. *Current address: Max Planck Institute of Molecular Cell Biology and Genetics Pfotenhauerstrasse 108, 01307 & Max Planck Institute for the Physics of Complex Systems Nothnitzerstrasse 38, 01187 Dresden, Germany To whom correspondence should be addressed; E-mail: [email protected], [email protected]. 7 1 0 2 r a M 9 ] h p - o i b . s c i s y h p [ 2 v 7 3 8 3 0 . 9 0 6 1 : v i X r a 1 Abstract The physical basis of flagellar and ciliary beating is a major problem in biology which is still far from completely understood. The fundamental cytoskeleton structure of cilia and flagella is the axoneme, a cylindrical array of microtubule doublets connected by passive crosslinkers and dynein motor proteins. The complex interplay of these elements leads to the generation of self-organized bending waves. Although many mathematical models have been proposed to understand this process, few attempts have been made to assess the role of dyneins on the nonlinear nature of the axoneme. Here, we investigate the nonlinear dynamics of flagella by considering an axonemal sliding control mechanism for dynein activity. This approach unveils the nonlinear selection of the oscillation amplitudes, which are typically either missed or pre- scribed in mathematical models. The explicit set of nonlinear equations are derived and solved numerically. Our analysis reveals the spatiotemporal dynamics of dynein populations and flag- ellum shape for different regimes of motor activity, medium viscosity and flagellum elasticity. Unstable modes saturate via the coupling of dynein kinetics and flagellum shape without the need of invoking a nonlinear axonemal response. Hence, our work reveals a novel mechanism for the saturation of unstable modes in axonemal beating. Keywords: Flagellar beating, Dynein, Spermatozoa, Self-organization. 2 1 Introduction Cilia and flagella play a crucial role in the survival, development, cell feeding and reproduction of microorganisms (1). These lash-like appendages follow regular beating patterns which en- able cell swimming in inertialess fluids (2). Bending deformations of the flagellum are driven by the collective action of ATP-powered dynein motor proteins, which generate sliding forces within the flagellar cytoskeleton, named axoneme (3). This structure has a characteristic '9+2' composition across several eukaryotic organisms, corresponding to 9 peripheral microtubule doublets in a cylindrical arrangement surrounding a central pair of microtubules (1). Additional proteins, such as the radial spokes and nexin crosslinkers, connect the central to the periph- eral microtubules and resist free sliding between the microtubule doublets, respectively. Each doublet consists of an A-microtubule in which dyneins are anchored at regular intervals along the length of the doublets, and a B-microtubule, where dynein heads bind in the neighbouring doublet (1, 4). In the presence of ATP, dyneins drive the sliding of neighbouring microtubule doublets, generating forces that can slide doublets apart if crosslinkers are removed (5). In the presence of crosslinkers, sliding is transformed into bending. Remarkably, this process seems to be carried out in a highly coordinated manner, in such a way that when one team of dyneins in the axoneme is active, the other team remains inactive (6). This mechanism leads to the propa- gation of bending undulations along the flagellum, as commonly observed during the movement of spermatozoa (4). Many questions still remain unanswered on how dynein-driven sliding causes the oscillatory bending of cilia and flagella (7). Over the last half a century, intensive experimental and theo- retical work has been done to understand the underlying mechanisms of dynein coordination in axonemal beating. Different mathematical models have been proposed to explain how sliding 3 forces shape the flagellar beat (8–16). Coordinated beating has been hypothesised consider- ing different mechanisms such as dynein's activity regulation through local axonemal curva- ture (9–11,16), due to the presence of a transverse force (t-force) acting on the axoneme (12) and by shear displacements (13–15). Other studies also examined the dynamics of flagellar beating by prescribing its internal activity (17, 18) or by considering a self-organized mechanism inde- pendent of the specific molecular details underlying the collective action of dyneins (13,14,19). In particular, the latter approach, although general from a physics perspective, it does not ex- plicitly incorporate dynein kinetics along the flagellum, which has been shown to be crucial in order to understand experimental observations on sperm flagella (20–22). Load-accelerated dis- sociation of dynein motors was proposed as a mechanism for axonemal sliding control, and was successfully used to infer the mechanical properties of motors from bull sperm flagella (15). In contrast, dynamic curvature regulation has been recently proposed to account for Chlamy- domonas flagellar beating (16). In the previous studies, linearized solutions of the models were fit to experimental data; however, it is unclear that such results still hold at the nonlin- ear level. Recent studies also investigated the emergence and saturation of unstable modes for different dynein control models (23, 24); however, saturation of such unstable modes was not self-regulated, but achieved via the addition of a nonlinear elastic contribution in the flagellum constitutive relation. Nevertheless, predictions on how dynein activity influences the selection of the beating frequency, amplitude and shape of the flagellum remain elusive. Here, we pro- vide a microscopic bottom-up approach and consider the intrinsic nonlinearities arising from the coupling between dynein activity and flagellar shape, regarding the eukaryotic flagellum as a generalized Euler-elastica filament bundle (25). This allows a close inspection on the onset of the flagellar bending wave instability, its transient dynamics and later saturation of unstable modes, which is solely driven by the nonlinear interplay between the flagellar shape and dynein kinetics. 4 We first derive the governing nonlinear equations using a load-accelerated feedback mech- anism for dynein along the flagellum. The linear stability analysis is presented, and eigenmode solutions are obtained similarly to Refs. (15, 24), to allow analytical progress and pedagogi- cal understanding. The nonlinear dynamics far from the Hopf bifurcation is studied numeri- cally and the resulting flagellar shapes are further analyzed using principal component analy- sis (26, 27). Finally, bending initiation and transient dynamics are studied subject to different initial conditions. 2 Continuum flagella equations We consider a filament bundle composed of two polar filaments subjected to planar deforma- tions. Each filament is modelled as an inextensible, unshearable, homogeneous elastic rod, for which the bending moment is proportional to the curvature and the Young modulus is E. The filaments are of length L and separated by a constant gap of size b, where b (cid:28) L (Fig. 1c). We define a material curve describing the shape of the filament bundle centerline as r(s, t). The positions of each polar filament forming the bundle read r± = r ± (b/2)n, with the orientation of the cross-section at distance s along its length defined by the normal vector to the centerline n = − sin φ ı + cos φ , being φ ≡ φ(s, t) the angle between the tangent vector s ≡ ∂sr ≡ rs and the ı direction (taken along the x axis). The subscripts (+) and (-) refer to the upper and lower filaments, respectively (Fig. 1c). The shape of the bundle is given at any time by the (cid:90) s 0 5 expression: r(s, t) = r(0, t) + (cos φ, sin φ)ds(cid:48) (1) denoted as sliding displacement: ∆(s, t) = (cid:90) s 0 The geometrical constraint of the filament bundle, induces an arc length mismatch ∆(s, t), (∂sr− − ∂sr+)ds(cid:48) = b(φ − φ0) (2) where φ0 ≡ φ(0, t). For simplicity, we have set any arc length incongruity between the two filaments at the base to zero and we will consider the filaments clamped at the base. A similar approach can be used to include basal compliance and other types of boundary conditions at the base (e.g. pivoting or free swimming head). Here we centre our study on the nonlinear action of motors along the flagellum. We aim to study the active and passive forces generated at each point along the arc length of the filament bundle. We define f (s, t) = f (s, t)s as the total internal force density generated at s at time t on the plus-filament due to the action of active and passive forces (see Fig. 1). By virtue of the action-reaction law, the minus-filament will experience a force density −f at the same point. Next, consider that N dyneins are anchored at each polar filament in a region lc around s, where lc is much smaller than the length of the flagellum L and much larger than the length of the regular intervals dyneins are attached to along the microtubule doublets. We shall call lc the 'tug-of-war' length. We define n±(s, t) as the number of bound dyneins in a region of size lc around s at time t which are anchored in the plus- or minus-filament respectively. We consider a tug-of-war at each point s along the flagellum with two antagonistic groups of N dyneins. The elastic sliding resistance between the two polar filaments exerted by nexin crosslinkers is assumed to be Hookean with an elastic modulus K. Thus, the internal force density f (s, t) reads: f (s, t) = ρ(n+F+ + n−F−) − K∆ (3) where ρ ≡ l−1 per motor each group of dyneins experiences due to the action of the antagonistic group. The is the density of tug-of-war units along the flagellum and F±(s, t) is the load c stresses on the filament bundle are given by a resultant contact force N (s, t) and resultant 6 Figure 1: (Online version in colour) Schematic view of the system a) Two-dimensional repre- sentation of a flagellum, where the centerline r(s) and tangent angle φ(s) of the flagellum are parametrized by the arc length parameter s. b) Passive (springs) and active (dynein motors) internal structures in the axoneme. Dyneins in the (+) and (-) filaments compete in a tug-of-war and bind/unbind from filaments with rates π± and ε± respectively. c) The flagellum as a two- filament bundle: two polar filaments are separated by a small gap of size b (cid:28) L. The presence of nexin crosslinkers and dynein motors generates a total force density f (s) along the bundle. contact moment M (s, t) acting at the point r(s, t). The internal force density f (s, t) only contributes to the internal moment of the bundle M (s, t) = M k, where k =  × ı, such that M (s, t) reads: M (s, t) = Ebφs − bF 7 (4) r+r⇡+xybr+rOffac(s)r(s)s=Ls=0b⇡r(s)b"+"sn where F (s, t) =(cid:82) L s f (s(cid:48), t)ds(cid:48), provided that ∂sφ± ≈ ∂sφ for bundles characterized by b (cid:28) L. The combined bending stiffness of the filament bundle is given by Eb = 2EI, where I is the second moment of the area of the external rods. Dynein kinetics is modelled by using a minimal two-state mechanochemical model with states k = 1, 2, corresponding to microtubule bound or unbound dyneins, respectively. Since the sum of bound and unbound motors at s remains constant at all times, we only study the plus and minus bound motor distributions n±(s, t). Dyneins bind with rates π± and unbind with rates ε± (Fig. 1b). The corresponding bound motor population dynamics reads: ∂tn± = π± − ε± (5) The binding/unbinding rates are given by π± = π0(N − n±), ε± = ε0n± exp(±F±/fc) where ε0 and π0 are constant rates and fc is the characteristic unbinding force. Motivated by previous studies on the collective action of molecular motors (15, 28), we assume an exponential depen- dence of the unbinding force on the resulting load. By considering that dyneins fulfill a linear velocity-force relationship with stall force f0 and velocity at zero load v0, the loads are given by F±(s, t) = ±f0(1 ∓ ∆t/v0). The stall force f0 is defined as the absolute value of the load a motor experience at stall (∆t = 0). Substituting the different definitions, the internal force density f (s, t) reads: (cid:18) (cid:19) n(cid:48)∆t v0 f (s, t) = f0ρ ¯n − − K∆ (6) where ¯n ≡ n+ − n− and n(cid:48) ≡ n+ + n−. For simplicity, we will derive the equations governing the tangent angle φ in the limit of small curvature (but possibly large amplitudes) such that tan- gential forces can be neglected. The derivation for arbitrary large curvature is also presented in the electronic Supplementary Text. Using resistive force theory in the limit of small curvature, we only consider normal forces along the flagellum obtaining ζ⊥φt = −Msss, where ζ⊥ is the 8 normal drag coefficient (8, 29). Combining the last expression with Eq. 4 we have: ζ⊥φt = −Ebφssss − bfss (7) Hereinafter we switch to dimensionless quantities while keeping the same notation. We non- dimensionalize the arc length with respect to the length scale L, time with respect to the correla- tion time of the system τ0 = 1/(ε0 +π0), motor number with respect to N, internal force density with respect to f0ρN and sliding displacement with respect to b. The correlation time defines how fast the motors will respond to a change in load. We also define Sp = L(ζ⊥/Ebτ0)1/4, µ ≡ Kb2L2/Eb, µa = bf0ρN L2/Eb and ζ ≡ b/v0τ0. The sperm number Sp characterizes the relative importance of bending forces to viscous drag. The parameter µ measures the relative importance of the sliding resistance compared with the bending stiffness (25). On the other hand, the parameter µa denotes the activity of dyneins, measuring the relative importance of motor force generation compared with the bending stiffness of the bundle. Finally ζ denotes the ratio of the bundle diameter and the characteristic shear induced by the motors. The dimen- sionless sperm equation in the limit of small curvature reads (2, 13, 18): where in our case the dimensionless internal force density f (s, t) takes the form: Sp4φt = −φssss − µafss f (s, t) = ¯n − ζn(cid:48)∆t − µ µa ∆ (8) (9) and ∆ = φ − φ0. Since the flagellar base is clamped, without loss of generality, we set φ0 = 0. Combining Eqs. 8 and 9 we obtain the nonlinear dynamics for the tangent angle: Sp4φt = −φssss + µφss − µa¯nss + µaζ[n(cid:48) ssφt + 2n(cid:48) sφts + n(cid:48)φtss] (10) In the absence of dynein activity, the last expression reduces to the dynamics of an elastic filament bundle with sliding resistance forces (25). Notice that this expression is obtained con- sidering the sliding mechanism and a linear velocity-force relationship for dyneins, but it is 9 independent of dynein kinetics. On the other hand, the dimensionless form of the bound motor population dynamics n± reads: ∂tn± = η(1 − n±) − (1 − η)n± exp[ ¯f (1 ∓ ζφt)] (11) where η ≡ π0/(π0 + ε0) is the duty ratio of the motors and ¯f ≡ f0/fc dictates the sensitivity of the unbinding rate on the load. In the Supplementary Text we include a table with a summary of all the variables and parameters in the model with their corresponding symbols. 3 Parameter values We present the choice of parameters based on experimental studies on sperm flagella and on the green algae Chlamydomonas. We first discuss the passive properties of a flagellum. The typical length of a human flagellum is L (cid:39) 50 µm and the axonemal diameter is found to be b (cid:39) 200 nm (4). The bending stiffness of the filament bundle has been reported to be Eb (cid:39) 0.9 · 10−21 N·m2 for sea-urchin sperm (4, 25) and Eb (cid:39) 1.7 · 10−21 N·m2 for bull sperm (15). On the other hand, the interdoublet elastic resistance from demembranated flagellar axonemes of Chlamy- domonas yields an estimated spring constant 2 · 10−3 N/m for 1 µm of axoneme (30), thus K (cid:39) 2· 103 N/m2. Finally, typical medium viscosities for sperm flagella range from ζ⊥ (cid:39) 10−3 Pa·s in low viscous media to ζ⊥ (cid:39) 1 Pa·s in high viscous media (4, 18). Next, we discuss the mechanochemical parameters associated to axonemal dynein. Axone- mal dynein are subdivided in inner and outer arms depending on its position in the axoneme, and can be found in heterodimeric and monomeric forms (22). For the sake of simplicity, we consider identical force generating dynein motor domains acting along the flagellum. The to- tal number of motor domains in a beating flagellum has been estimated to be (cid:39) 105 (31, 32). The stall force has been found in the range f0 (cid:39) 1 − 5 pN (33, 34). Following Ref. (15) we 10 choose the characteristic unbinding force for dynein such that ¯f = 2. Axonemal dynein is characterized by a low duty ratio estimated to be η (cid:39) 0.14 and speeds at zero load in the range v0 (cid:39) 5 − 7 µm/s (33, 35). The characteristic time scale of dynein kinetics sets the order of magnitude of the beating frequency in our model. We choose an estimated value of τ0 (cid:39) 50 ms (24, 35), which corresponds to a frequency of (cid:39) 10 Hz, comparable to the case of human sperm (4). Finally, we need to estimate ρ and N. Considering the length of the human sperm flagellum (L (cid:39) 50 µm) we obtain (cid:39) 2 · 103 motors/µm. In our description, we divide the axoneme in two regions with corresponding dynein teams. Therefore, we have ρN (cid:39) 103 mo- tors/ µm. In order to find ρ, we need to choose a criterion to decide the typical length scale or 'tug-of-war' length lc = ρ−1 in our coarse-grained description. The typical length scale in the system is given by lc ∼ L/√µa =(cid:112)Eb/bf0ρN (see Linear stability analysis). Using the previous parameters we get lc ∼ 1 µm and therefore ρ ∼ 1 µm−1 and N ∼ 103. Hence, we obtain Sp (cid:39) 5 − 20, µ (cid:39) 50 − 100, µa ∼ 103 and ζ ∼ 1. The motor activity is studied in a broad range µa (cid:39) (2− 6)· 103 since it plays the role of the main control parameter in our study. 4 Linear stability analysis In this section, we perform a linear stability analysis of Eqs. 10 and 11. The nonmoving state is characterized by φ = 0 and n± = n0 ≡ π0/(π0 + ε0e ¯f ). This means that the flagellum is aligned with respect to the x-axis and the number of plus and minus bound motors is constant in space and time. For the linear stability analysis, we consider the perturbed variables around the base state as φ = δφ and n± = n0 + δn±. Introducing the modulation δn ≡ δn+ = −δn− 11 around n0 and considering ¯f ζφt (cid:28) 1 we obtain: δnt = −¯τ−1δn + (1 − η)ζ ¯f e ¯f n0φt Sp4φt = −φssss + µφss + 2µa[ζn0φtss − δnss] (12) (13) where ¯τ ≡ n0/η. We use the ansatzs φ = φ(s)eσt + c.c and δn = δn(s)eσt + c.c, where σ is a complex eigenvalue and c.c accounts for complex conjugate. From Eq. 12 we get δn = χ(cid:48)(σ) φ, where χ(cid:48)(σ) is a complex response function: χ(cid:48)(σ) = ζ ¯f n0(1 − n0) σ 1 + σ¯τ (14) Using Eq. 9 and considering f = f (s)eσt + c.c, we obtain f = χ(σ) φ, where χ(σ) is a second complex response function: χ(σ) = 2ζn0 (cid:20) ¯f (1 − n0) σ − σ2¯τ 1 − (σ¯τ )2 − σ (cid:21) µ µa − (15) The latter response functions generalize the work in Ref. (15) for a complex eigenvalue σ and are equivalent to results presented in Ref. (24). With the ansatz φ ∼ δn ∼ eiqs in Eq. 13, we obtain the characteristic equation q4 − ¯χq2 + ¯σ = 0, where ¯χ ≡ µaχ, ¯σ ≡ σSp4 and ¯χ, ¯σ ∈ C. Solving the characteristic equation we obtain four possible roots qi, i = 1, . . . , 4 and the eigenfunctions read: 4(cid:88) φ(s) = Φjeiqj s (16) j=1 where qj, Φj ∈ C. Once φ is known, δn(s) = δN φ(s) exp(i∆θ) where δN = χ(cid:48) and ∆θ = arg(χ(cid:48)). Therefore, the evolution of δn is the same as for φ except for a phase shift ∆θ and an overall change on the amplitude δN, which depends on χ(cid:48)(σ). This result indicates the presence of a time delay between the action of motors and the response of the flagellum. Time delays commonly arise in systems where molecular motors work collectively (38). Indeed, the regulation of active forces by the time delay of the curvature was proposed as a mechanism to 12 Figure 2: (Online version in colour) Linear stability analysis. a,b) (Upper panels) Clamped head profile solutions corresponding to the marginal stability case (i.e. λ1 = 0) for Sp = 5, 10. The beating cycles are divided in 10 frames. (Lower panels) Tangent angle kymographs where T = 2π/ωc is the period. Amplitudes and angles are shown in arbitrary units. c) Marginal stability curve for the clamped condition. Points (a) and (b) in parameter space correspond to the profiles in a) and b). d) Curvature modulations as a function of the arc length for Sp = 5 (dashed line) and Sp = 12 (solid line). Inset: Wavenumber defined as k ≡ maxqi/2π as a function of Sp. µ = 50, ζ = 0.4, η = 0.14, ¯f = 2. 13 68101212345600.20.40.60.8112345681012123abc 12345601 12345601(cid:239)0.100.1(cid:239)0.100.1 12345601 12345601(cid:239)0.100.1(cid:239)0.100.1 12345601 12345601(cid:239)0.100.1(cid:239)0.100.1 12345601 12345601(cid:239)0.100.1(cid:239)0.100.1sst/Tt/TSp=5Sp=10φφd68101212345600.20.40.60.8112345681012123NoinstabilityTravellingwavesSpsφsµaSpkx 103(a)(b) generate travelling bending waves (9). In order to find φ(s), we need to impose the four bound- ary conditions, obtaining a linear system of equations for Φj, j = 1, . . . , 4. The procedure to obtain the set of boundary conditions for a clamped head is detailed in the Supplementary Text. The boundary conditions read φ(0) = 0, φsss(0) = − ¯χ φs(0), φs(1) = 0 and φss(1) = − ¯χ φ(1). By setting the determinant of the system to zero, we find the set of complex eigenvalues σn, with the corresponding growth rates λn = Re[σn] and frequencies ωn = Im[σn], which satisfy the boundary conditions, where λn, ωn ∈ R. We order the set of different eigenvalues according to its growth rate λn+1 < λn, such that the first one has the largest growth rate λ1. Defining u = (φ, δn)T , the general solution of the system reads: (cid:88) n (cid:18) φn (cid:19) δnn u(s, t) = An eλnteiωnt + c.c. (17) where An ∈ C are free amplitude parameters. For λn < 0, ∀n, solutions decay exponentially to the nonmoving state. On the other hand, when λ1 becomes positive, in the range of parameters studied, the system undergoes a Hopf bifurcation and solutions follow an exponential growth, oscillating with frequency ω1. Next, we study the marginal stable solutions, i.e. when the max- imum growth rate equals zero (λ1 = 0). For this, we define the critical frequency of oscillation as ωc ≡ ω1. Travelling waves propagate from tip to base, a feature already reported for the clamped type boundary condition (13, 14, 24). In Fig. 2c, the marginal stability curve in phase space is shown. Intuitively, as Sp is increased the travelling instability occurs for higher motor activity µa and the critical frequency of oscillation ωc follows a non-monotonic decrease (see Supplementary Figure S1). For low viscosity (Sp = 5) the wave propagation velocity is slightly oscillatory whereas for high viscosity (Sp = 10) it becomes more uniform (Fig. 2a and b, lower panels). These results are in agreement with studies on migrating human sperm, where in the limit of high viscosity waves propagated approximately at constant speed (36). For high viscosity, curvature tends to increase from base to tip, finally dropping to zero due to the zero 14 curvature boundary condition at the tail (see Fig. 2d and Supplementary Text). This modulation is consistent with experimental studies on human sperm, which show viscosity modulation of the bending amplitude (36). In the latter study; however, the effect is more pronounced possibly due to external elastic reinforcing structures found along the flagellum of mammalian species, as well as other nonlinear viscoelastic effects. Defining k ≡ maxqi/2π as the characteristic wavenumber, we obtain that it increases almost linearly with Sp (Fig. 2d, inset). Similar results can be obtained using other definitions for k, for example using the covariance matrix (see Prin- cipal component analysis section). 5 Nonlinear flagellar dynamics In this section, we study the nonlinear dynamics of the flagellum in the limit of small curvature by numerically solving Eqs. 10 and 11 using a second-order accurate implicit-explicit numer- ical scheme (see Supplementary Text). The unstable modes presented in Section 4 follow an initial exponential growth and eventually saturate at the steady state due to the nonlinearities in the system. In Fig. 3 two different saturated amplitude solutions are shown. Fig. 3a (left) corresponds to a case where the system is found close to the Hopf bifurcation, whereas Fig. 3a (right) corresponds to a regime far from the bifurcation. We notice that the marginal solution obtained in the linear stability analysis (Fig. 2b, upper panel) gives a very good estimate of the nonlinear profile close to the bifurcation point, although it does not provide the magnitude of φ or δn. Frequencies are ∼ 10 Hz and maximum amplitudes are found to be small, around (cid:39) 4% of the total flagellum length. However, the oscillation amplitude for high motor activity is more than double in respect to the case of low activity (Fig. 3a). The colour code in Fig. 3a indi- cates the value of the semi-difference of plus- and minus-bound motors δn. Plus-bound motors are predominant in regions of positive curvature (φs > 0) along the flagellum and vice-versa. 15 Despite the low duty ratio of dynein motors (35), ∼ 2% bound dyneins along the flagellum Figure 3: (Online version in colour) Nonlinear analysis. a) Nonlinear flagella steady state profiles for µa = 4300 (left) and µa = 5600 (right), considering the respective eigenmodes as initial conditions. The beating cycles are divided in 10 frames as in Fig. 2a,b (upper panels). b) φ (solid lines) and δn (dashed lines) evaluated at s = 3/4 for the profiles in (a) respectively. c) Maximum absolute tangent angle evaluated at s = 1/2 and dimensionless frequency ω as a function of the relative distance to the bifurcation ε = (µa − µc a. Sp = 10, µ = 50, η = 0.14, ζ = 0.4, ¯f = 2, L = 50 µm and τ0 = 50 ms. a)/µc 16 00.20.400.050.10.15εφmax(1/2)00.20.42.533.5εω100200300400500(cid:239)0.200.2(cid:239)101200300400500(cid:239)0.100.1(cid:239)505x 10(cid:239)3x 10(cid:239)2100200300400500(cid:239)0.200.2(cid:239)101200300400500(cid:239)0.100.1(cid:239)505x 10(cid:239)3x 10(cid:239)202040(cid:239)4(cid:239)2024 (cid:239)0.02(cid:239)0.0100.010.0202040(cid:239)1(cid:239)0.500.51 (cid:239)0.0100.01at(ms)t(ms)bδnX(µm)Y(µm)µa=5600µa=4300Y(µm)X(µm)δnc200400600800(cid:239)0.200.2(cid:239)101400600800(cid:239)0.100.1(cid:239)505x 10(cid:239)3x 10(cid:239)2200400600800(cid:239)0.200.2(cid:239)101400600800(cid:239)0.100.1(cid:239)505x 10(cid:239)3x 10(cid:239)2φ(3/4)δn(3/4)00.20.400.050.10.15εφmax(1/2)00.20.42.533.5εω are sufficient to produce micrometer-sized amplitude oscillations. The full flagella dynamics corresponding to Fig. 3a are provided in the Supplementary Movies 1 and 2. In Fig. 3b, the time evolution of φ and δn is shown at s = 3/4 for the cases in Fig. 3a, respectively. As mentioned in Section 4, the tangent angle φ is delayed respect to δn, and the time delay is not considerably affected by motor activity. Close to the instability thresh- old, both signals are very similar since the system is found near the linear regime; however, far from threshold, both signals greatly differ. For high motor activity, both the tangent angle and the fraction of bound dyneins at certain points along the flagellum exhibit cusp-like oscil- lations (Fig. 3b, right). This behaviour is typical of molecular motor assemblies working far from the instability threshold (37). Despite the signals S in Fig. 3b (right) are nonlinear, they conserve the symmetry S(t + T /2) = −S(t), where T is the period of the signal. This is a consequence of both plus and minus motor populations being identical, a property also found in spontaneous oscillations of motor assemblies (37, 38). Finally, in Fig. 3c we study how the amplitude and frequency of the oscillations vary with the relative distance from the bifurcation a. For small ε, the maximum absolute value of the tangent angle seems point ε ≡ (µa − µc to follow a square root dependence, characteristic of supercritical Hopf bifurcation; however, in a)/µc the strongly nonlinear regime the curve deviates from this trend. On the other hand, the beat- ing frequency decreases for increasing activity. The only possibility to increase µa keeping the other dimensionless parameters fixed is by increasing ρN; hence, the larger each dynein team becomes, the larger the activity. Consequently, the necessary time to unbind a sufficient number of dynein motors to drive the instability increases, leading to a lower beating frequency. 17 5.1 Principal component analysis In this section, we study the obtained nonlinear solutions using principal component analysis (26, 27). This technique treats the flagellar shapes as multi-feature data sets, which can be projected to a lower dimensional space characterized by principal shape modes. Here we will analyze the numerically resolved data following Ref. (26) to study sperm flagella. We discretize our flagella data with time-points ti, i = 1, . . . , p and M = 4 · 103 intervals corresponding to points sj = (j − 1)/M, j = 1, . . . , r along the flagellum, with r = M + 1. We construct a measurement matrix Φ of size p × r for the tangent angle where Φij = φ(sj, ti). This matrix represents a kymograph of the flagellar beat. We define the r×r covariance matrix as C = (Φ− ¯Φ)T (Φ − ¯Φ), where ¯Φ = [ ¯φ; . . . ; ¯φ] with all rows equal to ¯φ = [ ¯φ(s1), . . . , ¯φ(sr)], where ¯φ(si) is the mean tangent angle at si. The covariance matrix C is shown for µa = 4300 (Fig. 4a, left) and µa = 5600 (Fig. 4a, right). In Fig. 4a (left), we find negative correlation between tangent angles that are a distance λ/2 apart. Hence, a characteristic wavelength λ can be identified in the system, which manifests as a long-range correlation in the matrix C. On the other hand, strong positive correlations around the main diagonals correspond to short-ranged correlations mainly due to the bending stiffness of the bundle (26). The number of local maxima along the diagonals in C decreases from µa = 4300 to µa = 5600, and at the same time λ(cid:48) > λ. Hence, an increase in motor activity slightly increases the characteristic wavelength while decreasing the number of local maxima, which is related to the characteristic wavenumber. Employing an eigenvalue decomposition of the covariance matrix, we can obtain the eigenvectors v1, . . . , vr and their corresponding eigenvalues d1, . . . , dr, such that Cvj = djvj. Without loss of generality, we can sort the eigenvalues in descending order d1 ≥ . . . ≥ dr. We find that the first two eigenvalues capture > 99% variance of the data. This fact indicates that our flagellar waves can be suitably described in a two-dimensional shape space, since they can be regarded as single-frequency oscillators. Each flagellar shape Φi = [φ(s1, ti), . . . , φ(sr, ti)] can be expressed now as a linear 18 combination of the eigenvectors vk (26): Φi = ¯φ + r(cid:88) k=1 Bk(ti)vk (18) where Bk are the shape scores computed by a linear least-square fit. In Fig. 4b (left), the two first eigenvectors v1, v2 are shown for µa = 4300. In Fig. 4b (right), the flagellar shape at a certain time (thick solid line) is reconstructed (white line) by using a superposition of the two principal shape modes v1, v2 (solid and dashed lines, respectively) and fitting the scores B1, B2. Finally, in Fig. 4c we show the shape space trajectories beginning with small amplitude eigenmode solutions. While close to the bifurcation the limit cycle is elliptic (Fig. 4c, left), far from the bifurcation the limit cycle becomes distorted (Fig. 4c, right). Elliptic limit cycles were also found experimentally for bull sperm flagella (26). Hence, as found in Section 5, motor activity in the nonlinear regime significantly affects the shape of the flagellum when compared with the linear solutions, which only provide good estimates sufficiently close to the Hopf bifurcation. 5.2 Bending initiation and transient dynamics Finally, we study bending initiation and transient dynamics for two different initial conditions, in order to understand the selection of the unstable modes. In Fig. 5a,b the spatiotemporal transient dynamics are shown for the case of an initial eigenmode solution corresponding to the maximum eigenvalue (Fig. 5a) and an initial sine perturbation in φ, with equal constant bound motor densities (Fig. 5b). In case (b) travelling waves initially propagate in both directions and interfere at t = t(cid:48) (Fig. 5b,d and Supplementary Movie 3). However, in the steady state both the eigenmode and sine cases reach the same steady state solution, despite the sinusoidal initial condition being a superposition of eigenmodes. This result provides a strong evidence that the fastest growing mode is the one that takes over and saturates in the steady state. In Fig. 5c the transient dynamics for case (b) are shown for plus and minus-bound dynein populations close to the tail (s = 9/10). Both populations decay exponentially with characteristic time ¯τ to n0 19 Figure 4: (Online version in colour) Principal component analysis of flagellar beating. a) Co- variance matrix for µa = 4300 (left) and µa = 5600 (right). We can identify characteristic wavelengths λ, λ(cid:48) from negative long-range correlations in C. Notice λ(cid:48) > λ and the number of local maxima decreases when µa is increased. b) (Left) Two principal shape modes v1, v2 (solid and dashed lines, respectively), corresponding to the two maximum eigenvalues of the covariance matrix in Fig 4a (left). (Right) The flagellar shape at time t = 733 ms (thick solid line) is reconstructed (white line) by a superposition of the two principal shape modes v1, v2 in Fig. 4b (left) fitting the scores B1, B2. c) Flagellar dynamics in a reduced two-dimensional shape space for µa = 4300 (left) and µa = 5600 (right). Elliptic limit cycles are rescaled to better appreciate the distortion due to the nonlinear terms. Sp = 10, µ = 50, η = 0.14, ζ = 0.4. 20 ss 00.5100.51(cid:239)10010203000.51(cid:239)0.0500.05sφ00.51(cid:239)0.1(cid:239)0.0500.050.1sφv1v2abc(cid:239)505(cid:239)303B1B2(cid:239)20020(cid:239)10010B1B200.51(cid:239)0.0500.05sφ00.51(cid:239)0.1(cid:239)0.0500.050.1sφφsφsssλ/2ss 00.5100.510200400ssλ\x{FFFF}/2 Figure 5: (Online version in colour) Bending initiation and transient dynamics. a) Tangent an- gle kymograph for an eigenmode initial condition b) Tangent angle kymograph for a sinusoidal initial condition for the tangent angle with n+(0) = n−(0) = 0.1. c) Bound motor time evo- lution for the plus (solid line) and minus (dashed line) dynein populations at s = 9/10 for the case of a sinusoidal initial condition. Inset: Flagella profiles at different times in ms. d) Snap- shots of the flagellar shape for the sinusoidal initial condition up to t = t(cid:48)(cid:48) (white dashed line in Fig. 5b) at equal time intervals (20 ms). At t = t(cid:48), wave interference changes the direction of wave propagation. The full movie can be seen in Supplementary Movie 3. Sp = 5, µ = 100, µa = 2000, η = 0.14, ζ = 0.4, L = 50 µm and τ0 = 50 ms. Arrows indicate the direction of wave propagation. 21 0250500750051002040(cid:239)101 025050075000.51 025050075000.51(cid:239)0.200.2(cid:239)0.200.2 025050075000.51 025050075000.51(cid:239)0.200.2(cid:239)0.200.2φn+n−0150300450X(µm)Y(µm)s0204060(cid:239)0.4(cid:239)0.200.20.40.60204060(cid:239)0.4(cid:239)0.200.20.40.60204060(cid:239)0.4(cid:239)0.200.20.40.60204060(cid:239)0.4(cid:239)0.200.20.40.60204060(cid:239)0.4(cid:239)0.200.20.40.60204060(cid:239)0.4(cid:239)0.200.20.40.60204060(cid:239)0.4(cid:239)0.200.20.40.60204060(cid:239)0.4(cid:239)0.200.20.40.60.5 μm20 μmcabt = 0t = tʼt = tʼʼtʼʼtʼsineeigenmodeφsbound motors (%)dt(ms)t(ms)t(ms) and begin oscillating in anti-phase around this value, in a tug-of-war competition. 6 Discussion In this work, we presented a theoretical framework for planar axonemal beating by formulating a full set of nonlinear equations to test how flagellar amplitude and shape vary with dynein activity. We have shown how the nonlinear coupling of flagellar shape and dynein kinetics in a 'sliding-controlled' model provides a novel mechanism for the saturation of unstable modes in flagellar beating. Our study advances understanding of the nonlinear nature of the axoneme, typically studied at the linear level (13–16). The origin of the bending wave instability can be understood as a consequence of the antag- onistic action of dyneins competing along the flagellum (15, 39). The instability is then further stabilized by the nonlinear coupling between dynein activity and flagellum shape, without the need to invoke a nonlinear axonemal response to account for the saturation of the unstable modes, in contrast to previous studies (23,24). Moreover, the governing equations (Eqs. 10 and 11) contain all the nonlinearities in the limit of small curvature, and they are not the result of a power expansion to leading nonlinear order (19). Far from the Hopf bifurcation, linearized solutions fail to describe the flagellar shape and nonlinear effects arise in the system solely due to motor activity. At the nonlinear level, both the tangent angle and dynein population dynamics exhibit relaxation or cusp-like oscillations at some regions along the flagellum. Sim- ilar cusp-like shapes for the curvature have also been reported in sea urchin sperm (40). This phenomenology is characteristic of motor assemblies working in far-from-equilibrium condi- tions and has been found in other biological systems such as in the spontaneous oscillations of myofibrils (37, 41). Interestingly, despite the low duty ratio of axonemal dynein, a fraction of ∼ 2 % bound dyneins along the flagellum is sufficient to drive micrometer-sized amplitude 22 oscillations. Angular deflections are found to be ∼ 0.1 rad in the experimentally relevant activity range µa (cid:39) (2 − 6) · 103 and the order of magnitude seems not to be crucially affected by viscosity nor other parameters in the system. Hence, despite of the fact that our description provides an intrinsic mechanism for amplitude saturation, it is only able to generate small deflections, typ- ically an order of magnitude smaller than the ones reported for bull sperm flagella (15). Other structural constraints, such as line tension, are likely to influence the amplitude saturation, due to the elastohydrodynamic coupling with motor activity (see Supplementary Text). The pres- ence of tension on the self-organized beating of flagella was previously investigated at leading nonlinear order; however, deflections were also found to be small (19). Hence, we conclude that a 'sliding-controlled' mechanism may not be sufficient to generate large deflections. Our work adds to other recent studies were the 'sliding-controlled' hypothesis seems to lose support as the main mechanism responsible for flagellar beating (16, 24). Basal dynamics and elastic- ity are also likely to influence the amplitude saturation, and substantial further research is still needed to infer whether sliding-controlled regulation is the responsible mechanism behind flag- ellar wave generation. Principal component analysis allowed us to reduce the nonlinear dynamics of the flagel- lum in a two-dimensional shape space, regarding the flagellum as a single-frequency biological oscillator (32). Notice that a two-dimensional description would not hold for multifrequency oscillations, where an additional dimension is required (38). Interestingly, we found that as ac- tivity increases, the characteristic wavenumber of the system slightly decreases. Thus, dynein activity has an opposite effect on wavenumber selection when compared with the medium vis- cosity (see Fig. 2d, inset). We also showed that the steady state amplitude is selected by the fastest growing mode under the influence of competing unstable modes, provided that the initial 23 mode amplitudes are sufficiently small. An important aspect which is not studied explicitly in this work is the direction of wave propagation. For simplicity, we used clamped boundary conditions at the head which are known to induce travelling waves which propagate from tip to base in the sliding-controlled model (13, 24). It is beyond the scope of our study to assess the effects of different boundary conditions and the role of basal compliance at the head of the flagellum, which are known to crucially affect wave propagation (15). The present work also restricts to the case of small curvatures; however, the full nonlinear equations including the presence of tension could in principle be numerically solved as in previous studies (18, 19) (see Supplementary Text). Fi- nally, real flagella is subject to chemical noise due to the stochastic binding and unbinding of dynein motors. Recent studies have provided insights on this problem by investigating a noisy oscillator driven by molecular motors. However, their approach was not spatially extended (32). Our approach could be suitably extended to include chemical noise in the system through Eq. 11 by considering a chemical Langevin equation for the bound dynein populations including multiplicative noise (42). In particular, it can be easily deduced from our study that by consid- ering a force-independent unbinding rate, fluctuations of bound motors around the base state have mean N η and variance N η(1 − η), in agreement with the results in Ref. (32) where a different model was considered. The possibility to experimentally probe the activity of dyneins inside the axoneme is one of the most exciting future challenges in the study of cilia and flagella. These studies will be of vital importance to validate mathematical models of axonemal beating and the underlying mechanisms coordinating dynein activity and flagellar beating. 24 Ethics statement Not relevant to our work. Data accessibility Electronic Supporting information for this article includes a Supplementary Text and 3 Supple- mentary Movies. The data from the Supplementary Movies is available from the Dryad Digital Repository: http://dx.doi.org/10.5061/dryad.qs65q Competing interests We have no competing interests. Authors' contributions D.O. carried out analytical work, numerical simulations, data analysis and drafted the manuscript. H.G. and J.C. conceived of the study, coordinated the study and helped draft the manuscript. All authors gave final approval for publication. Acknowledgements No acknowledgements. Funding J.C. and D.O. acknowledge financial support from the Ministerio de Econom´ıa y Competi- tividad under projects FIS2010-21924-C02-02 and FIS2013-41144-P, and the Generalitat de Catalunya under projects 2009 SGR 14 and 2014 SGR 878. D.O. also acknowledges a FPU 25 grant from the Spanish Government with award number AP-2010-2503 and an EMBO Short Term Fellowship with ASTF number 314-2014. H.G. acknowledges support by the Hooke Fellowship, University of Oxford. References 1. Ginger M.L., Portman N. and McKean P.G. 2008 Swimming with protists: perception, motil- ity and flagellum assembly. Nat. Rev. Microbiol., 6, 838-850. (doi: 10.1038/nrmicro2009) 2. Lauga E. and Powers T. R. 2009 The hydrodynamics of swimming microorganisms. Rep. Prog. Phys, 72, 096601. (doi: 10.1088/0034-4885/72/9/096601) 3. Satir P. 1968 Studies on cilia. III. Further studies of the cilium tip and a 'sliding filament' model of ciliary motility. J. Cell Biol., 39, 77-94. (doi: 10.1083/jcb.39.1.77) 4. Gaffney E.A., Gadelha H., Smith D.J., Blake J.R. and Kirkman-Brown J.C. 2011 Mam- malian sperm motility: observation and theory. Annu. Rev. Fluid Mech., 43, 501-528. (doi: 10.1146/annurev-fluid-121108-145442) 5. Summers K.E. and Gibbons I.R. 1971 Adenosine triphosphate-induced sliding of tubules in trypsin-treated flagella of sea-urchin sperm. Proc. Natl. Acad. Sci. U.S.A., 68, 3092-3096. (doi: 10.1073/pnas.68.12.3092) 6. Mitchison T.J. and Mitchison H.M. 2010 How cilia beat. Nature, 463, 308-309. (doi: 10.1038/463308a) 7. Brokaw C.J. 2009 Thinking about flagellar oscillation. Cell Motil. Cytoskeleton, 66, 425- 436. (doi: 10.1002/cm.20313) 8. Machin K.E. 1958 Wave propagation along flagella. J. Exp. Biol., 35, 796-806. 26 9. Brokaw C.J. 1971 Bend propagation by a sliding filament model for flagella. J. Exp. Biol., 55, 289-304. 10. Hines M. and Blum J.J. 1978 Bend propagation in flagella. I. Derivation of equations of motion and their simulation. Biophys. J., 23, 41-57. (doi: 10.1016/S0006-3495(78)85431-9) 11. Hines M. and Blum J.J. 1979 Bend propagation in flagella. II. Incorporation of dynein cross-bridge kinetics into the equations of motion. Biophys. J., 25, 421-441. (doi: 10.1016/S0006-3495(79)85313-8) 12. Lindemann C.B. 1994 A "geometric clutch" hypothesis to explain oscillations of the ax- oneme of cilia and flagella. J. Theor. Biol., 168, 175-189. (doi: 10.1006/jtbi.1994.1097) 13. Camalet S., Julicher F. and Prost J. 1999 Self-organized beating and swimming of internally driven filaments. Phys. Rev. Lett., 82, 1590-1593. (doi: 10.1103/PhysRevLett.82.1590) 14. Camalet S. and Julicher F. 2000 Generic aspects of axonemal beating. New J. Phys., 2, 24. (doi: 10.1088/1367-2630/2/1/324) 15. Riedel-Kruse I.H., Hilfinger A., Howard J. and Julicher F. 2007 How molecular motors shape the flagellar beat. HFSP J. 1, 192-208. (doi:10.2976/1.2773861) 16. Sartori P., Geyer V. F., Scholich A., Julicher F., Howard J. 2016 Dynamic curvature regu- lation accounts for the symmetric and asymmetric beats of Chlamydomonas flagella. eLife 5:e13258 (doi:10.7554/eLife.13258) 17. Fu H.C., Wolgemuth C.W. and Powers T.R. 2008 Beating patterns of filaments in viscoelas- tic fluids. Phys. Rev. E, 78, 041913. (doi: 10.1103/PhysRevE.78.041913) 27 18. Gadelha H., Gaffney E.A., Smith D.J. and Kirkman-Brown J.C. 2010 Nonlinear instabil- ity in flagellar dynamics: a novel modulation mechanism in sperm migration? J. R. Soc. Interface, 7, 1689-1697. (doi: 10.1098/rsif.2010.0136) 19. Hilfinger A., Chattopadhyay A.K. and Julicher F. 2009 Nonlinear dynamics of cilia and flagella. Phys. Rev. E, 79, 051918. (doi: 10.1103/PhysRevE.79.051918) 20. Brokaw C.J. 1999 Computer simulation of flagellar movement. VII. conventional but functionally different cross-bridge models for inner and outer arm dyneins can explain the effects of outer arm dynein removal. Cell Motil. Cytoskeleton, 42, 134-148. (doi: 10.1002/(SICI)1097-0169(1999)42:2<134::AID-CM5>3.0.CO;2-B) 21. Brokaw C.J. 2014 Computer simulation of flagellar movement. X: Doublet pair splitting and bend propagation modeled using stochastic dynein kinetics. Cytoskeleton, 71, 273-284. (doi: 10.1002/cm.21168) 22. Roberts A.J., Kon T., Knight P.J., Sutoh K. and Burgess S.A. 2013 Functions and mechanics of dynein motor proteins. Nat. Rev. Mol. Cell Biol., 14, 713-726. (doi: 10.1038/nrm3667) 23. Bayly P.V. and Wilson K.S. 2014 Equations of interdoublet separation during flagella mo- tion reveal mechanisms of wave propagation and instability. Biophys. J., 107, 1756-1772. (doi: 10.1016/j.bpj.2014.07.064) 24. Bayly P.V. and Wilson K.S. 2015 Analysis of unstable modes distinguishes math- ematical models of flagellar motion. J. Roy. Soc. Interface, 12, 20150124. (doi: 10.1098/rsif.2015.0124) 25. Gadelha H., Gaffney E.A. and Goriely A. 2013 The counterbend phenomenon in flagellar axonemes and cross-linked filament bundles. Proc. Natl. Acad. Sci. U.S.A., 110, 12180- 12185. (doi: 10.1073/pnas.1302113110) 28 26. Werner S., Rink J.C., Riedel-Kruse I.H. and Friederich B.M. 2014 Shape mode analysis exposes movement patterns in biology: flagella and flatworms as case studies. PLoS ONE, 9:e113083. (doi: 10.1371/journal.pone.0113083) 27. Jolliffe I.T. 2002 Principal Component Analysis, Springer-Verlag, New York. 28. Muller M. J., Klumpp S., Lipowsky R. 2008 Tug-of-war as a cooperative mechanism for bidirectional cargo transport by molecular motors Proc. Natl. Acad. Sci. U. S. A., 105, 4609– 4614. (doi: 10.1073/pnas.0706825105) 29. Gray J. and Hancock G.J. 1955 The propulsion of sea-urchin spermatozoa. J. Exp. Biol., 32, 802-814. 30. Minoura I., Yagi T. and Kamiya. R. 1999 Direct measurement of inter-doublet elasticity in flagellar axonemes. Cell Struct. Funct., 24, 27-33. (doi:10.1247/csf.24.27) 31. Nicastro D., Schwartz C., Pierson J., Gaudette R., Porter M. E. and McIntosh J. R. 2006 The molecular architecture of axonemes revealed by cryoelectron tomography. Science, 313, 944-948. (doi: 10.1126/science.1128618) 32. Ma R., Klindt G.S., Riedel-Kruse I.H., Julicher F. and Friederich B.M. 2014 Active phase and amplitude fluctuations of flagellar beating. Phys. Rev. Lett., 113, 048101. (doi: 10.1103/PhysRevLett.113.048101) 33. Sakakibara H., Kojima H., Sakai Y., Katayama E. and Oiwa K. 1999 Inner-arm dynein c of Chlamydomonas flagella is a single-headed processive motor. Nature, 400, 586-590. (doi: 10.1038/23066) 29 34. Hirakawa E., Higuchi H. and Toyoshima Y. Y. 2000 Processive movement of single 22S dynein microtubules occurs only at low ATP concentrations. Proc. Natl. Acad. Sci. U.S.A., 97, 2533-2537. (doi: 10.1073/pnas.050585297) 35. Howard J. 2001 Mechanics of Motors Proteins and the Cytoskeleton, Sinauer Associates. 36. Smith D.J., Gaffney E.A., Gadelha H., Kapur N. and Kirkman-Brown J.C. 2009 Bend propagation in the flagella of migrating human sperm, and its modulation by viscosity. Cell Motil. Cytoskeleton, 66, 220-236. (doi: 10.1002/cm.20345) 37. Julicher F. and Prost J. 1997 Spontaneous oscillations of collective molecular motors. Phys. Rev. Lett., 78, 4510. (doi: 10.1103/PhysRevLett.78.4510) 38. Oriola D., Gadelha H., Blanch-Mercader C. and Casademunt J. 2014 Subharmonic oscilla- tions of collective molecular motors. EPL, 107, 18002. (doi: 10.1209/0295-5075/107/18002) 39. Howard J. 2009 Mechanical signaling in networks of motor and cytoskeletal proteins. Annu. Rev. Biophys., 38, 217-234. (doi: 10.1146/annurev.biophys.050708.133732) 40. Ohmuro J., Mogami Y. and Baba S.A. 2004 Progression of flagellar stages during ar- tificially delayed motility initiation in sea urchin sperm. Zool. Sci., 21, 1099-1108. (doi: 10.2108/zsj.21.1099) 41. Yasuda K., Shindo Y. and Ishiwata S. 1996 Synchronous behavior of spontaneous oscilla- tions of sarcomeres in skeletal myofibrils under isotonic conditions Biophys. J., 70, 1823- 1829. (doi: 10.1016/S0006-3495(96)79747-3) 42. Gillespie D.T. 2000 The chemical Langevin equation. J. Chem. Phys., 113, 297-306. (doi:10.1063/1.481811) 30
1203.0918
1
1203
2012-03-05T13:55:13
Accounting for the thickness effect in dynamic spherical indentation of a viscoelastic layer: Application to non-destructive testing of articular cartilage
[ "physics.bio-ph" ]
In recent years, dynamic indentation tests have been shown to be useful both in identification of mechanical properties of biological tissues (such as articular cartilage) and assessing their viability. We consider frictionless flat-ended and spherical sinusoidally-driven indentation tests utilizing displacement-controlled loading protocol. Articular cartilage tissue is modeled as a viscoelastic material with a time-independent Poisson's ratio. We study the dynamic indentation stiffness with the aim of formulating criteria for evaluation the quality of articular cartilage in order to be able to discriminate its degenerative state. In particular, evaluating the dynamic indentation stiffness at the turning point of the flat-ended indentation test, we introduce the so-called incomplete storage modulus. Considering the time difference between the time moments when the dynamic stiffness vanishes (contact force reaches its maximum) and the dynamic stiffness becomes infinite (indenter displacement reaches its maximum), we introduce the so-called incomplete loss angle. Analogous quantities can be introduced in the spherical sinusoidally-driven indentation test, however, to account for the thickness effect, a special approach is required. We apply an asymptotic modeling approach for analyzing and interpreting the results of the dynamic spherical indentation test in terms of the geometrical parameter of the indenter and viscoelastic characteristics of the material. Some implications to non-destructive indentation diagnostics of cartilage degeneration are discussed.
physics.bio-ph
physics
Accounting for the thickness effect in dynamic spherical indentation of a viscoelastic layer: Application to non-destructive testing of articular cartilage I. Argatov a, A.U. Daniels b, G. Mishuris a,∗, S. Ronken b, D. Wirz b aInstitute of Mathematics and Physics, Aberystwyth University, Wales, UK bLaboratory of Biomechanics & Biocalorimetry, University Basel, Switzerland Abstract In recent years, dynamic indentation tests have been shown to be useful both in identification of mechanical properties of biological tissues (such as articular carti- lage) and assessing their viability. We consider frictionless flat-ended and spherical sinusoidally-driven indentation tests utilizing displacement-controlled loading pro- tocol. Articular cartilage tissue is modeled as a viscoelastic material with a time- independent Poisson's ratio. We study the dynamic indentation stiffness with the aim of formulating criteria for evaluation the quality of articular cartilage in order to be able to discriminate its degenerative state. In particular, evaluating the dy- namic indentation stiffness at the turning point of the flat-ended indentation test, we introduce the so-called incomplete storage modulus. Considering the time differ- ence between the time moments when the dynamic stiffness vanishes (contact force reaches its maximum) and the dynamic stiffness becomes infinite (indenter displace- ment reaches its maximum), we introduce the so-called incomplete loss angle. Anal- ogous quantities can be introduced in the spherical sinusoidally-driven indentation test, however, to account for the thickness effect, a special approach is required. We apply an asymptotic modeling approach for analyzing and interpreting the results of the dynamic spherical indentation test in terms of the geometrical parameter of the indenter and viscoelastic characteristics of the material. Some implications to non-destructive indentation diagnostics of cartilage degeneration are discussed. Key words: Viscoelastic contact problem, cartilage layer, dynamic indentation test, asymptotic model ∗ Corresponding author. Introduction Joint cartilage is known to have very limited repair capabilities and poorly regenerates. Intensive recent research and development have brought many innovations and also first clinical results in cartilage repair. However, recent clinical, radiological and histological evaluation techniques show somehow contradictory results (Kusano et al., 2011) and this is why measuring stiffness parameters of cartilage, especially in vivo measurements, are of novel interest nowadays. Cartilage stiffness parameters can be measured in confined (Suh et al., 1995) or in unconfined (Armstrong et al., 1984) compression of cartilage specimen. However, both the confined and unconfined compression tests need sample preparation, usually cylindrically shaped specimens of cartilage, and therefore prohibit in vivo measurements. Furthermore, the mapping of the surface is limited by the sample size. These limitations are less restrictive than those usually encountered in indentation testing. The first mathematical model allowing to measure stiffness parameters of joint cartilage layer in indentation mode with flat-ended as well as with spherical indenters was developed by Hayes et al. (1972). In the case of a flat-ended indenter of radius a pressed against a sample of thickness h, the indentation stiffness defined as the ratio of the contact force P to the indenter displacement w is given by P w = 2aE 1 − ν2 κc. (1) Note that compared to Hayes et al. (1972), we replace the shear modulus G with E/(2(1+ nu)), where E is Young's modulus, ν is Poisson's ratio. The Hayes model (1) is based on Hooke's law and takes into account the thickness effect through the stalling factor κc. Because the widely used Hayes model assumes linear elasticity, it therefore does not take into consideration the fact that cartilage stiffness parameters are strain-rate dependent, and thus the Hayes model does not consider the dynamic nature of cartilage stiffness. (cid:113) E2 Recall that for time-dependent materials (Tschoegl, 1997), the dynamic stiffness is charac- terized by the complex dynamic modulus E∗ = E1 +iE2 consisting of the storage modulus E1 and the loss modulus E2 with i being the imaginary unit (i2 = −1). On the complex plane, E1, which is a real part of E∗, and E2, which is an imaginary part of as imag- inary E∗, represent the legs (catheti) of a right triangle with the hypotenuse of length E∗ = 2 . The loss angle, δ, results from the ratio of E2 and E1 through the relationship tan δ = E2/E1. It should be emphasized that the storage and loss moduli E1 and E2 represent the response of a material to a sinusoidal loading scheme and actually depend on the corresponding angular frequency of sinusoidal oscillations ω. In order to underline this fact we will write E1(ω) instead of E1 and so on. 1 + E2 For in vivo (or ex vivo) measurements of cartilage stiffness and mapping a cartilage surface, a mechanical model for cartilage has to be prioritized considering dynamic properties of cartilage and measuring in indentation mode (Ronken et al., 2011). And, application of a single indentation test during arthroscopy allows one to evaluate the quality of cartilage and to detect osteoarthritic degenerative changes (Korhonen et al., 2003). The readily available systems on the market for arthroscopic measurements such as the 'Artscan 1000' (Lyyra et al., 1995; Toyras et al., 2001) allow only stiffness measurements of the cartilage surface without explicit considering E1(ω) and δ(ω). To ascertain dynamic biomechanical properties of articular cartilage, Appleyard et al. (2001) investigated a handheld indentation probe with a flat-ended cylindrical indenter operating in a vibration mode at a single-frequency of 20 Hz. Using the theory of Hayes et al. (1972), the absolute value of the effective complex dynamic modulus can be evaluated as follows: E∗(ω) = . (2) (1 − ν2) 2aκc P0 w0 Here, P0 is the contact force amplitude, w0 is the displacement amplitude. Note that the complex dynamic modulus E∗(ω) is termed effective here, because in the case of a poroelastic material such as articular cartilage, the biomechanical response is dependent on the frequency ω and the boundary conditions for the sample as well. It was also observed (Appleyard et al., 2001) that when articular cartilage is indented at frequencies above 10 Hz there is marginal change in the effective parameters E∗(ω) and δ(ω) with the effective dynamic modulus being of similar magnitude to the 'instanteneous' elastic modulus generated during a rapid load step indentation test. Nevertheless, to the best of our knowledge, no study has analyzed thus far the relationship between the parameters of time-dependent materials measured in a vibration indentation test and in a single indentation test. In order to facilitate such a comparison, we consider sinusoidally-driven displacement-controlled indentation tests. Note that as a first approximation, the half- sinusoidal indentation history can be used for modeling impact tests. Measuring stiffness parameters of cartilage in indentation mode with spherical tipped indenters has advantages as well as drawbacks. At one hand, with spherical indenters the error obtained when hitting the surface not exactly perpendicular is much smaller than with flat-ended indenters, where the surface is touched with one edge of the indenter first. For example, when the surface with a spherical indenter will be hit with 80◦ instead of 90◦, the result for the stiffness will be underestimated by less than 2%. On the other hand, spherical indenters underestimate the inhomogeneity and changes in stiffness (Schinagl et al., 1997) as function of the indentation depth. This is why both theories, for flat-ended and for spherical tipped indenters are provided. As measuring mechanical properties gained a new importance in recent years, because dy- namic indentation tests have been shown to be helpful both in identification of mechanical properties of articular cartilage and assessing its viability (Bae et al., 2003; Broom and Flachsmann, 2003). Indentation stiffness is now accepted as a fundamental indicator of the functional mechanical properties of articular cartilage (de Freitas et al., 2006). The dynamic stiffness is defined as the ratio of input force, P (t), to output displacement, w(t). Thus, for a time-dependent material like articular cartilage, the indentation stiffness de- pends on the indentation protocol, and generally it is a function of time. It is also well known that the indentation stiffness depends on the indenter size as well as on the sample dimensions (see, e.g., Eq. (1)). This follows from a comparison of the stiffness dimension MT−2 with the dimension ML−1T−2 of Young's modulus. From a geometrical point of view, articular cartilage is usually considered as a layer of constant thickness, h. In view of the relative mechanical properties of cartilage and subchondral bone, it is assumed that the layer is firmly attached to a non-deformable base. Thus, the indentation scaling factor will depend on the aspect ratio α = a/h, where a is the radius if the contact area. The above simple analysis is applicable for the linear relationship between the contact force P (t) and the indenter displacement w(t), where the contact radius remains unchanged in time. In spherical indentation, the force-displacement relationship requires a more acute analysis. It will be shown that the results of the dynamic spherical indentation of a time- dependent material depend on the level of indentation. To a first approximation (Hayes and Mockros, 1971; Parsons and Black, 1977; Lau et al., 2008), cartilage tissue can be evaluated mechanically as a viscoelastic material with a time-independent Poisson's ratio, ν, such that the overall constitutive behavior of the material is expressed in terms of its complex modulus E∗(ω). Indentations tests for vis- coelastic materials were studied in a number of publications (Oyen, 2005; Cheng and Yang, 2009; Argatov and Mishuris, 2011). We consider flat-ended and spherical indentation tests utilizing displacement-controlled loading protocol with the indenter displacement modu- lated according to a sinusoidal law at an angular frequency ω = 2πf (rad/s), where f (Hz) is the loading frequency. We apply an asymptotic modeling approach for analyzing and interpreting the results of the dynamic spherical indentation test in terms of the ge- ometrical parameter of the indenter (indenter radius, R) and viscoelastic characteristics of the material. In particular, we examine the relationships between the storage modulus E1(ω) and loss angle δ(ω) and the so-called modified storage modulus E0 3/2(ω, 0) and the modified loss angle δ0 3/2(ω) in the displacement-controlled sinusoidally-driven indentation test. The rest of the paper is organized as follows. In Section 1, we consider cylindrical fric- tionless indentation of a viscoelastic layer. In particular, the linear force-displacement relationship is outlined in Section 1.1, while the indentation scaling factor for the cylin- drical indenter is considered in Section 1.2. Based on the analogy with the case of harmonic vibrations (considered in Sections 1.3 and 1.4), in 1.5, we introduce the incomplete storage modulus and loss angle as material characteristics that can be assessed directly from a single sinusoidally-driven indentation test. In Section 2, we study spherical indentation of a viscoelastic layer. Based on the general so- lution obtained by Ting (1968), in Sections 2.1 and 2.2, we write out the force-displacement relationship for the loading and unloading stages, respectively. The indentation scaling factor for the spherical indenter is introduced in Section 2.3. In Section 2.4, we introduce the so-called modified incomplete storage modulus and loss angle, and investigate their behavior in Section 2.5 for the standard viscoelastic solid model. In Section 3, we actually consider the thickness effect in spherical indentation of a vis- coelastic layer. By analogy with the elastic case, we introduce the quantity E0 3/2(ω, 0) (which is called the modified storage modulus, in view of its relation to the storage mod- ulus E1(ω)) while the modified loss angle δ0 3/2(ω) is introduced according to a standard interpretation of the time lag between the peak force and peak displacement. Some prop- erties of these parameters, which turned out to be dependent on the level of indentation, are illustrated for the standard viscoelastic solid model. Low- and high-frequency asymp- totic analysis of the quantities E0 β(ω) is presented in Sections 3.2 and 3.3, respectively. β(ω, 0) and δ0 Finally, in Sections 4 and 5, we outline a discussion of the results obtained and formulate our conclusions. 1 Cylindrical frictionless indentation of a viscoelastic layer 1.1 Liner force-displacement relationship We consider a viscoelastic layer bonded to a rigid substrate indented by a flat-ended cylin- drical indenter. For the sake of simplicity, we neglect friction and assume that Poisson's ratio, ν, of the layer material is time independent. Then, applying the elastic-viscoelastic correspondence principle (Christensen, 1971), one can arrive at the following equation between the applied force P (t) and the displacement of the indenter w(t) (Zhang and Zhang, 2004; Cao et al., 2010): P (t) = 2a 1 − ν2 κc(α) E(t − τ ) dw dτ (τ ) dτ. (3) Here, a is the radius of the contact area, h is the layer thickness, E(t) is the relaxation modulus, t is the time variable, t = 0− is the time moment just preceding the initial moment of contact, κc(α) is a dimensionless factor, which is determined from the solution of elastic contact problem for a cylindrical indenter, α is the relative radius of the contact area that is, i. e., α = . (4) Note that the dependence of κc(α) on Poisson's ratio is not indicated explicitly. Fig. 1a illustrates the behavior of κc(α) for different values of ν based on the numerical solution obtained by Hayes et al. (1972). Figure 1. (a) Indentation scaling factor for the cylindrical indenter as a function of relative contact radius; (b) Relative error of of the asymptotic approximation (7). Denoting E(t) = E∞Ψ(t), where E∞ is the relaxed elastic modulus (the limit of modulus E(t) at t → ∞), Ψ(t) is the relaxation function, we rewrite Eq. (3) in the following form: P (t) = 2aE∞ 1 − ν2 κc(α) Ψ(t − τ ) dw dτ (τ ) dτ. (5) t(cid:90) 0− a h t(cid:90) 0− 0124568Relative radius of the contact areaha/3706090120150Indentation factor for spherical indenter(a))(c3050.045.040.035.030.0012305101520100)()()(ccAMc50.045.040.035.030.000.10.20.30.40.6Relative radius of the contact areaha/0.5Relative error of asymptotic approximation, %(b)0.10.10.30.50.70.91.11.3 Inverting the relationship (5), we obtain w(t) = 1 − ν2 2aE∞ where Φ(t) is the creep function. t(cid:90) 1 κc(α) 0− Φ(t − τ ) dP dτ (τ ) dτ, (6) 1.2 Indentation scaling factor for the cylindrical indenter According to Vorovich et al. (1974); Argatov (2002), the following asymptotic model takes place for the indentation scaling factor κc(α): κc(α) = 1 + α 2a0 π (cid:34)(cid:18)2a0 + α2 (cid:19)4 π + α4 (cid:34)(cid:18)2a0 (cid:19)3 π + 8a1 3π (cid:35) (cid:18)2a0 (cid:19)2 π 32a0a1 3π2 + + α3 (cid:35) + O(α5). (7) Here, a0 and a1 are asymptotic constants depending on Poisson's ratio ν given by ∞(cid:90) (cid:104) (cid:105) 1 − L(λ) λ2m dλ. am = (−1)m 22m(m!)2 0 In the case of a layer bonded to a rigid base, we have 2κ sh 2λ − 4λ L(λ) = 2κ ch 2λ + 1 + κ2 + 4λ2 , where κ = 3 − 4ν is Kolosov's constant. To determine the range of validity of the asymptotic model (7), we compare its predictions with the numerical solution given by Hayes et al. (1972). As it could be expected (see Fig. 1b), the accuracy of the approximation κAM (α) given by (7) decreases as Poisson's ratio approaches 0.5 . Fig. 1b shows that asymptotic approximation (7) is quite accurate in the range α ∈ (0, 0.6), that is for the indenter diameter less than the layer thickness. Note here that the substrate effect on the incremental indentation stiffness was considered in the elastic case in (Argatov, 2010). c 1.3 Harmonic vibration Observe that Eq. (3) assumes that the layer material was at rest for t < 0. In order to study harmonic vibrations of a viscoelastic layer, we should replace Eq. (3) with the following one: P (t) = 2a 1 − ν2 κc(α) E(t − τ ) dw dτ (τ ) dτ. (8) t(cid:90) −∞ Substituting a harmonic displacement w(t) = Im{w0 exp(iωt)} with amplitude w0 and frequency ω into Eq. (8), one can arrive at the following equation: P (t) = 2aw0 1 − ν2 κc(α) Im{E∗(ω) exp(iωt)}. (9) Here, Im denotes the imaginary part of a complex number, E∗(ω) is the complex relaxation modulus given by E∗(ω) = iω E(s) exp(−iωs) ds. (10) ∞(cid:90) By convention (Pipkin, 1986; Tschoegl, 1997), we define the storage modulus, E1(ω), and the loss modulus, E2(ω), as the real and imaginary parts of E∗(ω), respectively, i. e., 0 E∗(ω) = E1(ω) + iE2(ω). From Eqs. (10) and (11), it follows that E1(ω) = ωE∞ E2(ω) = ωE∞ ∞(cid:90) ∞(cid:90) 0 Ψ(s) sin ωs ds, Ψ(s) cos ωs ds. Furthermore, according to Eq. (9), we can write 0 P (t) = P0 sin(ωt + δ), (11) (12) (13) (14) where P0 is the force amplitude, δ is the phase angle between the harmonic displacement and the force, given by the formulas P0 = 2a 1 − ν2 κc(α)E∗(ω)w0, E∗(ω) = E2(ω) E∗(ω), cos δ = E1(ω) E∗(ω) , sin δ = (cid:113) E1(ω)2 + E2(ω)2. (15) (16) Observe that the phase angle δ depends on the frequency ω (this is not indicated in notation for simplicity). Finally, note that the vibration indentation tests should be accomplished with a quasistatic preload to ensure a complete contact between the indenter's base and the layer surface, since tensile stresses are not allowed in frictionless indentation. 1.4 Determination of the complex relaxation modulus via vibration indentation tests We assume that the displacement and force amplitudes w0 and P0 as well as the phase angle δ are experimentally measurable quantities. Then, Eqs. (15) and (16) yield the following equations (Cao et al., 2010): E1(ω) = E2(ω) = 1 − ν2 2aκc(α) 1 − ν2 2aκc(α) P0 w0 P0 w0 cos δ, sin δ. (17) (18) Thus, for a given constant frequency ω, the vibration indentation test yields the storage and loss moduli E1(ω) and E2(ω), if the amplitude ratio P0/w0 and the phase angle δ are known from the experiment. Further, let tm denote the moment of time when the indentation speed w(t) vanishes, that is, when w(tm) = 0 and tm = π/(2ω) + πk/ω, k = 0,±1,±2, . . . . Considering the indentation process over a half of period t ∈ (0, π/ω), we will have tm = π/(2ω) and, correspondingly, w(tm) = w0 and P (tm) = P0 cos δ. Hence, taking Eq. (17) into account, we obtain the formula 1 − ν2 2aκc(α) where tm is a time moment such that w(tm) = 0. E1(ω) = P (tm) w(tm) , (19) Figure 2. (a) Displacement-controlled oscillation test; (b) Displacement-controlled indentation test. Thus, according to Eq. (19), the ratio P (tm)/w(tm) at the time moment of the displace- ment extremum determines the storage modulus. Let now t(cid:48) i. e., w(t(cid:48) tm = tm + π/(2ω).) Taking into account Eq. (18), we obtain m be the moment of time when the indentation displacement w(t) vanishes, m = πk/ω, k = 0,±1,±2, . . . . (Note that for harmonic vibrations m) = 0 and t(cid:48) E2(ω) = 1 − ν2 2aκc(α) ωP (t(cid:48) m) w(t(cid:48) m) , (20) where w(t(cid:48) m) is the indentation speed when the indentation displacement vanishes. Remark 1 For the sake of completeness, we provide below the dual-conjugate formulas for Eqs. (19) and (20). Let tM denote the moment of time when the derivative of the tMt′mtMtωωδ)()sin(tωmt′)sin(δω+t(a)00)(,)(PtPwtw)(tw)(tPtωπMt′~ωπ2=mtMt~ωωδ)(~00)(,)(PtPwtw)(tw)(tP)sin(tω1(b) contact force, k = 0,±1,±2, . . . . Let also t(cid:48) i. e., P (t(cid:48) Eqs. (17) and (18), the following relationships hold true: P (t), vanishes, that is, when P (tM ) = 0 and tM = (π/2 − δ)/ω + πk/ω, M be the time moment when the contact force P (t) vanishes, M = (π − δ)/ω + πk/ω, k = 0,±1,±2, . . . . Then, according to M ) = 0 and t(cid:48) E1(ω) 2aκc(α) E∗(ω)2 = 1 − ν2 E∗(ω)2 = −2aκc(α) 1 − ν2 E2(ω) , w(tM ) P (tM ) ωw(t(cid:48) M ) P (t(cid:48) M ) (22) Recall that the magnitude of the complex modulus, E∗(ω), is determined by the last formula (16). . (21) (24) 1.5 Indentation test with a sinusoidal displacement. Incomplete storage modulus and loss angle Let us first consider a single indentation test with a prescribed sinusoidal displacement according to the law (23) Here, w0 is the maximum depth of indentation, ω is a given quantity having the dimension of reciprocal time. The quantity w(t) = w0 sin ωt, t ∈ (0, π/ω). tm = π 2ω has a physical meaning of the time moment when the indentation displacement reaches its maximum. We emphasize that due to viscoelastic properties of the layer material, the duration of contact will be less than π/ω. According to Eq. (3), we get P (tm) = 2a 1 − ν2 κc(α)ωw0 tm(cid:90) 0 E(tm − τ ) cos ωτ dτ, where tm = π/(2ω). By analogy with Eq. (19), we define E1(ω) = 1 − ν2 2aκc(α) P (tm) w(tm) , (25) (26) where w(tm) = w0 (see Eq. (23)). It is clear that the quantity E1(ω), introduced for single indentation test, differs from the storage modulus E1(ω), introduced for vibration indentation test. In view of (25), Eq. (26) yields E1(ω) = ω π/(2ω) (cid:90) 0 E(s) sin ωs ds. (27) Recalling the notation E(t) = E∞Ψ(t), we rewrite Eq. (27) in the form (cid:90) π/(2ω) E1(ω) = ωE∞ Ψ(s) sin ωs ds. (28) 0 Comparing Eqs. (12) and (28), we see that their right-hand sides differ only by the integral upper limits. Now, let tM be the time moment when the contact force (3) corresponding to the inden- P (tM ) = 0. Then, by analogy with the case of tation law (23) reaches its maximum, i. e., linear harmonic vibrations (see Eq. (14)), we put δ(ω) = − ωtM . π 2 (29) The quantity δ(ω) is called the incomplete loss angle determined from the sinusoidally- driven displacement-controlled cylindrical indentation test (Argatov, 2012). In view of (24), formula (29) can be rewritten as follows: δ(ω) = π 2 (tm − tM ) tm . (30) The interrelations between the quantities E1(ω), δ(ω) and E1(ω), δ(ω) were investigated in (Argatov, 2012). It was shown that they asymptotically coincide, respectively, in both the low and high frequency limits, while within the intermediate range of ω, the differences depend on the viscoelastic model in question, that is on the properties of the relaxation modulus E(t). 2 Spherical frictionless indentation of a viscoelastic layer 2.1 Force-displacement relationship in the loading stage Applying the general solution obtained by Ting (1968) for a class of viscoelastic contact problems in terns of the corresponding elastic solutions, we will have P (t) = 4E∞h3 3(1 − ν2)R d dτ t(cid:90) 0 (cid:110) (cid:111) Ψ(t − τ ) dτ, α(τ )3F(α(τ )) Here, α(t) is the variable relative radius of the contact area, i. e. (cf. Eq. (4)) w(t) = h2α(t)2 R G(α(t)). α(t) = a(t) h , (31) (32) (33) while F(α) and G(α) are depending on Poisson's ratio ν dimensionless factors such that F(0) = G(0) = 1. We will use formulas (31) and (32) under the assumption that the relative contact radius α(t) monotonically increases in the time interval (0, tm), where tm is a certain moment of time. 2.2 Force-displacement relationship in the unloading stage Let us assume that the relative contact radius α(t) decreases to zero in the interval t ∈ (tm, tc), where tc is the time of contact of the indenter with the layer surface. Once again, making use of the general solution derived by Ting (1968) for the case when the variation of contact radius posses a single maximum, we obtain (34) (35) P (t) = 4E∞h3 3(1 − ν2)R t1(t)(cid:90) 0− (cid:110) (cid:111) Ψ(t − τ ) dτ, α(τ )3F(α(τ )) d dτ α(t)2G(α(t)) t(cid:90) Φ(t − τ ) ∂ ∂τ tm t1(τ ) τ(cid:90) (cid:110) (cid:111) Ψ(τ − η) dηdτ α(η)2G(α(η)) d dη . w(t) = h2 R − Here, t1(τ ) is the time moment prior to tm such that the contact radius a(τ ) is equal to the prior contact radius a(t1(τ )). The function t1(t) remains to be calculated. If the indenter displacement w(t) is prescribed, taking into account the relation a(t1) = a(t) for t1 ≤ tm ≤ t, we arrive at the equation w(t) = h2α(t)2 R G(α(t)), from which we get (cid:18) h2α(t)2 R (cid:19) G(α(t)) , t1(t) = w−1 t ∈ [tm, tc]. (36) (37) Finally, note that in the inner integral in (35), for η such that t1(τ ) ≤ η ≤ tm, the function α(η) should be calculated according to Eqs. (31) and (32). 2.3 Indentation scaling factor for the spherical indenter In the elastic case, according to the notation used in Eqs. (31) and (32), we have P = 4Eh3 3(1 − ν2)R α3F(α), w = α2G(α). h2 R Here, α is the relative radius of the contact area as defined by formula (4). Representing Eq. (39) in the form √ wR h = α (cid:113)G(α), and taking into account that G(0) = 1, we see that Eq. (39) can be inverted as where α =  = √ g(), wR h √ wR h . √ 4E R 3(1 − ν2) w3/2f (). Substituting the expression (41) into Eq. (38), we obtain the following relationship: (38) (39) (40) (41) (42) (43) (44) (45) (46) P = Here we introduced the notation f () = g()3F(g()). It is clear that f (0) = g(0) = 1. Finally, in view of Eq. (40), we can represent Eq. (43) as where we introduced the indentation scaling factor P = w3/2κs(α), √ 4E R 3(1 − ν2) (cid:16) κs(α) = f α (cid:113)G(α) (cid:17) . We emphasize that κs(α) is normalized in such a way that κs(0) = 1. Numerical values for √ κs(α) for a range of parameters α and ν are given in Table 1 based on the results obtained by Hayes et al. (1972). We note that κs(α) = (3/2) χκ, where χ and κ are parameters employed in their analysis. According to Argatov (2001), the following asymptotic expansion holds true: Table 1 Values of κs(α) for the spherical indenter. a/h ν = 0.30 ν = 0.35 ν = 0.40 ν = 0.45 ν = 0.50 0.04 0.06 0.08 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.034 1.052 1.072 1.091 1.197 1.315 1.445 1.585 1.734 1.892 2.058 2.228 2.405 1.25 2.863 1.5 3.337 1.75 3.821 2.0 4.311 2.25 4.808 2.5 5.309 2.75 5.810 3.0 6.316 1.035 1.055 1.075 1.095 1.208 1.333 1.472 1.622 1.784 1.955 2.135 2.322 2.517 3.023 3.554 4.100 4.654 5.217 5.787 6.363 1.037 1.058 1.080 1.102 1.221 1.356 1.507 1.672 1.851 2.044 2.245 2.459 2.681 3.268 3.891 4.542 5.212 5.903 6.604 7.317 6.939 8.040 (cid:34)320a3 1.040 1.063 1.086 1.109 1.240 1.389 1.557 1.744 1.949 2.171 2.410 2.664 2.932 3.659 4.455 5.311 6.221 7.179 8.180 9.215 1.044 1.069 1.094 1.120 1.266 1.435 1.629 1.849 2.096 2.370 2.671 2.998 3.354 4.359 5.532 6.886 8.427 10.163 12.113 14.288 10.289 16.699 (cid:35) (47) + O(α5). (48) Now, employing the forth-order asymptotic model constructed by Vorovich et al. (1974); Argatov (2002), we obtain f () = 1 +  2a0 π (cid:34)286a4 + 2 14a2 0 (cid:35) 3π2 + 3 64a0a1 0 27π3 + 32a1 15π + 4 0 9π4 + 5π2 + O(5). (cid:34) 16a3 27π3 − 32a1 15π 0 (cid:35) κs(α) = 1 + α 2a0 π (cid:34)22a4 0 3π4 + + α2 2a2 0 π2 − α3 (cid:35) 32a0a1 15π2 − α4 Figure 3. (a) Indentation scaling factor for the spherical indenter as a function of relative contact radius; (b) Relative error of of the asymptotic approximation (48). Fig. 3a shows details of the behavior of κs(α) for different values of Poisson'a ratio. The errors of the asymptotic approximation κAM (α) given by (48) are plotted as functions of a/h in Fig. 3b based on the data given in Table 1. s 2.4 Modified incomplete storage modulus and loss angle In the viscoelastic case, according to Eqs. (31), (32), (42), and (43), we will have P (t) = 4E∞h3 3(1 − ν2)R Here we used the notation (cf. (42)) (cid:111) Ψ(t − τ ) dτ. (τ )3/2f ((τ )) t(cid:90) 0 (cid:110) d dτ (cid:113) w(t)R h . (t) = (49) (50) (51) Equation (49) can be simplified for the case of a viscoelastic half-space when f ((τ )) ≡ 1 as follows: √ 4 R P (t) = 3(1 − ν2) w(τ )3/2(cid:111) (cid:110) t(cid:90) 0 d dτ E(t − τ ) dτ. We emphasize that Eqs. (49) and (51) are valid under the assumption that the indenter's displacement w(t) increases in the time interval (0, tm). Further, we consider again the same single indentation test with a prescribed sinusoidal displacement according to the indentation protocol (23). We may use Eqs. (49) and (51) in the time interval (0, tm) with tm = π/(2ω), that is up to the moment, when the indenter reaches its maximum indentation depth w0. 50.045.040.035.030.000.511.522.53Relative radius of the contact areaha/05101520Indentation factor for spherical indenter(a))(s00.10.20.30.40.5Relative radius of the contact areaha/30.035.040.045.050.0Relative error of asymptotic approximation, %(b)14118512100)()()(ssAMs By analogy with Eq. (26), we consider the quantity 3(1 − ν2) √ 4 R P (tm) w3/2 0 = E(t − τ ) d dτ tm(cid:90) 0 (cid:18) w(τ ) (cid:19)3/2 w0 dτ. (52) Note that the quantity on the right-hand side of Eq. (52) was previously considered in a number of studies on indentation of viscoelastic materials (Hu et al., 2001; Kren and Naumov, 2010). Substituting the expression (23) into the right-hand side of Eq. (52), we arrive at the following integral with β = 3/2: Eβ(ω) = (sin ωτ )βdτ. (53) π/(2ω) (cid:90) 0 (cid:18) π 2ω E (cid:19) d dτ − τ (cid:90) π/(2ω) Observe that here the parameter β was introduced to simplify notation. However, later we show (see Remark 2) that the notation Eβ(ω) is meaningful for different values of β. By changing the integration variable, the integral (53) may be cast in the form Eβ(ω) = ω E(s)β cosβ−1 ωs sin ωs ds. (54) 0 It is clear that for β = 1, the right-hand sides of (27) and (54) coincide. The quantity Eβ(ω) will be called the modified incomplete storage modulus. Remark 2 Recall (Galin, 1946; Borodich and Keer, 2004) that the force-displacement relationship in the elastic case for a rigid blunt indenter with the shape function z = Arλ is given by the equation P = [E/(1 − ν2)]A1−βKβwβ with β = (λ + 1)/λ and (with Γ(x) being the Gamma function) (cid:18) (cid:19)Γ ln(β − 1) (cid:18)1 + β − β2 22(β−1)(β − 1) (β − 1)2 (cid:19)2(1−β) (cid:18) 1 β − 1 Γ (cid:19)β−1 . 1 2(β − 1) Kβ = β exp For a spherical indenter, we have λ = 2 and β = 3/2. We refer to (Argatov, 2011) for complete details of this consideration in the elastic case. In the viscoelastic case, the force-displacement relationship in the loading phase is given by P (t) = EA1−βKβ 1 − ν2 (cid:110) w(τ )β(cid:111) t(cid:90) 0 d dτ E(t − τ ) dτ. Comparing this equation with Eq. (51), we see that the blunt indentation test yields the modified incomplete storage modulus Eβ(ω) introduced by formula (53). This explains the introduced notation. Further, assuming the variation of the indenter displacement in the form (23), we get the following variation of the contact force: √ 4 R P (t) = 3(1 − ν2) w3/2 0 t(cid:90) 0 E(t − τ ) d dτ (sin ωτ )3/2 dτ. (55) Now, replacing 3/2 with β in the exponent under the integral sign in (55), we obtain √ 4 R P (t) = 3(1 − ν2) w3/2 0 βω t(cid:90) 0 E(s)(sin ω(t − s))β−1 cos ω(t − s) ds. (56) Now, let tM be the time moment when the contact force (56) reaches its maximum, i. e., P (tM ) = 0. Then, by analogy with the case of linear harmonic vibrations, we put δβ(ω) = − ωtM . π 2 (57) The quantity δβ(ω) will be called the modified incomplete loss angle determined from the sinusoidally-driven displacement-controlled spherical indentation test. In view of (24), formula (57) can be rewritten in the form (30). 2.5 Modified incomplete storage modulus and loss angle. Standard viscoelastic solid model In order to fix our ideas, we assume that the layer's material follows a standard linear vis- coelastic solid model, which is described by the following normalized creep and relaxation functions: Φ(t) = 1 − (1 − ρ) exp(−t/τs), Ψ(t) = 1 − (1 − 1/ρ) exp(−t/(ρτs)). (58) Here, τs is the characteristic retardation or creep time of strain under applied step of stress, ρ is the ratio of E∞ to the unrelaxed elastic modulus E0 (modulus E(t) at t = 0), i. e., ρ = E∞/E0 < 1. The following relations are well known (Tschoegl, 1997): E1(ω) = E∞ + (E0 − E∞) ω2(ρτs)2 ω2(ρτs)2 + 1 , E2(ω) = (E0 − E∞) ωρτs , δ(ω) = arctan ω2(ρτs)2 + 1 (1 − ρ)ωρτs ρ + ω2(ρτs)2 . (59) (60) Fig. 4a shows the behavior of the modified incomplete storage modulus Eβ(ω) in com- parison with that of the storage modulus E1(ω) given by (59) for β = 1, 1.5, and 2. At high frequencies, the both dimensionless quantities E1(ω)/E∞ and Eβ(ω)/E∞ approach the limit value E0/E∞ = 1/ρ = 2. Fig. 4b shows the relative error of the approximation Figure 4. Modified incomplete storage modulus Eβ(ω). Figure 5. Modified incomplete loss angle δβ(ω). of E1(ω) by E3/2(ω) for different values of the dimensionless parameter ρ. In each case, it is assumed that the mean relaxation time ρτs is the same. Fig. 5a presents the behavior of the modified incomplete loss angle δM (ω) determined from the displacement-controlled indentation test in comparison with that of the loss angle δ(ω) given by (60) for β = 1, 1.5, and 2. The error of the approximation of the loss angle δ(ω) by δM (ω) is shown in Fig. 5b. Unfortunately, the deference between δM (ω) and δ(ω) does not vanish as ω → ∞. It will be shown that the limit value of the relative error is 33.33%. 01020304011.21.41.61.82Relative modified incomplete storage modulusRelative frequencys5.0(a)EE)(~125.1EE)(100.511.5211.11.21.31.41.501234567891015105051015201.03.05.07.09.0Relative frequencysconst)(sRelative error, %(b)100)()(~)(11EEE5.1051015200481216Relative frequencyIncomplete loss angle, deg.(a)15.1)()(~5.0220s02468100102030405060Relative error, %(b)1.03.05.07.09.0100)()(~)(5.1Relative frequencysconst)(s 3 Accounting for the thickness effect in spherical indentation of a viscoelastic layer 3.1 Dynamic parameters for assessing the mechanical properties and viability of articular cartilage by a spherical indentation test Finally, let us consider the quantity 3(1 − ν2) √ 4 R P (tm) 0 κs(αm) w3/2 = E0 3/2(ω, 0), (61) where κs(αm) is the indentation scaling factor corresponding to the maximum indentation depth am = a(tm), thus αm = am/h. According to Eqs. (46) and (49), the right-hand side of Eq. (61) is determined as follows: E0 3/2(ω, 0) = (sin ωτ )3/2f (0 sin ωτ ) dτ. (62) (cid:90) π/(2ω) (cid:18) π 2ω E (cid:19) d (cid:110) dτ − τ 1 f (0) 0 √ (cid:111) Here we introduced the notation (see Eq. (50)) √ 0 = . (63) w0R h Now, let t0 P (t0 M be the time moment when the contact force (49) reaches its maximum, i. e., M ) = 0. Then, by analogy with the modified incomplete loss angle δ3/2(ω), we put δ0 3/2(ω) = − ωt0 M . π 2 (64) In view of (24), formula (64) can be rewritten in the form (30). The quantities E0 simply called the modified storage modulus and modified loss angle. 3/2(ω, 0) and δ0 3/2(ω) (with explicit dependence on 0 hidden) will be It should be emphasized that the quantities E0 3/2(ω) depend on the layer thickness, although this fact is not reflected in the notation. Thus, in the viscoelastic case, the application of the indentation scaling factor κs(αm) in the same way as on the right-hand side of formula (61) does not completely accounts for the thickness effect. 3/2(ω, 0) and δ0 Fig. 6a presents the comparison of the quantity E0 3/2(ω, 0) with the modified incomplete storage modulus E3/2(ω) in the case of standard solid model. As it could be expected, the difference tends to zero as ω → 0, that is as the indentation protocol approaches the quasi-static limit. In the high-frequency range (as ω → ∞), it can be also established rigorously that E0 3/2(ω, 0) and E3/2(ω) both tend to E(0) as well as E1(ω) does. On M (ω) and δM (ω) does not vanish as ω → ∞ (see the contrary, the deference between δ0 Fig. 6b), and the limit value of the relative error depends on the value of 0. Figure 6. Difference between E0 indentation. 3/2(ω, 0), δ0 3/2(ω) and E3/2(ω), δ3/2(ω) for different levels of 3.2 Asymptotic analysis of E0 3/2(ω, 0) in the low- and high-frequency limits To fix our ideas, let us assume that the relaxation modulus E(t) is determined by the Prony series E(t) = E∞ + Ej exp (65) m(cid:88) j=1 (cid:18) − t ρj (cid:19) , where Ej and ρj are positive constants representing the relaxation strengths and relaxation times. Without loss of generality we may assume that ρ1 < ρ2 < . . . < ρm. Integrating by parts in (62) and changing the integration variable, we rewrite Eq. (62) as E(cid:48)(cid:18) z π/2(cid:90) (cid:19) ω 0 √ (cos z)βf (0 cos z) dz. (66) E0 β(ω, 0) = E(0) + 1 ωf (0) Here again it is assumed that β = 3/2. In the low-frequency limit, the behavior of E0 asymptotic behavior of the integral β(ω, 0) as ω → 0 will depend on the I j β(ω, 0) = exp (cos z)βf (0 √ cos z) dz. (67) π/2(cid:90) 0 (cid:19) (cid:18) − z ωρj (cid:90) Using the formula (Dwight (1961), formula (567.9)) xn exp(ax) dx = exp(ax) it can be easily shown that k=0 n(cid:88) (−1)n−kn! k!an+1−k xk, 012345678910Relative frequencysRelative error, %32100.51.52.5100)(~)(~)(~0EEE5.004.003.002.001.005.05.05.1(a)051015202530Relative error, %(b)100)(~)(~)(~000.40.81.21.62Relative frequencys5.004.003.002.001.005.05.05.1 π/2(cid:90) exp 0 (cid:18) − z ωρj (cid:19) (cid:18) zn dz = n!(ωρj)n+1 − exp (cid:18) − π 2ωρj (cid:19) n(cid:88) (cid:18) k=0 = n!(ωρj)n+1 + O (ωρj)n+1 exp (cid:18) π (cid:19)k 2 n!(ωρj)n+1−k (cid:19)(cid:19) k! − π 2ωρj , ω → 0. (68) Thus, making use of formula (68) and the two-term Taylor expansion 1 f (0) (cos z)βf (0 √ cos z) = 1 − z2 f(cid:48)(0) f (0) 0 4 + + O(z4), (cid:32) β 2 (cid:33) one can now expand the integral I j to ω2. In such a way, we arrive at the following asymptotic representation: β(ω, 0) determined by (68) in power series with respect E0 β(ω, 0) = E∞ + ω2 Ejρ2 j β + Here it was taken into account that j=1 E(0) = E∞ + m(cid:88) j=1 Ej. + O(ω4). (69) (70) (cid:32) m(cid:88) (cid:33) f(cid:48)(0) f (0) 0 2 m(cid:88) On the other hand, the following two-term asymptotic expansions holds true for the storage modulus (Tschoegl, 1997): E1(ω) = E∞ + ω2 Ejρ2 j + O(ω4), ω → 0. (71) j=1 Comparing asymptotic expansions (69) and (71), we see that the second terms on their right-hand sides coincide only if β = 1 and 0 = 0, when E0 β(ω, 0) coincides with the incomplete storage modulus E1(ω). β(ω, 0) as ω → ∞ depends on the smooth- In the high-frequency limit, the behavior of E0 ness properties of the function E(cid:48)(t) at the point t = 0. In view of (65), Eq. (66) readily yields where E(cid:48)(0) = −(cid:80)m j=1 Ej/ρj and E0 β(ω, 0) = E(0) + E(cid:48)(0) ω I 0 β(0) f (0) + O(ω−2), (72) π/2(cid:90) (cos z)βf (0 0 √ cos z) dz. I 0 β(0) = On the other hand, the following two-term asymptotic expansions holds true for the storage modulus (Argatov, 2012): E1(ω) = E(0) − E(cid:48)(cid:48)(0) ω2 + O(ω−4), ω → ∞. (73) Thus, based on the asymptotic expansions (72) and (73), it is established that E0 tends to E(0) as ω → ∞ as well as E1(ω) does. β(ω, 0) 3.3 Asymptotic analysis of δ0 β(ω) in the low- and high-frequency limits Differentiating both sides of Eq. (49) with respect to time and taking into account the sinusoidal protocol (23), we get 3(1 − ν2)R 4h33 0 P (t) = E(0) (sin ωt)βf (0 + (sin ωτ )βf (0 sin ωτ ) (cid:16) (cid:110) d dt d dτ t(cid:90) 0 (cid:17) √ sin ωt) √ (cid:111) E(cid:48)(t − τ ) dτ. Substituting now the value (see Eq. (64)) t0 M = (cid:18) π 2 1 ω (cid:19) − δ0 β into the equation P (t0 M ) = 0 in view of (74), we arrive at the following equation: (cid:16) cos δ0 β E(0)ω sin δ0 β (cid:17)β−1F 0 β(δ0 β) = (cid:90) π/2−δ0 β (cid:16) 0 × f (cid:26)(cid:16) (cid:113) d dz cos(δ0 β + z) 0 cos(δ0 β + z) (cid:17)β (cid:17)(cid:27) E(cid:48)(cid:18) z (cid:113) ω (cid:19) dz. (cid:17) . Here we introduced the notation F 0 β(δ0 β) = βf 0 (cid:16) (cid:113) cos δ0 β (cid:17) + 0 2 (cid:113) βf(cid:48)(cid:16) cos δ0 0 cos δ0 β (74) (75) (76) β, ω) and R(δ0 Let L(δ0 Following Argatov (2012), we construct solutions to Eq. (75), assuming that δ0 as ω → 0 and δ0 M (ω) (cid:28) 1 such that δ0 β, ω) denote the left and right hand sides of Eq. (75), respectively. M (ω) (cid:39) C0ω M (ω) (cid:39) C∞/ω as ω → ∞, where C0 and C∞ are constants. In both cases, L(δ0 β, ω) = E(0)ωδ0 βF 0 β(0) + O β → 0, δ0 , (77) (cid:16) β)3(cid:17) (δ0 where according to (76) we have F 0 β(0) = βf (0) + f(cid:48)(0). 0 2 In the low-frequency limit, making use of the asymptotic formula (68), we get m(cid:88) m(cid:88) R(δ0 β, ω) = ωδ0 βF 0 β(0) Ej + ω2F 0 β(0) Ejρj + O(ω3), ω → 0. (78) j=1 j=1 From (77) and (78), it follows that β(ω) (cid:39) ω δ0 E∞ m(cid:88) j=1 Ejρj + O(ω2), ω → 0. (79) Here the relation (70) was taken into account. We emphasize that the asymptotic representation (79) is in complete agreement with the leading term of the asymptotic expansion for the loss angle δ(ω) as ω → 0. In the high-frequency limit, we will have R(δ0 β, ω) = (cos z)βf (0 (cid:110) π/2(cid:90) 0 d dz (cid:111) E(cid:48)(0) dz + O(ω−1), ω → ∞. cos z) √ Now, from (77) and (80), it follows that β(ω) (cid:39) − E(cid:48)(0) δ0 E(0)ω f (0) F 0 β(0) , ω → ∞. (80) (81) On the other hand, the following asymptotic representation holds true for the loss angle: δ(ω) (cid:39) − E(cid:48)(0) E(0)ω , ω → ∞. (82) Comparing relations (81) and (82), we see that they coincide only if β = 1 and f (0) ≡ 1. In the general case, in view of (76), we have δ(ω) − δ0 δ(ω) β(ω) (cid:39) 2(β − 1)f (0) + 0f(cid:48)(0) 2βf (0) + 0f(cid:48)(0) , ω → ∞. (83) Thus, according to (83), in the case of a viscoelastic half-space, when f (0) ≡ 1, the β(ω) for δ(ω) approaches the value (β − 1)/β · 100%. relative error of the approximation δ0 4 Discussion The new material characteristics introduced above, that is the incomplete storage modulus E1(ω), the modified incomplete storage modulus Eβ(ω), and the modified storage modulus E0 β(ω, 0), can be represented as follows: π/(2ω) E1(ω) = − E(s) {cos ωs} ds, d ds Eβ(ω) = − (cid:90) π/(2ω) E(s) d ds (cid:26) (cid:110) (cos ωs)β(cid:111) ds, √ (84) (85) ds. (86) (cid:27) (cid:90) 0 (cid:90) 0 π/(2ω) ∞(cid:90) 0 β(ω, 0) = − E0 0 E(s) d ds (cos ωs)β f (0 cos ωs) f (0) In the same way, the storage modulus E1(ω) is recast as E1(ω) = − E(s) d ds {cos ωs} ds. (87) Thus, comparing formulas (84) -- (86) with (87), we see that E1(ω), Eβ(ω), and E0 β(ω, 0), represent a hierarchy of approximations for E1(ω). Applying an asymptotic modeling approach for analyzing the interrelations between the new quantities, we have shown that the modified storage moduli asymptotically coincide with the storage modulus in the low- and high-frequency ranges. The values β = 1, β = 3/2, and β = 2 correspond respectively to the cases of cylindri- cal, spherical, and conical indenters. The latter case also applies to pyramidal indenters (Giannakopoulos, 2006; Argatov, 2011). Fig. 7 illustrates the relationship between the parameters of viscoelastic materials mea- sured in a vibration indentation test and in a single indentation test with a flat-ended cylindrical indenter. Due to the nonmonotonic behavior of the modified incomplete storage modulus Eβ(ω) with respect to the storage modulus E1(ω) as it was shown in Fig. 4 (for the standard viscoelastic solid model), the relationship between the parameters measured in the vibration and indentation tests with a spherical indenter will be more complicated. Figure 7. Complex dynamic modulus schematic diagram. Measured in indentation test2E1E1~E~EMeasured in vibration test Now, let us consider the application of the developed theory to experimental data (Ronken et al., 2011). The experimental setup was described in detail elsewhere (Wirz et al., 2008). The modified moduli E0 3/2(ω) of swine hyaline cartilage at two different locations calculated using the following formulas (see, Eqs. (61) and (30)): 3/2(ω, 0) and loss angles δ0 E0 3/2(ω, 0) = δ0 3/2(ω) = π 2 P (tm) 0 f (0) 3(1 − ν2) √ w3/2 4 R (tm − tM ) . tm , (88) (89) Here, f (0) = κs(αm) is the indentation scaling factor corresponding to the maximum indentation depth and calculated according to Eq. (47). A Poisson's ratio of 0.5 was assumed. Note that the symbol tm now denotes the time moment of maximum indentation instead of the symbol tm, because in the impact tests the indentation variation w(t) does not follow the sine law (23) precisely. Table 2 Mean and standard deviation of the main parameters of two impact indentation tests (ten repetitions on one spot) for two swine cartilage samples with a spherical impactor of radius R = 0.5 mm and mass m = 1.9 g (Ronken et al., 2011). Sample thickness, h (mm) Initial indenter velocity, v0 (m/s) Time to maximum contact force, tM (ms) Maximum contact force, P (tM ) (N) Indentation duration, tm (ms) Maximum indentation, w0 (mm) Contact force at maximum indentation, P (tm) (N) Effective angular frequency, ω (×103 rad/s) Level of indentation, 0 Modified storage modulus, E0 Modified loss angle, δ0 3/2(ω, 0) (MPa) 3/2(ω) (rad) Coefficient of restitution, e∗ Lateral condyle Medial condyle 1.7 0.249 ± 0.002 0.68 ± 0.01 1.04 ± 0.02 0.74 ± 0.01 0.125 ± 0.002 1.02 ± 0.02 2.13 ± 0.02 0.147 ± 0.001 15.3 ± 0.5 0.114 ± 0.007 0.777 ± 0.006 1.9 0.266 ± 0.002 0.73 ± 0.01 0.96 ± 0.01 0.81 ± 0.01 0.145 ± 0.003 0.93 ± 0.01 1.93 ± 0.02 0.158 ± 0.002 11.0 ± 0.3 0.167 ± 0.008 0.722 ± 0.012 The data shown in the upper part of Table 2 was directly assessed in experiments, while the lower part of the table displays results evaluated according to the theory developed herein. The example illustrates the fact that the introduced characteristics E0 3/2(ω, 0) and δ0 3/2(ω) depend on the effective angular frequency ω, which in turn depends on the initial indenter velocity v0 as well as on the mechanical properties of the sample itself. As it could be expected, at high frequencies, the modulus E0 3/2(ω, 0) increases with increasing ω, while the angle δ0 3/2(ω) decreases (see Table 2). Note also that this example demonstrates a correlation in behavior of the modified loss angle δ0 3/2(ω) and the coefficient of restitution e∗. Of course, the impact indentation test requires a special consideration, but the observed characteristic behavior of E0 is quite typical. 3/2(ω, 0) and δ0 3/2(ω) with the change in ω Now, let us discuss the significance of the developed mathematical approach from the viewpoint of formulating criteria for evaluation the quality of articular cartilage. In the cylindrical (flat-ended) and spherical dynamic indentation tests, the following cartilage stiffness-related characteristics can be evaluated (see Eqs. (26) and (61), respectively): 1 − ν2 2aκc(α) 3(1 − ν2) √ 4 R P (tm) w0 , P (tm) 0 κs(αm) w3/2 . (90) (91) Here, w0 = w(tm) is the maximum indentation depth, P (tm) is the contact force corre- sponding to the time moment t = tm, when the indenter reaches its maximum indentation depth. The criteria (90) and (91) in their static form (with no attention paid to the dy- namic nature of indentation process) have been used in a number of experimental studies on detection of degenerative changes in joint cartilage. First of all, it should be noted that the indentation scaling factors κc(α) and κs(αm) were evaluated under the assumption of isotropy and homogeneity of articular cartilage layer. It is anticipated that the inhomogeneity effect will be smaller in spherical indentation. In view of the layered structure of articular cartilage, the the anisotropy effect requires tacking into account the adjusted value of indentation scaling factor and the corresponding Poisson's ratio. Because Poisson's ratio ν enters formulas (90) and (91) not only through the factor 1− ν2 but also through the dependence of κc(α) and κs(αm) on ν, the question of the appropriate choice of ν for the criteria (90) and (91) is more than academic, and it still remains open. Second, in dynamic indentation testing, the criteria (90) and (91) will depend on the loading protocol employed, because articular cartilage exhibits viscoelastic and poroelastic properties. Thus, in order to increase the sensitivity of the measurements with a hand- held indentation probe, the indentation protocol should be reproducible as well as the indentation time tm should be kept the same. Third, in the framework of linear viscoelasticity, the criteria (90) and (91) are interpreted as the incomplete storage modulus E1(ω) and the modified storage modulus E0 3/2(ω, 0). Due to the linearity of the theory under consideration, the criterium (90) does not depend on the level of indentation. Hence, this fact could be used to check whether the linearity assumption is appropriate for small indentation depths, when w0/h ≤ 0.1 or even less. On the other hand, the criterium (91) does depend on the level of indentation determined by the parameter (see Eq. (63)) √ w0R h . 0 = However, since the main manifestation of the thickness effect has been taken into account by means of the indentation scaling factor κs(αm) evaluated at the maximum indentation depth, the dependence of E0 3/2(ω, 0) on 0 is rather weak as it is predicted by the standard viscoelastic solid model (see Fig. 6a). Further, in both indentation tests, one can also measure a dimensionless quantity that is directly related to time-dependent energy dissipation due to viscoelastic and poroelastic relaxation. Namely, the incomplete loss angle δ(ω) and the modified loss angle δ0 3/2(ω) were introduced based on the time shift between maxima of of the input (indentation dis- placement) and the output (contact force) through Eq. (30). It is important to emphasize that the quantity δ(ω), which is measured in the flat-ended indentation test, does not depend on the thickness of the articular cartilage layer. At the same time, the thickness effect plays a crucial role in manifestation of the time-dependent response to indentation with a spherical indenter (see Fig. 6b). Finally, the developed viscoelastic models of cylindrical and spherical dynamic indentation tests allow one to compare the diagnostics criteria (that is diagnostics characteristics) experimentally measured by different indentation probes utilizing different indentation protocols (e.g., more closely approximating the actual movement of an operator's hand) as well as operating in different modes (vibration, dynamic indentation, impact testing). In view of the established fact that the incomplete storage modulus E1(ω) and the modified storage modulus E0 3/2(ω, 0) asymptotically coincide with the storage modulus E1(ω) in the low- and high-frequency ranges, the interrelationships between the different tests will be of particular interest in the middle frequency range. 5 Conclusions We considered frictionless flat-ended and spherical sinusoidally-driven indentation tests utilizing displacement-controlled loading protocol. In order to perform a rigorous analysis, we modeled the deformational behavior of articular cartilage tissue in the framework of viscoelasticity with a time-independent Poisson's ratio. In the linear case of flat-ended indentation test, evaluating the dynamic indentation stiffness at the test turning point t = tm, we introduced the incomplete storage modulus E1(ω) for the effective frequency ω = π/(2tm). Considering the time difference tm − tM between the time moments when the contact force reaches its maximum (dynamic stiffness vanishes at t = tM ) and the indenter displacement reaches its maximum (dynamic stiffness becomes infinite at t = tm), we introduced the so-called incomplete loss angle δ(ω). Analogous quantities were introduced in the nonlinear case of spherical sinusoidally-driven indentation test. First, when the sample thickness effect can be neglected, we introduced the modified incomplete storage modulus E3/2(ω) and the modified incomplete loss angle δ3/2(ω) (we use the same notation as in the linear case). Second, to account for the thick- ness effect, we introduced the indentation scaling factor κs(α) for the spherical indenter depending on Poisson's ratio and the relative contact radius α = a/h. Making use of the indentation scaling factor corresponding to the maximum indentation depth κs(αm), we introduced the modified storage modulus E0 3/2(ω, 0) which depends on the level of in- 3/2(ω) dentation characterized by the parameter 0 = was introduced in the same way. w0R/h. The modified loss angle δ0 √ We applied an asymptotic modeling approach for analyzing the interrelations between the new quantities E0 3/2(ω) and the classical characteristics E1(ω), δ(ω) in the 3/2(ω, 0), δ0 low- and high-frequency ranges. It was shown that the modified storage modulus asymp- 3/2(ω, 0) (cid:39) totically coincides with the storage modulus in the both limit cases, that is E0 E1(ω) as ω → 0 and ω → ∞. However, coinciding with the loss angle δ(ω) in the low- frequency range, the modified loss angle δ0 β(ω) markedly differs from δ(ω) in the high- frequency limit. We illustrated these facts for the standard viscoelastic solid model. Finally, the present study suggests that the use of dynamic indentation tests is largely twofold: the criteria (90) and (91) yield dimensional diagnostics characteristics, which could be related to some integral measure of material properties of the tested biological tissue; the second aspect of dynamic indentation diagnostics hinges on the importance of continuous monitoring of the tissue response to indentation. It is believed that both characteristics E1(ω) (evaluated according to the criterium (90)) and δ(ω), which are as- sociated with flat-ended indentation tests, can elicit perceptions of the mechanical quality of articular cartilage. In the dynamic non-destructive testing with a spherical indenter, in view of the fact that the thickness of a biological tissue sample is supposed to be unknown, the pronounced effect of the sample thickness on the modified loss angle δ0 3/2(ω) observed for different levels of indentation can be used as an indicator of the importance of the thickness effect for the modified storage modulus E0 3/2(ω, 0). Acknowledgements One of the authors (I.A.) gratefully acknowledges the support from the European Union Seventh Framework Programme under contract number PIIF-GA-2009-253055. References Appleyard, R.C., Swain, M.V., Khanna, S., Murrell, G.A.C., 2001. The accuracy and reli- ability of a novel handheld dynamic indentation probe for analysing articular cartilage. Phys. Med. Biol. 46, 541 -- 550. Argatov, I.I., 2001. The pressure of a punch in the form of an elliptic paraboloid on an elastic layer of finite thickness. J. Appl. Math. Mech. 65, 495 -- 508. Argatov, I.I., 2002. Characteristics of local compliance of an elastic body under a small punch indented into the plane part of its boundary. J. Appl. Mech. Techn. Phys. 43, 147 -- 153. Argatov, I.I., 2010. Frictionless and adhesive nanoindentation: Asymptotic modeling of size effects. Mech. Mater. 42, 807 -- 815. Argatov, I., 2011. Depth-sensing indentation of a transversely isotropic elastic layer: Second-order asymptotic models for canonical indenters. Int. J. Solids Struct. 48, 3444 -- 3452. Argatov, I., 2012. Sinusoidally-driven flat-ended indentation of time-dependent materials: Asymptotic models for low and high rate loading. Mech. Mater. 48, 56 -- 70. Argatov, I., Mishuris, G., 2011. An analytical solution for a linear viscoelastic layer loaded with a cylindrical punch: Evaluation of the rebound indentation test with application for assessing viability of articular cartilage. Mech. Res. Comm. 38, 565 -- 568. Armstrong, C.G., Lai, W.M., Mow, V.C., 1984. An analysis of the unconfined compression of articular cartilage. J. Biomech. Eng. 106, 165 -- 173. Bae, W.C., Temple, M.M., Amiel, D., Coutts, R.D., Niederauer, G.G., Sah, R.L., 2003. Indentation testing of human cartilage: Sensitivity to articular surface degeneration. Arthritis Rheum. 48, 3382 -- 3394. Borodich, F.M., Keer, L.M., 2004. Contact problems and depth-sensing nanoindentation for frictionless and frictional boundary conditions. Int. J. Solids Struct. 41, 2479 -- 2499. Broom, N.D., Flachsmann, R., 2003. Physical indicators of cartilage health: The relevance of compliance, thickness, swelling and fibrillar texture. J. Anat. 202, 481 -- 94. Cao, Y., Ma, D., Raabe, D., 2009. The use of flat punch indentation to determine the viscoelastic properties in the time and frequency domains of a soft layer bonded to a rigid substrate. Acta Biomaterialia 5, 240 -- 248. Cheng, Y.-T., Yang, F., 2009. Obtaining shear relaxation modulus and creep compli- ance of linear viscoelastic materials from instrumented indentation using axisymmetric indenters of power-law profiles. J. Mater. Res. 24, 3013 -- 3017. Christensen, R.M., 1971, Theory of Viscoelasticity. Academic Press, New York. de Freitas, P., Wirz, D., Stolz, M., Gopfert, B., Friederich, N.-F., Daniels, A.U., 2010. Pulsatile dynamic stiffness of cartilage-like materials and use of agarose gels to validate mechanical methods and models. J. Biomed. Mater. Res. Part B: Appl. Biomater. 78B, 347 -- 357. Dwight, H.B., 1961. Tables of Integrals and Other Mathematical Data. The Macmillan Company, New York. Galin, L.A., 1946. Spatial contact problems of the theory of elasticity for punches of circular shape in planar projection. J. Appl. Math. Mech. (PMM) 10, 425 -- 448 (in Russian). Giannakopoulos, A.E., 2006. Elastic and viscoelastic indentation of flat surfaces by pyra- mid indentors. J. Mech. Phys. Solids 54, 1305 -- 1332. Hayes, W.C., Keer, L.M., Herrmann, G., Mockros, L.F., 1972. A mathematical analysis for indentation tests of articular cartilage. J. Biomech. 5, 541 -- 551. Hayes, W.C., Mockros, L.F., 1971. Viscoelastic properties of human articular cartilage. J. Appl. Physiol. 31, 562 -- 568. Hu, K., Radhakrishnan, P., Patel, R.V., Mao, J.J., 2001. Regional structural and vis- coelastic properties of fibrocartilage upon dynamic nanoindentation of the articular condyle. J. Struct. Biol. 136, 46 -- 52. Korhonen, R.K., Saarakkala, S., Toyras, J., Laasanen, M.S., Kiviranta, I., Jurvelin, J.S., 2003. Experimental and numerical validation for the novel configuration of an arthro- scopic indentation instrument. Phys. Med. Biol. 48, 1565 -- 1576. Kren, A.P., Naumov, A.O., 2010. Determination of the relaxation function for viscoelastic materials at low velocity impact. Int. J. Impact Eng. 37, 170 -- 176. Kusano, T., Jakob, R.P., Gautier, E., Magnussen, R.A., Hoogewoud, H., Jacobi, M., 2011. Treatment of isolated chondral and osteochondral defects in the knee by autologous matrix-induced chondrogenesis (AMIC). Knee Surg. Sports Traumatol. Arthrosc. DOI 10.1007/s00167-011-1840-2. Lau, A., Oyen, M.L., Kent, R.W., Murakami, D., Torigaki, T., 2008. Indentation stiffness of aging human costal cartilage. Acta Biomater. 4, 97 -- 103. Lyyra-Laitinen, T., Niinimaki, M., Toyras, J., Lindgren, R., Kiviranta, I., Jurvelin, J.S., 1999. Optimization of the arthroscopic indentation instrument for the measurement of thin cartilage stiffness. Phys. Med. Biol. 44, 2511 -- 2524. Oyen, M.L., 2005. Spherical indentation creep following ramp loading. J. Mater. Res. 20, 2094 -- 2100. Parsons, J.R. and Black, J., 1977. The viscoelastic shear behavior of normal rabbit artic- ular cartilage. J. Biomech. 10, 21 -- 29. Pipkin, A.C., 1972. Lectures on Viscoelasticity Theory. Springer, Berlin. Ronken, S., Arnold, M.P., Ardura Garc´ıa, H., Jeger, A., Daniels, A.U., Wirz, D., 2011. A comparison of healthy human and swine articular cartilage dynamic indentation mechanics. Biomech. Model. Mechanobiol. DOI: 10.1007/s10237-011-0338-7. Schinagl, R.M., Gurskis, D., Chen, A.C., Sah, R.L., 1997. Depth-dependent confined com- pression modulus of full-thickness bovine articular cartilage. J. Orthop. Res. 15, 499 -- 506. Suh, J.-K., Li, Z., Woo, S.L.-Y., 1995. Dynamic behavior of a biphasic cartilage model under cyclic compressive loading. J. Biomech. 28, 357 -- 364. Ting, T.C.T., 1968. Contact problems in the linear theory of viscoelasticity. J. Appl. Mech. 35, 248 -- 254. Tschoegl, N.W. 1997. Time dependence in material properties: An overview. Mech. Time- Depend. Mat. 1, 3 -- 31. Toyras, J., Lyyra-Laitinen, T., Niinimaki, M., Lindgren, R., Nieminen, M.T., Kiviranta, I., Jurvelin, J.S., 2001. Estimation of the Young's modulus of articular cartilage using an arthroscopic indentation instrument and ultrasonic measurement of tissue thickness. J. Biomech. 34, 251 -- 256. Vorovich, I.I., Aleksandrov, V.M., Babeshko, V.A. 1974. Non-classical Mixed Problems of the Theory of Elasticity. Nauka, Moscow [in Russian]. Wirz, D., Kohler, C., Keller, K., Gopfert, B., Hudetz, D., Daniels, A.U., 2008. Dynamic stiffness of articular cartilage by single impact micro-indentation (SIMI). J. Biomech. 41, Suppl. 1, P. S172. Zhang, C.Y., Zhang, Y.W., 2004. Extracting the mechanical properties of a viscoelastic polymeric film on a hard elastic substrate. J. Mater. Res. 14, 3053 -- 3061.
1708.02876
2
1708
2017-11-29T20:26:05
Elucidating distinct ion channel populations on the surface of hippocampal neurons via single-particle tracking recurrence analysis
[ "physics.bio-ph" ]
Protein and lipid nanodomains are prevalent on the surface of mammalian cells. In particular, it has been recently recognized that ion channels assemble into surface nanoclusters in the soma of cultured neurons. However, the interactions of these molecules with surface nanodomains display a considerable degree of heterogeneity. Here, we investigate this heterogeneity and develop statistical tools based on the recurrence of individual trajectories to identify subpopulations within ion channels in the neuronal surface. We specifically study the dynamics of the K$^+$ channel Kv1.4 and the Na$^+$ channel Nav1.6 on the surface of cultured hippocampal neurons at the single-molecule level. We find that both these molecules are expressed in two different forms with distinct kinetics with regards to surface interactions, emphasizing the complex proteomic landscape of the neuronal surface. Further, the tools presented in this work provide new methods for the analysis of membrane nanodomains, transient confinement, and identification of populations within single-particle trajectories.
physics.bio-ph
physics
Elucidating distinct ion channel populations on the surface of hippocampal neurons via single-particle tracking recurrence analysis Grzegorz Sikora1,2, Agnieszka Wy loma´nska1 and Janusz Gajda1, Laura Sol´e3, Elizabeth J. Akin3, Michael M. Tamkun3,4, and Diego Krapf2,5 1Faculty of Pure and Applied Mathematics, Hugo Steinhaus Center, Wroc law University of Science and Technology, 50-370 Wroc law, Poland 2Department of Electrical and Computer Engineering, Colorado State University, Fort Collins, CO 80523, USA 3Department of Biomedical Sciences, Colorado State University, Fort Collins, CO 80523, USA 4Department of Biochemistry and Molecular Biology, Colorado State University, Fort Collins, CO 80523, USA and 5School of Biomedical Engineering, Colorado State University, Fort Collins, CO 80523, USA∗ (Dated: December 1, 2017) Protein and lipid nanodomains are prevalent on the surface of mammalian cells. In particular, it has been recently recognized that ion channels assemble into surface nanoclusters in the soma of cultured neurons. However, the interactions of these molecules with surface nanodomains display a considerable degree of heterogeneity. Here, we investigate this heterogeneity and develop statistical tools based on the recurrence of individual trajectories to identify subpopulations within ion channels in the neuronal surface. We specifically study the dynamics of the K+ channel Kv1.4 and the Na+ channel Nav1.6 on the surface of cultured hippocampal neurons at the single-molecule level. We find that both these molecules are expressed in two different forms with distinct kinetics with regards to surface interactions, emphasizing the complex proteomic landscape of the neuronal surface. Further, the tools presented in this work provide new methods for the analysis of membrane nanodomains, transient confinement, and identification of populations within single-particle trajectories. I. INTRODUCTION One of the most striking features of mammalian cells lies in their ability to perform extremely intricate func- tions with a limited number of protein-coding genes. This number is much smaller than originally estimated [1, 2]. For example, human and mouse genomes have merely 19,817 and 21,968 protein-coding genes (GENCODE 26 and GENCODE M13 [3]). In order to reach the diversity and complexity required by cells in any mammal, genes can produce multiple protein forms, which can further be chemically modified at several locations. As a con- sequence, cells can employ the same protein for remark- ably different functions. A particular type of proteins that exhibit exceptional diversity are integral membrane proteins, such as receptors and ion channels. It is es- timated that approximately 26% of the human protein- coding genes code for membrane proteins [4]. Biological systems are often characterized by both static [5] and dynamic [6] heterogeneities. However, such disorder cannot be usually probed by ensemble-averaged measurements. On the other hand, single-molecule tech- niques are ideal for observing functional heterogeneities and to extract information on the distribution of molec- ular properties. Single-molecule experiments have pro- vided information on functional heterogeneities in enzy- matic turnover [7], RNA folding [8], Holliday junctions [9], and helicase activity [10], to name a few examples. As single-molecule techniques advance, it is becoming clear ∗ [email protected] that functional heterogeneity is ubiquitous in the com- plex realm of biological systems [11–13]. In the plasma membrane functional heterogeneities can be employed to exploit the same protein in multiple cel- lular functions or to regulate physiological processes by altering intermolecular interactions. For example, be- sides regulating action potential waveform in neurons, the ion channel Kv2.1 has a non-traditional structural role by which it induces endoplasmic reticulum/plasma membrane contact sites [14] and alters membrane protein trafficking [15]. Identifying heterogeneities and quantify- ing the distribution of molecular properties are important steps in cell biology. Single-particle tracking provides unique advantages for the investigation of the dynam- ics of individual molecules [16–18]. However, observing heterogeneous dynamics can be challenging due to the in- herent thermal fluctuations and experimental noise [19]. Some types of heterogeneous dynamics that have been recognized in trajectories in the plasma membrane in- clude hop-diffusion between actin-delimited membrane compartments [20–22], confinement in nanoscale mem- brane domains [23–25], and transient tethering to intra- cellular scaffolds [26, 27]. Thus, tools that allow both to identify heterogeneous dynamics in single-particle trajec- tories and to distinguish particle-to-particle variations in terms of their dynamics, are necessary. Different meth- ods have been developed to identify transition points within intermittent trajectories. For example, a system- level maximum-likelihood method has been employed to identify periods of confined motion within trajectories ex- hibiting Gaussian diffusion [28]. This method is very ef- fective when dealing with Gaussian-based models. Alter- natively, universal model-free methods enable the iden- tification of change points in an individual trajectory by considering a local functional that transforms the trajec- tory into a new time series. This new time series can then be used to characterize intermittent behavior [29, 30]. Examples of local functionals that have been employed include the diffusivity [31], convex hull [29], anomalous exponent of the local MSD [32], and directional changes [22, 33]. In particular, the local MSD exponent and the convex hull have been used to detect confinement zones. The advantage of local functional methods lies in the fact that they can be applied without prior knowledge of the model. In this paper we study the heterogeneous dynamics of two voltage-gated ion channels in the somatic plasma membrane of hippocampal neurons, the K+ channel Kv1.4 and Na+ channel Nav1.6. These channels are ob- served to be transiently confined in nanoscale domains, but while some molecules remain confined for minutes, others escape in less than 1 s. We introduce a local func- tional method based on recurrence analysis to identify regions of confinement in the path. Then, we classify tra- jectories employing a three-step protocol. First, a regime variance test quantifies heterogeneity in particle dynam- ics. Second, a silhouette analysis is used to identify the exact number of trajectory classes. And third, a k-means algorithm is used to set thresholds and separate trajecto- ries into different classes. We find that there are two dif- ferent classes of trajectories for both Kv1.4 and Nav1.6. These classes of trajectories have very different residence times within the confined domains. While populations that exhibit weak interactions have sojourn times with exponential tails, the populations with strong interac- tions appear to have heavy tails. These results highlight the complexity of the neuronal surface and provide tools for the study of static and dynamic heterogeneities in the plasma membrane. II. MATERIALS AND METHODS A. Cell culture, transfection, and labeling Rat hippocampal neurons were cultured and imaged in glass-bottomed plates as previously described [25, 34]. Animals were used according to protocols approved by the Institutional Animal Care and Use Committee of Colorado State University (Animal Welfare Assurance Number A3572-01). Nav1.6 and Kv1.4 constructs were each modified to contain an extracellular biotin accep- tor domain (BAD) in an extracellular loop. These con- structs (Nav1.6-BAD and Kv1.4-BAD) were previously functionally validated [34]. Neuronal transfections were performed after 6 days in culture for Nav1.6 and 7 days in culture for Kv1.4, using Lipofectamine 2000 (LifeTech- nologies, Grand Island, NY). Cells were co-transfected with 1 µg of either Kv1.4-BAD or Nav1.6-BAD and 1 µg pSec-BirA (bacterial biotin ligase) to biotinylate the 2 channel. Labeling of the surface channel was performed before imaging at DIV10. Neurons were rinsed with neu- ronal imaging saline (NIS), to remove the Neurobasal me- dia. Cells were incubated for 10 min with streptavidin- conjugated CF640R (Biotium, Hayward, CA) diluted 1:1000 in NIS. Streptavidin-CF64R labeling was done at 37◦C in the presence of 1% bovine serum albumin (cat. A0281, Sigma, St Louis, MO). Excess label was removed by rinsing with neuronal imaging saline. B. Imaging Total internal reflection fluorescence (TIRF) images were acquired at 20 frames per second. Before TIRF imaging, differential interference contrast (DIC) and wide- field fluorescence imaging were used to distinguish transfected neurons from the relatively flat glia. Neurons were readily identified based on the characteristic soma morphology and localization of Nav1.6 to the axon ini- tial segment. All imaging was performed at 37◦C using objective and stage heaters. C. Image processing and single-molecule tracking Images were background subtracted and filtered using a Gaussian kernel with a standard deviation of 0.6 pixels in ImageJ. Tracking of individual fluorophores was then performed in MATLAB using the U-track automated algorithm [35]. Manual inspection confirmed accurate single-molecule detection and tracking. D. Identification of transient confinement periods In order to identify periods of transient confinement within individual trajectories we developed an algorithm based on trajectory recurrence analysis where we eval- uate the total number of visits to the current site [36]. When a particle is confined within a nanoscale domain, it moves in small area unavoidably visiting the same sites multiple times in a short period. In contrast, during free unconfined motion, the random walk is less compact and its exploration region in the same time is wider. In the re- currence analysis algorithm, at each particle position we calculated the distance to the subsequent point and con- structed a circle with diameter equal to this distance, cen- tered midway between the two consecutive points. Next, the number of times the walker position lies within circle, V ′ j denotes the number of visits to site j, where j = 1, 2, ..., N −1 and N is number of data points in the trajectory. The method by which Vj is found is illustrated in Figs. 1A and B for two simulated trajec- tories. To improve the algorithm reliability and enhance the differences between confinement and free states, we first segment the data into disjoint windows of size n = 3 and then sum over the three consecutive V ′ j values within j , is calculated. Thus, V ′ 1 +V ′ each window. For example, V1 = V2 = V3 = V ′ 2 +V ′ 3 . The identification of states is performed according to Vj remaining either above or below a given threshold (Vth). The threshold is selected taking under consideration the behavior of the analyzed data and can vary for diverse data sets. In our experimental data, the threshold was selected to be Vth = 11; and in synthetic data Vth = 6. The procedure for threshold selection is further detailed in Section III. The statistic Vj is susceptible to statistical noise. Namely, there is a finite, albeit small, probability that Vj crosses the threshold in a single window even though there is no real change of behavior in the data. Nev- ertheless, the probability that such events take place in two consecutive windows is much smaller. Thus we elimi- nate most false-positive point changes by considering the dwell time within each state. If a particle crosses the threshold but the dwell time within the new state is only a single time window, i.e., three points, the time series is considered to remain within the same state. E. Determination of number of classes among particles We determined the number of different trajectory classes based on the time spent in the confined state. After segmentation of trajectories into free and confined states, we calculated the total fraction of time φ each tra- jectory resides within the confined state. For a trajectory i with m confined sojourn times τik, the fraction of time in the confined state is φi = 1 Ti m X k=1 τik, (1) where Ti is the observation time. Then, we evaluated if there exist at least two types of trajectories by em- ploying the regime variance test described in Ref. [37]. Briefly, given M trajectories with φ1, φ2, ..., φM fractions of time, we first visually examine trajectory-to-trajectory fluctuations by constructing the successive summation of φ2, 3 on the basis of the Ck statistic. To find the switch- ing point k′, we fit two regression lines to the arrays C1, · · · , Ck and Ck+1, · · · , CM and calculate the squared sums of residuals for both lines. The switching point k′ is obtained by minimizing the total squared sum of resid- uals. Next, the data {φi} are divided into two arrays: φ1, · · · , φk′ and φk′ +1, · · · , φM . Then, given a desired confidence level γ, the quantiles (1 − γ)/2 and γ/2 of the squared data φ2 for the first array (i = 1, · · · , k′) are i calculated. Here, for the sake of simplicity, we assume the variance of the first array is smaller than the second one; otherwise the quantiles are computed for the second array. The core of the regime variance test is the number of observations B from the data φ2 j in the second array (j = k′ + 1, · · · , M ) that fall into the constructed quan- tiles interval. The null hypothesis of both regions having the same distribution implies that B has binomial dis- tribution P (B = k′) = (cid:0)M −k′ −B, where p = γ. Therefore the p-value of the test is equal to the cumulative distribution function of this distribution eval- uated at B. A large p-value of the test (greater than the confidence level γ) indicates the null hypothesis is not rejected. B (cid:1)pB(1 − p)M −k′ After confirming the existence of at least two types of trajectories with respect to sojourn times in the confined states, we determined the number of classes on the basis of the silhouette criterion [38]. For a fixed number of classes c the silhouette statistic assigns value sc(i) to the observation φi, given by sc(i) = b(φi) − a(φi) max{a(φi), b(φi)} , i = 1, 2, ..., M, (3) where a(φi) is the average distance to all values in the allocated class and b(φi) is the distance to the nearest neighbor class. For each number of classes c, all possible divisions into c classes are considered and the optimal division is the one that maximizes the silhouette statistic sc. The silhouette criterion then takes a value silc = M X i=1 sc(i)/M, (4) Ck = k X i=1 φ2 i , k = 1, 2, ..., M. (2) that varies from 0 to 1. The optimal number of classes maximizes silc. If the fractions φi correspond to a single type of trajecto- ries, then Ck is a linear function with respect to k, oth- erwise a piecewise linear behavior with different slopes indicates there are at least two different regimes. Note that the values φi do not need to be ordered and the Ck statistics is non-decreasing. The regime variance test was shown in Ref. [37] to be effective when dealing both with Gaussian and L´evy-stable random variables. The null hypothesis of the regime variance test corre- sponds to the case with a single regime. To test this hy- pothesis, first the most likely switching point k′ is found F. Classification of trajectories In order to classify trajectories according to their frac- tions of time being in the confined state φi, we used a clustering method based on the k-means algorithm [39] implemented in MATLAB. The k-means clustering par- titions the set of M observations into k clusters in a way that each observation belongs to the cluster with the nearest mean. The assignment of φi into a cluster is thus based on the minimization of the average Euclidean distance between the points in that cluster and the clus- ter mean. This method yields a partitioning of the data space into Voronoi cells. G. Statistics In the experimental data analysis we compared the distributions of different characteristics corresponding to confinement states for classified trajectories. In order to evaluate if two data sets have the same distribution we used the Kolmogorov-Smirnov (KS) test for two samples [40]. The KS statistic for two data sets with cumulative distribution functions F1(x) and F2(x) is KS = supxF1(x) − F2(x), (5) where supx is the supremum. A large p-value of the KS test indicates the H0 hypothesis is not rejected and the two data sets have the same distribution. III. VALIDATION OF CONFINEMENT IDENTIFICATION METHOD In this section we evaluate the effectiveness of the con- finement identification method. As the toy model we an- alyze intermittent fractional Brownian motion (FBM). FBM is a stochastic process driven by stationary Gaus- sian, but power-law correlated noise [41, 42]. It is one of the classical anomalous diffusion process for which the mean square displacement (MSD) hx2(t)i = Kαtα, with generalized diffusion coefficient Kα and anomalous expo- nent α. In terms of the commonly used Hurst exponent, H = α/2. The process is superdiffusive when α > 1 and subdiffusive when 0 < α < 1. As α decreases, the ran- dom walk becomes more compact. In particular, when α is close to zero, the FBM resembles confinement in a domain with a small drift. In order to illustrate the confinement recurrence anal- ysis, we present two short FBM trajectories (N = 50 points) with α = 0.1 and α = 0.9 in Fig. 1A and B, respectively. Given that a FBM with α = 0.1 is a very compact random walk, it resembles motion in a confined domain. FBM with α = 0.9 is a good model for uncon- fined subdiffusion. We expect the number of visits Vj to have larger values in the regions with α = 0.1. Thus, the number of visits provide a metric to segment the tra- jectory according to its recurrence. For both cases two consecutive observations are marked in Fig. 1A and B. The constructed circle in the trajectory with α = 0.1 encloses 17 points and the circle in the trajectory with α = 0.9 encloses only one point. We analyze an intermittent FBM where the anoma- lous diffusion exponent alternates between α = 0.1 and α = 0.9. For simplicity the random walk is defined as a renewal process where the process correlations are re- set when α changes. We simulate the intermittent FBM with five segments of different lengths. The first, third 4 and fifth segments correspond to α = 0.1 while the sec- ond and fourth to α = 0.9. Fig. 1C shows four simulated intermittent FBM realizations together with the results of the recurrence analysis method. The parts of the tra- jectories identified as confined motion are marked in red. The recurrence analysis takes under consideration two- dimensional trajectories but we present one-dimensional time series for clarity. The two dimensional trajectories are shown to the right of the time traces. The vertical dashed lines correspond to the switching points between the two FBMs. The time series of the number of visits Vj for these four simulated trajectories are presented in Fig. 1 D. Again, the vertical dashed lines correspond to the true switching points between the two regimes of in- termittent FBMs. As seen in Fig. 1D, the time series Vj remains for long times at low values that correspond to free diffusion and high Vj values corresponding to con- fined motion. Thus it is possible to discriminate between different phases of motion by employing a threshold on the Vj time series. The choice of the threshold value for segmentation of the trajectories depends on the charac- ter of the data but it can be effectively chosen by visual inspection. Here, we chose a threshold Vth = 6. However, as can be seen in Fig. 1D the method is prone to statisti- cal noise: in the free regions we observe falsely identified short periods of confinement and vice versa. Therefore, there is need to correct the method in order to overcome the falsely identified regions with short dwell times. As explained in the methods, this correction is introduced by eliminating all transitions where the dwell time is a single time window, that is three points. IV. RESULTS A. Detection of Kv1.4 transient confinement We have imaged hippocampal neurons expressing Kv1.4-CF640R and Nav1.6-CF640R and tracked their motion on the somatic surface. Figure 2A shows 92 Kv1.4-CF640R trajectories obtained in a typical cell. The trajectories are highly heterogeneous with some tra- jectories being very compact while others explore large regions. However, this heterogeneity does not appear to be related to the location of the molecules within the cell. Figure 2B shows a zoom on the trajectory indicated by an arrow. As we have previously reported [32], Kv1.4 ion channels exhibit intermittent behavior with periods of confinement and periods of free diffusion. We employ recurrence analysis based on the number Vj of visits to site j, to segment the trajectory according to being in either a confined or free state. Figure 2C shows an histogram of the number of visits Vj at each site as defined in our algorithm for identification of confinement. The trajectories are next segmented using a threshold Vth = 11. Figure 2D shows the time series Vj of the trajectory shown in panel B. We add the number of visits within three consecutive circles and thus our temporal A y 3 2 1 0 −1 −2 B 2 0 y −2 −4 −6 −1 0 x 1 2 −6 −4 −2 x 0 200 C y 150 100 50 0 −50 0 400 D 800 time 1200 1600 −50 50 0 x V 200 0 200 0 200 0 200 0 0 400 800 time 1200 1600 FIG. 1. Method for identification of confined motion based on number of visits to current site. The method is validated using numerical simulations. (A) FBM with α = 0.1, i.e., Hurst exponent H = 0.05. At each successive points pair such as those indicated by the arrows a circle is drawn as indicated in the figure. In this case there are 17 points (visits) found within the circular area. (B) FBM with α = 0.9. Here the tracer visits the region within the selected circle only once. (C) y- coordinate of two-dimensional intermittent FBM with α = 0.1 (first, third and fifth segment) and α = 0.9 (second and fourth segment). The vertical dashed lines show the true switching points. The parts of the trajectories identified as confined (low α) are marked in red. The two-dimensional trajectories are presented on the right. (D) Vj statistic, i.e., number of recurrent visits, as a function of time. Again, the vertical dashed lines show the true switching points. The threshold for finding confined regions in this example is Vth = 6. resolution is 150 ms in this analysis. The confined regions are found as the periods with Vj ≥ Vth and are colored in red in Figs. 2B and D. The x(t) and y(t) time series of the same trajectory are shown in Figs. 2E and F, also with the confined regions colored in red. A B 5 200 nm 2 µm 10 C 5 0 0 , ) 0 0 0 1 x ( s t n u o c D 80 60 V 40 200 400 600 800 V 20 Vth =11 0 0 1 2 confined free 4 5 3 t (s) E 3.8 3.6 3.4 3.2 ) m µ ( x F 13.8 13.4 13.0 12.6 ) m µ ( y 3 0 1 2 4 5 3 t (s) 12.2 0 1 2 4 5 3 t (s) FIG. 2. Segmentation of Kv1.4 trajectories. (A) 92 individual Kv1.4 trajectories in one hippocampal neuron. (B) Zoom of a trajectory where the periods of confinement are shown in red. (C) Histogram of the number of visits to current site, Vj (117,195 sites; 649 trajectories). As explained in the text, the sites are determined by the exploration in 150 ms. This metric is used to evaluate how compact the random walk is and to identify regions of confinement. (D) Time series of number of visits Vj for the trajectory in panel B. When the number of visits is above the threshold Vth = 11, the particle is considered to be in the confined state. At 4.6 s, Vj drops below the threshold during three points in the time series. However, because the dwell time is not longer than three points, i.e., a single time window, this region is considered to remain in the confined state. This segment is marked in blue and indicated by an arrow in panel B. (E-F) Time series of localization along x and y for the trajectory shown in panel B. The regions that are detected to be confined are shown in red and the identified switching point marked by dashed vertical line. B. Kv1.4 are classified according to their surface interactions We have observed that Kv1.4 channels exhibit peri- ods of transient confinement and the instantaneous state of the protein can be determined by recurrence analy- sis. Further, visual examination of the trajectories in Fig. 2A suggests the data are markedly heterogeneous. Thus we study whether there are more than one class of particles using the regime variance method accord- ing to the fraction of time φ that each particle spends in the confined state. From a physiological perspective such different types of molecules could be the result of post-translational modifications that alter molecular in- teractions. Figure 3A shows an histogram of the fractions of time spent in the confined state where the counts indi- cate number of trajectories. These fractions of time vary from φ = 0.02 up to φ = 1. Figure 3B shows the regime variance statistic Ck vs. trajectory number k (Eq. (2)). Using this metric with a confidence level γ = 0.05, we find that there are at least two distinct classes of trajectories in the Kv1.4 data (p = 10−3). Using the silhouette crite- rion, we find that there are two classes of trajectories as silc is maximized by c = 2 (sil2 = 0.9 and sil3 = 0.8). As a simple control of the regime variance test, we apply it to the simulated trajectory set of intermittent FBM that was presented in Fig. 1. The regime variance test for the simulated trajectories does not reject the hypothesis of a single class (p = 0.29, inset of Fig. 3B). Kv1.4 trajectories are classified according to their frac- tion of time in the confined regime φ. The k-means al- gorithm yields class division according to φ < 0.69. Fig- ures 3C and D show examples of trajectories in each of the classes, where the confined states are marked in red. To characterize the differences in the behavior of parti- cles belonging to each class we study the distributions of residence times within the confined state. Figure 3E shows the complementary cumulative distribution func- tion (CCDF) of the residence time, i.e., P [T > t] for par- ticles in each of the classes. The Kolmogorov-Smirnov two-sample test rejects the null hypothesis of the same distribution of the residence time for two classes with p = 10−108. For the trajectories with φ < 0.69 we find that the sojourn times have an exponential distribution tail with a characteristic decay time τ = 0.43 ± 0.05 s. C. Characterization of Kv1.4 confining domains The confining domains were found from the periods within the trajectories that particles exhibit confined mo- tion. However, only regions where the particle remains confined for at least 10 frames were analyzed. In the cases that the particle was confined for more than 20 frames, only the first 20 points are employed in the anal- ysis to avoid any potential problems related to drift of the confining domain. Radii of gyration were found along the major and minor principal axes and the domain was approximated as an ellipse with these major and minor semi-axes, respectively, as shown in the inset of Fig. 4A. The radius of gyration along the u direction is defined for a trajectory of N points as R2 k (uk − u)2, where uk is the distance of the kth point to the corresponding principal axis. Figures 4 A and B show the cumulative distribution function and box plots of the elliptical area of the confining domains for both classes of trajectories. Figure 4 C and D show the characterization of the same domains in terms of the major semi-axis. The majority u = PN 300 A s t n u o c 200 100 120 B k C 60 0 0 0.2 0.4 0.6 0.8 1 0 0 200 6 200 100 0 0 400 k 400 600 Class 1 C 0.5 µm Class 2 D 0.5 µm = 0.76 = 0.88 = 0.34 = 0.48 = 0.92 = 0.97 = 0.30 = 0.60 E ) t ( F - 1 1 0.1 0.01 10-3 0 Class 1 Class 2 2 4 6 8 10 t (s) i=1 φ2 FIG. 3. Determination of number of distinct classes among Kv1.4 trajectories (649 trajectories, 9 cells). (A) Histogram of fraction of the observation time that a particle spends in the confined state. (B) Regime variance test statistic Ck = Pk i . The regime variance test rejects the hypoth- esis of a single regime with p = 10−3. The inset shows the regime variance test applied to 500 realizations of numerical simulations of the type shown in Fig. 1. In these simulations, the test does not reject the hypothesis of a single class of tra- jectories (p = 0.29). (C-D) Examples of Kv1.4 trajectories in the two distinct classes. The fraction φ is given for each ex- ample. (E) Complementary cumulative distribution function of the residence times in confined states for each of the two classes (class 1: 1,659 sojourn times, 575 trajectories; class 2: 575 sojourn times, 74 trajectories). of domain sizes are much larger that the particle localiza- tion uncertainty. These data show that even though the interactions of distinct pools of Kv1.4 with these domains are different according to their sojourn times, the confin- ing domains for both classes share similar morphological characteristics. D. Characterization of Nav1.6 motion Nav1.6 ion channels were previously also found to exhibit periods of transient confinement [25]. Nav1.6 trajectories in a representative cell are shown in Fig. 5A. Regime variance test also shows that there exist at least two classes of Nav1.6 trajectories and the silhou- 0.2 B A 2 µm C1 C2 1 C F D C 0.5 0 0 1 E F D C 0.5 0 0 1 A F D C 0.5 0 0 1 C F D C 0.5 0 0 Class 1 Class 2 50 nm 0.04 0.08 0.12 Ellipsoidal area (µm2) Class 1 Class 2 100 200 300 Major semi-axis (nm) ) 2 m µ ( a e r a l i a d o s p i l l E ) m n ( s i x a - i m e s r o j a M 0.1 0 600 D 400 200 0 C1 C2 FIG. 4. Confining regions are the same for the particles in both classes. (A) Cumulative distribution function and (B) box plots for the confining regions area for particles in both classes. Box plots show 5-95% quantiles as whiskers together with quartiles and median (Class 1: 1,165 domains; Class 2: 113 domains). The inset in panel A shows the characteriza- tion of the confining domain in a sample trajectory using the radii of gyration along the principal axes. The first 20 points of the confining domain are employed to find the radii of gy- ration. The obtained confining ellipse is shown together with the principal axes. (C) Cumulative distribution function and (D) box plots for the confining major semi-axes. ette criterion indicates the number of classes equals two (sil2 = 0.89 and sil3 = 0.75). Application of a k-means algorithm yields a fraction of times threshold φ = 0.74 for classifying trajectories. The distributions of sojourn times in the confined state are shown Fig. 5B. Again, as seen for Kv1.4 channels, the distributions of times in the two states are markedly different. The sojourn times in the class with φ < 0.74 are exponentially distributed with characteristic decay time τ = 0.93 ± 0.04 s. The characterization of confining domains for Nav1.6 according to the radii of gyration along the principal axes is shown in Figs. 5 C-F. The Kolomogorov-Smirnov two- sample test rejects the null hypothesis of the same distri- bution of confinement sizes for two classes with p = 0.007 but the characteristics of both populations are similar. V. DISCUSSION AND CONCLUSIONS The dynamics of membrane proteins is often char- acterized by a high degree of heterogeneity. In gen- eral, these fluctuations can arise from two very differ- 7 Class 1 Class 2 4 6 t (s) 8 10 2 1 0.1 B ) t ( F - 1 0.01 10-3 0 Class 1 Class 2 ) 2 m µ ( a e r a l i a d o s p i l l E ) m n ( s i x a - i m e s r o j a M D 0.2 0.1 0 400 F 300 200 100 0 C1 C2 C1 C2 0.05 0.1 0.15 Ellipsoidal area (µm2) Class 1 Class 2 100 200 300 Major semi-axis (nm) FIG. 5. Nav1.6 channels in the soma of hippocampal neu- rons (386 trajectories, 4 cells). (A) Trajectories of Nav1.6 in one example cell (79 trajectories). (B) Residence times in confining regions for the two classes found using the silhou- ette method (Class 1: 915 times, 307 trajectories; class 2: 706 times, 79 trajectories). (C) Cumulative distribution function (CDF) and (D) box plots of the confining regions area for particles in both classes. Box plots show 5-95% quantiles as whiskers together with quartiles and median (Class 1: 605 domains; class 2: 171 domains). (E) CDF and (F) box plots of the confining major semi-axes. ent mechanisms. In the first situation, proteins per- form a random walk in a heterogeneous landscape while, in the second, proteins undergo post-translational mod- ifications so that they interact in substantially differ- ent ways with the same complexes. The first situa- tion has been studied both experimentally and theo- retically. Heterogeneous diffusion landscapes can yield intriguing results that involve population splitting and non-ergodicity [43, 44]. Besides analysis of individual trajectories, the diffusion landscape of membrane pro- teins has been studied using single-particle tracking pho- toactivated localization microscopy (sptPALM) [45] and universal points-accumulation-for-imaging-in-nanoscale- topography (uPAINT) [46], which yield high-density sur- face maps. Furthermore, these high-density maps can be accurately evaluated using Bayesian inference tools, which provide information on both diffusion and energy landscapes [47]. We have previously employed sptPALM in combina- tion with Bayesian inference tools to show that Nav1.6 channels are clustered into nanoscale domains [25]. How- ever, one of the interesting aspects of those observations lies in the fact that Nav channels exhibit a marked hetero- geneity in their interaction with the nanoclusters. There- fore we set to study the molecule-to-molecule heterogene- ity in the neuronal surface. We raise the question, can we unravel distinct molecule subpopulations according to interactions with nanoclusters? To this end we develop a methodology by which we segment the trajectories into regions of transient confinement and regions of free mo- tion and we identify two distinctive subpopulations both in the Nav1.6 and in the Kv1.4 dynamics. These subpop- ulations exhibit different types of interactions with their respective membrane nanodomains. In one population the trajectories are mostly in the free state while in the second population, the trajectories exhibit long periods under transient confinement. The size of the confined re- gions are characterized from the motion of the molecules and it is found these regions have a mean diameter that is ten times the localization accuracy. Therefore the trap- ping events cannot be considered to be immobilization due to binding as is the case for Kv2.1 channels in HEK cells [27, 48]. The interactions of one of the populations with the con- fining nanodomains exhibit a 'normal' type of statistics with sojourn times that are exponentially distributed. Therefore, the system can be considered to be Marko- vian, i.e., to have no memory. Surprisingly, the sec- ond population exhibits a heavy-tail, non-exponential sojourn-time distribution. This behavior brings up the hypothesis of complex behavior with the possibility of ergodicity breaking and aging in the dynamics of the 8 ion channels. Consistent with these observations, we have recently found that the dynamics of the majority of Kv1.4 and Nav1.6 trajectories in the somatic plasma membrane exhibit non-ergodic dynamics according to dy- namical functional tests [32]. The tools developed in this work can be employed in the study of membrane nanodomains, which are widespread among mammalian cells. Further these do- mains can play important physiological roles. In B and T lymphocytes, reorganization of signaling nanodomains leads to cell activation [49]. In neurons, nanoclustering of membrane proteins has key functions in synaptic trans- mission [50]. We apply four time-series analysis tools to extract specific information on heterogeneous inter- actions with nanodomains: (i) A recurrence analysis is used to find transitions in the diffusive behavior. This analysis has the advantages of providing high temporal resolution that can be applied to both Markovian and non-Markovian processes. (ii) A regime variance test quantifies the heterogeneity in the sojourn times. (iii) A silhouette algorithm finds the number of different classes according to protein dynamics. (iv) A k-means algorithm is used to set thresholds and separate trajectories into dif- ferent classes. By using the algorithms provided in the Supplemental Materials [36] we identified and character- ized distinct behaviors of the same proteins expressed in the neuronal soma, which had not been distinguished with previous analyses. ACKNOWLEDGMENTS This work was supported by the National Science Foundation under Grant 1401432 (to DK), NCN OPUS Grant No. UMO-2016/21/B/ST1/00929 (to AW), NCN Maestro Grant No. 2012/06/A/ST1/00258 (to JG), and the National Institutes of Health grant RO1NS085142 (to MMT). [1] Elizabeth Pennisi, "ENCODE project writes eulogy for junk DNA," Science 337, 1159–1161 (2012). [2] Iakes Ezkurdia, David Juan, Jose Manuel Rodriguez, Adam Frankish, Mark Diekhans, Jennifer Harrow, Jesus Vazquez, Alfonso Valencia, and Michael L Tress, "Mul- tiple evidence strands suggest that there may be as few as 19 000 human protein-coding genes," Hum. Mol. Gen. 23, 5866–5878 (2014). [3] Jennifer Harrow, Adam Frankish, Jose M Gonzalez, Elec- tra Tapanari, Mark Diekhans, Felix Kokocinski, Bron- wen L Aken, Daniel Barrell, Amonida Zadissa, Stephen Searle, et al., "GENCODE: the reference human genome annotation for The ENCODE Project," Genome Res. 22, 1760–1774 (2012). [4] Linn Fagerberg, Kalle Jonasson, Gunnar von Heijne, Mathias Uhl´en, and Lisa Berglund, "Prediction of the human membrane proteome," Proteomics 10, 1141–1149 (2010). [5] Qifeng Xue and Edward S Yeung, "Differences in the chemical reactivity of individual molecules of an en- zyme," Nature 373, 681–683 (1995). [6] Robert Zwanzig, "Dynamical disorder: Passage through a fluctuating bottleneck," J. Chem. Phys. 97, 3587–3589 (1992). [7] H Peter Lu, Luying Xun, and X Sunney Xie, "Single- molecule enzymatic dynamics," Science 282, 1877–1882 (1998). [8] Sergey V Solomatin, Max Greenfeld, Steven Chu, and Daniel Herschlag, "Multiple native states reveal persis- tent ruggedness of an RNA folding landscape," Nature 463, 681–684 (2010). [9] Changbong Hyeon, Jinwoo Lee, Jeseong Yoon, Sungchul Hohng, and D Thirumalai, "Hidden complexity in the isomerization dynamics of Holliday junctions," Nat. Chem. 4, 907–914 (2012). [10] Bian Liu, Ronald J Baskin, and Stephen C Kowal- czykowski, "DNA unwinding heterogeneity by RecBCD results from static molecules able to equilibrate," Nature 500, 482–485 (2013). [11] Taekjip Ha, "Single-molecule fluorescence resonance en- ergy transfer," Methods 25, 78–86 (2001). [12] Jeffrey R Moffitt, Yann R Chemla, Steven B Smith, and Carlos Bustamante, "Recent advances in optical tweez- ers," Annu. Rev. Biochem. 77, 205–228 (2008). [13] Michael Hinczewski, Changbong Hyeon, and D. Thiru- malai, "Directly measuring single-molecule heterogeneity using force spectroscopy," Proc. Natl. Acad. Sci. U.S.A. 113, E3852–E3861 (2016). [14] Philip D Fox, Christopher J Haberkorn, Elizabeth J Akin, Peter J Seel, Diego Krapf, and Michael M Tamkun, "Induction of stable ER–plasma-membrane junctions by Kv2.1 potassium channels," J. Cell Sci. 128, 2096–2105 (2015). [15] Emily Deutsch, Aubrey V Weigel, Elizabeth J Akin, Phil Fox, Gentry Hansen, Christopher J Haberkorn, Rob Lof- tus, Diego Krapf, and Michael M Tamkun, "Kv2.1 cell surface clusters are insertion platforms for ion channel delivery to the plasma membrane," Mol. Biol. Cell 23, 2917–2929 (2012). [16] Ralf Metzler, Jae-Hyung Jeon, Andrey G Cherstvy, and Eli Barkai, "Anomalous diffusion models and their prop- erties: non-stationarity, non-ergodicity, and ageing at the centenary of single particle tracking," Phys. Chem. Chem. Phys. 16, 24128–24164 (2014). [17] Diego Krapf, "Mechanisms underlying anomalous diffu- sion in the plasma membrane," Curr. Top. Membr. 75, 167–207 (2015). [18] Carlo Manzo and Maria F Garcia-Parajo, "A review of progress in single particle tracking: from methods to bio- physical insights," Rep. Prog. Phys. 78, 124601 (2015). [19] Dino Ott, Poul M Bendix, and Lene B Oddershede, "Re- vealing hidden dynamics within living soft matter," ACS Nano 7, 8333–8339 (2013). [20] Ken Ritchie, Ryota Iino, Takahiro Fujiwara, Kotono Murase, and Akihiro Kusumi, "The fence and picket structure of the plasma membrane of live cells as revealed by single molecule techniques (Review)," Mol. Membr. Biol. 20, 13–18 (2003). [21] Nicholas L Andrews, Keith A Lidke, Janet R Pfeiffer, Alan R Burns, Bridget S Wilson, Janet M Oliver, and Diane S Lidke, "Actin restricts FcεRI diffusion and fa- cilitates antigen-induced receptor immobilization," Nat. Cell Biol. 10, 955–963 (2008). [22] Sanaz Sadegh, Jenny L Higgins, Patrick C Mannion, Michael M Tamkun, and Diego Krapf, "Plasma mem- brane is compartmentalized by a self-similar cortical actin meshwork," Phys. Rev. X 7, 011031 (2017). [23] Maxime Dahan, Sabine Levi, Camilla Luccardini, Philippe Rostaing, Beatrice Riveau, and Antoine Triller, "Diffusion dynamics of glycine receptors revealed by single-quantum dot tracking," Science 302, 442–445 (2003). [24] Maria F Garcia-Parajo, Alessandra Cambi, Juan A Torreno-Pina, Nancy Thompson, and Ken Jacobson, "Nanoclustering as a dominant feature of plasma mem- brane organization," J. Cell Sci. 127, 4995–5005 (2014). [25] Elizabeth J. Akin, Laura Sole, Ben Johnson, Mohamed el Beheity, Jean-Baptiste Masson, Diego Krapf, and Michael M. Tamkun, "Single-molecule imaging of Nav1.6 on the surface of hippocampal neurons reveals somatic 9 nanoclusters," Biophys. J. 111, 1235–1247 (2016). [26] Daniel Choquet and Antoine Triller, "The dynamic synapse," Neuron 80, 691–703 (2013). [27] Aubrey V Weigel, Michael M Tamkun, and Diego Krapf, "Quantifying the dynamic interactions between a clathrin-coated pit and cargo molecules," Proc. Natl. Acad. Sci. U.S.A. 110, E4591–E4600 (2013). [28] Peter K Koo and Simon GJ Mochrie, "Systems-level ap- proach to uncovering diffusive states and their transi- tions from single-particle trajectories," Phys. Rev. E 94, 052412 (2016). [29] Yann Lanoisel´ee and Denis S Grebenkov, "Unraveling in- termittent features in single-particle trajectories by a lo- cal convex hull method," Phys. Rev. E 96, 022144 (2017). [30] Thorsten Wagner, Alexandra Kroll, Chandrashekara R Haramagatti, Hans-Gerd Lipinski, and Martin Wie- mann, "Classification and segmentation of nanoparti- cle diffusion trajectories in cellular micro environments," PloS one 12, e0170165 (2017). [31] Fredrik Persson, Martin Lind´en, Cecilia Unoson, and Johan Elf, "Extracting intracellular diffusive states and transition rates from single-molecule tracking data," Nat. Methods 10, 265 (2013). [32] Aleksander Weron, Krzysztof Burnecki, Elizabeth J. Akin, Laura Sol´e, Micha l Balcerek, Michael M. Tamkun, and Diego Krapf, "Ergodicity breaking on the neuronal surface emerges from random switching between diffusive states," Sci. Rep. 7, 5404 (2017). [33] Eugene A Katrukha, Marina Mikhaylova, Hugo X van Brakel, Paul M van Bergen en Henegouwen, Anna Akhmanova, Casper C Hoogenraad, and Lukas C Kapitein, "Probing cytoskeletal modulation of pas- sive and active intracellular dynamics using nanobody- functionalized quantum dots," Nat. Commun. 8, 14772 (2017). [34] Elizabeth J Akin, Laura Sol´e, Sulayman D Dib-Hajj, Stephen G Waxman, and Michael M Tamkun, "Pref- erential targeting of Nav1.6 voltage-gated Na+ channels to the axon initial segment during development," PloS one 10, e0124397 (2015). [35] Khuloud Jaqaman, Dinah Loerke, Marcel Mettlen, Hiro- taka Kuwata, Sergio Grinstein, Sandra L Schmid, and Gaudenz Danuser, "Robust single-particle tracking in live-cell time-lapse sequences," Nat. Methods 5, 695–702 (2008). [36] See Supplemental Material at [URL will be inserted by publisher] for MATLAB codes. [37] Janusz Gajda, Grzegorz Sikora, and Agnieszka Wy loma´nska, "Regime variance testing–a quantile ap- proach," Acta. Phys. Pol. B 44, 1015–1035 (2013). [38] Peter J Rousseeuw, "Silhouettes: a graphical aid to the interpretation and validation of cluster analysis," J. Com- put. Appl. Math. 20, 53–65 (1987). [39] Anil K Jain, "Data clustering: 50 years beyond K- means," Pattern Recogn. Lett. 31, 651–666 (2010). [40] Ralph B DAgostino and Michael A Stephens, Goodness of Fit Techniques, Statistics: a Series of Textbooks and Monographs, Vol. 68 (Marcel Dekker, New York NY, 1986). [41] Jan Beran, Statistics for long-memory processes, Vol. 61 (Chapman & Hall/CRC Monographs on Statistics & Ap- plied Probability, Boca Raton FL, 1994). [42] Benoit B Mandelbrot and John W Van Ness, "Fractional Brownian motions, fractional noises and applications," 10 SIAM Rev. 10, 422–437 (1968). Biophys. J. 99, 1303–1310 (2010). [43] Andrey G Cherstvy and Ralf Metzler, "Population split- ting, trapping, and non-ergodicity in heterogeneous dif- fusion processes," Phys. Chem. Chem. Phys. 15, 20220– 20235 (2013). [44] Carlo Manzo, Juan A Torreno-Pina, Pietro Massignan, Gerald J Lapeyre Jr, Maciej Lewenstein, and Maria F Garc´ıa-Parajo, "Weak ergodicity breaking of receptor motion in living cells stemming from random diffusivity," Phys. Rev. X 5, 011021 (2015). [45] Suliana Manley, Jennifer M Gillette, George H Pat- terson, Hari Shroff, Harald F Hess, Eric Betzig, and Jennifer Lippincott-Schwartz, "High-density mapping of single-molecule trajectories with photoactivated localiza- tion microscopy," Nat. Methods 5, 155 (2008). [46] Gregory Giannone, Eric Hosy, Florian Levet, Au- drey Constals, Katrin Schulze, Alexander I Sobolevsky, Michael P Rosconi, Eric Gouaux, Robert Tamp´e, Daniel Choquet, et al., "Dynamic superresolution imaging of en- dogenous proteins on living cells at ultra-high density," [47] Jean-Baptiste Masson, Patrice Dionne, Charlotte Sal- vatico, Marianne Renner, Christian G Specht, Antoine Triller, and Maxime Dahan, "Mapping the energy and diffusion landscapes of membrane proteins at the cell surface using high-density single-molecule imaging and Bayesian inference: application to the multiscale dynam- ics of glycine receptors in the neuronal membrane," Bio- phys. J. 106, 74–83 (2014). [48] Aubrey V Weigel, Blair Simon, Michael M Tamkun, and Diego Krapf, "Ergodic and nonergodic processes coexist in the plasma membrane as observed by single-molecule tracking," Proc. Natl. Acad. Sci. U.S.A. 108, 6438–6443 (2011). [49] Paola Pizzo and Antonella Viola, "Lipid rafts in lympho- cyte activation," Microbes Infect. 6, 686–692 (2004). [50] Adish Dani, Bo Huang, Joseph Bergan, Catherine Dulac, and Xiaowei Zhuang, "Superresolution imaging of chem- ical synapses in the brain," Neuron 68, 843–856 (2010).
1107.1442
3
1107
2012-03-14T18:08:48
Footprint traversal by ATP-dependent chromatin remodeler motor
[ "physics.bio-ph", "q-bio.SC" ]
ATP-dependent chromatin remodeling enzymes (CRE) are bio-molecular motors in eukaryotic cells. These are driven by a chemical fuel, namely, adenosine triphosphate (ATP). CREs actively participate in many cellular processes that require accessibility of specific segments of DNA which are packaged as chromatin. The basic unit of chromatin is a nucleosome where 146 bp $\sim$ 50 nm of a double stranded DNA (dsDNA) is wrapped around a spool formed by histone proteins. The helical path of histone-DNA contact on a nucleosome is also called "footprint". We investigate the mechanism of footprint traversal by a CRE that translocates along the dsDNA. Our two-state model of a CRE captures effectively two distinct chemical (or conformational) states in the mechano-chemical cycle of each ATP-dependent CRE. We calculate the mean time of traversal. Our predictions on the ATP-dependence of the mean traversal time can be tested by carrying out {\it in-vitro} experiments on mono-nucleosomes.
physics.bio-ph
physics
Footprint traversal by ATP-dependent chromatin remodeler motor Ashok Garai,1 Jesrael Mani,1 and Debashish Chowdhury∗1, 2 1Department of Physics, Indian Institute of Technology, Kanpur 208016, India 2Max-Planck Institute for Physics of Complex Systems, 01187 Dresden, Germany (Dated: October 10, 2018) ATP-dependent chromatin remodeling enzymes (CRE) are bio-molecular motors in eukaryotic cells. These are driven by a chemical fuel, namely, adenosine triphosphate (ATP). CREs actively participate in many cellular processes that require accessibility of specific segments of DNA which are packaged as chromatin. The basic unit of chromatin is a nucleosome where 146 bp ∼ 50 nm of a double stranded DNA (dsDNA) is wrapped around a spool formed by histone proteins. The helical path of histone-DNA contact on a nucleosome is also called "footprint". We investigate the mecha- nism of footprint traversal by a CRE that translocates along the dsDNA. Our two-state model of a CRE captures effectively two distinct chemical (or conformational) states in the mechano-chemical cycle of each ATP-dependent CRE. We calculate the mean time of traversal. Our predictions on the ATP-dependence of the mean traversal time can be tested by carrying out in-vitro experiments on mono-nucleosomes. PACS numbers: 87.16.ad,87.16.Sr,87.16.dj I. INTRODUCTION A deoxyribonucleic acid (DNA) molecule is a linear heteropolymer whose monomeric subunits, called nu- cleotides, are denoted by the four letters A, T, C and G. The sequence of these nucleotides in a DNA molecule chemically encodes genetic informations. In the nucleus of an eukaryotic cell, DNA is stored in a hierarchically organized structure called chromatin [1 -- 4]. The primary repeating unit of chromatin at the lowest level of the hi- erarchical structure is a nucleosome [5]. The cylindrically shaped core of each nucleosome consists of an octamer of histone proteins around which 146 base pairs (i.e.,∼ 50 nm) of the the double stranded DNA is wrapped about two turns (more precisely, 1.7 helical turns); the arrange- ment is reminiscent of wrapping of a thread around a spool. There are 14 equispaced sites, at intervals of 10 base pairs (bp), on the surface of the cylindrical spool. Electrostatic attraction between these binding sites on the histone spool and the oppositely charged DNA seems to dominate the histone-DNA interactions which stabi- lize the nucleosomes. Throughout this paper, the helical curve formed by the histone-DNA overlap will be called the "footprint". The DNA stores the genetic blueprint of an organism. If nucleosomes were static, segments of DNA buried in nucleosomes would not be accessible for various functions involving the corresponding gene [6 -- 8]. However, in real- ity, nucleosomes are dynamic. Spontaneous dynamics of nucleosomes are usually consequences of thermal fluctua- tions whereas the active dynamic processes are driven by special purpose molecular machines called chromatin re- modeling enzymes (CRE) fuelled by ATP [9 -- 14]. Various aspects of chromatin dynamics has received some atten- ∗Corresponding author(E-mail: [email protected]) tion of theoretical modelers, including physicists, over the last few years [15 -- 27]. In general, "Chromatin remodelling" refers to a range of enzyme mediated structural transitions that occur dur- ing gene regulation in eukaryotic cells. To make DNA, which is wrapped around histone octamer, accessible for various DNA-dependent processes, it is always necessary to rearrange or mobilize the nucleosomes. In principle, there are at least four different ways in which a CRE can affect the nucleosomes [28]: (i) sliding the histone oc- tamer, i.e., repositioning of the entire histone spool, on the dsDNA; (ii) exchange of one or more of the histone subunits of the spool with those in the surrounding solu- tion (also called replacement of histones) (iii) removal of one or more of the histone subunits of the spool, leaving the remaining subunits intact, and (iv) complete ejection of the whole histone octamer without replacement. Our theoretical work here is closely related to sliding. In the next section we describe a scenario in which either a CRE motor (or, other ATP-dependent motors that translocate along dsDNA) traverse the "footprint". Because of the stochasticity of the underlying mechano- chemical kinetics, the footprint traversal time (FTT) is a fluctuating random variable. Extending an earlier model developed by Chou [29], we analytically calculate the mean FTT (MFTT) of the CRE motor. To our surprise, we found that the ATP-dependence of the various ATP-driven activities of CREs have not been studied systematically in the published literature. In par- ticular, we address the question: how does the MFTT of a CRE motor vary with the variation of the concentra- tion of ATP? This rate is not necessarily directly pro- portional to the rate of ATP hydrolysis by the CRE be- cause the mechanical sliding of the nucleosome need not be tightly coupled with the hydrolysis of ATP by the CRE. Therefore, we develope here an analytical theory predicting the ATP-dependence of the ATP-dependent footprint-traversal by CRE. We hope our result will stim- ulate systematic experimental investigations on the ATP- dependence of ATP-dependent CREs. II. CRE: PHENOMENOLOGY AND MOTIVATION FOR THIS WORK Chromatin is not a frozen static aggregate of DNA and proteins. Spontaneous thermal fluctuations can cause a transient unwrapping and rewrapping of the nucleosomal DNA from one end of the nucleosome spool; the cor- responding rates for an isolated single nucleosome are, typically, 4s−1 and 20 − 90s−1, respectively [30]. In other words, once wrapped fully, the nucleosomal DNA remains in that state for about 250ms before unwrapping again spontaneously; however, it waits in the unwrapped state only for about 10 − 50ms before rewrapping again spon- taneously. Surprisingly, the accessibility of the nucleoso- mal DNA is only modestly affected if instead of a single nucleosome the experiment is repeated with an array of homogeneously distributed nucleosomes [31, 32]. More- over, folding of an array of nucleosomes makes the linker DNA about 50 times less accessible [31, 32]. Further- more, on the nucleosomal DNA, the farther is a site from the entry and exit points, the longer one has to wait to access it by the rarer spontaneous fluctuation of suffi- ciently large size [33]. Thus, nucleosomal DNA far from both the entry and exit sites is practically inaccessible by spontaneous thermal fluctuations. Can a nucleosome slide spontaneously by thermal fluc- tuations thereby exposing the nucleosomal DNA? In- terestingly, spontaneous repositioning of nucleosome on DNA strands is a well known phenomenon [34]. How can one reconcile accessibility of nucleosomal DNA by such repositioning [34] with the difficulty of access by unwrap- ping from either end of the spool [33]? If the DNA were to move unidirectionally along its own superhelical contour on the surface of the histone, at every step it would have to first transiently detach simultaneously from all the 14 binding sites and then reattach at the same sites after its contour gets shifted by 10 bp (or multiples of 10 bp). But, the energy cost of the simultaneous detachment of the DNA from all the 14 binding sites is prohibitively large because the total energy of binding at the 14 sites is about 75kBT [17, 18, 21]. But, why can't the cylindrical spool simply roll on the wrapped nucleosomal DNA thereby repositioning itself? If the nucleosome rolls by detaching DNA from one end of the spool, cannot it compensate this loss of binding energy by simulataneous attachment with a binding at the other end? If such an energy compensation were possible, detachment from only one binding site would be required at a time. But, the cylindrical spool has a finite size on which only a finite number (14) of binding sites for DNA are accomodated. Therefore, by rolling over the DNA the spool would not offer any vacant binding site to the DNA with which it can bind. This rolling mechanism would successfully lead to spontaneous sliding 2 of the nucleosome only if the histone spool were infinite with an infinite sequence of binding sites for DNA on its surface [21]. We now describe a plausible mechanism for sponta- neous sliding of a nucleosome [16 -- 18]. In the process of normal "breathing", most often the spontaneously un- wrapped flap rewraps exactly to its original position on the histone surface. However, if the rewrapping of a un- wrapped flap takes place at a slightly displaced location on the histone spool a small bulge (or loop) of DNA forms on the surface of the histones. Since the successive bind- ing sites are separated by 10bp, the length of the loop is quantized in the multiples of 10bp [16]. Such a spon- taneously created DNA loop, can diffuse in an unbiased manner on the surface of the histone spool. In the begin- ning of each step DNA from one end of the loop detaches from the histone spool, but the consequent energy loss is made up by the attachment of DNA at the other end of the loop to the histone spool before the step is completed. Consequently, by this diffusive dynamics, the DNA loop can traverse the entire length of the 14 binding sites on the histone spool of a nucleosome which will manifest as sliding of the nucleosome by a length that is exactly equal to the length of DNA in the loop. The diffusing DNA bulge can be formed by a "twist", rather than bending, of DNA [35 -- 37]. Spontaneous sliding of a nucleosome, however, is too slow to support intranuclear processes which need access to nucleosomal DNA. It is now widely agreed that ATP-dependent active re- modeling of nucleosome can account for the fast sliding of nucleosomes. Nevertheless, bulging DNA loop is ex- pected to play a key role in the remodeling process [27]. Our model describes how a CRE motor can wedge itself at the fork between the histone spool and a transiently detached segment of dsDNA and, by exploiting the spon- taneously diffusing loop by an ATP-dependent ratchet- ing, traverse the footprint in a directed manner. Because of the intrinsic stochasticity of the mechano-chemistry of the CRE and that of the diffusive motion of the DNA loop, the overall motion of the CRE is noisy and the time it takes to traverse the footprint is random. The main question we address in this paper is the fol- lowing: if < T > is the MFTT, what is the dependence of < T > on the concentration of ATP in the surrounding aqueous medium? To our knowledge, in the published literature neither systematic experiments nor any ana- lytical theory has addressed this question. In this paper, by extending Chou's model [29] of CRE, we capture the role of ATP explicitly and derive an analytical expres- sion for the dependence of < T > on the concentration of ATP. A CRE may be regarded as a molecular motor where input energy is derived from ATP hydrolysis and the output is mechanical work. The directed movement of the CRE may be caused either by a power stroke or by a Brownian ratchet mechanism [2]. The kinetics of the ATP-dependent CRE motor is formulated in our model in terms of a set of master equations. The kinetic scheme can be interpreted in terms of both power stroke and Brownian ratchet mechanisms. One interesting question [28] in the context of CRE is whether the CRE translocates along the DNA by mov- ing around the nucleosome, or whether the CRE anchors on the histone octamer and "pumps" DNA by pulling around the octamer. From the perspective of physicists, these two alternative scenarios can be viewed as merely a change of frame of reference- one is fixed with respect to the DNA whereas the other is fixed with respect to the CRE. Therefore, we describe the operation of the CRE with respect to a reference frame with respect to which the CRE translocates along the DNA; but, the model can be reformulated by a coordinate transformation so as to capture the alternative scenario where the CRE pumps the DNA. III. THE MODEL FIG. 1: A schematic representation of the isolated nucleo- some. We model a mono nucleosome where a dsDNA is wrapped one-and-three-fourth turn around a disc-shaped spool made of histone proteins (see Fig.1(a)). Follow- ing Chou [29], we consider the scenario where the CRE "wedges itself underneath the histone". The sites of histone-DNA contact along the DNA chain is represented as a one-dimensional lattice. Therefore, the lattice constant is, typically, 10 bp ((see Fig.1(b)). The total number n of lattice sites is equal to the total number of histone-DNA contact in a single nucleosome. A. Flap, loop and diffusive sliding of histone spool 3 n* α β n k k n−1 ....... 2 k k 1 β α bk 0* 0 ku FIG. 2: A schematic representation of the position of the ther- mally generated flap and it's diffusion along wrapped DNA. momentarily unwrap from the histone spool at a rate ku. For energetic reasons, the most likely length of such a segment would be one lattice spacing, i.e., about 10bp. Following Chou [29], we call such unwrapped segments a "flap". The rate of the reverse transition, in which re- binding of the DNA flap with the histone, takes place at a rate kb. A flap need not re-make the original histone-DNA con- tact. Instead, by pulling in an extra segment of the DNA, its next segment can bind with the last binding site on the histone spool, with rate α thereby forming what Chou [29] referred to as a "loop'. While located at either end of the lattice, a loop can revert to a flap at a rate β. The rates α and β are well approximated by [29] α ∼ kbe−Ebend/(kB T ), and β ∼ ku (1) where Ebend is the energy cost of bending the DNA into the shape of the loop. A loop can step forward or backward. In the absence of any CRE, the rates of the forward and backward step- pings of the loop are equal (denoted by k), provided the size of the loop Lloop remains unaltered (see Fig.2); in each forward step it unwraps one segment of DNA from the histone in the direction of its hop and re-wraps an- other equally long segment behind it. Therefore, one can approximate k by [29] k ∼ ku(cid:18) kb (kb + ku)(cid:19) (2) When a loop, after entering the lattice from one end, makes an eventual exit from the other end, it completes the "sliding" of the histone spool by a distance Lloop along the DNA in the opposite direction. Therefore, from the perspective of the sliding histone spool, its effective rate of hopping by a step of size Lloop along the dsDNA strand is the same as the rate pn at which a DNA loop of length Lloop traverses the lattice of n sites from one end to the other. Suppose Pj(t) denotes the probability that the loop is located at j (0 ≤ j ≤ 1). Following Chou's arguments, based on master equations for Pj(t), one gets [29] In this subsection we present a summary of Chou's ideas [29] which we need in the next subsection where we extend Chou's model. Here we consider the simple situation when no CRE is present and the kinetics of the system is governed solely by spontaneous thermal fluctu- ations. Because of these fluctuations, from either end of the histone-DNA contact region, small segments of DNA pn = αkku (n − 1)βkb + k(α + 2kb) (3) In the absence of a CRE, the traversal of a DNA loop of length Lloop from left to right is as likely as that from right to left. Therefore, the histone spool can slide for- ward or backward, with equal rate pn, by a step of size Lloop. As we'll see in the next subsection, peeling off of the DNA from the histone spool by a CRE motor keeps decreasing the effective value of n which, in turn, in- creases the effective sliding rate pn. B. Kinetics of CRE-driven directed sliding of histone spool Next, we consider the effect of DNA loop diffusion on the ATP-dependent translocation kinetics of a CRE. The model and results presented in this subsection are exten- sions of Chou's work [29] by incorporating explicitly a Brownian ratchet mechanism for CRE motors. As in ref.[29], we assume that the step size of the CRE motor is identical to the length of the thermally generated DNA loop. Therefore, the mechanical movements of the CRE motor can be described as that of a "particle" on the one-dimensional lattice on which the equi-spaced sites denote the histone-DNA contact points. We denote the position of the CRE motor on this lattice by the integers m. We now extend Chou's model [29] by exploiting a su- perficial similarity with the Garai-Chowdhury-Betterton (GCB) model [38] for the Brownian ratchet mechanism of monomeric helicase motors. A DNA helicase unwinds a dsDNA and translocates along one of two strands. At any arbitrary instant of time, the configuration of the system looks very similar to that shown in Fig.1(b) except that the surface of the DNA spool and the dsDNA would be replaced by the two strands of the dsDNA itself. The lattice constant is 1bp in the case of a helicase whereas it is about 10 bp in Fig.1(b). In the Brownian ratchet mechanism, momen- tary local unwinding of a segment, typically, 1 bp long, takes place at the fork by spontaneous thermal fluctu- ation; the opportunistic advance of the helicase merely prevents closure of the segment. Similarly, in the Brown- ian ratchet mechanism of the CRE, the CRE is assumed to "wedge" itself just in front of the DNA-histone fork. The CRE motor can move forward only if the segment in front of it is unwrapped by thermal fluctuation. FIG. 3: A schematic representation of the transition of the CRE between its two states. 4 2 1 m−1 ω 21 2 1 m ω 12 2 f ω 12 1 m+1 FIG. 4: A schematic representation of the position of the motor with two state of the model. ω 21 ω 12 ω f 12 Flat Potential Asymmetric Periodic Potential FIG. 5: A schematic representation of the Brownian ratchet mechanism. The mechano-chemical cycle of the CRE is captured in our model exactly the same way in which that of the helicase was formulated in the GCB model [38]. We as- sume that sequence of states in each mechano-chemical cycle of a CRE can be combined into two distinct groups which we label by the integers 1 and 2 (see Fig.3). The allowed transitions and the corresponding rate constants are shown in Fig.4. The physical processes captured by these rate con- stants can be motivated by a comparison with the ab- stract Brownian ratchet mechanism, illustrated in Fig.5. ATP hydrolysis by the CRE drives its transition from the state 1 to the state 2 at a rate ω21. Let us assume that the motor experiences two different types of potentials in the states 1 and 2. Let us further assume that initially the pe- riodic potential, with asymmetric sawtooth-like period, is kept on for sometime and during this time the motor set- tles at a position that coincides with one of the minima of this potential. Now if this potential is switched off then the probability distribution of the position of the motor will spread as a symmetric Gaussian. After sometime this Gaussian profile is broad enough to overlap with the next well (shadded region in the Fig.5), in addition to the original well. Now if the sawtooth potential is again switched on then, with a non-zero probability (that is proportional to the area of the shaded region) the motor will find itself in the next well. Our model accounts for this possibility with the transition associated with the rate constant ωf 12. There is also a finite probability that the particle stays back in its original well; this is captured by the transition with the rate constant ω12. The CRE motor would step forward at the rate ωf 12 if the next site in front is cleared. But if the next site is not cleared and it has to wait for the unwrapping of the DNA segment by thermal fluctuation. Consequently, its effective hoping rate 12 = ku(cid:18) ωf ωf 12 12 + kb)(cid:19) (ωf (4) is reduced from the free hopping rate ωf depends on both ku and kb. 12 by a factor that When a diffusing loop reaches in front of the motor it momentarily creates a flap of two bond segments. Three different transitions are now possible (see Fig.6): (i) the motor's position remains unaltered while the two open segments close, (ii) the motor moves forward by one step while one segment of the flap closes; (iii) the motor moves forward by two steps and the flap cannot close. The rate for the process (i) is kb/2 irrespective of the "chemical" state of the motor. However, the rates of the processes (ii) and (iii) depend on whether the motor was in the "chemical" state 1 or 2. If the motor is in the state 2, the rate of the process (ii) is given by ((ωf 12)−1 + k−1 b )−1 and that of the process (iii) is given by ((ωf 12)−1 + (ω21)−1 + (ωf 12)−1). Therefore, 5 (a) (b) f0 = kb 2λf , f1 = ωf 12kb 12 + kb)λf (ωf , f2 = with the normalization constant ωf 12ω21 ,(5) λf (2ω21 + ωf 12) λf = kb 2 + ωf 12kb (ωf 12 + kb) + ωf 12ω21 (2ω21 + ωf 12) (6) where the symbols f0, f1 and f2 are the probabilities of the processes (i), (ii) and (iii) above when the motor is in the "chemical" state 2. Similarly, g0 = g2 = kb 2λh , g1 = kbω21ωf 12 (ω21ωf 12 + kbωf 12 + kbω21)λh , ω21ωf 12 12 + ω21)λh 2(ωf , (7) FIG. 6: Schematic representations of the possible transitions when the motor is in front of the open flap. In (a) and (b) the motor is in the chemical state 1 and 2, respectively. with the normalization constant λh = kb 2 + kbω21ωf 12 12 + kbωf 12 + kbω21 ω21ωf + ω21ωf 12 12 + ω21) 2(ωf , (8) For n≥N+1 are the corresponding probabilities, when the motor is in the "chemical" state 1. Suppose, N is the maximum number of histone-DNA contacts possible in the nucleosome. Let m denote the instantaneous position of the motor. n is the distance be- tween the motor and the far end of histone-DNA contact. The master equations for the probabilities P (m, n, t) are as follows: dP1(m, n)/dt = ω12P2(m, n) − ω21P1(m, n) 12P2(m − 1, n + 1) + ωf + pN [P1(m, n + 1) + P1(m, n − 1) − 2P1(m, n)] (9) and dP2(m, n)/dt = ω21P1(m, n) − ω12P2(m, n) 12P2(m, n) − ωf + pN [P2(m, n + 1) + P2(m, n − 1) − 2P2(m, n)] (10) For n=N dP1(m, N )/dt = ω12P2(m, N ) − ω21P1(m, N ) 12P2(m − 1, N + 1) + ωf + pN [P1(m, N + 1) − P1(m, N )] + pN [f1P2(m − 1, N ) + g1P1(m − 1, N )] + pN −1g0P1(m, N − 1) (11) + b n 2 2 r n S n + a n 1 n+1 1 n − b n F n a − n 6 2 1 n−1 FIG. 7: A schematic representation of the position of the DNA-histone contact with two state of the model. and dP2(m, N )/dt = ω21P1(m, N ) − ω12P2(m, N ) and 12P2(m, N ) − ωf + pN [P2(m, N + 1) − P2(m, N )] + f0pN −1P2(m, N − 1) (12) dP2(m, 1)/dt = ω21P1(m, 1) − ω12P2(m, 1) − kuP2(m, 1) (18) For 3≤ n ≤ N dP1(m, n)/dt = ω12P2(m, n) − ω21P1(m, n) 12P2(m − 1, n + 1) − pnP1(m, n) + ωf + pn[f1P2(m − 1, n) + g1P1(m − 1, n)] + pn+1[f2P2(m − 2, n + 1) + g2P1(m − 2, n + 1)] + g0pn−1P1(m, n − 1) (13) C. Footprint traversal time We define Pµ,n(t) = Pm Pµ(m, n, t) as the probability that the n histone-DNA contacts are intact at time t, irrespective of the position of the CRE motor. From equations (9)-(18), summing over m, we get the following equations: For n≥(N+1) and dP2(m, n)/dt = ω21P1(m, n) − ω12P2(m, n) − (ωf 12 + pN )P2(m, n) + f0pn−1P2(m, n − 1) (14) and dP1,n/dt = pN [P1,n+1 − 2P1,n + P1,n−1] + ωf 12P2,n+1 + ω12P2,n − ω21P1,n (19) For n=2 dP1(m, 2)/dt = ω12P2(m, 2) − ω21P1(m, 2) dP2,n/dt = pN [P2,n+1 − 2P2,n + P2,n−1] + ω21P1,n − ω12P2,n − ωf 12P2,n (20) 12P2(m − 1, 3) − p2P1(m, 2) + ωf + p2[g1P1(m − 1, 2) + f1P2(m − 1, 2)] + p3[g2P1(m − 2, 3) + f2P2(m − 2, 3)] and dP2(m, 2)/dt = ω21P1(m, 2) − ω12P2(m, 2) 12P2(m, 2) − p2P2(m, 2) − ωf For n=1 dP1(m, 1)/dt = ω12P2(m, 1) − ω21P1(m, 1) (15) (16) 12P2(m − 1, 2) − kuP1(m, 1) + ωf + p2[f2P2(m − 2, 2) + g2P1(m − 2, 2)] (17) For n=N dP1,N /dt = −pN P1,N + ω12P2,N + ωf 12P2,N +1 +g0pN −1P1,N −1 + g1pN P1,N + pN P1,N +1 + f1pN P2,N − ω21P1,N (21) and dP2,N /dt = −(ωf 12 + pN )P2,N + ω21P1,N − ω12P2,N +pN P2,N +1 + f0pN −1P2,N −1 (22) For 3≤ n ≤ N dP1,n/dt = −pnP1,n + ωf 12P2,n+1 +ω12P2,n − ω21P1,n + g0pn−1P1,n−1 + g1pnP1,n + g2pn+1P1,n+1 + f1pnP2,n + f2pn+1P2,n+1 (23) and dP2,n/dt = −(ωf 12 + pN )P2,n + ω21P1,n −ω12P2,n + f0pn−1P2,n−1 (24) For n=1 dP1,1/dt = −kuP1,1 + ωf 12P2,2 + f2p2P2,2 + g2p2P1,2 + ω12P2,1 − ω21P1,1 (25) We define the survival probability Sµ,n(t) to be the probability that the CRE has not yet reached the far end of the footprint till time t, given that initially (at t = 0) there were n intact contacts between the histone spool and the DNA on the footprint in front of the CRE motor. Obviously, Sµ,n(t) is the solution of the equations for Pµ,n(t) with the initial condition Sµ,n(0) = 1. Interestingly, the time-evolution of Sµ,n(t) can be re- cast as 7 and dP2,1/dt = −kuP2,1 + ω21P1,1 − ω12P2,1 (26) For n=2 dP1,2/dt = −p2P1,2 + ωf 12P2,3 +g1p2P1,2 + f1p2P2,2 + g2p3P1,3 +f2p3P2,3 + ω12P2,2 − ω21P1,2 (27) and dP2,2/dt = −ω12P2,2 + ω21P1,2 − p2P2,2 −ωf 12P2,2 (28) For n ≥ (N + 1) dS1,n/dt = a+ n (S1,n+1 − S1,n) + a− +rn(S2,n − S1,n) n (S1,n−1 − S1,n) (29) dS2,n/dt = b+ n (S2,n+1 − S2,n) + b− +sn(S1,n − S2,n) + Fn(S1,n−1 − S2,n)(30) n (S2,n−1 − S2,n) where the transition rates a± on the value of n as follows: n , b± n , rn, sn and Fn depend Fn = ωf 12, a+ n = pN , a− n = pN , b+ n = pN , b− n = pN , rn = ω21, sn = ω12 For n=N Fn = ωf 12, a+ n = g0pN , a− n = pN , b+ n = f0pN , b− n = pN , rn = ω21, sn = (ω12 + f1pN ) For 3 ≤ n < N Fn = (ωf 12 + f2pn), a+ n = g0pn, a− n = g2pn, b+ n = f0pn, b− n = 0, rn = ω21, sn = (ω12 + f1pn) For n = 2 F2 = ωf 12 + f2p2, a+ 2 = 0, a− 2 = g2p2, b+ 2 = 0, b− 2 = 0, r2 = ω21, s2 = ω12 + f1p2 For n = 1 F1 = 0, a+ 1 = 0, a− 1 = ku, b+ 1 = 0, b− 1 = ku, r1 = ω21, s1 = ω12 The master equations (29)-(30) together, effectively, cor- respond to the kinetic scheme shown in the Fig. 7. Using this scheme, the MFTT for the single CRE motor can be calculated analytically by extending the theoretical framework developed in ref.[39] for calculating the mean first-passage time of random walks. by Tµ,n = Z ∞ 0 Sµ,n(t)dt (31) Following Pury and Caceres [39], the MFTT is given Since Sµ,n(∞) = 0 and Sµ,n(0) = 1, integrating the equa- tions (29) and (30) with respect to t, we get Using (43) in (44) we, finally, get 8 − 1 = a+ n [T1,n+1 − T1,n] + a− n [T1,n−1 − T1,n] + rn[T2,n − T1,n] (32) − 1 = b+ n (T2,n+1 − T2,n) + b− n (T2,n−1 − T2,n) + sn(T1,n − T2,n) + Fn(T1,n−1 − T2,n) (33) Defining ∆µ,n = Tµ,n+1 − Tµ,n, δn = T2,n − T1,n equations (32) and (33) can be expressed as − 1 = a+ n ∆1,n − a− n ∆1,n−1 + rnδn (34) (35) −1 = b+ n ∆2,n −b− n ∆2,n−1 −snδn −Fn(∆1,n−1 +δn) (36) Now, in the special case ∆1,n = ∆2,n = ∆n, equations (35) and (36) become − 1 = a+ n ∆n − a− n ∆n−1 + rnδn − 1 = b+ n ∆n − (b− n + Fn)∆n−1 − (sn + Fn)δn (37) (38) (39) Next, multiplying equation (38) by (sn + Fn) and Eq. (39) by rn, and then adding the resulting equations, we get − (rn + sn + Fn) = {rnb+ − {rn(b− n + (sn + Fn)a+ n + Fn) + (sn + Fn)a− n }∆n n }∆n−1 (40) Eq. (40) can be re-written as − Cn = Bn∆n − An∆n−1 (41) where, Cn = (rn + sn + Fn), Bn = {rnb+ An = {rn(b− n + (sn + Fn)a+ n + Fn) + (sn + Fn)a− n } n }, We can rewrite Eq. (41) as follows ∆n−1 = Bn An ∆n + Cn An (42) (43) For a fully wrapped histone, the MFTT td is given by td = N Xn=1 △n (44) N td = [Cn/An + Xn=1 ∞ Xi=1 Cn+i/An+i i−1 Yk=0 Bn+k/An+k] (45) Since it is not easy to get an intuitive feeling for the implications of the expression (45), we anaylyze its spe- cial simpler forms in some limiting cases. In the limit of extremely slow motor, i.e., ωf 12 → 0, as expected, the expression (45) for the MFTT td diverges. For ensuring high-speed of the CRE motor, we need simultaneously wf 12/kb >> 1 and w21/kb >> 1. If, for simplicity, we make the additional assumption that wf 12 is the slower of the two, i.e., ω21 >> ωf 12, we have f0 = f1 = g0 = g1 ≃ 0 and f2 ≃ 1, and g2 ≃ 1. Hence, in this limit, Bn = 0 for n ≤ N and, therefore, td = N Xn=1 Cn An ≃ N/ku (46) which is identical to the corresponding limiting value of td reported in ref.[29]. This is a consequence of the fact that in the limit of extremely fast motor, because of the assumption of very large value of ω21, the 2-state model reduces to an effectively 1-state model. We make a nu- merical estimate of td in this limit by computing an ap- proximate value of ku. Defining K = kb/ku (47) as the flap binding constant, we can rewrite the equation (46) as td = N K kb (48) Range of typical values of K has been used earlier by Chou [29]. Using this range of values for K, one can estimate ku, provided a typical value of kb is known. Therefore, we now estimate the typical numerical val- ues of kb following Schiessel and coworkers [16 -- 18]. Sup- pose, L0(≃ 50 nm) be the length of the DNA that wraps around the histone spool. Let L + dL be the contour length of the loop induced by spontaneous thermal fluc- tuations where (see Fig. 8) L is the exposed arc length on the histone spool that was covered by the DNA seg- ment prior to the loop formation and dL is a small seg- ment of the linker dsDNA that has been pulled into the loop. ′ ′ We assume that the life time of a loop (τ ) is much shorter than the average time required to form a loop. Following Schiessel et al. [16 -- 18] we write down the rate of loop formation as α ≃ L0exp(−Ebend/(kBT )) τ L′ Comparing Eqs. (1) and (49) we obtain kb = L0 τ L′ . (49) (50) ′ L = L + dL ′ L 9 K=20 K=10 K=2 K=1 K=0.2 2e+06 1 - N α d t 10000 100 2 0.1 10 60 1 ωf 12/α FIG. 8: Top view of the histone octamer bound with DNA. Loop formation involving length dL of linker chain being in- corporated into the nucleosome, with length L′ of the exposed surface. (adapted from Fig. 2 of ref. [16]). FIG. 9: Normalized MFTT αN −1td plotted against the nor- malized motor speed ωf 12/α for different values of K with ω12/α = ω21/α = 0.5. Since τ −1 = k characterizes the rate of unbiased diffusion of the loop around the histone spool [16] τ ≃ L2 D where D is the corresponding diffusion constant. From Stokes- Einstein relation D = kBT /ζ, where ζ ≃ ηL [16] and η is the effective viscosity of the aqueous medium. Combining all the results and substituting these into Eq. (50) we finally obtain 0 ′ kb = kBT ηL0(L′ )2 (51) ′ The estimation can be completed only if an estimate of L is available. Following ref. [16] (Eq. (2a) of [16]), we get ′ L ≃ ( 20π4κ λR2 0 )1/6(dL/R0)1/3R0, (52) where κ is the bending elastic constant of the semi- flexiable DNA chain, λ is the adsorption energy per unit length and R0 is the radius of the histone spool. Us- ing the reasonable values quoted in ref.[16 -- 18], namely, R0 = 5nm, κ = 207.10pN − nm2, dL = 3.40nm and λ = 5.92pN we obtain from Eq. = 16.43nm. using this estimate of L , together with 1kBT = 4.142 pN-nm, L0 = 500A, and η = 1 Centipoise, we obtain from Eq. (51) kb = 306877.4s−1. (52) L ′ ′ With the above estimated value of kb, and N = 15 from Eq. (48) we get the estimates td = 0.000024s for K = 0.5 and td = 0.0005s for K = 10. Such small values of td, estimated from Eq.46, arise from the fact that the approximate form (46) is valid only in the limit of extremely fast motor. Therefore, this limiting formula provides only a lower bound and does not correspond to real CRE motors under physiological conditions. In Fig.9 we plot the normalized MFTT tdα/N as a function of the normalized motor speed ωf 12/α for ω12/α = ω21/α = 0.5 and a few fixed values of the pa- rameter K. For any fixed value of K, the normalized MFTT decreases monotonically with the increase of the normalized motor speed and saturates to the value given by equation (46) in the limit ωf a given value of ωf increases the MFTT increases. 12/α → ∞. Moreover, for 12/α, as the flap binding constant K In Fig.10, td is plotted against ωf 12/α for (a) K = 10, ω12/α = 0.8, and (b) K = 10, ω12/α = 0.1, each for a few distinct values of ω21/α. The MFTT decreases as ω21/α increases. This is a consequence of the fact that ω21 depends on the ATP concentration. For small ω12/α reduces the amplitude of peeling time. In Fig.11 we demonstrate that for large value of ω21/α, which effectively speeds up the motor, reduces the mag- nitude of the MFTT. Although the qualitative trends of variations of td with ωf 12/α in our model is similar to that in Chou's model [29], wide range of variation of td is possible in our model by controlling ω21 which, in turn, can be controlled by the ATP concentration. In order to explore the dependence of td on the con- centration of ATP, we first assume that ω21 = ω0 21[AT P ] (53) Assuming a typical value ω0 21 = 106M −1s−1, we have plotted the normalized MFTT against the ATP concen- tration for two different normalized values of the unhin- dreed motor speed keeping the other parameters fixed. With the increase of ATP concentration, the MFTT de- creases and, gradulally saturates. When ATP concentra- tion is sufficiently high, the step with rate constant ω21 is no longer rate-limiting. We also find that, for a given ATP concentration, the higher is the value of ωf 12/α the shorter is the MFTT td. The linear dependence of ω21 on ATP concentration, as envisaged in (53), may be valid only at sufficiently low concentration of ATP. In general, ω21 may follow the usual Michaelis-Menten equation for the rate of enzy- matic reactions (because ω21 represents the rate of ATP hydrolysis catalyzed by the CRE motor) [40]. In that case ω21 itself would saturate with the increase of ATP concentration, instead of increasing linearly with [ATP]. (a) (a) 10 1 - N α d t 1e+08 1e+06 1000 2 1e+06 100000 10000 1000 100 10 1 - N α d t 21/α=0.1 ω 21/α=0.3 ω 21/α=0.5 ω 21/α=0.7 ω ω 21/α=0.9 21/α=1.0 ω ω 12/α=0.8 K=10 100 1 ωf 10 12/α (b) ω ω ω ω ω ω 21/α=0.1 21/α=0.3 21/α=0.5 21/α=0.7 21/α=0.9 21/α=1.0 K=10 ω 12/α=0.1 10000 1000 100 10 1 - N α d t 2 0.2 2e+08 1e+07 100000 1000 1 - N α d t 12/α=0.1 ω 12/α=0.3 ω 12/α=0.5 ω 12/α=0.7 ω ω 12/α=0.9 12/α=1.0 ω ω 21/α=0.8 K=10 10 1 ωf 12/α (b) 12/α=0.1 ω ω 12/α=0.3 ω 12/α=0.5 12/α=0.7 ω 12/α=0.9 ω 12/α=1.0 ω ω 21/α=0.1 K=10 1 0.1 1 ωf 10 12/α 100 20 0.3 1 ωf 12/α 10 FIG. 10: Normalized MFTT αN −1td plotted against the nor- malized motor speed ωf 12/α for different values of ω21/α with (a) ω12/α = 0.8, and (b) ω12/α = 0.1, K = 10. FIG. 11: Normalized MFTT αN −1td plotted against the nor- malized motor speed ωf 12/α for different values of ω12/α with (a) ω21/α = 0.8, and (b) ω21/α = 0.1, K = 10. IV. CONCLUSION In this paper we have studied the process of ATP- dependent chromatin remodeling. For simplicity, we have considered only a single nucleosome consisting of a ds- DNA strand wrapped one and three-fourth turns around a cylindrical spool made of histone proteins. We have extended Chou's model [29] by assigning two distinct "chemical" states to the CRE and postulating a minimal mechano-chemical kinetic scheme for capturing the ef- fects of ATP hydrolysis explicitly. Our theoretical frame- work has been developed exploiting a close analogy with the unzipping of a double-stranded DNA by a helicase [38]. We have written down the master equations for the postulated kinetic scheme. This model of footprint traversal by ATP-dependent CRE can be easily inter- preted as an implementation of a Brownian ratchet mech- anism. From an analytical treatment of this stochastic kinetic model, we have derived analytical expression for the MFTT of the ATP-dependent CRE. We make explicit analytical predictions on the dependence of the MFTT on (i) the unhindred speed of the CRE, as well as on (ii) the ωf ωf 12/α=1.0 12/α=0.5 ω 12/α=0.8 K=10 70 60 50 40 30 20 10 1 - N α d t 0 1e-06 1e-05 0.0001 0.001 [ATP] (M) FIG. 12: Normalized MFTT αN −1td plotted against the ATP concentration [ATP] for ω12/α = 0.1, K = 10.0 and two dif- ferent values of ωf 12/α. concentration of ATP. In principle our theoretical predic- tions can be tested by carrying out in-vitro experiments with a single nucleosome. Acknowledgements: DC thanks Michael Poirier and Tom Chou for useful discussion and correspondence, re- spectively. DC acknowledges support of the Visitors Program of the Max-Planck Institute for the Physics of Complex Systems in Dresden where parts of this work were carried out during two separate visits. DC also thanks Frank Julicher for discussions in the initial stages of this work and Anirban Sain for a critical reading of the manuscript. This work is also supported, in part, by IIT Kanpur through the Dr. Jag Mohan Chair professorship (DC) and by a research grant from CSIR, India (DC). AG thanks UGC, India, for a senior research fellowship. 11 [1] A. Wolfee, Chromatin: structure and function, (Aca- [22] W. Mobius, R.A. Neher and U. Gerland, Phys. Rev. Lett. demic Press, 1998). 97, 208102 (2006). [2] J. Widom, Annu. Rev. Biophys. Biomol.Struc. 27, 285 (1998). [3] H.Schiessel, J. Phys. Condens. Matter, 15, R699 (2003). [4] C. Lavelle and A. Benecke, Eur. Phys. J. E 19, 379 (2006). [5] R.D. Kornberg and Y. Lorch, Cell 98, 285 (1999). [6] K.J. Polach and J. Widom, J. Mol. Biol. 254, 130 (1995). [7] J.D. Anderson, A. Thastrom and J. Widom, Mol. Cell. Biol. 22, 7147 (2002). [8] H.Y. Fan, X. He, R.E. Kingston and G.J. Narlikar, Mol. Cell 11, 1311 (2003). [9] A. Flaus and T. Owen-Hughes, Biopolymers 68, 563 (2003). [10] A. Flaus and T. Owen-Hughes, FEBS J. 278, 3579 (2011). [23] J. Langowski, Eur. Phys. J. E 19, 241 (2006). [24] A. Lense and J.M. Victor, Eur. Phys. J. E 19, 279 (2006). [25] C. Vaillant, B. Audit, C. Thermes and A. Arneodo, Eur. Phys. J. E 19, 263 (2006). [26] P. Ranjith, J. Yan and J.F. Marko, PNAS 104, 13649 (2007). [27] G. Lia, E. Praly, H. Ferreira, C. Stockdale, Y.C. Tse- Dinh, D. Dunlap, V. Croquette, D. Bensimon and T. Owen-Hughes, Mol. Cell 21, 417 (2006). [28] B.R. Cairns, Nat. Struct. and Mol. Biol. 14, 989 (2007). [29] T. Chou, Phys. Rev. Lett., 99, 058105 (2007). [30] G. Li, M. Levitus, C. Bustamante and J. Widom, Rapid spontaneous accessibility of nucleosomal DNA, Nat. Struct. Mol. Biol. 12, 46-53 (2004). [31] M.G. Poirier, M. Bussiek, J. Langowski, J. Mol. Biol. [11] S.E. Halford, A.J. Welsh and M.D. Szczelkun, Annu. 379, 772-786 (2008). Rev. Biophys. Biomol. Struct. 33, 1 (2004). [32] M.G. Poirier, E. Oh, H.S. Tims and J. Widom, Nat. [12] A. Saha, J. Wittmeyer and B.R. Cairns, Nat. Rev. Mol. Struct. Mol. Biol. 16, 938 (2009). Cell Biol. 7, 437 (2006). [33] H.S. Tims, K. Gurunathan, M. Levitus and J. Widom, [13] C.R. Clapier and B.R. Cairns, Annu. Rev. Biochem. 78, J. Mol. Biol. 411, 430 (2011). 273 (2009). [34] G. Meersseman, S. Pennings and E.M. Bradbury, EMBO [14] L. R. Racki and G. J. Narlikar, Curr. Opin, Genet. Dev. J. 11, 2951 (1992). 18, 137 (2008). [15] T. Sakaue, K. Yoshikawa, S.H. Yoshimura and K. [35] A. Lusser and J.T. kadonaga, BioEssays 25, 1192 (2003). [36] P.B. Becker and W. Horz, Annu Rev. Biochem. 71, 247- Takeyasu, Phys. Rev. Lett. 87, 078105 (2001). 273 (2002). [16] H. Schiessel, J. Widom, R.F. Bruinsma and W.M. Gel- [37] G. Langst and P.B. Becker, Biochim. Biophys. Acta bert, Phys. Rev. Lett. 86, 4414 (2001). 1677, 58 (2004). [17] I.M. Kulic and H. Schiessel, Phys. Rev. Lett. 91, 148103 [38] A. Garai, D. Chowdhury, M. Betterton, 77, 061910 (2003). (2008). [18] I.M. Kulic and H. Schiessel, Biophys. J. 84, 3197 (2003). [19] H. Schiessel, Eur. Phys. J. E 19, 251 (2006). [20] F. Mohammad-Rafiee, I.M. Kulic and H. Schiessel, J. Mol. Biol. 344, 47 (2004). [21] R. Blossey and H. Schiwssel, FEBS J. 278, 3619 (2011). [39] P.A. Pury and M. O. Caceres, J. Phys. A 36, 2695 (2003). [40] M. Dixon and E.C. Webb, Enzymes (Academic Press, 1979).
1707.07620
1
1707
2017-05-13T19:00:14
Comparison of Decision Tree Based Classification Strategies to Detect External Chemical Stimuli from Raw and Filtered Plant Electrical Response
[ "physics.bio-ph", "cs.LG", "physics.data-an", "stat.AP", "stat.ML" ]
Plants monitor their surrounding environment and control their physiological functions by producing an electrical response. We recorded electrical signals from different plants by exposing them to Sodium Chloride (NaCl), Ozone (O3) and Sulfuric Acid (H2SO4) under laboratory conditions. After applying pre-processing techniques such as filtering and drift removal, we extracted few statistical features from the acquired plant electrical signals. Using these features, combined with different classification algorithms, we used a decision tree based multi-class classification strategy to identify the three different external chemical stimuli. We here present our exploration to obtain the optimum set of ranked feature and classifier combination that can separate a particular chemical stimulus from the incoming stream of plant electrical signals. The paper also reports an exhaustive comparison of similar feature based classification using the filtered and the raw plant signals, containing the high frequency stochastic part and also the low frequency trends present in it, as two different cases for feature extraction. The work, presented in this paper opens up new possibilities for using plant electrical signals to monitor and detect other environmental stimuli apart from NaCl, O3 and H2SO4 in future.
physics.bio-ph
physics
Sensors and Actuators B: Chemical Comparison of Decision Tree Based Classification Strategies to Detect External Chemical Stimuli from Raw and Filtered Plant Electrical Response Shre Kumar Chatterjee1, Saptarshi Das1, Koushik Maharatna1, Elisa Masi2, Luisa Santopolo2, Ilaria Colzi2, Stefano Mancuso2 and Andrea Vitaletti3,4 1 School of Electronics and Computer Science, University of Southampton, Southampton SO17 1BJ, UK 2 Department of Agri-food Production and Environmental Science, University of Florence, FIorence, Italy 3 DIAG, Sapienza University of Rome, via Ariosto 25, 00185 Rome, Italy Email: [email protected], [email protected] (S. Das*) Phone: +44(0)7448572598, Fax: 02380 593045 Abstract Plants monitor their surrounding environment and control their physiological functions by producing an electrical response. We recorded electrical signals from different plants by exposing them to Sodium Chloride (NaCl), Ozone (O3) and Sulfuric Acid (H2SO4) under laboratory conditions. After applying pre-processing techniques such as filtering and drift removal, we extracted few statistical features from the acquired plant electrical signals. Using these features, combined with different classification algorithms, we used a decision tree based multi-class classification strategy to identify the three different external chemical stimuli. We here present our exploration to obtain the optimum set of ranked feature and classifier combination that can separate a particular chemical stimulus from the incoming stream of plant electrical signals. The paper also reports an exhaustive comparison of similar feature based classification using the filtered and the raw plant signals, containing the high frequency stochastic part and also the low frequency trends present in it, as two different cases for feature extraction. The work, presented in this paper opens up new possibilities for using plant electrical signals to monitor and detect other environmental stimuli apart from NaCl, O3 and H2SO4 in future. Index Terms Decision tree, multiclass classification, discriminant analysis, Mahalanobis distance classifier, statistical features, plant electrical signal processing, time series analysis Introduction I. Plants, such as Mimosa pudica (Touch-me-not) and Helianthus annuus (Sunflower), show some form of physical changes due to external stimuli in the form of touch and sunlight 1 Sensors and Actuators B: Chemical respectively [1]. The wilting of general plants due to dry environmental conditions is also commonly found. For many years, researchers have tried to establish the relationship between these reactions of the plants and the surrounding environmental conditions [1]. It has been found that the underlying phenomenon behind this is the plant electrophysiological mechanism which may be traced in the electrical response of the plant to the external stimulus [1]. These electrical signals, which control various physiological functions in the plants, hold useful information about the external stimulus (which causes the electrical signal in the plant) contained within its deterministic and stochastic parts to different extents. Analysis using low frequency (trend) part of the plant electrical signal to study the external chemical or light stimulus has been reported in Chatterjee et al. [2], [3]. Also, other studies on plant electrical signal processing have been reported in [4]–[14], in particular use of classification techniques to find out the applied external stimuli, through various statistical features computed from the recorded plant electrical signal, was reported first in [2]. Since the statistical features in [2] were extracted from raw plant signals (with low frequency trends or drifts), a background (pre- stimulus) information subtraction method was adopted in the classification process to focus only on the incremental values in each feature due to the application of the stimulus. In this paper, we initially focus on the information contained in the stochastic part of the plant electrical signals by applying a high pass filtering on the raw signals to remove the inconsistent trends or drifts. We also used raw signals with the trends (using the background information subtraction method as reported in [2]) to show a comparative analysis between the classification performance of the filtered and raw plant signals. Thus, we here explore, if there is any improvement in the classification process while using only the detrended random part rather than the raw signal containing small local fluctuations superimposed on relatively larger change in the trends. In order to develop a classification strategy for detecting the external chemical stimuli, here we used 15 features out of which 11 features have been reported in [2]. In addition to these 11 features, four additional statistical features have been explored along with independent testing of the classifiers with a much larger dataset. In this paper, we report the use of discriminant analysis and Mahalanobis distance based classification algorithms to establish a decision tree based classification system using both One-Versus-One (OVO) and One- Versus-Rest (OVR) strategy [15], [16]. We also report the validation scheme of the classifiers in two different ways – 1) Leave One Out Cross Validation (LOOCV) on ~73% of the available data (retrospective study) and 2) independent testing of the remaining ~27% of the data (prospective study). Datasets from experiments on plants using three different stimuli – NaCl, H2SO4 and O3 have been used in this exploration and can be found on the EU FP7 funded project PLants 2 Sensors and Actuators B: Chemical Employed As SEnsor Devices (PLEASED) website. The work presented in this paper may help in taking a step closer to the concept of a plant electrical signal based external stimuli sensing platform which the PLEASED project aims to develop. Such a device, if successful, will aid in monitoring a large geographical area for multiple environmental stimuli or pollutants of interest. In order to proceed towards the realization of such a stimulus classification scheme using plant electrical signals, the steps shown in Figure 1 were followed. Figure 1: Steps for classification of environmental stimuli using plant electrical signal. The work presented in this paper presents the following salient contributions over previous approaches: • The present work uses filtered (containing only stochastic part of the signals) as well as raw plant signal (containing both deterministic and stochastic parts of the signals) for classifying the external stimuli for a comparative analysis between two pre-processing techniques affecting the final classification results. • An exhaustive set of experimental data with 28,070 data blocks were used for training the classifiers, which is ~7.4 times higher than the data used in our previous work [2]. • Apart from Tomato and Cucumber plants which were used for the experiments reported in [2], Cabbage was included in the present work for extracting the experimental data. This helped to analyse the electrical response in a pool of three different species due to chemical stimulus. • Features from NaCl of different concentration (5ml and 10 ml) were combined to be labelled as NaCl in the present classification work, whereas in the previous work [2], these two stimuli were considered as separate classes. 3 Sensors and Actuators B: Chemical • In the previous study [2], the classification results were obtained using individual (univariate analysis) or feature pairs (bivariate analysis), whereas multivariate feature analysis have been carried out here. • Apart from the previously explored 11 statistical features in [2], here we explore four additional features of the plant electrical signal for classification of the external stimuli viz. higher order central moments (hyperskewness and hyperflatness), fano factor, and correlation dimension. • The classification results reported in the previous study [2] was average results for six binary stimuli combinations that in a way averages the detectability of one class with the others. Whereas a systematic decision tree has been developed in the present study that helps answering the question which chemical stimulus can be easily detected from the plant signals and which stimuli are hard to differentiate from the rest. • Along with reporting the retrospective classification accuracy employing LOOCV, a separate held out dataset was used for prospective validation in this work, whereas only retrospective LOOCV results were reported in the previous work [2], on a much smaller dataset. II. Recording Electrical Signal from Plants Raw electrical signals from different experiments involving O3, NaCl and H2SO4 as external stimuli, under laboratory conditions, were acquired from different plants. Each experiment was conducted on a new plant, thereby eliminating the risk of any residual effect of the previous experiments infiltrating the current electrophysiological condition of the plants. The experiments were conducted inside a plastic transparent box placed in a Faraday cage kept in a dark room, as shown in Figure 2, to minimize any external electrical or light interference. Artificial lights using LED lamps were provided in a 12 hours darkness / 12 hours light cycle to cater for the plant's photosynthetic needs. The plastic enclosure was used to introduce ozone and this enclosure being transparent is important so that the plant gets light for photosynthesis under normal condition. During chemical stress, the room was made dark as light can generate a different type of action potential thus masking the effect of chemical stimuli. For each experiment, three stainless steel electrodes from Bionen S.A.S, each of 0.35 mm in diameter and 15 mm in length, were used. The polarization effect of using steel electrodes may be studied in the future which is not considered here. These electrodes, similar to those used for Electromyography (EMG), were inserted in the middle and at the top of the stem of the plant for data acquisition and in the base of the stem for reference. The other end of the electrodes was connected into a two channel high impedance electrometer (DUO 773, WPI, USA). Data recording was carried out at a rate of 10 samples/sec through a data acquisition system, 4 Sensors and Actuators B: Chemical provided by Labtrax, WPI. More detailed information on the experimental setup can be found in [2]. For introducing O3 as a stimulus, an inlet silicone tube was placed in the plastic transparent enclosure, while a second outlet tube throws the O3 from the box to a chemical hood. O3 was produced using a commercial ozone generator (mod. STERIL, OZONIS) and the concentration inside the box was measured using a suitable sensor. Similarly, regarding H2SO4 or NaCl stimulus, a silicone tube inserted into the plant soil and connected to a syringe placed outside the Faraday cage for injecting the solution. Figure 2 shows the description of the experimental setup to administer the stimuli. Before applying the stimulus, a period of 45 minutes was allowed to the plant to recover after the insertion of the electrodes into the stem. Figure 2: Experimental setup to extract plant electrical signal using different external stimuli. Table 1: Details of the experiments with different chemical stimuli Stimulus Concentration Stimulus applied during each experiment Ozone (O3) 16 ppm/ 13.07 ppm Multiple Sulfuric acid (H2SO4) Sodium Chloride (NaCl) 5 ml of 0.05/ 0.025 M solution Once/twice 5/10ml of 3M solution Once 5 Plant species used Tomato Cucumber Cabbage Tomato Cabbage Tomato Number of independent time series 8 4 20 8 20 16 Sensors and Actuators B: Chemical Table 1 gives details about the experiments conducted to extract the datasets which has been used for the exploration presented in this paper. For experiments with H2SO4, the stimulus was applied once or twice to the plant in the form of 5 ml solution of 0.025M or 0.05M concentration of H2SO4. For experiments with O3, the gas was injected into the box for 1 min every hour and the maximum concentration ranged between 13-16 ppm. NaCl treatment was carried out through addition of 5 or 10 ml of 3M NaCl solution to the soil. In [2], experiments using 5ml and 10ml of NaCl solutions were used as two separate classes. In the present work, we have combined both these concentrations of stimulus as a single class (i.e. NaCl), as here we want to explore the effect of NaCl as a whole while neglecting the slight variation due to its different concentrations. Also, as is evident from Table 1, the experiments were conducted on multiple species of plants - Tomato (Solanum lycopersicum), Cucumber (Cucumis sativus) and Cabbage (Brassica oleracea) so that we can build a generalized classifier which is not specific to a particular plant species but is able to pick up the common signature in the electrical response of various different species. Whether some plant species are more sensitive to certain type of stimulus, can be explored in a future study. Therefore in this paper, we aim in making a generalized and robust detection technique for three chemical stimuli - H2SO4, NaCl, and O3, rather than focusing on quantification of its variability due to the different species and concentration of the stimuli. III. Pre-processing and Data Segmentation A. Designing Optimum Filter for Drift Removal In general it is not known which range of frequencies of the plant electrical signal contains useful information about the external stimulus which can be fed to the classifier. Therefore the first aim was to calibrate the pre-stimulus parts (due to different amplitude level for each experiment being different at the onset) and then bringing them at a common platform, so that the change in the post-stimulus part, upon application of a stimulus, could be quantified. We tuned the parameters of four digital high-pass Infinite Impulse Response (IIR) filters – Butterworth, Chebyshev type I, Chebyshev type II, and Elliptic filter [17], such that it minimizes the distance between the centroids of the clusters, formed using the distribution of wavelet packet energy along different wavelet basis for the pre-stimulus signals. We selected IIR over FIR filter structures due to lesser number of parameters which needed to be tuned. The objective function selected for the optimization is the energy contents of different leaf nodes in wavelet packet decomposition for the pre-stimulus parts of the plant signals which were acquired under laboratory settings, details of which have been reported in [17]. The IIR filter parameters were tuned in such a way that it produced almost overlapping clusters in the 6 Sensors and Actuators B: Chemical domain of wavelet packet energy along different wavelet basis for the pre-stimulus signals in different experiments, without enforcing the same for the post-stimulus signals (as this may mask the natural stimulated response of the plant). Therefore under such tuning of the filter as a pre-processing step for benchmarking of the background information (before any stimulus was applied), the post stimulus response can now be analysed to investigate any gross change in the statistical characteristics of the plant electrical signals. Through the exploration reported in [17], it has been found that a 6th order Chebyshev type-II high-pass filter yields the minimum cost function with a cut-off frequency of 0.77 Hz and stop-band ripple of 100 dB. Due to the fact that plant signals are slow in nature [1] and the bandwidth of the data is already constrained due to the present 10 Hz sampling rate, the recorded signals did not need to be low-pass filtered for the high frequency noise attenuation which might have come due to a faster sampling rate. However plant signals has significant drift in the low frequency part of the spectrum, removal of which has been addressed in [17]. In Figure 3, the top panels show example of raw signals, the middle panels show the high-pass filtered signals and the bottom panels show the Welch power spectrum estimate for the post-stimulus part. The dotted vertical lines indicate the time instant for application of the stimulus. Figure 3: Plant electrical signals due to three different chemical stimuli. Representative examples of the raw and filtered plant signals using this optimum filter are shown in Figure 3 for the three different chemical stimuli under consideration. The high-pass filtering removes the inconsistent low frequency drifts as evident from the almost horizontal detrended mean and the drop in power at low frequency in the Welch power spectrum estimates for the respective post-stimulus parts. It is worth mentioning that in most cases the drift is present in recorded plant electrical signals, like many other biological signals, due to 7 Sensors and Actuators B: Chemical variation in the instrumentation condition, environmental artefacts and changes in biological conditions over time. The possible reasons, effect and removal of the drift is discussed in a detailed manner in [17]. Due to the presence of the drift, either the signal has to be filtered to minimise its effect or the background information embedded in the features needs to be removed. We here compare both of these methods that yields better classification results. From Figure 3, apparently it may appear that the power spectrums of the filtered plant signals under three different stimuli look quite similar (except few glitches in certain frequencies). This is in agreement with the findings in [2] and this study later reported that a simple spectral power as a feature cannot distinguish between these signals and thereby not provide good classification performance. Although the power spectrums look similar, there may be different information embedded within the plant electrical signals. These information's can be captured using higher order moments and other nonlinear statistical features which may not be captured with spectral power alone. B. Data Windowing and Validation Scheme for the Classification Our previous work [2], [3] shows that plant electrical signals are slow in nature, therefore we here hypothesize that there will be sufficient information about the external stimulus within ~1.5 min of data segment (with sampling rate = 10 samples/second). In [17], a segment size of 256, 512, 1024 samples were chosen which are equivalent to ~0.42, 0.85, 1.7 min of streaming data respectively, in order to show the spectral energy distribution of the plant signals. The Fourier domain representation of the autocorrelation function i.e. the power spectral densities are explored in [17] and also in Figure 3 which show that the unfiltered signals have significantly high power in the low frequency due to the drifts which are reduced due to the optimum high-pass filtering. A very large sample size selection may introduce slow oscillations of the plant signal, thus corrupting the extracted features. It is important to note that after the optimum filtering only the high frequency components remain dominant as shown in Figure 3 bottom panels, denoting a fast decaying autocorrelation sequence. Table 2: Blocks (of 1024 samples) for each stimulus in different validation schemes of the classifiers Validation Scheme NaCl H2SO4 O3 352 Prospective study (independent testing) 276 Retrospective study (LOOCV) 1340 26378 148 9692 of such Next, we divided the entire filtered signal into blocks of 1024 samples. In total blocks were obtained (each block containing 1024 samples) from the post stimulus parts of different experimental datasets. Out of these total blocks, blocks (~27% of the total data 3.8 10 410 ≈ × 4 8 Sensors and Actuators B: Chemical blocks) were set aside for prospective study i.e. independent testing of the classifier settings. These data blocks for prospective study were obtained from completely separate experiments which were never seen by the classifiers. Figure 4: Dividing the filtered plant electrical signal into blocks of 1024 samples (shown by black circles) and computing feature values from each of these blocks (shown by red/green/blue boxes). The data segmentation with non-overlapping window is shown in Figure 4 for the three stimuli, where the black circles refer to the blocks of 1024 samples from which the feature values are computed. The number of blocks used for each stimuli are shown in Table 2 from which we can see the main imbalance in the signal length was caused due to Ozone as a class, which contributed more data blocks due to the longer duration of the experimental recording than NaCl and H2SO4 [2]. The best classifier and feature combinations was selected by employing the LOOCV [18] on ~73% of the total data. The reason behind using only 73% of the data to find out the best feature-classifier combination were a) to avoid selecting an over-fitted model of the total data, and b) test the obtained best classifier-feature combinations on a section of independent data i.e. to estimate the performance of this combination in classifying completely unseen data. 9 Sensors and Actuators B: Chemical C. Rationale of segmentation of long plant signals for feature extraction and classification In an ideal scenario of cheap and easy experimental data collection, it is often argued to train the classifier on a portion of the available dataset and then testing the model on a completely different experimental data commonly known as the held-out validation. Whereas we here employ a different approach, also known as the cross-validation scheme, commonly used when the independent realizations are small due to experimental constraints/costs involved. The problem with the held-out testing (particularly for smaller number of independent dataset) is that the results may introduce some bias which is typical to that particular segment of the held out dataset. In such a scenario, a smaller segment of the held out data in the testing phase may not always represent the common statistical properties of the whole dataset. Whereas a cross-validation reports average classification accuracy over all the segments of the data and is a much more robust measure of accuracy for judging the classification performance while working with small number of independent experimental datasets. Due to such scarcity of independent time series data coming from different experiments (as shown in Table 1), we first use resampling on the limited number of long recorded signals by windowing them in smaller segments. This resampling approach increases the number of data points in each class, so that a portion of the data can be held out to test the standard way of cross- validation in training phase, followed by independent held-out validation in testing phase for judging the classifier's performance. While working with limited number of independent experimental dataset, this resampling and data segmentation method has been widely used before and it is based on the assumption that the segmentation and resampling in the same long time series has similar statistical characteristics that we want to uncover using a classifier. This approach also assumes that different data segments coming from the same or limited number of long time series capture the effect of a particular stimulus in a similar way so that they can be grouped together using a classifier. Huerta et al. [19] and Vembu et al. [20] reported dynamic support vector machine (SVM) with linear autoregressive (AR) kernel in similar time series classification problems. This is a viable option for time series classification, if the underlying data generation process is almost linear and smooth nonlinear. But for highly nonlinear and nonstationary time series, meaningful feature extraction that can capture these complex behaviours, becomes a necessity. For example in our study, the Hurst exponent, detrended fluctuation analysis (DFA) signify the long-range correlations and power law characteristics in the data, higher moments (skewness, kurtosis, hyperskewness, hyperflatness) signifying different forms of non-Gaussian attributes, correlation dimension and entropy signifying the extent of superimposed deterministic chaotic and stochastic noisy attributes in the data etc. [21]. 10 Sensors and Actuators B: Chemical The above mentioned method of using held out data may be a valid approach when the number of time series are significantly large where each independent time series can be represented as a single data point. In the case of relatively less number of available independent experimental dataset, a resampling statistics is a viable option, wherein the same long time series are chopped in several smaller segments to increase the number of data points. This resampling approach considers the data in each segment being independent and identically distributed (IID) and consider correlation only within the segment. Dividing long time series signals (with fewer independent realisations) in to smaller segments ignores the correlation lost between successive segments. In other words, it is assumed that the auto- correlated samples lie within a short range i.e. within the segment of 1024 samples but there is negligible autocorrelation between samples far apart, exceeding the window length. The choice of the window length and its effect on having a uniform background to compare the stimulated response has been discussed in [17]. Here, as a pre-processing step, we carried out signal detrending using the optimum high pass filter in [17] which significantly reduces the long distance autocorrelation of the samples so that a shorter window based feature extraction and classification can be applied effectively. This is a valid assumption for practical time series classification problem if the goal is to detect the cause behind the change in time series patterns by only looking at some features of the smaller segments. Here the segment size was chosen as 1024 samples (equivalent to 1.7 min) to extract the features. For any practical detection, this is a sufficiently long window for monitoring the chemical stimuli in the environment, rather than extracting features on the whole duration of the time series. This approach of resampling long time series in smaller segments has been previously used in many other biological time series analysis problems e.g. Electrocardiogram (ECG) beat classification [22], [23] and sleep stage classification using smaller epochs from a long Electroencephalogram (EEG) recordings [24], [25], with relatively less number of independent experiments or subjects, as in the present case as well. In our previous work [17], analyses with different segment size of the plant signals (256, 512, 1024 samples) were carried out to tune the parameters of the drift removal high-pass filter. It is understandable that too small a segment size may result in the nonstationary signal behaving as a stationary signal [26], however most of the features of long range correlation like Hurst exponent and DFA will not be very meaningful in such a scenario. Also, within a very short time window, there may not be significant change in the statistical characteristics of the plant signal, as the inherent biological oscillations in plants are slow in nature. Therefore we have chosen the window heuristically as 1024 samples that allows taking sufficient samples for both short and long range feature extraction from these segmented time series. 11 Sensors and Actuators B: Chemical The data segmentation is a valid approach with a stationary and ergodic assumption of the signal in both the pre and post-stimulus parts. A stationary ergodic process implies that the random process will not exhibit any significant change in its statistical properties with time and the statistical measures such as various moments (e.g. mean, variance etc.) of the stochastic process will be similar. As in the present scenario, where the plant electrical responses are typically non-stationary biological signals, we took two different approaches of pre-processing in order to transform them from non-stationary to stationary signals/features. This transformation assumes that the effect of the stimulus is embedded in the signals in a similar way between different segments. This also implies that the stimulus has modified one or more statistical features to a significant extent between the pre and post-stimulus parts whereas such variation between segments coming from the same pre or post-stimulus parts are not that significant. This criteria is enforced by the optimum high-pass IIR filter as designed in [17] since it brings the energy distribution of the background signals to a common reference point (centroid of the wavelet packet energies across different nodes). The two pre-processing steps for enforcing this stationarity assumption is given below: • When using raw plant signals for feature extraction followed by classification – the features were derived from small epochs of fixed sample size (1024 samples), from both pre and post stimulus parts of the signals. The mean values of each feature from the pre-stimuli parts, were then calculated and subtracted from the corresponding features of the post- stimulus parts. This ensured that the resultant pre-stimulus mean subtracted features of the post-stimulus parts reflect only the change in the statistical features due the introduction of the stimulus. Hence any bias due to the difference in the pre-stimulus signal's starting amplitude or slow time varying drift can be minimised. This approach has also been adopted in [2] for a similar classification problem. • When using filtered plant signals, an optimum high-pass filter was applied on the raw signal to remove the low frequency trends and the remaining stochastic parts were used to divide the long signals in multiple epochs of fixed sample size (1024 samples). Features were calculated from each of these epochs from the post-stimulus parts of the signals since the high pass filter has already been tuned using the pre-stimulus information as reported in [17]. In both of these two methods, each segmented epochs were assumed to be independent and identically distributed (IID) belonging to the same class as of the original duration of the plant electrical signals. This segmentation may break the long range correlations in the data but resampling of long time series and using each segment as a different data point is a well- established method in other bio-electrical signal processing application as well e.g. ECG beat classification [22], [23] and dividing EEG signals into several epochs for sleep stage 12 Sensors and Actuators B: Chemical classification [24], [25], amongst many other similar time series analysis problems as discussed before. Resampling and windowing are a routine part of biomedical data (ECG/EEG) analysis like heart-beat segmentation and event related potential (ERP) epoch extraction from EEG and finally using them in classification. The field of plant electrophysiology based stimulus classification is new yet growing field and as such there is no benchmark exists for routine data analysis. We here propose data windowing and statistical feature extraction, as some standard pre-processing techniques to detect the stimulus to the plant. IV. Feature Extraction and Significance In total, 15 features were extracted from each block containing 1024 samples. Out of these 15 features, 11 features were previously reported in [2] which includes basic measures of descriptive statistics like mean (µ), variance (σ2), skewness (γ), kurtosis (β), Interquartile range (IQR). It also includes features capturing nonlinear and non-stationary behaviour [27], e.g. Hjorth mobility, Hjorth complexity, detrended fluctuation analysis (DFA), Hurst exponent, wavelet packet entropy and also frequency domain features like the average spectral power etc. Apart from these 11, four additional features have been considered in the present paper. Two out of these four are higher standardized moments capturing the non-Gaussian nature of the signal given in (1). NS = 5,6 = ( ( ) x µ σ − ) N (1) 6N = in (1) which is described as hyperskewness when respectively [28]. The third additional feature is the Fano factor or the index of dispersion which also characterizes the shape of the underlying probability described as 5N = and hyperflatness when in 2 ) F σ µ = ( density function for the data generation process [29]. The fourth feature considered is the correlation dimension (D) which takes a fractional value for chaotic trajectories while taking integer values for regular objects like line or surface [30]. For large number of samples in a time series data, the correlation dimension (D) converges to the Hausdorff or fractal dimension (Df). From a measured signal using the Takens' time delay embedding, the correlation sum is first computed and then inspected for self-similarity in a log-log plot and the slope of this graph gives the correlation dimension (D) [30]. These 15 features, extracted from blocks of 1024 samples, were normalized to lie within a minimum {0} and maximum {1} value by using (2).  x = ( x − x min ) ( x max − x min ) (2) where, x is the feature value, minx and maxx are the minimum and maximum values of the feature vector respectively. 13 Sensors and Actuators B: Chemical The same 15 features were also extracted from raw signals (background information subtracted) and normalized, so that we could compare the results obtained using both filtered and raw signals. The univariate histogram plots of the features, extracted from the filtered and raw data, are shown in Figure 5 and Figure 6 respectively to explore the class separation and overlaps using individual features. From these figures, we note that in most cases, the individual feature does not allow a straightforward separation of the classes. Figure 5: Histograms for 15 features (computed from filtered data), showing separation of different classes. 14 Sensors and Actuators B: Chemical Figure 6: Histograms for 15 features (computed from raw data), showing separation of different classes. It is worth mentioning that here the purpose of the study is not interpretation of these statistical features with changing biological behaviour of the plant under chemical stress. Rather the goal here is whether we can reliably detect the chemical stimuli by only observing these features. The classification results shown in the next sections indicate that the individual features alone may not be very useful to detect these changes in the underlying physiological characteristics of the plant under chemical stimulation and act as a clear biomarker as revealed by many overlapping univariate histograms in Figure 5 and Figure 6. But a particular combination of these features may indeed provide useful results to detect these external stimuli. V. Classification Methodology We have used four discriminant analysis classifier variants – LDA, QDA, diaglinear, diagquadratic and the Mahalonobis distance classification algorithms as also previously explored in [2], but in two different decision tree settings. The diaglinear and diagquadratic classifiers, which are also known as Naïve Bayes classifier, uses a simple linear and quadratic kernel along with only diagonal estimates of the covariance matrix (thereby neglecting any cross-terms or feature correlations). The Mahalanobis distance classifier modifies the distance measure instead of the standard Euclidean version [15], [16]. We have restricted the study using only these five classifier variants due to their low computational complexity on large datasets, compared to nonlinear kernel based classifiers like the support vector machines. This will assist in our final goal of implementing the best classification algorithm on an electronic embedded system within a resource constrained environment that can be deployed in real field. A. Architecture of the Decision Tree We explored two different types of decision tree architectures – OVR and OVO for the present multi-class classification problem involving three chemical stimuli O3, NaCl and H2SO4. These are schematically shown in Figure 7 and Figure 8 respectively. In both the configurations, we explored the best feature-classifier combination which would produce the best accuracy in terms of both cross-validation and independent testing. Thus we have reduced a multiclass classification problem to one or more binary classification problems [31], [32] exploring two distinct decision tree methods, as described in the following sections using OVR [33] and OVO [34] schemes. B. One vs. Rest (OVR) Scheme In the OVR scheme [35], [36], the three classes (stimuli) were classified in a two-node set up. In the first node, the best separable class (out of the three) was found out (depicted as Class A in Figure 7) along with the best features-classifier combination. In the second node, the 15 Sensors and Actuators B: Chemical remaining two classes were classified. In our case, after the retrospective study, it was possible to associate the three chemical stimuli with the classes A, B and C, as shown in Figure 7. In order to find out the best separable class out of the three classes, three binary classification settings were setup for the first node. In each binary setting, one of the stimuli was considered as one class and the other two were clubbed as the rest. For each of these settings, we could rank features using Fisher Discriminant Ratio (FDR) [15], [16] given by (3). FDR µ µ σ σ 2 = − 1 2 1 + ( 2 ) 2 ( 2 ) (3) 2µ and 2σ are mean where 1µ and 1σ is mean and standard deviation of features of one class, and standard deviation of features of the rest. This mean and standard deviation in (3) correspond to each feature for each class and therefore should not be confused with the signal means (raw or filtered), described before as potential features. A systematic feature selection method, employing Sequential Forward Search (SFS) algorithm [15], [16] on the FDR based ranked features, for each of the three binary classification settings (i.e. Class A vs. rest, Class B vs. rest and Class C vs. rest) was carried out using the five classifier variants. Thus the best feature-classifier combination was found out for the best separable class. The best feature- classifier combination for the remaining two classes were also found out using a similar approach (i.e. ranking the features using FDR, then employing SFS along with five different classifiers). Figure 7: Decision tree incorporating OVR configuration for multi-class classification. C. One vs. One (OVO) Scheme In the case of OVO configuration, three binary classification settings were simultaneously carried out (as shown in Figure 8). If two classifiers affirm the presence of a particular class, then only that particular class was predicted [37]. In the case of contradictory decisions by two 16 Sensors and Actuators B: Chemical classifiers, such an assignment was discarded. For each binary classification settings, the best feature-classifier combination was found out. The feature ranking for each of the three settings were again recalculated using FDR, as explained earlier for different pairs of classes. SFS was employed again on these FDR based ranked features for all the five classifier variants to find out the best feature-classifier combination for each of the three binary classifiers shown in Figure 8. Figure 8: Decision tree incorporating OVO configuration for multi-class classification. D. Retrospective Study (Using LOOCV) The results for identifying each class in a binary classification are obtained in terms of sensitivity and specificity given by the confusion matrix in Table 3 [15], [16]. Table 3: Confusion matrix Predicted class Positive (p) Negative (n) Measures For cases, where P N≈ of accuracy in (4). Positive (P) True Positive (TP) False Negative (FN) Actual class Negative (N) False Positive (FP) True Negative (TN) Sensitivity=TP/(TP+FN) Specificity=TN/(TN+FP) , accuracy of the classification is obtained using the traditional notion Accuracy = ( TP TN + ) ( P N + ) (4) N or vice versa), the But in the case of unbalanced data [38] in the two classes (i.e. P balanced accuracy [39]–[42] is used to determine the accuracy of the classification as given by (5). The derivation of (5) is given in [38] for a single run of the classifier and for a fixed threshold in the classifier's class assignment. Accuracy balanced ( ) = ( sensitivity + specificity ) 2 (5) 17 Sensors and Actuators B: Chemical Since we have an unbalanced dataset due to different duration of exposure of the plants to different stimuli, the balanced accuracy has been used here which henceforth will be referred as accuracy, in the remainder of this paper. Identifying the difficult class within an unbalanced data-set has been the focus of several recent works like [43], as also attempted here. We thus explored the best combination of features and classifiers to get the optimum classification results within the decision tree framework in the following way: • The features for every binary classification setting (OVR and OVO) were ranked using FDR as given by (3). Depending on the OVR/OVO scheme, the feature ranking may get changed even though the feature descriptions are constant. • Using the ranked features in each binary classification setting, a SFS algorithm was employed with all the five classifiers to find the best result. The ranked features for each binary classification settings are provided in the supplementary material, for brevity. E. Visualization of Class Separability on the LDA Basis The LDA basis (Fisher-faces) [15], [16] vectors are found by solving the generalized eigenvalue problem as given by (6). 1 − S w S W Wλ b = (6) The terms bS and WS are called between-class and within-class scatter matrix respectively for the original n-dimensional feature vector extracted from the plant signals. Figure 9: LDA basis showing separation of three stimuli using filtered signal. 18 Sensors and Actuators B: Chemical The eigenvectors corresponding to the largest eigenvalue gives the optimal projection where the variance between the features within a class is maximized. The three dimensions of the LDA basis are obtained from eigenvectors corresponding to the three largest eigenvalues. The separability of the three stimuli on 3-dimensional LDA basis using filtered and raw signals are shown in Figure 9 and Figure 10 respectively. The data shown in Figure 9 and Figure 10 are those which are used for training (~73% of total data) the classifiers during LOOCV retrospective study. Figure 10: LDA basis showing separation of three stimuli using raw signal. VI. Results and Discussions A. Retrospective Study Using Filtered Plant Signals We here report the results obtained using the filtered plant signals. The classification accuracy obtained using the SFS algorithm on the ranked features and using all the five variants of classifiers are shown in Figure 11 (OVR) and Figure 12 (OVO), from which the classifier and feature combination giving the best accuracy could be easily identified (as the maxima of the curves) for each binary classification scenario. These results are obtained by carrying out LOOCV on the ~73% training dataset for the retrospective study as explained earlier. In Table 4, along with the best results in terms of percentage of accuracy, the top features and classifiers have also been reported. As can be seen from Table 4, the results are obtained for different binary classification scenario. For every such scenario, the features have been ranked using FDR given in (3), and hence not all the cases will have the same features at the same ranking. 19 Sensors and Actuators B: Chemical From Table 4 we can observe that 100% classification is achieved between NaCl and H2SO4 using top 6 features and QDA/Mahalanobis classifier. This result is quite significant since both the stimuli are administered through the soil (and hence the uptake is through the roots). A similar result is obtained for NaCl and O3 although using top 15 features. Here O3 is administered through a spray all over the plant, so the absorption is mainly through the leaves. The requirement of 15 dimensional feature space to distinguish between the two stimuli probably means the effect of how the stimuli is applied to the plant is irrelevant, i.e. whether through the root or leaf. However, perhaps it also shows that the electrophysiological effect in the plant due to NaCl is different to a great extent than the other two stimuli as it produces a 100% classification result. Figure 11: Accuracy vs. increment in features (SFS) for OVR setting (using filtered signal). Table 4: Best classification accuracy for the filtered signal (features + classifier combinations) Scheme Stimuli H2SO4 NaCl O3 OVO NaCl O3 OVR Rest 100% (top 6 features), QDA / Mahalanobis 76.53% (top 2 features), Mahalanobis 76.53% (top 2 features), Mahalanobis - * 100% (top 15 features), QDA / Mahalanobis 100% (top 14 features), QDA / Mahalanobis - 76.34% (top 15 features), Mahalanobis 20 Sensors and Actuators B: Chemical The best result between O3 and H2SO4 is approximately 76% which is not as good as that obtained using NaCl stimulus. This observation possibly points to the fact that the electrophysiological effect due to O3 is not as separable from the effect of H2SO4, especially using low dimensional discriminant analysis classifiers. Figure 12: Accuracy vs. increment in features (SFS) for OVO setting (using filtered signal). B. Constructing the Decision Tree for Multiple Class Using Binary Classification Results of the Filtered Plant Signals The decision tree was constructed using two configurations – OVR and OVO, as explained in earlier sections. From Table 4, we can see that NaCl vs. Rest gives the best classification result of 100%, using top 14 features (as ranked using FDR) along with QDA/Mahalanobis distance classifier. So in the first node of Figure 7, we set class A = NaCl as the best separable stimuli from the rest. Thus the decision tree will first test if an incoming feature vector belongs to NaCl or not. If it is tested to be not NaCl, then the remaining binary classification to be evaluated in the second node is between O3 vs. H2SO4 (= class B and class C respectively in Figure 7). As seen from Table 4, the best result achieved for classification between O3 and H2SO4 is 76.53% using top two features along with the Mahalanobis distance classifier. Thus, following the retrospective study we found out that the decision tree in OVR configuration will require top 14 features (NaCl vs. Rest) along with QDA/Mahalanobis classifier in the first node to test an incoming feature vector to be classified as NaCl. If it is 21 Sensors and Actuators B: Chemical found to be not from NaCl, then in the second node, the feature vector is tested for O3 or H2SO4 using top two features and Mahalanobis classifier. In the OVO configuration, three binary classifier settings were used. As can be seen from Table 4, NaCl vs. O3 and NaCl vs. H2SO4 achieve 100% accuracy using top six and top 15 features respectively, along with either QDA or Mahalanobis distance classifier. The O3 vs. H2SO4 achieves the best result of 76.53% using top two features along with Mahalanobis distance classifier. Therefore after the retrospective study, we see that the decision tree in OVO configuration will test an incoming feature vector for three binary classification settings simultaneously. The classifier setting of NaCl vs. O3 will require top six features along with QDA/Mahalanobis classifier. The next classifier setting of NaCl vs. H2SO4 will require top 15 features along with QDA/Mahalanobis classifier. The last classifier setting of O3 vs. H2SO4 will require top two features along with Mahalanobis classifier. C. Prospective Study Using Features from the Filtered Signal One experimental dataset from each stimuli were set aside for making up the independent test dataset for the prospective study. The results, which were obtained using the retrospective study with LOOCV, were then used to design the decision tree (in OVR/OVO configuration). These two decision trees were then used to test the streams of features extracted from the signals constituting the independent test dataset. These features, as mentioned in section III, were extracted from a block of 1024 samples from the post-stimulus section of the signals belonging to the test dataset. For each of the two decision tree configurations – OVR and OVO, a decision criteria based on the value of the classifier function interpreted which stimuli the features point towards. The answers were then measured against the known class labels of the test dataset, thereby determining the accuracy of the configuration. During the prospective study, if the decision tree could not give a definite class for the input feature vector (computed from the independent test data), then the resulting class was termed as Unknown. The results for each of the three stimuli were collected according to (7). Accuracy prospective = ( c a + c b + c c ) ( n a + n b + n c ) (7) , , b c } c c c denotes the correct number of blocks belonging to each of the three classes a Here{ (stimuli) detected by the decision tree. The total number of blocks belonging to the three stimuli are denoted by{ n n n and can be referred from Table 2. From the results of retrospective a study in Table 4, we found out how many ranked features were required for the best results for classification of each stimuli within the training dataset (with LOOCV). Here it needs to be } c , b , 22 Sensors and Actuators B: Chemical noted that the feature and classifier settings achieved through retrospective study, did not yield similar good results during prospective study i.e. in OVR configuration, using top 14 features to detect NaCl in the first node and top 2 features to detect O3 / H2SO4 in the second node along with the Mahalanobis distance classifier (in both nodes) managed to produce an accuracy of only 33%. It failed to detect any blocks from H2SO4 and only one block from O3. However changing the classifier to LDA in the first node (keeping Mahalanobis in the second node) improved the accuracy to around 74% through increased detection of H2SO4 (134 out of 148 blocks) and O3 (5862 out of 9692 blocks). However from Figure 11, we clearly see that LDA in the first node produces an accuracy of less than 70% for NaCl vs. Rest. A similar poor result was obtained during prospective study when using OVO configuration. Thus we decided to check the effect of increasing features (using SFS algorithm) in each node simultaneously to see which features and classifier combination produces good results for both retrospective and prospective studies. Figure 13: Retrospective vs. prospective study results for OVR configuration, using five different classifiers and SFS. To achieve a good classification accuracy for both training and testing datasets, we compared the results of correctly detected blocks from each stimulus for both retrospective and prospective study by incrementing the ranked features using SFS. In other words, for each of the two classifiers (at two nodes) in OVR, we increased the features simultaneously. This is best understood by referring back to Figure 7, where we now know that at node 1, the binary classification to be tested is NaCl vs. rest and at node 2 it is H2SO4 vs. O3 (as established through retrospective study). So we started with best feature in both the nodes and also noted the number of correctly classified blocks from NaCl, H2SO4 and O3. Then we incremented the 23 Sensors and Actuators B: Chemical number of features by one and noted the results and it continued till all the 15 features were employed in SFS. This process was repeated for all the five variants of the discriminant analysis classifiers. Thus we now have classification results involving independent test data, for incrementing features and five different classifiers. Figure 14: Retrospective vs. prospective study results for OVO configuration, using five different classifiers and SFS. Similarly for OVO, we started with top one feature for all three classifiers (for three different binary classification) and noted the correctly classified blocks from each of the three stimuli. Then we incremented the feature by one and noted the results and it continued till all the features were employed in the SFS. Again, this process was repeated for all the five classifiers. We also had results from the retrospective study where the correctly classified block from each stimulus was noted for every increment in feature in the SFS for each of the five discriminant classifiers. From the results obtained for both retrospective and prospective study in OVR configuration as shown in Figure 13, we could clearly observe that using top three features through top seven features and Mahalanobis classifier produced the best classification results of greater than 82% in the prospective study. Similarly, using top three and above features along with Mahalanobis classifier produces a classification result of greater than 92% in the retrospective study. Among the two naïve Bayes classifiers, diagquadratic classifier performed well using top one feature providing accuracy greater than 82% in both retrospective and prospective study whereas diaglinear classifier provided a mediocre performance of around 59% during the 24 Sensors and Actuators B: Chemical prospective study. Among other classifiers, QDA provided an accuracy of greater than 82% in retrospective and 83% in prospective study by using top one feature, whereas LDA classifier produced a classification accuracy of greater than 80% and 59% in retrospective and prospective study respectively. Next, we looked at the results from OVO configuration as shown in Figure 14. The best results of classification accuracy were greater than 88% in retrospective and greater than 83% in prospective study and was obtained using top five features and Mahalanobis classifier. The LDA produced the best classification accuracy of ~86% during retrospective study and ~67% during prospective study using top eleven features. The QDA on the other hand produced a classification accuracy of ~85% and ~83% during retrospective and prospective studies respectively, using the first ranked feature. Among the naïve Bayes classifiers, diagquadratic classifier again performed well using top one feature providing accuracy of around 85% during retrospective and ~83% during prospective studies whereas diaglinear classifier provided a mediocre performance of around 60% during prospective study and somewhat acceptable performance of ~75% during retrospective study. As a summary, we can conclude that using top four features computed from the stochastic part of plant electrical response, along with Mahalanobis distance classifier, provides a best multiclass classification accuracy of ~93% in retrospective and ~89% in prospective study in an OVR setting. The detailed list of features has been provided in the supplementary material for convenience. In the next subsection, we compared these results with those obtained using features from the raw signals (with background information subtracted) as also explored in [2]. Here we used an extended feature set and more exhaustive experimental data for calculating both independent and cross-validation accuracy. D. Retrospective Study Using the Raw Plant Signals We adopted the same methodology for classification using features from raw signals as we did for the filtered plant signal. The classification accuracy obtained using the SFS algorithm on ranked features and using all the five variants of discriminant analysis classifiers are shown in Figure 15 (OVR) and in Figure 16 (OVO). From these, we identified the classifier and feature combination giving the best accuracy for each binary classification scenario. These results were obtained by carrying out LOOCV on the ~73% training dataset for the retrospective study as explained earlier. The feature ranks for each binary classification scenario is given in the supplementary material. The best results obtained in retrospective study (~73% dataset, LOOCV) in OVR/OVO setting is given in Table 5 from which we found out that Class {A = NaCl, B = H2SO4, C = O3} respectively in the OVR setting. Comparing these results with those obtained with the filtered signals, we observe that any binary classification results involving NaCl performs much better 25 Sensors and Actuators B: Chemical than other two chemical stimuli. As can be compared from Table 4 and Table 5, 100% accuracy is achieved for any binary classification involving NaCl when using filtered signals compared to ~90% achieved using raw signals. Classification result for O3 and H2SO4 using raw signal during retrospective study is around ~78% which is marginally higher than that obtained using the filtered signal. Figure 15: Accuracy vs. increment in features (SFS) for OVR setting (using raw signal). Table 5: Best classification accuracy for the raw signal (features + classifier combinations). Scheme Stimuli H2SO4 NaCl Ozone OVO NaCl O3 OVR Rest 89.43% (top 14 features), LDA 78.11% (top 13 features), QDA 78.04% (top 13 features), QDA - * 90.85% (top 15 features), LDA 90.47% (top 15 features), QDA - 79.58% (top 10 features), QDA Also in the case of filtered signals, a higher dimensional feature space was required for achieving the best classification accuracy for NaCl vs. Ozone (15 features), NaCl vs. rest (14 features) and Ozone vs. rest (15 features). Rest of the binary combinations required comparatively lower dimensional feature space. When this was compared to the classification results using raw signals, we saw that for discriminating Ozone vs. rest we required top 10 26 Sensors and Actuators B: Chemical features for achieving the best classification results. All other binary classification combinations required 13 or more features. Therefore the separability of any two stimuli in the feature space is comparatively better for filtered signal than the raw signals. Figure 16: Accuracy vs. increment in features (SFS) for OVO setting (using raw signal). E. Constructing the Decision Tree for Multiple Classes Using Binary Classification Results from the Raw Plant Signals Thus, following the retrospective study we found out that the decision tree in OVR configuration using raw signals will require top 15 features (NaCl vs. Rest) along with QDA classifier in the first node to test an incoming feature vector for belonging to NaCl class. If it is found to be not from NaCl, then in the second node, the feature vector is tested for belonging to either O3 or H2SO4 using top 13 features and QDA classifier. In OVO configuration, three binary classifier settings have been used as explained before for the filtered signals. As can be seen from Table 5, NaCl vs. O3 and NaCl vs. H2SO4 achieve ~90% and ~89% accuracy using top 15 and top 14 features respectively, along with the LDA classifier. The O3 vs. H2SO4 achieves the best result of 78.11% using top 13 features along with QDA classifier. Therefore after the retrospective study, we noted that the decision tree in OVO configuration using raw plant electrical signals will test an incoming feature vector for three binary classification settings simultaneously. The classifier setting of NaCl vs. O3 will require top 15 features along with LDA classifier. The next classifier setting of NaCl vs. H2SO4 will require top 27 Sensors and Actuators B: Chemical 14 features along with LDA classifier. The last classifier setting of O3 vs. H2SO4 will require top 13 features along with QDA classifier. F. Prospective Study Using Features from Raw Data However, when the configurations obtained from retrospective study were used during prospective study, classification accuracy was as low as 30% which was similar to as found when using filtered signals. The results of prospective and retrospective study with respect to increment in ranked features (using SFS) obtained through five different classifiers are shown in Figure 17. Figure 17: Retrospective vs. prospective study results for OVR configuration, using five different classifiers and SFS (features computed from raw signals). Here again it needs to be noted that we checked the effect of incrementing ranked features (using SFS algorithm) in order to see which features and classifier combination produced good results for both retrospective and prospective studies rather than using those classifier-feature combinations which yielded the best results only during the retrospective study. The reason behind this approach is that using different partitions of the data may result in inconsistent results between the prospective and retrospective analysis, particularly using higher number of features and more complex classifiers. However a good practice should include keeping the number of features to be low and verifying the accuracy of the classifier on the held-out data based prospective study while using the best classifier-feature settings obtained from the retrospective study with LOOCV. 28 Sensors and Actuators B: Chemical We observed that Mahalanobis classifier produced the best results for both retrospective (~87%) and prospective (~90%) study in an OVR setting. Although the prospective results were around 77% when using the naïve Bayes (diaglinear and diagquadratic) classifiers along with the top 1 feature, the retrospective results were limited to below ~70%. The LDA and QDA also perform mediocre as can be seen from Figure 17. Looking at the OVO settings as shown in Figure 18, we conclude that using the top five features computed from raw plant electrical signals, Mahalanobis classifier produces the best results for both retrospective (~92%) and prospective (~90%) study in an OVO setting. LDA, QDA and the naïve Bayes (diaglinear and diagquadratic) classifiers produce above 75% results using top one feature for both retrospective and prospective study. Increasing the features produce good classification results for retrospective study but deteriorates the performance for the prospective study. Figure 18: Retrospective vs. prospective study results (raw signal with background information subtracted) for OVO configuration, using five different classifiers and SFS. From the above exploration the following recommendations could be made: • The Mahalanobis distance classifier performs the best for both raw and filtered signals. • However when compared between raw and filtered signals, the results are better for raw signals in OVO settings using the top five features. • These top five features, as can be found in the supplementary material, are different for different binary combinations and include IQR, hyper-flatness, kurtosis, variance, DFA, hyper-skewness, wavelet entropy and average spectral power. 29 Sensors and Actuators B: Chemical the recommended the decision feature-classifier combinations and Therefore tree architecture, presented in this paper could be considered as a thorough exploration to classify three externally applied chemical stimuli using plant electrical signals. The optimal architecture balances both the cross-validated training and independent held-out testing accuracies of the classifier tree. In addition, a reduced number of features have been recommended here even though it slightly sacrifices the classification accuracy in order to facilitate electronic implementation of the classification algorithm in a resource constrained environment in a future study [44]. Regarding the adopted data partitioning approach to test various classification methods on limited number of data-points, as such there is no consensus in the contemporary literature and many different approaches have been proposed. As a summary, Molinaro et al. [45] have provided a detailed comparison of different resampling methods for classification tasks including LOOCV, N-fold cross validation (N = 2, 5, 10), Monte Carlo cross validation, 0.632 bootstrap with and without replacement, split or hold out (1/2, 1/3 partition) methods. In this study, we first split the data in training with LOOCV (73%) and independent hold out testing (27%) which is in agreement with the recommendations of [45] and also enjoys the benefit of an acceptable bias-variance trade-off, although our method increases the computational burden due to multiple LOOCV runs in each node of the decision tree. A less computationally expensive option at the cost of less accurate solution can be using combination of N-fold cross validation and partitioning/split methods. Here we not only select the best models based on LOOCV alone but also use independent held out validation results which makes our findings even more substantial. VII. Conclusion The paper reports an exhaustive exploration of designing a decision tree classifier based on five different discriminant analysis classifier and 15 statistical features extracted from plant electrical signals. We employ two multiclass classification strategies OVO and OVR along with retrospective and prospective testing of the classifier to establish its generalization capability. It is found that amongst the three chemical stimuli - NaCl, in general is best separable compared to O3 and H2SO4. Future scope of work could be implementing the decision tree algorithm along with the statistical features in an electronic hardware platform to develop a plant based novel biosensor that the EU FP7 project PLEASED aims to develop. We also acknowledge that in this paper, we did not explicitly consider the underlying mechanism of plant physiology for understanding how different signals are generated by plants as an effect of applying different chemical stimuli. In other words, we here aim to develop a mapping in which we simply observe the electrical signals generated by the plants 30 Sensors and Actuators B: Chemical when subjected to specific chemical stimuli under controlled laboratory conditions. Considering some elements of the underlying plant electrophysiological knowledge in future works may provide deeper insights to implement perhaps more effective classification strategies. ACKNOWLEDGMENTS The work reported in this paper was supported by project PLants Employed As SEnsor Devices (PLEASED), EC grant agreement number 296582. The experimental data are available in the PLEASED website at http://pleased-fp7.eu/?page_id=253. REFERENCES [1] A. G. Volkov, Plant Electrophysiology: Theory and Methods. Springer Science & Business Media, 2007. [2] S. K. Chatterjee, S. Das, K. Maharatna, E. Masi, L. Santopolo, S. Mancuso, and A. Vitaletti, "Exploring strategies for classification of external stimuli using statistical features of the plant electrical response," Journal of The Royal Society Interface, vol. 12, no. 104, p. 20141225, 2015. [3] S. K. Chatterjee, S. Ghosh, S. Das, V. Manzella, A. Vitaletti, E. Masi, L. Santopolo, S. Mancuso, and K. Maharatna, "Forward and inverse modelling approaches for prediction of light stimulus from electrophysiological response in plants," Measurement, vol. 53, pp. 101–116, 2014. [4] L. Wang and Q. Li, "Weak electrical signals of the jasmine processed by RBF neural networks forecast," in 2010 3rd International Conference on Biomedical Engineering and Informatics, 2010, vol. 7, pp. 3095–3099. [5] Y. Liu, Z. Junmei, L. Xiaoli, K. Jiangming, and Y. Kai, "The Research of Plants' Water Stress Acoustic Emission Signal Processing Methods," in 2011 Third International Conference on Measuring Technology and Mechatronics Automation, 2011, vol. 3, pp. 922–925. [6] E. F. Cabral, P. C. Pecora, A. I. C. Arce, A. R. B. Tech, and E. J. X. Costa, "The oscillatory bioelectrical signal from plants explained by a simulated electrical model and tested using Lempel-Ziv complexity," Computers and electronics in agriculture, vol. 76, no. 1, pp. 1–5, 2011. [7] J. Lu and W. Ding, "The feature extraction of plant electrical signal based on wavelet packet and neural network," in Automatic Control and Artificial Intelligence (ACAI 2012), International Conference on, 2012, pp. 2119–2122. [8] L. Jingxia and D. Weimin, "Study and evaluation of plant electrical signal processing method," in Image and Signal Processing (CISP), 2011 4th International Congress on, 2011, vol. 5, pp. 2788–2791. [9] L. Wang and L. LI HX, "Studies on the plant electric wave signal by the wavelet analysis," in Journal of Physics: Conference Series, 4th International Symposium on Instrumentation Science and Technology (ISIST'06). Bristol & Philadelphia: IOP 31 Sensors and Actuators B: Chemical Publishing, 2006, pp. 520–525. [10] L. Tian, Q. Meng, L. Wang, J. Dong, and H. Wu, "Research on the effect of electrical signals on growth of Sansevieria under Light-Emitting Diode (LED) lighting environment," PloS one, vol. 10, no. 6, p. e0131838, 2015. [11] L. Wang and J. Ding, "Processing on information fusion of weak electrical signals in plants," in 2010 Third International Conference on Information and Computing, 2010, vol. 2, pp. 21–24. [12] L. Huang, Z.-Y. Wang, L.-L. Zhao, D. Zhao, C. Wang, Z.-L. Xu, R.-F. Hou, and X.-J. Qiao, "Electrical signal measurement in plants using blind source separation with independent component analysis," Computers and Electronics in Agriculture, vol. 71, pp. S54–S59, 2010. [13] X. Zhang, N. Yu, G. Xi, and X. Meng, "Changes in the power spectrum of electrical signals in maize leaf induced by osmotic stress," Chinese Science Bulletin, vol. 57, no. 4, pp. 413–420, 2012. [14] Y. Wu, T. Guo, and J. Jiang, "Application of Blind Sources Separation in plant leaves classification," in Intelligent Control and Automation (WCICA), 2012 10th World Congress on, 2012, pp. 4174–4179. [15] C. M. Bishop, Pattern Recognition and Machine Learning. Springer, New York, 2007. [16] S. Theodoridis, A. Pikrakis, K. Koutroumbas, and D. Cavouras, Introduction to pattern recognition: a matlab approach. Academic Press, 2010. [17] S. Das, B. J. Ajiwibawa, S. K. Chatterjee, S. Ghosh, K. Maharatna, S. Dasmahapatra, A. Vitaletti, E. Masi, and S. Mancuso, "Drift removal in plant electrical signals via IIR filtering using wavelet energy," Computers and Electronics in Agriculture, vol. 118, pp. 15–23, 2015. [18] N. Uniyal, H. Eskandari, P. Abolmaesumi, S. Sojoudi, P. Gordon, L. Warren, R. N. Rohling, S. E. Salcudean, and M. Moradi, "Ultrasound RF time series for classification of breast lesions," IEEE transactions on medical imaging, vol. 34, no. 2, pp. 652–661, 2015. [19] R. Huerta, S. Vembu, M. K. Muezzinoglu, and A. Vergara, "Dynamical svm for time series classification," in Joint DAGM (German Association for Pattern Recognition) and OAGM Symposium, 2012, pp. 216–225. [20] S. Vembu, A. Vergara, M. K. Muezzinoglu, and R. Huerta, "On time series features and kernels for machine olfaction," Sensors and Actuators B: Chemical, vol. 174, pp. 535– 546, 2012. [21] S. Das, I. Pan, and S. Das, "Effect of random parameter switching on commensurate fractional order chaotic systems," Chaos, Solitons & Fractals, vol. 91, pp. 157–173, 2016. [22] P. De Chazal, M. O'Dwyer, and R. B. Reilly, "Automatic classification of heartbeats using ECG morphology and heartbeat interval features," IEEE Transactions on Biomedical Engineering, vol. 51, no. 7, pp. 1196–1206, 2004. 32 Sensors and Actuators B: Chemical [23] C. Ye, B. V. Kumar, and M. T. Coimbra, "Heartbeat classification using morphological and dynamic features of ECG signals," IEEE Transactions on Biomedical Engineering, vol. 59, no. 10, pp. 2930–2941, 2012. [24] L. Fraiwan, K. Lweesy, N. Khasawneh, H. Wenz, and H. Dickhaus, "Automated sleep stage identification system based on time-frequency analysis of a single EEG channel and random forest classifier," Computer methods and programs in biomedicine, vol. 108, no. 1, pp. 10–19, 2012. [25] M. Sinn, K. Keller, and B. Chen, "Segmentation and classification of time series using ordinal pattern distributions," The European Physical Journal Special Topics, vol. 222, no. 2, pp. 587–598, 2013. [26] D. G. Manolakis, V. K. Ingle, and S. M. Kogon, Statistical and adaptive signal processing: spectral estimation, signal modeling, adaptive filtering, and array processing, vol. 46. Artech House Norwood, 2005. [27] D. Kugiumtzis and A. Tsimpiris, "Measures of Analysis of Time Series (MATS): A MATLAB Toolkit for Computation of Multiple Measures on Time Series Data Bases," Journal of Statistical Software, vol. 33, no. i05, 2010. [28] B. Lindgren, A. V. Johansson, and Y. Tsuji, "Universality of probability density distributions in the overlap region in high Reynolds number turbulent boundary layers," Physics of Fluids, vol. 16, no. 7, pp. 2587–2591, 2004. [29] M. C. Teich, "Fractal character of the auditory neural spike train," IEEE Transactions on Biomedical Engineering, vol. 36, no. 1, pp. 150–160, 1989. [30] K. J. Blinowska and J. Zygierewicz, Practical Biomedical Signal Analysis Using MATLAB. CRC Press, 2011. [31] A. FernáNdez, V. LóPez, M. Galar, M. J. Del Jesus, and F. Herrera, "Analysing the classification of imbalanced data-sets with multiple classes: Binarization techniques and ad-hoc approaches," Knowledge-based systems, vol. 42, pp. 97–110, 2013. [32] E. L. Allwein, R. E. Schapire, and Y. Singer, "Reducing multiclass to binary: A unifying approach for margin classifiers," Journal of machine learning research, vol. 1, no. Dec, pp. 113–141, 2000. [33] S. Hashemi, Y. Yang, Z. Mirzamomen, and M. Kangavari, "Adapted one-versus-all decision trees for data stream classification," IEEE Transactions on Knowledge and Data Engineering, vol. 21, no. 5, pp. 624–637, 2009. [34] R. Debnath, N. Takahide, and H. Takahashi, "A decision based one-against-one method for multi-class support vector machine," Pattern Analysis and Applications, vol. 7, no. 2, pp. 164–175, 2004. [35] A. Rocha and S. K. Goldenstein, "Multiclass from binary: Expanding one-versus-all, one-versus-one and ECOC-based approaches," IEEE Transactions on Neural Networks and Learning Systems, vol. 25, no. 2, pp. 289–302, 2014. [36] M. Galar, A. Fernández, E. Barrenechea, H. Bustince, and F. Herrera, "An overview of ensemble methods for binary classifiers in multi-class problems: Experimental study on one-vs-one and one-vs-all schemes," Pattern Recognition, vol. 44, no. 8, pp. 1761– 33 Sensors and Actuators B: Chemical 1776, 2011. [37] J. A. SáEz, M. Galar, J. Luengo, and F. Herrera, "Tackling the problem of classification with noisy data using Multiple Classifier Systems: Analysis of the performance and robustness," Information Sciences, vol. 247, pp. 1–20, 2013. [38] M. Galar, A. Fernandez, E. Barrenechea, H. Bustince, and F. Herrera, "A review on ensembles for the class imbalance problem: bagging-, boosting-, and hybrid-based approaches," IEEE Transactions on Systems, Man, and Cybernetics, Part C (Applications and Reviews), vol. 42, no. 4, pp. 463–484, 2012. [39] V. Garcia, R. A. Mollineda, and J. S. Sánchez, "Index of balanced accuracy: A performance measure for skewed class distributions," in Iberian Conference on Pattern Recognition and Image Analysis, 2009, pp. 441–448. [40] M. Sokolova, N. Japkowicz, and S. Szpakowicz, "Beyond accuracy, F-score and ROC: a family of discriminant measures for performance evaluation," in Australasian Joint Conference on Artificial Intelligence, 2006, pp. 1015–1021. [41] C. X. Ling, J. Huang, and H. Zhang, "AUC: a better measure than accuracy in comparing learning algorithms," in Conference of the Canadian Society for Computational Studies of Intelligence, 2003, pp. 329–341. [42] A. Maratea, A. Petrosino, and M. Manzo, "Adjusted F-measure and kernel scaling for imbalanced data learning," Information Sciences, vol. 257, pp. 331–341, 2014. [43] M. Galar, A. Fernández, E. Barrenechea, and F. Herrera, "Empowering difficult classes with a similarity-based aggregation in multi-class classification problems," Information Sciences, vol. 264, pp. 135–157, 2014. [44] S. Das and K. Maharatna, "Machine Learning Techniques for Remote Healthcare," in Systems Design for Remote Healthcare, Springer, 2014, pp. 129–172. [45] A. M. Molinaro, R. Simon, and R. M. Pfeiffer, "Prediction error estimation: a comparison of resampling methods," Bioinformatics, vol. 21, no. 15, pp. 3301–3307, 2005. 34
1710.02346
1
1710
2017-10-06T10:53:39
Particle transport across a channel via an oscillating potential
[ "physics.bio-ph", "cond-mat.soft", "q-bio.SC" ]
Membrane protein transporters alternate their substrate-binding sites between the extracellular and cytosolic side of the membrane according to the alternating access mechanism. Inspired by this intriguing mechanism devised by nature, we study particle transport through a channel coupled with an energy well that oscillates its position between the two entrances of the channel. We optimize particle transport across the channel by adjusting the oscillation frequency. At the optimal oscillation frequency, the translocation rate through the channel is a hundred times higher with respect to free diffusion across the channel. Our findings reveal the effect of time dependent potentials on particle transport across a channel and will be relevant for membrane transport and microfluidics application.
physics.bio-ph
physics
Particle transport across a channel via an oscillating potential Yizhou Tan, Jannes Gladrow, and Ulrich F. Keyser Cavendish laboratory, University of Cambridge, Cambridge CB3 0HE, United Kingdom Departamento de Fisica, Universidad Autonoma Metropolitana-Iztapalapa, 09340 Mexico City, Mexico Leonardo Dagdug Stefano Pagliara∗ Living Systems Institute, University of Exeter, Exeter EX4 4QD, United Kingdom and Cavendish laboratory, University of Cambridge, Cambridge CB3 0HE, United Kingdom (Dated: October 4, 2018) Membrane protein transporters alternate their substrate-binding sites between the extracellular and cytosolic side of the membrane according to the alternating access mechanism. Inspired by this intriguing mechanism devised by nature, we study particle transport through a channel coupled with an energy well that oscillates its position between the two entrances of the channel. We optimize particle transport across the channel by adjusting the oscillation frequency. At the optimal oscilla- tion frequency, the translocation rate through the channel is a hundred times higher with respect to free diffusion across the channel. Our findings reveal the effect of time dependent potentials on particle transport across a channel and will be relevant for membrane transport and microfluidics application. I. INTRODUCTION Transport proteins are ubiquitously expressed in all kingdoms of life and allow for the continuous exchange of ions and nutrients across cell membranes [1]. A fea- ture common to all transporters is their capability to bind their substrate. The number, position and strength of the substrate-binding sites can be optimized in or- der to maximize substrate exchange across the cell mem- brane [2]. The physical mechanisms underlying transport optimization have been extensively investigated experi- mentally [3 -- 9], by molecular dynamics simulations [10], and independently rationalized by a continuum diffu- sion model based on the Smoluchowski equation [11], a discrete stochastic model [12] and a general kinetic model [13]. However, these studies do not take into account a fun- damental hallmark shared by several transporters that is their capability to alternate their substrate-binding sites between the extracellular and cytosolic side of the mem- brane according to the alternating access mechanism pro- posed fifty years ago [14]. A simplified alternated particle transport mechanism can be achieved by modulating the energy landscape in which particles diffuse. Indeed, opti- cal potentials modulated in time have been employed to study particle diffusion [15, 16], to induce thermal ratch- ets [17, 18] to direct [19] and sort Brownian particles [20 -- 24], to study particle escape and synchronization [25], to investigate stochastic resonance and resonant activa- tion [26, 27]. However, to the best of our knowledge, the effect of oscillating potentials on particle transport across one-dimensional (1D) channels remains to be in- vestigated. ∗ [email protected] In this letter, inspired by the naturally occurring al- ternating access mechanism, we use our previously in- troduced experimental model system [7, 8, 28] to couple a modulated potential in a quasi 1D microfluidic chan- nel. Specifically, we use holographic optical tweezers (HOTs) [29, 30] to create an optical potential that oscil- lates in time between the two entrances of the channel. We find that (i) there is an optimal oscillation frequency that maximizes the particle transport rate through the channel; (ii) at this oscillation frequency, the particle transport rate is two orders of magnitude larger with re- spect to free diffusion; (iii) channel occupancy increases with oscillation frequency and (iv) the optimal oscillation frequency is the one that resonates with particle diffusion across the region between the centres of the energy well positions. II. EXPERIMENTAL METHODS Our microfluidic devices are fabricated as previously reported [31, 32]. They consist of two 3D reservoirs with a depth of 12 µm separated by a polydimethylsiloxane barrier and connected by an array of microfluidic chan- nels. Each channel has a cross section of around 0.9×0.9 µm2 and a length of 2L = 4.8 µm. The reservoirs are filled with spherical polystyrene particles of diame- ter (510±10) nm. We use a laser line trap generated by HOTs to create an attractive potential well that extends from the centre of the channel to 1.7 µm in the left reser- voir (Fig. 1(a), (c) and dotted line in Fig. 1(d)). After a time interval TΩ, we switch off this laser line and simul- taneously switch on a second laser line trap that extends, for a time interval TΩ, from the centre of the channel to 1.7 µm in the right reservoir (solid line in Fig. 1(d)). In this way, we produce an attractive potential that oscil- 7 1 0 2 t c O 6 ] h p - o i b . s c i s y h p [ 1 v 6 4 3 2 0 . 0 1 7 1 : v i X r a 2 FIG. 1. (a) Schematics illustrating the oscillation of the po- sition of a laser line trap between the left and right hand side of a channel at a frequency f. (b) Bright-field image of a 510 nm polystyrene particle diffusing in a microfluidic channel. (c) Corresponding dark-field image showing the intensity dis- tribution of a laser line trap positioned at the left entrance of the microfluidic channel. The dashed lines highlight the chan- nel contours. (d) Oscillating energy potential created when the laser trap is positioned at the left (dotted line) and at the right channel entrance (solid line). The potential extension, depth and shape are estimated from the intensity distribu- tion of the laser traps. The vertical lines indicate the channel entrances. (e) Schematics illustrating particle return (e) and translocation (f) events. lates at frequency f = (2TΩ)−1 between the two channel entrances (Fig. 1(a,d) and Video 1). We estimate the extension, depth and shape of the energy wells from the intensity distribution of the line traps generated by HOTs (Fig. 1(d)) [33, 34]. The deduced potentials are validated by measuring the velocity of a particle dragged through the channels by moving the sample stage at a constant speed [35]. Experiments are performed over a range of os- cillation frequencies and particle concentrations c in the reservoirs. We record videos of particles undergoing Brownian mo- tion in the channels and reservoirs and extract particle trajectories [36, 37]. We define an attempt as the event for which a particle enters into the channel from either reservoirs and explores it for at least 33 ms, one frame time of the CCD camera that we use [7]. Once a parti- cle has entered the channel, it can either go back to the same reservoir, defined as a return event (Fig. 1(e)), or translocate through the channel and exit to the opposite reservoir, defined as a translocation event (Fig. 1(f)). We determine the attempt rate JA, the translocation rate JT and the translocation probability PT , defined as JT /JA. Average rate values for each oscillation frequency are ob- tained from at least five experiments of one hour duration each. In order to collect statistically sufficient samples for the translocation time, we use HOTs for automated drag-and-release experiments in which we trap a single particle in one of the two reservoirs and place it in one of the two channel entrances. At t = 0 s, we release the par- ticle and simultaneously switch on the oscillating optical FIG. 2. Dependence of (a) attempt rate JA, (b) translo- cation probability PT and (c) translocation rate JT on the potential oscillation frequency with a particle concentration of 0.07 (circles), 0.22 (triangles) and 1.01 nM (squares) in the reservoirs. Lines are two-term exponential fitting of the data and allow identifying the following peak frequencies: 0.05, 0.10 and 0.15 Hz for the translocation probability and 0.14, 0.16 and 0.19 Hz for the translocation rate at c=0.07, 0.22 and 1.01 nM, respectively (fitting details in Appendix C). potential (sketched in Appendix A and Video 2). III. RESULTS AND DISCUSSION A. Dependence of translocation rate and probability on the frequency of the oscillating potential For c = (0.07 ± 0.01) nM, the attempt rate increases with the oscillation frequency up to (318 ± 30) particles (h−1) at a frequency of 0.5 Hz (circles in Fig. 2(a)). The translocation probability instead sharply decreases down to 0.02 for f = 0.5 Hz (circles in Fig. 2(b)). These two ef- fects cancel each other and as a consequence the translo- cation rate has a weak dependence on the oscillation fre- quency (circles in Fig. 2(c)). At higher particle concentrations, the frequency of the oscillating potential strongly affects particle transport across the channel. For c = (0.22 ± 0.06) nM, the at- tempt rate increases with frequency up to a maximum of (1733 ± 163) particles (h−1) for f = 1 Hz (triangles in Fig. 2(a) and Fig. 3(a)). The translocation probabil- ity and the translocation rate instead first increase with frequency and peak at oscillation frequencies of 0.1 Hz and 0.16 Hz, respectively (triangles in Fig. 2(b,c) and Fig. 3(b,c)). At even higher particle concentrations c = (1.01±0.07) 3 Comparison of (a) attempt rate JA, (b) translo- FIG. 3. cation probability PT and (c) translocation rate JT between the alternating access configuration (triangles), a static po- tential with two energy wells with the same extension but 42% smaller depth with respect to the oscillating potential (dotted line) and a channel without optical potential coupled (dashed line). The reduced depth avoids channel jamming (Appendix D) in the presence of a static potential. nM, the attempt rate is not significantly affected by the frequency of the oscillating potential (squares in Fig. 2(a)). The translocation probability and rate first increase with frequency peaking at an optimal oscillation frequency of 0.15 Hz and 0.19 Hz, respectively, and then decrease at higher frequencies (squares in Fig. 2(b) and (c)). We find that the attempt and translocation rate in- crease with the concentration of the particles in the reser- voirs for all tested oscillation frequencies (Fig. 2(a,c)), but interestingly, this is not the case for the translocation probability (Fig. 2(b)). Moreover, for c = 0.22 nM the translocation rate at the optimal oscillation frequency f = 0.1 Hz is 102 times higher than the one measured in free diffusion (dashed line in Fig. 3(c) and Appendix B) and twice the one measured for a static potential con- stantly switched on (dotted lines in Fig. 3(c)). For c = 1.01 nM at f = 0.1 Hz, the translocation rate (squares in Fig. 2(c)) is 65 times higher than in free diffusion and 4 times higher than in the presence of the static double well potential. FIG. 4. Dependence of the experimentaly measured prob- ability to find (a) no particle p(0), (b) one particle p(1) , (c) two particles p(2) or (d) more than two colloidal parti- cles p(≥ 3) on the oscillation potential frequency and particle concentration. Circles, triangles and squares represent data for a particle concentration of 0.07 nM, 0.22 nM and 1.01 nM, respectively. Dotted and dashed lines represent data for a static potential and particle free diffusion in the channel, respectively. simultaneously find n particles in the channel. At high particle concentrations, we measure that the probability to find one particle in the channel p(1) is at a maxi- mum for frequencies close to the optimal oscillation fre- quency (triangles and squares for c = 0.22 and 1.07 nM, respectively, in Fig. 4(b)). Notably, the channel is pre- dominantly empty at low frequencies (e.g. p(0)=0.58 for c = 0.22 nM and f = 0.05 Hz, triangles in Fig. 4(a)) whereas at high frequency and concentration the channel is crowded (Fig. 4(c,d)), e.g. p(≥ 3) = 0.94 for c = 1.07 nM and f = 1 Hz. Therefore, the optimal oscillation fre- quency is the one that allows for populating the channel without overcrowding it [7]. C. Dependence of the channel translocation time on the frequency of the oscillating potential We perform drag-and-release experiments to measure the translocation time of a particle across the channel (Video 2). This experiment is repeated at least 300 times for each f and performed at c = 0.01 nM. In free diffusion the translocation time can be calculated as previously reported [38]: B. Dependence of channel occupancy on the frequency of the oscillating potential Ttr = 2 3 L2 Dc (1) In order to gain more insight on the presence of an optimal oscillation frequency, we measure the channel occupation probability p(n), which is the probability to where L is half the length of the channel and Dc is the particle diffusion coefficient. In the presence of an external potential U (x) in the channel, the translocation 4 FIG. 6. Schematics illustrating a representative transloca- tion from left to right in the presence of the oscillating po- tential. (a) A particle enters the left entrance of the channel when the laser line is on at the left channel entrance and dif- fuses through region I to the attractive well energy minimum. (b) The left laser line is switched off and simultaneously the right one is switched on, the particle freely diffuses through region II. (c) The particle diffuses through region III into the minimum of the right attractive energy well. (d) The laser line at the right entrance is switched off and the one on the left switched on and the particle diffuses through region IV and out of the channel. reservoir. Upon entering the channel, the particle may diffuse to the minimum of the left attractive potential well, while the left laser line is switched on (Region I in Fig. 6(a)). The particle is trapped close to this position until, at TΩ = (2f )−1, the left laser line is switched off and the right line is turned on. At this time the particle is free to diffuse either towards the left or right entrance of the channel. In the most efficient scenario in terms of particle transport, the particle travels in free diffusion across region II (Fig. 6(b)) and region III where reaches the right-hand side potential minimum when the right laser line is still on (Fig. 6(c)). Finally, when this line is switched off the particle is free to diffuse through region IV out of the channel (Fig. 6(d)). We perform drag-and- release experiments to evaluate the transition time across each of the four regions above. The particle's first transi- tion time by free diffusion in channel portions of different length is plotted in Fig. 7(a). The first transition time through region II and III is TII&III = 2.8 s at the oscilla- tion frequency f = 0 Hz (Fig. 7(b), a particle is released from the HOTs at the left entrance of the channel with the right-hand side potential constantly on). TII = 1.61 s at f = 0 Hz is smaller than the corresponding Ttr = 1.95 s calculated according to Eq. 1 due to the presence of the external potential. TIII = 1.02 s at f = 0 Hz is in agreement with the value calculated according to Eq. 2 (Ttr = 1.13 s) by using the experimentally measured po- tential. TII&III is close to TΩ = 5 s indicating that the optimal oscillation frequency is the one that matches the transition time through regions II and III. Notably, for f higher than the optimal oscillation frequency, particle's FIG. 5. (a-e) Histograms reporting the normalized distri- bution of translocation times measured in drag-and-release experiments at c = 0.01 nM (red bars) and via Brownian dynamics simulations (blue bars) for different potential oscil- lation frequencies f . (f) Experimental (triangles) and sim- ulated (circles) translocation probability PT normalized to their maximum values. Experimental PT is the mean of six sets of 50 independent measurements obtained in drag-and- release experiments. time is given by [39]: Text = 1 Dc (cid:82) b a ((cid:82) x a eβU (y)dy)((cid:82) b (cid:82) b a eβU (y)dy x eβU (y)dy)e−βU (x)dx (2) where [a, b] is the transition length, β = 1/(kBT ) with kB and T denoting the Boltzmann constant and absolute temperature. We measure the distribution of translocation times for each oscillation frequency. We find resonance-like peaks with the first maximum located at 40.5, 20.5, 10.5 and 4.5 s for f = 0.025, 0.05, 0.1 and 0.33 Hz, respectively (red bins in Fig. 5(a-e) and Appendix E). Due to the oscillating nature of our potential, Eq. (1) and Eq. (2) can not fully describe our experimental data. However, we performed 1D Brownian dynamics simulations by using the experimentally measured oscillating potential (details in Appendix F) and find translocation time values that favourably compare with the experimental values (Fig. 5 (a-e)). Moreover, our Brownian dynamics simulations confirm the frequency dependence of the translocation probability with an optimal oscillation frequency of f = 0.125 Hz close to the experimentally measured one (black dots and red triangles, respectively, in Fig. 5(f)). D. Optimal oscillation frequency is defined by diffusion between potential wells By measuring the transition times across portions of the channel, we provide an intuitive explanation of the optimal oscillation frequency. Firstly, let us consider a representative translocation from the left to the right 5 FIG. 7. (a) Transition times of a freely diffusing particle through portions of the channel of different lengths. The ex- perimental values (circles) are mean and standard errors of the values obtained in 200 experiments. The theoretical val- ues (line) are calculated according to Eq. 1 by using experi- mentally measured parameters. (b) Distribution of transition times through regions II and III for f = 0 Hz, measured by performing drag and release experiments at c = 0.1 nM. The peak position is obtained by fitting the data with a Gaussian function. that allows synchronizing alternating access with particle diffusion across the region of the channel between the two oscillating energy well positions. We anticipate that our findings will stimulate further investigation on mimicking the functioning of membrane protein transporters [41], on synchronized oscillations [27, 42 -- 44] and on the use of modulated potentials for particle control in microfluidics applications [21]. transition through regions II and III is interrupted by the potential oscillation. For f lower than the optimal oscillation frequency, a particle has a higher chance to exit the channel through region I resulting in a return event, although the chance for a particle to be trans- ported through regions II and II is increased. Overall for frequencies different from the optimal frequency, particle diffusion through regions II and III does not synchronize with the time scale defined by the oscillation frequency. This explains the observed decrease in translocation rate and probability at f lower and higher than the optimal frequency (Fig. 2). Optical potentials modulated in time have been exten- sively employed to direct particle motion [15 -- 27], includ- ing in microfluidic applications [21]. However, these have yet to be implemented for enhancing particle transport across a quasi 1D microfluidic channel connecting two reservoirs. In this paper we create a modulated optical potential consisting of a laser line that alternates its po- sition between the two entrances of a microfluidic chan- nel. We optimize the oscillation frequency (Fig. 2) and explain the physical mechanism underlying the optimal oscillation frequency (Fig. 4-7). Particle transport in the presence of the modulated optical potential is two orders of magnitudes higher than in free diffusion (Fig. 3). Our experiment indicates that oscillating potentials may be an additional avenue for enhancing transport across synthetic channels or pores. In order to mimic membrane transport in living cells, we are planning to scale our synthetic platform down to the nanoscale [40] where the characteristic diffusion time is closer to the one observed in protein transporter. Furthermore, it is possi- ble to explore asymmetric systems with charged particles only in one of the two reservoirs. Thus our system will be mimicking electrochemical gradients and even exhibit ef- fects like charge polarisation under applied external driv- ing forces. Finally, an avenue with exciting phenomena will involve mixing several types of particles with dif- ferent sizes and surface charges, exploring competitive effects and thus mimicking the mechanisms of secondary active transporters, where the transport of a substrate is coupled to the transport of a second substrate. IV. SUMMARY ACKNOWLEDGMENTS We studied the effect of a time dependent potential on particle transport through a microfluidic channel. In- spired by the alternating access mechanism, we coupled an energy well that oscillates between the two entrances of a microfluidic channel. We found that particle trans- port through the channel can be maximized by optimiz- ing the oscillation potential frequency. Importantly, the optimal oscillation frequency makes the alternating ac- cess channel more efficient in terms of transport com- pared to static channels where particles are either in free diffusion or can simultaneously bind to the ends of the channel. We found that the optimal frequency is the one We thank Shayan Lameh for proofreading. This work was supported by a Royal Society Research Grant, a Wellcome Trust Strategic Seed Corn Fund and a Start up Grant from the University of Exeter awarded to S.P. U.F.K was funded by an ERC Consolidator Grant (De- signerpores 67144). Y.T. was supported by scholarship from Cavendish-NUDT, Lundgren and Pannett Fund, Churchill College. J.G. acknowledges the support of the Winton Programme for the Physics of Sustainabil- ity and the European Union's Horizon 2020 research and innovation programme under ETN grant 674979- NANOTRANS. Appendix A: Drag and release experiment FIG. 8. Schematics of the drag-and-release experiment. (a) A particle is trapped in one of the two reservoirs. (b) The par- ticle is dragged and positioned at one of the two channel en- trances. (c) The particle is released from the optical trap and simultaneously a laser line is switched on at the same chan- nel entrance. (d) After TΩ this laser line is switched off and a laser line is switched on at the opposite channel entrance. The laser line position is oscillated at a frequency f = 1/(2 ∗ TΩ) until the particle leaves the channel. This experiment is re- peated at least 300 times for each f and performed at particle concentration c = 0.01 nM in the reservoirs. 6 TABLE I. Fitting of attempt rate data c (nM) a 0.07 0.22 1.01 9853436.5 83037687.0 3013.6 b -2.1 -0.3 0.65 c -9853424.6 -83037511.7 10014.3 d -2.1 -0.3 -73.7 TABLE II. Fitting of translocation rate data c (nM) 0.07 0.22 1.01 a 26.7 77.4 137.6 b -1.9 -2.0 -1.0 c -19.0 -86.7 -187.9 d -12.7 -14.5 -17.0 TABLE III. Fitting of translocation probability data c (nM) 0.07 0.22 1.01 a 0.2 0.2 0.05 b -3.9 -3.9 -1.8 c -0.1 -0.2 -0.1 d -45.0 -23.9 -20.7 Appendix B: Transport of particles through the channel in free diffusion Appendix D: Channel jamming with static double well potential FIG. 9. Dependence of the attempt (triangles) and translo- cation rate (squares) with respect to particle concentration for particles in free diffusion through the microfluidic chan- nel described in the main text. Data and error bars are the mean and standard deviation of the values measured in five different one hour long experiments. FIG. 10. Bright field image of 510 nm polystyrene particles jamming the entrances of the channel permeated with static energy wells as deep as the potential wells used for the oscilla- tion. For this reason in the experiments reported in Fig. 3 and Fig. 4, for the static potential we used a depth 42% smaller with respect to the well depth used for the oscillating poten- tial. Appendix C: Fitting attempt and translocation rate Appendix E: Transition and translocation times Attempt rate, translocation rate and translocation probability in Fig. 2 and 3 are fitted by a two-term ex- ponential model f (x) = a ∗ exp(b ∗ x) + c ∗ exp(d ∗ x) via the nonlinear least-squares method, where a, b, c, d are the fitting parameters. The values for these parameters estimated by the fitting are reported in Tables I-III. For a particle in free diffusion in the channel the translocation time calculated according to Eq. 1 using L = 4.8/2 µm and Dc = 0.25 µm2/s [37] , is 15.36 s. This is also in agreement with the value obtained via Brownian simulation (14.99 s). 7 Appendix F: Brownian dynamics simulation 1D channel with length L. We carried out Brownian dynamics simulations in a Particle trajectories start at −L/2 and are terminated at their first contact with each of the perfect absorbing boundaries set at −L/2 and L/2. In simulations we track the fraction of particles that end at L/2 as well as their transit time defined as the time that a particle takes to reach L/2 for the first time. The actual particle's position, xn+1, is given by xn+1 = xn + xran + βDcF ∆t, where xn is the previous posi- tion, xran is a pseudo random number generated with a Gaussian distribution with average position displacement µ = 0 and standard deviation σ = 2Dc∆t, and Force F was derived from the potential depicted in Fig. 1(d). When running simulations we set ∆t = 1 × 10−4 s and we average over 100 millions random walkers. √ FIG. 11. Histograms reporting the measured distributions of translocation times for f = 0.05 Hz, 0.1 Hz and 1 Hz (from left to right) and c = 0.07 nM, 0.22 nM and 1.01 nM (from top to bottom). The distributions are fitted with Gaussian functions (solid lines) allowing for the extrapolation of harmonic peaks. The insets report the harmonic peak positions obtained by fitting Gaussian functions. [1] B. Alberts, D. Bray, J. Lewis, M. Raff, K. Roberts, J. D. Watson, and A. Grimstone, Molecular Biology of the Cell (3rd edn) (Garland Science, 1995). [2] J. J. Kasianowicz, T. L. Nguyen, and V. M. Stanford, Proceedings of the National Academy of Sciences of the United States of America 103, 11431 (2006). [3] R. Benz, A. Schmid, T. Nakae, and G. H. Vos- Scheperkeuter, Journal of bacteriology 165, 978 (1986). [4] J. M. Ward and H. Sze, Plant Physiology 99, 925 (1992). [5] J. E. W. Meyer and G. E. Schulz, Protein Science 6, 1084 (1997). R. P. A. Dullens, Physical Review E 93, 012608 (2016). [17] L. P. Faucheux, L. S. Bourdieu, P. D. Kaplan, and A. J. Libchaber, Physical Review Letters 74, 1504 (1995). [18] S.-H. Lee, K. Ladavac, M. Polin, and D. G. Grier, Phys- ical Review Letters 94, 110601 (2005). [19] S. Bleil, P. Reimann, and C. Bechinger, Physical Review E 75, 1 (2007). [20] L. Gorre-Talini, S. Jeanjean, and P. Silberzan, Physical Review E 56, 2025 (1997). [21] M. P. MacDonald, G. C. Spalding, and K. Dholakia, Nature 426, 421 (2003). [6] L. Kullman, M. Winterhalter, and S. M. Bezrukov, Bio- [22] A. Jon´as and P. Zem´anek, Electrophoresis 29, 4813 physical Journal 82, 803 (2002). (2008). [7] S. Pagliara, S. L. Dettmer, and U. F. Keyser, Physical [23] K. Xiao and D. G. Grier, Physical Review E 82, 051407 Review Letters 113, 048102 (2014). (2010). [8] S. Pagliara, C. Schwall, and U. F. Keyser, Advanced [24] K. Ladavac, K. Kasza, and D. G. Grier, Physical Review materials 25, 844 (2013). E 70, 010901 (2004). [9] A. Horner, F. Zocher, J. Preiner, N. Ollinger, C. Siligan, S. A. Akimov, and P. Pohl, Science Advances 1, 1 (2015). [10] M. Ø. Jensen, S. Park, E. Tajkhorshid, and K. Schulten, Proceedings of the National Academy of Sciences of the United States of America 99, 6731 (2002). [25] A. Simon and A. Libchaber, Physical Review Letters 68, 3375 (1992). [26] D. Babic, C. Schmitt, I. Poberaj, Europhysics Letters 67, 158 (2004). and C. Bechinger, [27] C. Schmitt, B. Dybiec, P. Hanggi, and C. Bechinger, [11] A. M. Berezhkovskii and S. M. Bezrukov, Biophysical Europhysics Letters 74, 937 (2006). Journal 88, L17 (2005). [12] A. Kolomeisky, Physical Review Letters 98, 048105 (2007). [13] A. Zilman, Biophysical Journal 96, 1235 (2009). [14] O. Jardetzky, Nature 211, 969 (1966). [15] S.-H. Lee and D. G. Grier, Physical Review Letters 96, 190601 (2006). [16] M. P. N. Juniper, A. V. Straube, D. G. A. L. Aarts, and [28] K. Misiunas, S. Pagliara, E. Lauga, J. R. Lister, and U. F. Keyser, Physical Review Letters 115, 038301 (2015). [29] M. Padgett and R. Di Leonardo, Lab on a Chip 11, 1196 (2011). [30] R. W. Bowman, G. M. Gibson, A. Linnenberger, D. B. Phillips, J. a. Grieve, D. M. Carberry, S. Serati, M. J. Miles, and M. J. Padgett, Computer Physics Communi- 8 cations 185, 268 (2014). [38] S. Redner, A guide to first-passage processes (Cambridge [31] S. Pagliara, C. Chimerel, R. Langford, D. G. a. L. Aarts, University Press, 2001). and U. F. Keyser, Lab on a chip 11, 3365 (2011). [32] S. Pagliara, S. L. Dettmer, K. Misiunas, L. Lea, Y. Tan, and U. F. Keyser, The European Physical Journal Special Topics 223, 3145 (2014). [33] M. Pelton, K. Ladavac, and D. G. Grier, Physical Review [39] A. M. Berezhkovskii, L. Dagdug, and S. M. Bezrukov, The Journal of Physical Chemistry B 121, 5455 (2017). [40] N. A. Bell, C. R. Engst, M. Ablay, G. Divitini, C. Ducati, T. Liedl, and U. F. Keyser, Nano letters 12, 512 (2011). [41] Y. Caspi, D. Zbaida, H. Cohen, and M. Elbaum, Nano E 70, 031108 (2004). Letters 8, 3728 (2008). [34] Y. Hayashi, S. Ashihara, T. Shimura, and K. Kuroda, [42] C. R. Doering and J. C. Gadoua, Physical Review Letters Optics Communications 281, 3792 (2008). 69, 2318 (1992). [35] M. P. N. Juniper, R. Besseling, D. G. a. L. Aarts, and [43] K. Hayashi, S. de Lorenzo, M. Manosas, J. M. Huguet, R. P. a. Dullens, Optics express 20, 28707 (2012). and F. Ritort, Physical Review X 2, 1 (2012). [36] D. G. G. John C. Crocker, Journal of Colloid and Inter- face Science 310, 298 (1996). [37] S. L. Dettmer, S. Pagliara, K. Misiunas, and U. F. Keyser, Physical Review E 89, 062305 (2014). [44] M. P. N. Juniper, A. V. Straube, R. Besseling, D. G. A. L. Aarts, and R. P. A. Dullens, Nature communications 6, 7187 (2015).
1003.2092
2
1003
2012-06-05T11:22:00
Modeling symbiosis by interactions through species carrying capacities
[ "physics.bio-ph", "nlin.AO", "physics.soc-ph", "q-bio.PE" ]
We introduce a mathematical model of symbiosis between different species by taking into account the influence of each species on the carrying capacities of the others. The modeled entities can pertain to biological and ecological societies or to social, economic and financial societies. Our model includes three basic types: symbiosis with direct mutual interactions, symbiosis with asymmetric interactions, and symbiosis without direct interactions. In all cases, we provide a complete classification of all admissible dynamical regimes. The proposed model of symbiosis turned out to be very rich, as it exhibits four qualitatively different regimes: convergence to stationary states, unbounded exponential growth, finite-time singularity, and finite-time death or extinction of species.
physics.bio-ph
physics
Modeling symbiosis by interactions through species carrying capacities V.I. Yukalov1,2,∗ , E.P. Yukalova1,3, and D. Sornette1,4 1Department of Management, Technology and Economics, Swiss Federal Institute of Technology, Zurich CH-8032, Switzerland 2Bogolubov Laboratory of Theoretical Physics, Joint Institute for Nuclear Research, Dubna 141980, Russia 3Laboratory of Information Technologies, Joint Institute for Nuclear Research, Dubna 141980, Russia 4Swiss Finance Institute, c/o University of Geneva, 40 blvd. Du Pont d’Arve, CH 1211 Geneva 4, Switzerland Abstract We introduce a mathematical model of symbiosis between different species by taking into account the influence of each species on the carrying capacities of the others. The modeled entities can pertain to biological and ecological societies or to social, economic and financial societies. Our model includes three basic types: symbiosis with direct mutual interactions, symbiosis with asymmetric interactions, and symbiosis without direct interactions. In all cases, we provide a complete classification of all admissible dynamical regimes. The proposed model of symbiosis turned out to be very rich, as it exhibits four qualitatively different regimes: convergence to stationary states, unbounded exponential growth, finite-time singularity, and finite-time death or extinction of species. Keywords: Mathematical models of symbiosis, Nonlinear differential equations, Dy- namics of symbiotic systems, Functional carrying capacity PACS: 02.30.Hq, 87.10.Ca, 87.10.Ed, 87.23.Cc, 87.23.Ge, 87.23.Kg, 89.65.Gh 2 1 0 2 n u J 5 ] h p - o i b . s c i s y h p [ 2 v 2 9 0 2 . 3 0 0 1 : v i X r a ∗Corresponding author: V.I. Yukalov Bogolubov Laboratory of Theoretical Physics Joint Institute for Nuclear Research, Dubna 141980, Russia E-mail: [email protected] Tel: +7(496) 21 63947, Fax: +7(496) 21 65084 1 1 Introduction The term symbiosis describes close and usually long-term interactions between different bio- logical species. Symbiotic relationships are well known in biological and ecological societies. Numerous examples can be found in the books [1–5]. Arguably, the symbiosis that is the most important to us as humans is the one between the typical human body and the mul- titudes of commensal organisms representing members of five of the six kingdoms of life. Specifically, the microbiome, with more than 2000 bacterial species and some 1014 microor- ganisms (about ten times the number of cells of the human body), is a key partner to the human immune system, and to the human metabolism as it participates in the synthesis of essential vitamins and amino acids, as well as in the degradation of otherwise indigestible plant material, and of certain drugs and pollutants in the guts [6]. Let us also mention the associations between plant roots and fungi, that are a feature of many terrestrial ecosystems, and without which the kingdom of plants would not exist as we know it. In biology, one distinguishes three main types of symbiotic relations: mutualism, com- mensalism, and parasitism. Below, we briefly define these types of symbiosis, as they are described in the books [1–5], and give some examples illustrating these relations. Mutualism is any relationship between individuals of different species where both indi- viduals derive a benefit. Generally, only lifelong interactions involving close physical and biochemical contact can properly be considered symbiotic. A large percentage of herbivores have mutualistic gut fauna that help them digest plant matter, which is more difficult to digest than animal prey. Coral reefs are the result of mutualisms between coral organisms and various types of algae that live inside them. Most land plants and land ecosystems rely on mutualisms between the plants, which fix carbon from the air, and mycorrhizal fungi, which help in extracting minerals from the ground. An example of mutual symbiosis is the relationship between the ocellaris clownfish that dwell among the tentacles of Ritteri sea anemones. The territorial fish protects the anemone from anemone-eating fish, and in turn the stinging tentacles of the anemone protect the clownfish from its predators. A special mu- cus on the clownfish protects it from the stinging tentacles. See details and more examples in Refs. [1–5]. Commensalism is a class of relationship between two organisms where one organism benefits but the other is neutral (there is no harm or benefit). An example of commensalism: cattle egrets foraging in fields among cattle or other livestock. As cattle, horses and other livestock graze on the field, they cause movements that stir up various insects. As the insects are stirred up, the cattle egrets following the livestock catch and feed upon them. The egrets benefit from this relationship because the livestock have helped them find their meals, while the livestock are typically unaffected by it. Another example of commensalism is between tigers and golden jackals. In India, lone golden jackals expelled from their pack have been known to form commensal relationships with tigers. These solitary jackals will attach themselves to a particular tiger, trailing it at a safe distance in order to feed on the big cat’s kills. Tigers have been known to tolerate these jackals though having no profit from them. One more example of commensalism: birds following army ant raids on a forest floor. As the army ant colony travels on the forest floor, they stir up various flying insect species. As the insects flee from the army ants, the birds following the ants catch the fleeing insects. In this way, the army ants and the birds are in a commensal relationship because the birds benefit while the army ants are unaffected [1–5]. Parasitism is a type of symbiotic relationship between organisms of different species where 2 one organism, the parasite, benefits at the expense of the other, the host. Traditionally parasite referred to organisms with lifestages that needed more than one host. These are now called macroparasites (typically protozoa and helminths). The word parasite now also refers to microparasites, which are typically smaller, such as viruses and bacteria, and can be directly transmitted between hosts of the same species. Parasites that live on the surface of the host are called ectoparasites (e.g. some mites). Those that live inside the host are called endoparasites (including all parasitic worms). A typical example of endoparasites are bacteria and viruses [1–5]. Generally speaking, symbiosis is not restricted to biological systems. Many relations in social and economic societies can also be interpreted as examples of symbiosis. For instance, the economic and intellectual levels of human societies can be considered as symbiotic to each other. The interconnections between basic and applied research are symbiotic as well [7–9]. Economic and financial relations between different firms or between firms and banks can be treated as symbiotic. The notion of symbiosis has a very wide applicability to various relations, whether in biology, ecology, economy, or finance. Therefore, when we shall refer to societies, we keep in mind different kinds of societies, including biological, human, economic, and others, whose interactions can be considered from the point of view of symbiotic relations. Extending the notion of symbiosis to social and economic systems, one meets the same types of symbiosis in many different forms and structures. As in biology, there can occur relations between different more or less independent parts or species, illustrated by the relations between plant roots and fungi [10], or in economics, the relations between firms and banks, or between economies and arts. Social systems are also often characterized by the existence of complicated relations, e.g., corresponding to the relations between a general system and a particular subsystem. Examples are the relations between the economy of a country and a particular branch of the economy, or between culture and language. Biological counterparts are widespread, as illustrated by the endosymbiotic origin of mitochondria and plastids of eukaryotic cells, by some marine annelid worms lacking mouth, guts and anus, and which rely on multiple extracellular bacterial endosymbionts for their excretory system [11], and more generally between the microbiome network and the mammal body [12]. Recently, the novel term industrial symbiosis has appeared to describe a certain type of eco-industrial development within the larger framework of industrial ecology. Industrial ecology is a relatively new field that is based on the ideology of nature. It claims that industrial ecosystem may behave similar to the natural ecosystem. In the same way as in nature, where symbiosis refers to an association between at least two different species, industrial symbiosis is an association between two or more industrial facilities or companies. For example, the wastes or byproducts of one company can become the raw materials for another [13–15]. Tyre shred, plastic pellets or waste steam from a factory are examples of outputs that can be sold on to other businesses. Mutual collaboration between different firms in many cases can be characterized as mutualistic symbiosis [16]. In this way, the notion of symbiosis is now applicable to biological as well as to economical societies. Interactions between different co-existing species are usually modeled by equations of the Lotka-Volterra type [17, 18]. Such equations are adequate for predator-prey relations between different species, whether one studies the simple co-existence of two species [17, 18], or more complicated cases corresponding to high-dimensional dynamical systems, such as multiply connected food webs [19, 20], or the inter-relations between different cells and pathogens inside biological organisms [21, 22]. However, the symbiotic relations are known to be principally different from the predator-prey relations [1–5]. Hence, they require another 3 mathematical representation. To be precise, we keep in mind the dynamical representation of symbiotic relations and the evolution of coexisting symbiotic species characterized by their concentrations. There are particular models describing the biological co-evolution of species by considering equations for some distributions, whether over size, or over age, or over phenotypes [23–26]. But our aim is to obtain the equations for the concentrations themselves, so that these equations could describe the variety of possible symbiotic relations. Monod [27, 28] considered the microbial growth in a chemostat, which is a liquid phase of a chemical substrate nutrient. The Monod equations have been widely used for describing fermentation processes [29, 30]. These equations can also describe the situation when two or more bacterial species compete for the same growth substrate [31–34]. When the populations of microorganisms, inhabiting a common nonliving environment, compete for nutrients, this is, of course, an example of coexistence and coevolution. This, however, is not symbiosis in its direct sense, though it may share many features of the latter. The usual end of the evolution, when two species compete for the same food substrate, is that the one attaining the higher growth rate, under the given conditions, competes more successfully and ultimately displaces the slower-growing competitor [32–34]. It is well known [1–5, 47] that competition processes are quite different from symbiosis. Therefore several bacterial species competing for the same food substrate [31–34], strictly speaking, do not illustrate symbiosis or, in the best case, represent its very particular form. The Holling [35] second type functional relation between predators and prey describes the situation when predators meet, not the amount of prey proportional to their total number, but an effective number of attacked prey that is proportional to the prey density D over 1 + cD , with a parameter c > 0. This relation takes into account that predators, in order to consume prey, need to search for it, chase, kill, eat, and digest. This is why predators attack not all prey but a limited number of them, which saturates to a constant when the prey density increases [36]. Such predator-prey relations, clearly, do not characterize symbiosis. The main aim of the present paper is to suggest a mathematical model of symbiosis characterizing the overall dynamics of this process for various systems. We concentrate in the present paper on the general mathematical properties of the considered model and on the classification of admissible types of symbiotic behavior in the frame of this model. We stress that our aim is not the reinvention of a novel qualitative classification of symbiotic relations. Such a classification is well developed and widely employed in the biological literature [1–5]. And we accept this classification, as most other authors do. Our aim is to suggest a model of symbiosis that would fit well the known classification and would provide the possibility for studying dynamical as well as mathematical consequences of different symbiosis types. The basic point that allows us to suggest a model of symbiosis is the idea that in symbiotic relations it is not the species that interact directly with each other, as in the Lotka-Volterra equation, but that symbiotic species act on the carrying capacities of each other. The carrying capacity of a biological species in an environment is generally understood as the maximum population size of the species that the environment can sustain indefinitely, given the food, habitat, water and other necessities available in the environment. In popula- tion biology, carrying capacity is defined as the environment maximal load [37]. Historically, carrying capacity has been treated as a given fixed value [38, 39]. But later it has been understood that the carrying capacity of an environment may vary for different numbers of species and may change over time due to a variety of factors, including food availability, 4 water supply, environmental conditions, living space, and, the most important, population activity. Mutual coexistence and symbiosis of several species strongly influence the carrying ca- pacities of the species, with the changes being, to a first approximation, proportional to the species numbers. For example, humans have increased the carrying capacity of the environ- ment for a few other species, including those with which we live in a mutually beneficial symbiosis. Those companion species include more than about 20 billion domestic animals such as cows, horses, pigs, sheep, goats, dogs, cats, and chickens, as well as certain plants such as wheat, rice, barley, maize, tomato, and cabbage. Clearly, humans and their se- lected companions have benefited greatly through active management of mutual carrying capacities [40]. Interactions between two or more biological species are known to essentially influence the carrying capacity of each other, by either increasing it, when species derive a mutual benefit, or decreasing it, when their interactions are antagonistic [41–43]. The same applies to economic and financial interactions between firms, which also form a kind of symbiosis, where the interacting firms develop the carrying capacity of each other also roughly proportionally to their sizes [16]. When species coexist, their carrying capacities are influenced by the species mutual interactions, either facilitating the capacity development or damaging it. Being functions of the species populations, such nonequilibrium carrying capacities can be naturally represented as polynomials over the population numbers [44]. Thus, it is now generally accepted that symbiotic organisms influence the carrying capac- ities of each other, hence the carrying capacities of symbiotic species are not fixed quantities, but should be considered as functions of population sizes [45, 46]. In the present paper, we suggest a mathematical formulation of this idea and investigate its consequences. The paper proceeds as follows. Section 2 presents the general structure of the model. In Section 3, we specify the classification of different types of symbiotic interactions, in the frame of the suggested mathematical model. Sections 4, 5, and 6 focus on the structure of the model of symbiosis in the presence of mutual, respectively, asymmetric, and in the absence of direct interactions. Section 7 concludes, explaining the principal difference of our model from the Lotka-Volterra model, and proposes a way to avoid the finite-time singu- larities arising under parasitic interactions. Also, we described several concrete examples demonstrating that the suggested mathematical models can be directly applied to different types of symbiotic relations, whether in biological symbiosis or in economic and industrial symbiosis. A discussion is given on the relation of the suggested approach to the processes describing the coexistence of microbial populations. 2 General model of symbiosis Our basic suggestion, making our approach principally new, is to describe symbiotic relations by the mutual influence of the co-existing species on their respective carrying capacities. In- deed, in the presence of symbiotic relations, the species act on the livelihood of each other by creating resources that others exploit, by using resources created by others without feed- backs, or by destroying the livelihood in the case of parasitism. In mathematical language, the livelihood is nothing but the carrying capacity. This is the principal point of the ap- proach we suggest, as compared to the predator-prey models, where the species directly predate or compete. Such direct interactions are appropriate for describing the predator- 5 prey relations, when predators eat prey. But this is not what one usually calls symbiosis. Predation and competition are commonly treated as different processes as compared to sym- biosis [47]. The standard description of symbiosis [1–5] corresponds exactly to the mutual influence of species on the carrying capacities of each other. For example, in the tree-fungi symbiosis, neither of the species eats another, but they do influence the carrying capacities of each other, producing the chemical elements helping the growth of both species. Let us consider several species, or system parts, quantified by their population number Ni , with the index i enumerating the species. For instance, in biology, Ni can be the number of individuals of the i-th species or, in finance, this can be the amount of money or, in economics, this could be the quantity of some goods. To study the mutual influence of varying carrying capacities, we start with the logistic-type equations dNi dt = γiNi − CiN 2 i Ki , (1) for the variables Ni = Ni (t) as functions of time t ≥ 0. Here γi is the birth rate of the biological species i or the growth rate in economic systems. The coefficient Ci characterizes the intensity of mutual competition between the agents of the i-th species. We will thus consider only the case where γi > 0 , Ci > 0 . (2) The principal difference from the standard logistic equation is that the carrying capacity is here considered to be a function Ki = Ai + BiSi ({N1 , N2 , . . .}) of the quantities Ni , in order to account for the co-existing symbiotic species. The first term Ai is the carrying capacity of the given surrounding livelihood. The second term characterizes the carrying capacity produced by other species. We thus capture the typical feature of symbiotic relations via the mutual influence of each species on the carrying capacities of others. The symbiotic coefficient Bi defines the intensity of producing, or destroying, the carrying capacity in the process of symbiotic relations. We refer to Bi as the production coefficient, when Bi is positive, or as the destruction coefficient, when Bi is negative. Thence, we shall consider that (3) Bi ∈ (−∞, ∞) . Since the nature of the mutual interactions is characterized by the sign of the symbiotic coefficient Bi , the symbiosis function Ai > 0 , (4) Si ({N1 , N2 , . . .}) ≥ 0 , describing the type of symbiotic relations, can be taken non-negative. Accepting that the symbiotic functions depend on the species populations, we exemplify this in what follows by the simplest form of such a dependence, assuming that the effective carrying capacity is a linear combination of the natural carrying capacity, provided by na- ture, and by the carrying capacity produced (or destroyed) by the mutual species activity. Generally, we assume that the symbiotic functions are analytical functions, hence, they can be expanded in power series over the species populations. We shall consider the following particular cases of such expansions, corresponding to different types of mutual influence of (5) 6 two symbiotic species. When the carrying capacity of an i-species is influenced by the mutual interactions with a j -species, the effective carrying capacity is represented as Ki = Ai + BiNiNj . And if the carrying capacity of the i-species is influenced by the j -species without direct interactions, as it happens in the case of commensalism, then the effective carrying capacity is given by the form Ki = Ai + BiNj . The natural carrying capacity Ai is supposed to be nonzero, which implies that the species could exist without their symbionts, though the existence of the latter can drastically change the species behavior. In general, there can happen situations when one symbiont is obligatory for another, so that one of the species cannot survive without their counterpart. Such a situation would correspond to zero carrying capacity. But this rare case is not treated here. The mathematical structure of Eq. (1) with (3), describing symbiosis, is principally different from equations of the Lotka-Volterra type and cannot be reduced to the latter, as we shall show in the Discussion section. As we mentioned in the Introduction, in biology and ecology, three main categories of symbiotic relations can be distinguished (mutualism, parasitism, and commensalism), depending on whether the influence of one species on another is positive, negative, or neutral [1–5]. This classification is straightforwardly linked to our model. Below, we explain this for the case of two species, when i = 1, 2, which will be treated in what follows. Mutualism implies the relations in which both species extract some benefit from their relationship. In our model, this is equivalent to B1 > 0 , B2 > 0 (mutualism) . Parasitism means that one of the species can benefit, while the other is harmed in the process. Generally, one can say that parasitism is the relation in which at least one of the species is harmed. This means that one of the following inequalities is valid: either or or B1 > 0 , B2 < 0 , B1 < 0 , B2 > 0 , B1 < 0 , B2 < 0 (parasitism) . Commensalism is the relation in which one of the species benefits, while another is not affected. Hence, one of the following inequalities is satisfied: either B1 > 0 , B2 = 0 , or B1 = 0 , B2 > 0 (commensalism) . In addition to this classification, it is possible to distinguish different kinds of mutual in- teractions embodied in the form of the symbiosis function Si ({Ni}) obeying inequality (5), which will be specified later. 7 Before specifying this function Si ({Ni}), it is convenient to introduce dimensionless units. For this purpose, we may choose two characteristic scales, Nef f and Zef f , to serve as the measuring units for the species variables N1 and N2 , respectively. We thus define the dimen- sionless species characteristics N1 Nef f , x ≡ N2 Zef f z ≡ and the dimensionless carrying capacities y1 ≡ γ1K1 C1Nef f , y2 ≡ γ1K2 C2Zef f . Let us measure time in units of 1/γ1 and introduce the ratio γ2 γ1 . α ≡ (6) (7) (8) If the life times of the co-existing species were essentially different, then their mutual influence would be rather limited, inducing just local in time perturbations. The most interesting case is when the species co-exist in the long time scale comparable to their life times. This implies that the most important case is when the ratio (8) is close to one, which we shall take into account in what follows. Then, for two symbiotic species, Eqs. (1) take the form dx dt = x − x2 y1 , dz dt = z − z 2 y2 . (9) These equations, controlling the time evolution of the functions x = x(t) and z = z(t), must be complemented by the initial conditions x(0) = x0 , z(0) = z0 . (10) By definition, the quantities Ni , measuring the amount of the corresponding species, are positive. Hence, we are interested only in the non-negative solutions of Eqs. (9): (11) x(t) ≥ 0 , z(t) ≥ 0 . We then need to specify the forms of the symbiosis functions Si ({Ni}) obeying inequality (5) and, respectively, the forms of the carrying capacities (3). The forms of functions (5) can vary, depending on the kind of symbiotic relations between the considered species. At the end of this section, formulating the general idea of our approach, it is worth stressing again the meaning of the chosen type of the equations. We start with Eqs. (1) having the form of logistic equations, with all terms enjoying the known straightforward underpinning. The quantity Ki here is the carrying capacity of the i−th species. There can be numerous interpretations of what in particular cases could be this capacity, depending on whether biological, or social, or financial systems are considered. For concreteness, let us talk about food. So, the carrying capacity is the amount of food available for the species. Our basic point is that the carrying capacity is not a fixed number, characterizing the amount of given food, but it consists of two parts, as in Eq. (3). The first part is the naturally provided amount of food, while the second part is the additional food produced during 8 the interaction of symbiotic species. It is possible to give an infinite number of examples illustrating this. The most classical example is the symbiosis of trees with mushrooms, or, more generally, plants with fungi. In such interactions, as is well known, the elements are produced that serve as additional food for both symbiotic species. That is, there appears additional food created by the collaboration of the species. We consider the simple case where more agents in species lead to the increase of food production during their interactions. In a reduced formalism at the macro-level, the simplest and most robust way to account for this effect is to take the symbiotically produced amount of food to be proportional to the number of agents in each of the species. This is equivalent to writing the additionally produced food for the 1−th species as B1N1N2 , where Ni is the number of agents in the i−th species group. Similarly, for the second species the additional food is B2N2N1 . From this definition, it immediately follows that the symbiotic coefficient Bi is the amount of food produced in the process of symbiotic relations by the collaboration of an agent from the first type of species with an agent from the second species type. The idea that the amount of food can be created in the process of symbiotic relations, or destroyed, when these relations are parasitic, is the pivotal new point we have advanced. This is exactly what happens in symbiotic relations. And this provides a novel mathematical approach to describing symbiotic relations. As we demonstrate in the following sections, our model allows for the description of all types of symbiosis. 3 Classification of forms of symbiotic relations There can exist three main types of symbiotic relations. 3.1 Influencing livelihood through mutual interactions A common case is when the symbiotic species influence the livelihood of each other by means of mutual interactions. Known biological examples are plant trees and mushrooms and many other plants and fungi [1–5]. In the social sciences, this could be the relationship between the economic level of a country and the intellectual level of society. Or this can be the relation between two firms, producing different kinds of goods in close collaboration with each other. Another example is the relation between basic and applied research [7–9]. The simplest form of the carrying capacities, representing livelihoods that are influenced by mutual interactions, can be written as K1 = A1 + B1N1N2 , K2 = A2 + B2N2N1 . (12) The first terms A1 and A2 define the given carrying capacities, which in dimensionless units are a1 ≡ a2 ≡ The second terms characterize the produced carrying capacities, corresponding to the mutual influence of symbiotic species on the livelihood of each other. The dimensionless symbiotic coefficients are (13) , γ1A1 C1Nef f γ1A2 C2Zef f . γ1B1Zef f C1 b ≡ γ1B2Nef f C2 . (14) , g ≡ 9 The dimensionless carrying capacities (7) become y1 = a1 + bxz , y2 = a2 + gxz . The given carrying capacities are taken positive, a1 > 0 , a2 > 0 , (15) (16) which means that each of the two species can survive in the absence of the other one. The interaction coefficients b and g can be of any sign, depending on whether the relations are beneficial or parasitic. The scaling units Nef f and Zef f can be chosen arbitrarily. It is convenient to choose them as Nef f = , Zef f = γ1A1 C1 γ1A2 C2 . Then one has and the symbiotic coefficients are a1 = a2 = 1 , γ 2 1 B1A2 C1C2 , b ≡ γ 2 1 A1B2 C1C2 . g ≡ Therefore, the carrying capacities (15) become y1 = 1 + bxz , y2 = 1 + gxz . (17) (18) (19) (20) 3.2 Influencing livelihood through asymmetric interactions Another possibility is when the interactions of symbiotic species are not symmetric, such that the carrying capacity of one of them is influenced by mutual interactions, while the carrying capacity of another is influenced solely by the other species, not involving their mutual interactions. This type of symbiosis is common to many biological as well as social systems, when one of them is a subsystem of a larger one. For example, the relation between the economic level of a country and the development level of a particular branch of economics, say between the country gross domestic product and the level of science; or the relation between culture and language; or the relation between the size of social groups and the size of brain [48, 49]. In that asymmetric case, the carrying capacities are K1 = A1 + B1N1N2 , K2 = A2 + B2N1 . (21) In dimensionless units, a1 and a2 have the same form as in Eq. (13), while the symbiotic coefficients are b ≡ g ≡ The carrying capacities (7) are given by the expressions , γ1B1Zef f C1 γ1B2Nef f Zef f . (22) (23) y1 = a1 + bxz , y2 = a2 + gx . 10 Opting for the scaling units (17) yields normalization (18), with the symbiotic coefficients b ≡ g ≡ Thus, the dimensionless carrying capacities (23) become , γ 2 1 B1A2 C1C2 γ 2 1 A1B2 C1A2 . y1 = 1 + bxz , y2 = 1 + gx . (24) (25) 3.3 Influencing livelihood without direct interactions This is, probably, the most common of biological and ecological symbiosis. Numerous exam- ples can be found in the books [1–5]. In this case, the livelihood of each species is influenced by the presence of another species without involving their direct interactions. The carrying capacities are correspondingly given by K1 = A1 + B1N2 , K2 = A2 + B2N1 . (26) The terms A1 and A2 are the a priori given carrying capacities, in the absence of the other species. Their dimensionless forms are the same as in Eq. (13), but the symbiotic coefficients are γ1B2Nef f γ1B1Zef f g ≡ b ≡ C1Nef f C2Zef f The dimensionless carrying capacities (7) take the forms , y1 = a1 + bz , y2 = a2 + gx . . (27) (28) Choosing again the same scaling units (17) gives the same normalization (18), but with the symbiotic coefficients γ1B1A2 g ≡ b ≡ A1C2 This results in the dimensionless carrying capacities , γ1A1B2 C1A2 y1 = 1 + bz , y2 = 1 + gx . . (29) (30) Note that here, as well as in the previous cases, the symbiotic coefficients b and g can take different values, depending on the strength of the mutual influence between the sym- biotic species. And these coefficients can be of different signs, describing either beneficial or parasitic relations. In the following sections, we give a detailed analysis of each type of symbiosis classified above. 4 Symbiosis with mutual interactions We study the equations (9), with the carrying capacities (20), leading to the coupled system of two ordinary differential equations: x2 z 2 dz dx = z − = x − dt 1 + bxz dt 1 + gxz This system of equations is symmetric with respect to the change x → z and b → g . Three kinds of qualitatively different solutions are found for these equations, which we describe in turn. (31) , . 11 4.1 Convergence to Stationary States Formally, there are five fixed points for Eqs. (31). However, as expressed by Eq. (11), we are looking for non-negative solutions. For each of the non-negative solutions, we accomplish the Lyapunov stability analysis and select the stable solutions. We assume that the reader is sufficiently qualified in the Lyapunov technique, so that we do not discuss the details of the stability analysis and present only the final results of the analysis for the stable stationary solutions that are non-negative. Equations (31) can possess a single stable fixed point given by the expressions 1 − b + g − p(1 + b − g )2 − 4b 1 + b − g − p(1 + b − g )2 − 4b 2g 2b This point is stable in the stability region shown in Fig. 1 characterized by one of the conditions, either x∗ = z ∗ = (32) , . or b < 0 , −∞ < g < ∞ , g ≤ gc ≡ (cid:16)√b − 1(cid:17)2 g ≤ 0 . b ≥ 1 , On the boundaries of the stability region, we have 0 ≤ b < 1 , or ≤ 1 , (33) (34) (35) x∗ = 1 , z ∗ = (b = 0, g < 1) , 1 1 − g z ∗ = 1 , x∗ = x∗ = (b < 1, g = 0) , 1 1 − b 1 1 √b 1 − √b , The solution to Eqs. (31) tends to the stationary point (32), provided that the symbiosis parameters b and g are in the stability region defined above, and when the initial conditions are in the attraction basin of this fixed point. For positive b and g , the attraction basin can be found only numerically while, if at least one of these parameters is negative, the attraction basin is characterized by one of the conditions, when either (0 < b < 1, g = gc) . z ∗ = (36) x0 z0 < 1 b (b < 0, g > 0) , (37) or when x0 z0 < or when (b > 0, g < 0) , 1 g x0z0 < min (cid:26) 1 g (cid:27) 1 b The approach to the stationary solution (32) can be either monotonic or non-monotonic, from above or from below, depending on the parameters b and g and on the initial conditions x0 and z0 . In Figs. 2, 3, and 4, the typical behavior of the solutions is shown for different (b < 0, g < 0) . (39) (38) , 12 parameters and initial conditions: when x0 > x∗ and z0 > z ∗ (Fig. 2), when x0 > x∗ but z0 < z ∗ (Fig. 3), and when x0 < x∗ with z0 < z ∗ (Fig. 4). The case when x0 < x∗ but z0 > z ∗ is similar to the case when x0 > x∗ but z0 < z ∗ , with changing x by z and b by g . It is instructive to compare the behavior of the symbiotic solutions, described by the coupled Eqs. (31), with that of the solutions of the decoupled equations (b = g = 0) dx dt dz dt = z − z 2 = x − x2 , The solutions for these non-symbiotic species tend to the stable fixed point x∗ = z ∗ = 1. The comparison, explicitly showing the role of symbiosis, is demonstrated in Figs. 5 and 6. These two figures illustrate the general property that beneficial (respectively, parasitic) symbiosis leads to the increase (respectively, decrease) of the stationary solutions. Another important message is that stationary states can exist even in the presence of parasites. (b = g = 0) . (40) 4.2 Unbounded Exponential Growth The solutions to the symbiotic equations (31) grow to infinity with increasing time t → ∞, when there are no stable fixed points, that is, when either or when 0 < b < 1 , g > gc , b > 1 , g > 0 . (41) (42) A similar exponential divergence at infinity occurs when the stable fixed points exist, but the initial conditions are taken outside of the attraction basin, so that either x0 z0 > 1 b (b < 0, g > 0) , (43) or when x0 z0 > or when (b > 0, g < 0) , 1 g g (cid:27) x0 z0 > max (cid:26) 1 1 b The typical behavior of such increasing solutions is shown in Fig. 7. When the initial condi- tions are outside of the attraction basin, this means that at least one of them is sufficiently large, that is, the initial population is so large that even the existence of parasites cannot suppress the development of this species. (b < 0, g < 0) . (44) (45) , 4.3 Finite-Time Death and Singularity A specific behavior occurs under parasitic symbiotic relations, when stable fixed points can exist, but the initial conditions are taken outside of the attraction basin, so that < x0 z0 < 1 b 1 g 13 (b < g < 0) . (46) Then, at the critical time tc , defined as the solution of the equation 1 b the first variable x(t) sharply drops to zero, while the second one rises to infinity: x(tc )z(tc ) = , x(t) → 0 , z(t) → +∞ , Here the overdot means time derivative. The behavior is inverted for the case when x(t) → −∞ (t → tc ) , z (t) → +∞ (t → tc ) . 1 1 b g for which the critical time is given as the solution of the equation (g < b < 0) , < x0 z0 < 1 g At this time, the first solution rises to infinity, while the second one quickly dies: x(tc )z(tc ) = . (47) (48) (49) (50) x(t) → +∞ (t → tc ) , x(t) → +∞ , z (t) → −∞ (t → tc ) . z(t) → 0 , The situation corresponding to the case (46) is illustrated in Fig. 8. The meaning of this phenomenon, when one of the species dies, while the other rises, is easily understandable. Under the mutually parasitic relations, each species destroys the livelihood of the other one. That species, whose livelihood is destroyed faster, dies out, while the other species grows at the expense of the former one. (51) 5 Symbiosis with asymmetric interactions If the symbiotic relations between species are characterized by asymmetric interactions with the carrying capacities (25), then Eqs. (9) take the form dx dt = x − x2 1 + bxz , dz dt = z − z 2 1 + gx . (52) The following qualitatively different dynamic regimes can occur. 5.1 Convergence to stationary states There can exist just one stable fixed point, given by 1 − b − p(1 − b)2 − 4bg 2bg x∗ = , z ∗ = 1 + b − p(1 − b)2 − 4bg 2b . (53) 14 Let us introduce the critical parameter value gc ≡ The fixed point (53) is stable and non-negative if either . (1 − b)2 4b or if or if or when b ≤ −1 , g ≥ gc , − 1 < b ≤ 0 , g > −1 , 0 < b < 1 , −1 ≤ g ≤ gc , (54) (55) (56) (57) −1 < g < 0 . b ≥ 1 , The overall region of stability is depicted in Fig. 9. The marginal values of the fixed point, occurring at the boundary of the stability region, (58) are x∗ = 1 , z ∗ = 1 + g (b = 0, g > −1) , (b < 1, g = 0) , (59) x∗ = z ∗ = 1 if g 6= gc , and 1 1 − b 2 1 − b if g = gc and either b < −1 or 0 < b < 1. The solutions tend to the fixed point, provided that the related initial conditions are in the basin of attraction. When one of the symbiotic coefficients is negative, the basin of attraction is defined so that either 1 + b 2b (g = gc) , x∗ = z ∗ = (60) , , or if or when x0 z0 < 1 b x0 < 1 g (b < 0, g > 0) , (b > 0, g < 0) , (61) (62) , x0 < 1 1 b g The convergence to the stationary state can be monotonic or non-monotonic, as illustrated in Figs. 10 and 11 for various initial conditions and different symbiotic coefficients. The convergence of solutions to the stable fixed point is qualitatively similar to that documented in the case of symmetric interactions. (b < 0, g < 0) . x0z0 < (63) 15 5.2 Exponential growth at infinity Another regime, which is similar to that found for symmetric interactions, is the exponential growth at increasing time t → ∞. This happens when there are no stable fixed points and either (64) g > gc , 0 < b < 1 , or when b > 1 , g > 0 , (65) with the critical parameter gc given in Eq. (54). The same kind of exponential growth arises when there exist stable fixed points, but the initial conditions are outside of their basin of attraction, which occurs for 0 < b < 1 , 0 < g < gc , or when 1 b The temporal behavior is qualitatively similar to that of Fig. 7. (b < 0, g > 0) . x0 z0 > (66) (67) 5.3 Finite-time singularity A different regime occurs when the parameter g is negative, and no stable fixed point exists, which occurs when or if or if 0 < b < ∞ , g < −1 , − 1 < b < 0 , g < −1 , g < gc < −1 . b ≤ −1 , When the initial conditions obey the inequalities x0 > , 1 g x0 z0 > − 1 b (b > 0, g < 0) , (68) (69) (70) (71) or the inequalities x0 > 1 1 g b then the solutions live only a finite life until the critical time tc . The value of this critical point can only be determined numerically. As time approaches tc , the first variable remains finite, while the second one diverges: (b < 0, g < 0) , x0z0 > , (72) x(t) → x(tc ) (t → tc ) , x(t) → x(tc ) > 0 , z (t) → ∞ (t → tc ) . z(t) → ∞ , Here x(tc ) and x(tc ) imply finite values of the corresponding functions at the critical time tc . The main difference with the case of Eq. (48) is the finiteness of the solution x(tc ) at the critical time. The typical behavior of the two variables in this regime is shown in Fig. (73) 16 12. The explosive divergence of one of the species announces a change of regime: beyond tc , the system of two species will change their characteristics and, possibly, the nature of their interactions. The same kind of behavior also happens when the stable fixed points exist for b and negative g in the region b > −1 , −1 < g < 0 , or if b < −1 , but initial conditions are taken outside of the basin of attraction, so that x0 , z0 are defined by Eqs. (71) or (72), correspondingly. gc < g < 0 , 5.4 Finite-time death An interesting behavior appears for some parameters, in particular, when the parameter g is negative, meaning parasitism, so that no stable fixed points exist or when fixed points exist but the initial conditions are outside their domain of attraction. For g < 0, when the symbiosis coefficients are taken outside of the stability region presented in Fig. 9, and initial conditions obey the following inequalities, either x0 < , 1 g x0 z0 > − 1 b (b > 0, g < 0) , (74) or if or if x0z0 < (b < 0, g < 0) , 1 1 b g 1 1 b g 1 1 b g then the solutions have a finite life and one of the species dies or exhibits a gradient catas- trophe at the death time td (see below) found as the solution to the equation (b < 0, g < 0) , (b < 0, g < 0) , or when x0z0 < x0 < x0z0 > , , , (75) (76) (77) x0 < x0 > g x(td) = 1 , at which either z(t) = −∞ or x(t) = −∞ for t → td . When the parameters are taken outside of the stability region and the initial conditions obey inequalities (74) or (75), then (78) x(t) → 1 g z (t) → −∞ (t → td ) . z(t) → z(td ) , When the initial conditions obey inequality (76), then x(t) → x(td ) (t → td ) , , (79) x(t) → , 1 g x(t) → +∞ (t → td ) , 17 z(t) → g z (t) → −∞ (t → td) , b while, when the initial conditions satisfy inequality (77), then , (80) , , x(t) → −∞ (t → td ) , 1 x(t) → g z(t) → g z (t) → +∞ (t → td) . b In the cases (74) and (75), the solutions (79) remain non-zero at t → td , but the time derivative of z(t) tends to −∞. Such a behavior corresponds to the so-called gradient catastrophe. The value z(td ) is close to zero, as shown in Fig. 13. This kind of behavior can be interpreted as the extinction of the species described by the variable z(t). Therefore, after td , we can set (81) (82) (t ≥ td ) . z(t) ≡ 0 Then the continuation of the evolution of the population x(t) after td is characterized by the single equation x = x − x2 . The death of the species z(t) is due to the large negative (parasitic) symbiosis coefficient g , which leads the species x to suppress the population z of the other species, until its complete demise occurring at the time td . A behavior similar to that, described by Eqs. (80) and (81), occurs for g < 0, when the symbiosis coefficients are in the stability region, but the initial conditions obey inequalities (76) or (77). When {x0 , z0} obey Eq. (76), then again there is a death time td , given by the same Eq. (78). Here z(td ) is not necessarily close to zero, but it experiences the same gradient catastrophe as in Eq. (79). Contrary to the case of Eq. (79), the derivative x(t) tends to +∞. Strictly speaking, the coupled system of Eqs. (52) has no solution after the time td . However, we can again interpret the gradient catastrophe, with z (t) → −∞ as the death of the species z(t), extending the solution x after td using condition (82), as is shown in Fig. 14. In the case when the initial conditions obey Eq. (77), the evolution also ends at the death time td , defined by the same Eq. (78). At this point, conditions (81) hold true. The difference with the case of Eqs. (80) is that here the gradient catastrophe happens with x(t) → −∞ while z (t) → +∞ for t → td . Hence, interpreting td as the time of extinction of the species x that are killed by the species z , the subsequent evolution of the dynamics after td is obtained by setting x(t) ≡ 0 The overall behavior is presented in Fig. 15. Comparing Figs. 14 and 15, we see that the dynamics essentially depends on the initial conditions. (t > td ) . (83) 6 Symbiosis without direct interactions When symbiosis is characterized by the influence of species on the livelihood of each other without direct mutual interactions, the corresponding carrying capacities are given by Eqs. (30). Then, the symbiotic Eqs. (9) take the form dx dt = x − x2 1 + bz , dz dt = z − z 2 1 + gx . (84) 18 These equations are symmetric with respect to the replacement b → g and x → z . The different possible regimes exhibited by Eqs. (84) are described as follows. 6.1 Convergence to stationary states There can exist the single stable fixed point 1 + b 1 − bg which is positive and stable if either x∗ = , z ∗ = 1 + g 1 − bg , − 1 ≤ b < 0 , g ≥ −1 (85) (86) or if 1 0 ≤ g < gc ≡ b ≥ 0 , b The corresponding stability region is shown in Fig. 16. On the boundary of the stability region, the fixed point degenerates to one of the values (87) . z ∗ = 1 x∗ = 0 , (b = −1, g > −1) , x∗ = 1 , z ∗ = 0 (b > −1, g = −1) , x∗ = 1 − z ∗ , 0 ≤ z ∗ ≤ 1 (b = g = −1) . (88) For b > 0 and g > 0, the basin of attraction is the whole region in the b − g plane, where inequalities (87) are valid. When one of the parameters b or g is negative, or both are negative, then the attraction basin is defined by one of the following conditions, either or by or by x0 < z0 < 1 g 1 b (b > 0, g < 0) , (b < 0, g > 0) , (89) (90) 1 1 b g The convergence to the stationary solution can be monotonic or non-monotonic, depending on the symbiosis parameters and initial conditions, similarly to Figs. 2, 3, 4, and 10, 11. (b < 0 g < 0) . x0 < z0 < (91) , 6.2 Exponential growth at infinity This behavior, which is analogous to that displayed in Fig. 7, occurs when 1 g > gc ≡ b Then, both solutions x(t) and z(t) exponentially increase as t → ∞. b > 0 , . (92) 19 6.3 Finite-time singularity When one or both of the symbiosis parameters are negative, implying parasitic symbiotic relations, a finite-time singularity can occur. For the initial conditions 1 g there exists a critical time tc , such that x0 > , z0 > − 1 b (b > 0, g < 0) , (93) x(t) → x(tc ) (t → tc ) , x(t) → x(tc ) , z (t) → +∞ (t → tc ) . z(t) → +∞ , This behavior is similar to that shown in Fig. 12. Because of the equation symmetry, there is the opposite case, occurring for (94) x0 > − 1 g , z0 > 1 b (b < 0, g > 0) , when x(t) → +∞ (t → tc ) , x(t) → +∞ , (t → tc ) . z (t) → z (tc) z(t) → z(tc ) , If both symbiosis parameters are negative, and (95) (96) 1 1 b g there exists a finite-time singularity of the same type as above, with the particular behavior depending on the parameter values and initial conditions. (b < 0, g < 0) , x0 > z0 > (97) , 6.4 Finite-time death The occurrence of a gradient catastrophe, of the type described previously in Sec. 5, depends on the values of the symbiotic parameters and initial conditions. If the symbiotic parameters are outside of the stability region and the initial conditions are such that 1 g then there exists a finite time td , determined as the solution of the equation (b > 0, g ≤ −1) , z0 > − x0 < 1 b , g x(td) = 1 , (98) (99) where 1 x(t) → (t → td ) , x(t) → x(td ) g z (t) → −∞ (t → td ) . z(t) → z(td ) , (100) The gradient catastrophe occurs for the species z , whose gradient tends to −∞. This can be interpreted as an abrupt collapse of z to 0 and thus as the extinction of this species. The , 20 continuation of the solution x is then obtained by setting z ≡ 0 for t > td . The extinction of the species z is caused by the strong parasitic action of the species x onto species z , represented by the negative value of the symbiotic parameter g . Under the condition 1 b the role of x and z is interchanged and the finite-time extinction is observed for the species x. At the finite time td , given as the solution to the equation (b ≤ −1, g > 0) , x0 > − (101) z0 < 1 g , we have bz(td ) = 1 , (102) x(t) → −∞ (t → td ) x(t) → x(td ) , 1 z(t) → z (t) → z(td ) (t → td ) . , b Again, interpreting the gradient catastrophe for x as the extinction of x, the continuation of the solution for z is obtained by setting x ≡ 0 for t > td . The extinction of the species x is caused by the strong parasitic action of z on x, due to the negative symbiotic parameter b. The gradient catastrophe can occur even when the symbiotic parameters are inside the stability region, but with initial conditions that are outside of the basin of attraction. Thus, if (103) 1 1 b g then there exists a finite time td , given as the solution of the equations (b < 0, g < 0) , x0 < z0 > , (104) (105) g x(td) = bz(td ) = 1 , at which , x(t) → 1 g 1 z → −∞ (t → td) . z(t) → b Again, here the gradient catastrophe for z can be understood as the extinction of z . Con- versely, when the initial conditions are such that x → +∞ (t → td ) , (106) , 1 1 b g then there is the finite time td , given by the same Eqs. (105), when (b < 0, g < 0) , x0 > z0 < , (107) , x(t) → 1 g 1 z (t) → +∞ (t → td ) . z(t) → b This can be interpreted as the extinction of the species x, after which it can be set zero for t > td . The overall behavior is presented in Fig. 17. x(t) → −∞ (t → td ) , (108) , 21 7 Discussion 7.1 Summary of the main results and outlook We have proposed an approach to describe the dynamics of symbiotic relations between several species. We have argued that the notion of symbiosis can be applied not merely to biological systems but can also be generalized to social systems of different nature. The principal point of our approach is the description of symbiosis through the influence of each species on the carrying capacities of the others. This is in agreement with the common understanding that symbiotic species act on the livelihood of each other, either improving it, under beneficial relations, or destroying it, under parasitic relations. The general symbiotic model can be represented by several variants characterizing dif- ferent types of symbiotic relations. We have considered three basic types of such relations, symbiosis with mutual interactions, symbiosis with asymmetric interactions, and symbiosis without direct interactions. In all cases, we have provided a complete classification of all admissible dynamical regimes. The functional dependence of the carrying capacities on the species variables x and z constitutes the principal difference between our model and the logistic equation, where the carrying capacity is fixed. We have chosen here the simplest functional form of the carrying capacity characterizing symbiotic interactions. The carrying capacity has been taken as a combination of terms not exceeding the bilinear order in terms of the species variables. In general, it would be possible to describe the carrying capacity as an expansion in increasing powers of x and z . Then, we could use an effective summation of the expansion, obtaining a more complicated expression, e.g., in the form of the self-similar exponentials, as has been done for the model describing the dynamics of a nonequilibrium financial system [50]. Another modification could be by including the effects of delays into the carrying capacity, as has been done for the model of punctuated evolution [51]. Even in its simplest form, the proposed model of symbiosis turned out to be sufficiently rich, exhibiting four qualitatively different regimes: convergence to stationary states, un- bounded exponential growth, finite-time singularity, and finite-time death or extinction of species. The suggested model of symbiosis can be applied to a variety of systems, biological, social, economic, financial, and so on. We have just mentioned some of the possible applications. The main goal of the present paper has been to advance a general parsimonious model of symbiosis and to analyze its main dynamical regimes. Here, we limit ourselves by the mathematical side of the problem. Particular applications require separate investigations and will be studied in future publications. 7.2 Recipe to avoid finite-time singularities We have shown that singularities appear under parasitic relations, when at least one of the symbiosis coefficients is negative. The characteristics of real systems, of course, cannot diverge. Therefore, the occurrence of finite-time singularities should be understood as the manifestation of an instability developing in the system, which signals a change of regime into a new structural phase described by different mechanisms and thus different equations [52]. In some cases, the singularity can be avoided by slightly modifying the model, e.g, by taking 22 into account higher-order powers of the variables x and z . Such higher orders can remove the finite-time singularity [50]. In any case, when a finite-time singularity does happen, this can be understood as the signal that the system behavior is drastically changing at this point, somewhat similar to the changes occurring under phase transitions [53, 54]. In order to be more specific, let us delineate how the model can be modified in order to avoid the appearance of finite-time singularities, leading to a more realistic behavior. Recall that finite-time singularities appear under parasitic relations, when at least one of the symbiosis coefficients is negative. The divergence occurs when the effective carrying capacity either becomes zero or is negative owing to the choice of initial conditions. Hence, mathematically, to avoid the occurrence of such divergences, it would be sufficient to have always positive-definite carrying capacities. This also would be reasonable for the ma jority of biological systems, though for economic and financial systems, a transient effective negative carrying capacity can have sense, representing the leverage of an economy over-stretching its borrowing level beyond its capacity for reimbursement, leading to crises and bankruptcies as illustrated by the sovereign default issues in Greece, Ireland, and in Europe that were revealed in 2010. With the goal of removing finite-time singularities, let us consider the case of Sec. 4, where the finite-time singularity happens at a critical time, as is described in subsection 4.3, when either the carrying capacity y1 or y2 becomes zero. We may treat the carrying capacities y1 = 1 + bxz and y2 = 1 + gxz as the first terms of the general expansions y1 = y2 = bn (xz)n , an (xz)n , ∞ ∞ Xn=0 Xn=0 with a0 = 1, a1 = b, b0 = 1, and b1 = g . To find effective limits of such expansions, we can resort to some resummation, or renormalization, procedure. Probably, the most general procedure of this kind is based on the self-similar approximation theory [55–62]. Invoking the variant of this theory, employing self-similar exponential approximants [63, 64] yields renormalized effective sums that are always sign defined. For instance, the first-order exponential approximants for sums (109) read as (109) y ∗ 1 = exp(bxz) , y ∗ 2 = exp(gxz). These forms are evidently positive for any signs of b and g , which makes it straightforward to avoid the finite-time singularities caused by the occurrence of zero values of yi at some finite time. The same method of avoiding the finite-time singularities can be used for other cases, where such divergences arise. This method, e.g., was employed in Ref. [50] for re- moving the finite-time singularities in the dynamical models of financial markets. The use of the renormalized expressions for the effective carrying capacities makes the mathematical treatment more involved and requires separate consideration. 7.3 Comparison of symbiosis model with predator-prey Lotka- Volterra model It is worth emphasizing that the suggested symbiosis model is basically different from the predator-prey Lotka-Volterra model. For two species, the general form of the latter can be 23 written as dx2 dt dx1 dt (110) = γ2x2 − a2x2 = γ1x1 − a1x2 1 + b12x1x2 , 2 + b21x2x1 , where the coefficients a1 and a2 are positive. One gets the standard Lotka-Volterra model [17, 18] for prey, with the concentration x1 , and predators, with the concentration x2 , when γ1 > 0, a1 = 0, and b12 < 0, while γ2 < 0, a2 = 0 and b21 > 0. In the general case, Eqs. (110) correspond to the competitive Lotka-Volterra model, if the coefficients b12 and b21 are negative, while if they are positive, one has the mutualistic Lotka-Volterra model [65, 66]. Let us compare the mathematical structure of these equations (110) with the symbiosis equations studied above, say, with Eqs. (31). It is easy to see that their nonlinearity is of rather different forms, which makes the behavior of their solutions principally different. One could ask the question whether there are, maybe, some particular cases when these equations are close to each other. For example, could it be that Eqs. (31) would reduce to Eqs. (110) at small values of xz? Expanding the expressions 1/y1 and 1/y2 in powers of xz , and retaining the minimal terms containing the symbiosis coefficients, we get dx dz dt ≃ z − z 2 + gxz 3 . dt ≃ x − x2 + bx3 z , These equations, clearly, are very different from Eqs. (110). But, maybe, the symbiosis and Lotka-Volterra equations could be equivalent for the dynamics close to stationary points, when the symbiosis equations could be reduced to polynomial forms? To check this, let us consider small deviations from fixed points x∗ and z ∗ , defined as x2 ≡ z − z ∗ . x1 ≡ x − x∗ , To compare with Eqs. (110), we need to represent Eqs. (31) in the polynomial form of second order with respect to the small deviations x1 and x2 . This results in the following equations 1 + B12x2 = A11x1 + A12x2 + B11x2 2 + C12x1x2 , 1 + B22x2 = A21x1 + A22x2 + B21x2 2 + C21x2x1 . dx1 dt dx2 dt Again these equations are principally different from Eqs. (110). Of course, if we would limit ourselves to only small deviations of first order in Eqs. (31) and (110), we would come to linear equations. All linear equations are formally similar to each other. However, the linearized forms obtained from Eqs. (31) and (110) will have absolutely different coefficients and will correspond to different fixed points, with different stability properties. We should also mention the existence of a fundamental difference at the conceptual level. The Lotka-Volterra equations of the type (110) represent the dynamics of species as resulting from pairwise interactions, associated with the quadratic and bilinear terms. In contrast, the symbiosis equation (31) expresses the interactions through the impact of species concentrations on their carrying capacities. In principle, modeling explicitly the dynamics of the carrying capacities as genuine coupled dynamical variables provides a richer conceptual and technical approach to complex ecologies. The carrying capacities act as important relevant dynamical variables. 24 Therefore we have to conclude that the symbiosis and prey-predator equations are prin- cipally different and cannot be reduced to each other. They have different mathematical structure and different solutions. The mathematical difference of the suggested symbiosis model from the predator-prey models reflects the actual difference between the processes of symbiosis and predation [1–5, 47]. 7.4 Examples of symbiotic types In conclusion, we feel it is useful to return back to concrete examples of particular symbiotic relations treated in the paper. This would give the reader the feeling that the considered mathematical models are closely connected to real cases of symbiosis. Probably, one of the best known examples of symbiosis is the mutualistic interaction between trees and fungi. Trees ability to generate large amounts of biomass or store carbon is underpinned by their interactions with soil microbes known as mycorrhizal fungi that are a crucial part of all forest ecosystems. Mycorrhizas are symbiotic relationships between certain fungi and the roots of plants. The fine fungal threads (called hyphae) either ensheathe or penetrate the host plant’s roots. The fungus helps the plant to extract nutrients and water from the soil. Fungi excel at procuring necessary, but scarce, nutrients such as phosphate and nitrogen. They also protect their hosts against harmful organisms. In return, the fungus receives sugars via the plant’s photosynthesis. The fungus within the root is protected from competition with other soil microbes and gains preferential access to carbohydrates within the plant. As with most mutualistic relationships, each partner grows better in association with the other than it would individually. For instance, birch (Betula spp.) has a number of these partnerships, the most familiar being with the red and white fly agaric (Amanita muscaria), as well as with the chanterelle (Cantharellus cibarius). Scots pine has mycorrhizal associations with over 200 species of fungi in Scotland, including another kind of chanterelle (Cantharellus lutescens). In fact, the ma jority of plants forests benefit from mycorrhizal relationships, and it is thought that mycorrhizas helped plants to colonize the land, millions of years ago [67]. In the frame of our approach, such a symbiotic relation is described by the carrying capacities (12). There, the carrying capacity A1 is the amount of natural resources available for trees, and the carrying capacity A2 , the amount of natural resources of fungi. In their mutual interactions, trees and fungi produce additional resources that enlarge their carrying capacities by providing more food for each of them. Thus, B1N1N2 and B2N2N1 are the additional amounts of food for trees and fungi, respectively, produced in the process of their interaction. As is explained in the Introduction, the ideas of biological symbiosis nowadays are widely used in economic relations, leading to the appearance of such terms as industrial symbiosis and economic symbiosis. For example, the relations between employers and employees can be described in the same way as symbiotic relations between some biological species [68]. The approach, known as Partner Fidelity Feedback, holds that, similarly to biological species, social symbionts have evolved to help their hosts because a healthy host automatically feeds back benefits to the symbionts. A cheating symbiont would seem to be treated like any other environmental setback, such as infertile soil, and a mutualistic symbiont elicits the same sort of investments that are triggered by the availability of new resources, like a patch of sunlight [68]. In the frame of our model, the symbiotic relations between employers and employees 25 should be characterized by the carrying capacities (21). Then, A1 and A2 are the natu- ral carrying capacities available irrespectively of interactions. Employers provide tools to employees for producing additional capacity B1N1N2 . As a result of the growing firm or company, the employees also profit getting the increased capacity B2N1 , say, in the form of increased salaries. Another example of industrial symbiosis between different enterprises is the process of sharing of services, utility, and by-product resources among diverse industrial actors in order to add value, reduce costs, and improve the environment. Industrial symbiosis is a subset of industrial ecology, with a particular focus on material and energy exchange. Industrial ecology is a relatively new field that is based on a natural paradigm, claiming that an industrial ecosystem may behave in a similar way to the natural biological ecosystems [15]. A straightforward case is when companies collaborate to utilize each other’s by-products and otherwise share resources, enlarging by this their carrying capacities. In our approach, this situation corresponds to the carrying capacities (26). Then the capacities A1 and A2 are the resources available to each of the industrial symbionts. Their additional resources come from the exchange of the companies by-products, increasing their capacities by the amounts B1N2 and B2N1 , respectively. These by-products are produced by each of the companies, without direct interactions between them. And without this exchange, the by-products would be lost, while due to the exchange, they increase the carrying capacities of the symbiotic companies. These examples illustrate that our main idea of modeling symbiotic relations by taking into account the mutual influence of symbiotic species on there carrying capacities allows us to describe various types of symbiosis for biological or social symbiotic systems. 7.5 Symbiosis in microbial systems The symbiosis of microbial species is of special interest for several reasons: • it is so much widespread in nature, • it can be observed and studied in the laboratory, which is often impossible for other wild species, • it can be modeled in artificial synthetic-biology games [69, 70], and • the specific features of microbial coexistence, such as resource enrichment [71, 72], fluctuation enhancement [73], quorum sensing [74, 75], group selection [76, 77], devel- opment of cooperation [77–79], and evolutionary race [80, 81], can be illuminating for understanding, characterizing, and organizing both technological as well as social sys- tems [82]. Microbial symbiosis has been a survival feature of bacteria since their origin. The best example of this is the presence of the energy factories known as mitochondria in eukaryotic cells. Mitochondria arose because of the symbiosis between an ancient bacterium and a eukaryote. Over evolutionary time, the symbiosis became permanent, and the bacterium became part of the host. However, even to the present day, the differences in constitution and arrangement of the genetic material of mitochondria and the host cell’s nucleus attest to the symbiotic origin of mitochondria. 26 There are many well-known examples of bacterial mutualism. The first example is the presence of huge numbers of bacteria in the intestinal tract of warm-blooded animals such as humans [83]. About 10 percent of the dry weight of a human consists of bacteria. The bacteria act to break down foodstuffs, and so directly participate in the digestive process. As well, some of the intestinal bacteria produce products that are crucial to the health of the host. For example, in humans, some of the gut bacteria manufacture vitamin K, vitamin B12, biotin, and riboflavin. These vitamins are important to the host but are not produced by the host. The bacteria benefit by inhabiting an extremely hospitable environment. The natural activities and numbers of the bacteria also serve to protect the host from colonization by disease-causing microorganisms and to educate the host immune system. The importance of this type of symbiosis is exemplified by the adverse health effects to the host that can occur when the symbiotic balance is disturbed by antibiotic therapy. The skin is colonized by a number of different types of bacteria, such as those from the genera Staphylococcus and Streptococcus [84]. The bacteria are exposed to a ready supply of nutrients, and their colonization of the skin helps protect that surface from colonization by less desirable microorganisms. Bacteria themselves coexist with bacteriophages that are viruses attacking bacteria. Therefore, bacteria and virulent phages are often treated as prey and predators, respec- tively. Classical predator-prey systems are modeled by the Lotka-Volterra equations, whose solutions display oscillations in the populations of the competing species [85]. However, the experimental studies of bacteria-phage biology [86, 87] reveal the existence of stationary non-oscillating equilibrium states of populations and also the occurrence of extinction phe- nomena, when one of the species becomes completely extinct. It has therefore been necessary to modify the classical predator-prey equations for accounting for the non-oscillatory stable coexistence of bacteria and phages [88–92]. In this regard, we would like to emphasize that, in our approach, the coexisting equilib- rium states, corresponding to stable fixed points, appear naturally for all three models (31), (52), and (84). Under strong parasitic relations, dynamics with death occurring in finite time are also present, characterizing the phenomenon of species extinction, as is shown in Figs. 13, 14, 15, and 17. Similar extinction of bacteria E .coli, attacked by a large population of bacteriophages T 4, has been observed in experiments [87]. A detailed description of coexisting bacterial hosts and virulent phages requires to take into account the renewable environmental resources, satiation effects, the host lysis, when each infected host releases many phages, and also spatial heterogeneity. Including these processes into the consideration would need to add several more equations, essentially com- plicating the dynamical system, which is not the aim of our paper. Our main goal has been to suggest a new mechanism, not treated earlier, that of the mutual influence of symbionts on the carrying capacities of each other and to demonstrate that it is not less, but, probably, even more important than other known factors. Taking into account only this mutual influ- ence makes it possible to get a rich variety of admissible types of the symbiotic behavior, including the stable coexistence of species as well as extinction of one of them, caused by parasitic relations. Acknowledgements 27 We acknowledge financial support from the ETH Competence Center “Coping with Crises in Complex Socio-Economic Systems” (CCSS) through ETH Research Grant CH1-01-08-2 and the ETH Zurich Foundation. 28 References [1] D. Boucher, The Biology of Mutualism: Ecology and Evolution, Oxford University, New York, 1988. [2] A.E. Douglas, Symbiotic Interactions, Oxford University, Oxford, 1994. [3] J. Sapp, Evolution by Association: A History of Symbiosis, Oxford University, Oxford, 1994. [4] V. Ahmadjian, S. Paracer, Symbiosis: An Introduction to Biological Associations, Ox- ford University, Oxford, 2000. [5] C.R. Townsend, M. Begon, J.D. Harper, Ecology: Individuals, Populations and Com- munities, Balckwell Science, Oxford, 2002. [6] P.J. Turnbaugh, R.E. Ley, M. Hamady, C.M. Fraser-Liggett, R. Knight, J.I. Gordon, Nature 449 (2007) 804-810. [7] E. Von Hippel, The Sources of Innovation, Oxford University, Oxford, 1988. [8] G.M. Grossman, E. Helpman, Innovation and Growth in the Global Economy, MIT, Massachusets, 1991. [9] R.R. Richard, National Innovation Systems: A Comparative Analysis, Oxford Univer- sity, Oxford, 1993. [10] J. Whitfield, Nature 449 (2007) 136-138. [11] T. Woyke, H. Teeling, N.N. Ivanova, M. Huntemann, M. Richter, F.O. Gloeckner, D. Boffelli, I.J. Anderson, K.W. Barry, H.J. Shapiro, E. Szeto, N.C. Kyrpides, M. Muss- mann, R. Amann, C. Bergin, C. Ruehland, E.M. Rubin, N. Dubilier, Nature 443 (2006) 950-955. [12] L. Dethlefsen, M. McFall-Ngai, D.A. Relman, Nature 449 (2007) 811-818. [13] M.R. Chertow, Annu. Rev. Energy Environ. 25 (2000) 313-337. [14] T. Graedel, B. Allenby, Industrial Ecology, Prentice Hall, Englewood Cliffs, 2003. [15] J.M. Pearce, Renewable Energy 33 (2008) 1101-1108. [16] K. Press, A Life Cycle for Clusters, Physica, Heidelberg, 2006. [17] A.J. Lotka, Elements of Physical Biology, Williams and Wilkins, Baltimore, 1925. [18] V. Volterra, Nature 118 (1926) 558-560. [19] T. Gross, U. Feudel, Phys. Rev. E 73 (2006) 016205. [20] T. Gross, L. Rudolf, S.A. Levin, U. Dieckmann, Science 325 (2009) 747-750. [21] D. Sornette, V.I. Yukalov, E.P. Yukalova, J.Y. Henry, D. Schwab, J.P. Cobb, J. Biol. Syst. 17 (2009) 225-267. 29 [22] V.I. Yukalov, D. Sornette, E.P. Yukalova, J.Y. Henry, J.P. Cobb, Concepts. Phys. 6 (2009) 179-194. [23] P.R. Painter, A.G. Marr, Ann. Rev. Microbiol. 22 (1968) 519-548. [24] P.A. Rikvold, J. Math. Biol. 55 (2007) 653-677. [25] P.A. Rikvold, Ecolog. Complex. 6 (2009) 443-452. [26] M. Doebeli, I. Ispolatov, Scince 328 (2010) 494-497. [27] J. Monod, Ann. Rev. Microbiol. 3 (1949) 371-394. [28] J. Monod, Ann. Inst. Pasteur 79 (1950) 390-410. [29] N. van Uden, Ann. Rev. Microbiol 23 (1969) 473-486. [30] A.G. Fredrickson, R.G. Megee, H.M. Tsuchiya, Adv. Appl. Microbiol. 13 (1970) 419-465. [31] E.O. Powell, J. Gen. Microbiol. 18 (1958) 259-268. [32] H.W. Jannasch, R.I. Mateles, Adv. Microb. Physiol. 11 (1974) 165-212. [33] A.G. Fredrickson, Ann. Rev. Microbiol. 31 (1977) 63-88. [34] A.G. Fredrickson, G. Stephanopoulos, Science 213 (1981) 972-979. [35] C.S. Holling, Canad. Entomol. 91 (1959) 293-320. [36] V. Krivan, J. Eisner, Theor. Popul. Biol. 70 (2006) 421-430. [37] C. Hui, Ecolog. Modell. 192 (2006) 317-320. [38] K.S. Zimmerer, Ann. Assoc. Am. Geogr. 84 (1994) 108-125. [39] N.F. Sayre, Ann. Assoc. Am. Geogr. 98 (2008) 120-134. [40] M. Begon, J.L. Harper, C.R. Townsend, Ecology: Individuals, Populations and Com- munities, Blackwell Science, London, 1990. [41] D.H. Boucher, S. James, K.H. Keeler, Annu. Rev. Ecol. Systemat. 13 (1982) 315-347. [42] R.M. Callaway, Botan. Rev. 61 (1995) 306-349. [43] J.J. Stachowicz, BioScience 51 (2001) 235-246. [44] P.J. Richerson, R. Boyd, Human Ecology Rev. 4 (1998) 85-90. [45] K. deLaplante, B. Brown, K.A. Peacock, eds., Philosophy of Ecology, Elsevier, Oxford, 2011. [46] L.J. Goff, ed., Algal Symbiosis, Cambridge University, Cambridge, 2011. [47] F. Brauer, C. Castillo-Chavez, Mathematical Models in Population Biology and Epi- demiology, Springer, Berlin, 2000. 30 [48] R.I.M. Dunbar, Evolut. Anthropol. 6 (1998) 178-190. [49] W.X. Zhou, D. Sornette, R.A. Hill, R.I.M. Dunbar, Proc. Roy. Soc. B 272 (2005) 439- 444. [50] V.I. Yukalov, D. Sornette, E.P. Yukalova, J. Econ. Behav. Org. 70 (2009) 206-230. [51] V.I. Yukalov, E.P. Yukalova, D. Sornette, Physica D 238 (2009) 1752-1767. [52] A. Johansen, D. Sornette, Physica A 294 (2001) 465-502. [53] V.I. Yukalov, Phys. Rep. 208 (1991) 395-492. [54] D. Sornette, Critical Phenomena in Natural Sciences, Springer, Berlin, 2004. [55] V.I. Yukalov, Phys. Rev. A 42 (1990) 3324-3334. [56] V.I. Yukalov, Physica A 167 (1990) 833-860. [57] V.I. Yukalov, J. Math. Phys. 32 (1991) 1235-1239. [58] V.I. Yukalov, J. Math. Phys. 33 (1992) 3994-4001. [59] V.I. Yukalov, E.P. Yukalova, Physica A 198 (1993) 573-592. [60] V.I. Yukalov, E.P. Yukalova, Physica A 206 (1994) 553-580. [61] V.I. Yukalov, E.P. Yukalova, Physica A 225 (1996) 336-362. [62] V.I. Yukalov, E.P. Yukalova, Ann. Phys. (NY) 277 (1999) 219-254. [63] V.I. Yukalov, S. Gluzman, Phys. Rev. E 58 (1998) 1359-1382. [64] S. Gluzman, V.I. Yukalov, Phys. Rev. E 58 (1998) 4197-4209. [65] S. Smale, J. Math. Biol. 3 (1976) 5-7. [66] M. Hirsch, SIAM J. Math. Anal. 21 (1990) 1225-1234. [67] L. Margulis, The Symbiotic Planet, Phoenix, London, 1998. [68] E.G. Weyl, M.E. Frederickson, D.W. Yu, N.E. Pierce, Proc. Natl. Acad. Sci. USA 107 (2010) 15712-15716. [69] B.S. Chen, C.H. Chang, H.C. Lee, Bioinformatics 25 (2009) 1822-1830. [70] B.E. Beckmann, P.K. McKinley, Proc. of Conference on Genetic and Evolutionary Com- putation, ACM, New York, 2009, p. 97-104. [71] M. Rozenzweig, Science 171 (1971) 385-387. [72] R.J. Doyle, Biofilms, Academic, New York, 1991. [73] C. Boettiger, J. Dushoff, J.S. Weitz, Theor. Popul. Biol. 77 (2010) 6-13. 31 [74] M.R. Parsek, E.P. Greenberg, Trends Microbiol. 13 (2005) 27-33. [75] D. An, T. Danhorn, C. Fuqua, M.R. Parsek, Proc. Natl. Acad. Sci. USA 103 (2006) 3828-3833. [76] J.S. Weitz, Y. Mileyko, R.I. Joh, E.O. Voit, Biophys. J. 95 (2008) 2673-2680. [77] J.A. Damore, J. Gore, J. Theor. Biol. 299 (2012) 31-41. [78] J. Gore, H. Youk, A. van Oudenaarden, Nature 459 (2009) 253-256. [79] Z. Wang, N. Goldenfeld, Phys. Rev. E 84 (2011) 020902. [80] J.A. Damore, J. Gore, Evolution 65 (2011) 2391-2398. [81] D.N.L. Menge, F. Ballantyne, J.S. Weitz, Theor. Ecol. 4 (2011) 163-177. [82] K.R. Foster, K. Parkinson, C.R.L. Thompson, Trends Genetics 23 (2007) 74-80. [83] F. Backhed, R.E. Ley, J.L. Sonnenburg, D.A. Peterson and J.I. Gordon, Science 307 (2005) 1915-1920. [84] E.A. Grice, H.H. Kong and S. Conlan Science, 324 (2009) 1190-1192. [85] A.A. Berryman, Ecology 73 (1992) 1530-1535. [86] M. Alexander, Ann. Rev. Microbiol. 35 (1981) 113-133. [87] B.J.M. Bohannan, R.E. Lenski, Ecology 78 (1997) 2303-2315. [88] B.R. Levin, F.M. Stewart, L. Chao, Am. Naturalist, 111 (1977) 3-24. [89] G.W. Harrison, Bull. Math. Biol. 48 (1986) 137-148. [90] J.S. Weitz, H. Hartman, S.A. Levin, Proc. Natl. Acad. Sci. USA 102 (2005) 9535-9540. [91] J.S. Weitz, J. Dushoff, Theor. Ecol. 1 (2008) 13-19. [92] Z. Wang, N. Goldenfeld, Phys. Rev. E 82 (2010) 011918. 32 Figure Captions Fig. 1. Region of stability (shaded) in the parameter plane b − g for the fixed points in the case of symbiosis with mutual interactions. Fig. 2. Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed-dotted line), as functions of time, in the case of symmetric mutual interactions, when x0 > x∗ and z0 > z ∗ , for different parameters b and g and initial conditions {x0 , z0}: (a) b = −0.1, g = 1, {0.75, 3.8}1, {1, 3.8}2 , {2.5, 3.8}3 ; the fixed points being x∗ = 0.730, z ∗ = 3.702. (b) b = 0.5, g = 0.01 < gc = 0.086, {2.5, 1.5}1 , {4, 1.5}2 , {7, 1.5}3 ; x∗ = 2.043, z ∗ = 1.021. (c) b = −0.5, g = −0.1, {0.8, 1}1 , {0.8, 1.5}2 , {0.8, 2.499}3 , x∗ = 0.681, z ∗ = 0.936. (d) b = 0.5, g = −0.1 < gc = 0.086, {3, 1}1 , {3, 3}2 , x∗ = 1.742, z ∗ = 0.852. Fig. 3. Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed-dotted line), as functions of time, in the case of symmetric mutual interactions, when x0 > x∗ but z0 < z ∗ , for different parameters b and g and initial conditions {x0 , z0}: (a) b = 0.05, g = 0.5 < gc = 0.603, {4, 0.05}1 , {4, 0.5}2 , {4, 2}3 ; the fixed points being x∗ = 1.130, z ∗ = 2.298. (b) b = 0.7, g = 0.02, {6, 0.05}1 , {6, 0.5}2 , {6, 0.9}3 ; x∗ = 4.258, z ∗ = 1.093. (c) b = −1, g = 0.1, {0.45, 0.01}1, {1, 0.01}2 , {1.5, 0.01}3 , x∗ = 0.438, z ∗ = 1.281. (d) b = −0.01, g = 1, {3, 0.05}1 , {3, 2}2 , {3, 3.33}3 , x∗ = 0.730, z ∗ = 3.702. Fig. 4. Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed-dotted line), as functions of time, in the case of symmetric mutual interactions, when x0 < x∗ and z0 < z ∗ , for different parameters b and g and initial conditions {x0 , z0}: (a) b = 0.1, g = 0.45 < gc = 0.468, {0.05, 2}1 , {0.5, 2}2 , {1.1, 2}3 ; the fixed points being x∗ = 1.333, z ∗ = 2.5. (b) b = 0.7, g = 0.025 < gc = 0.027, {0.1, 0.01}1 , {1.4, 0.01}2, {4.9, 0.01}3 ; x∗ = 5, z ∗ = 1.143. (c) b = −1, g = 0.1, {0.01, 0.1}1 , {0.2, 0.1}2 , {0.487, 0.1}3, x∗ = 0.4874, z ∗ = 1.051. (d) b = −0.1, g = −1, {0.01, 0.001}1, {0.01, 0.3}2, {0.01, 0.51}3, x∗ = 0.951, z ∗ = 0.513. Fig. 5. Comparison of the symbiotic solutions x(t) (solid line) and z(t) (dashed-dotted line) with the solutions x(t) = z(t) (dashed line) of the decoupled equations (40) for the same initial conditions x0 = z0 = 0.1 < 1, but for different symbiotic parameters b and g : (a) b = 0.25, g = 0.1 < gc = 0.25; the stationary points of the symbiotic equations being x∗ = 1.411, z ∗ = 1.164. (b) b = 2, g = −0.5, the fixed points of the symbiotic equations being x∗ = 3.562, z ∗ = 0.360. (c) b = −1, g = 2, the symbiotic fixed points being x∗ = 0.293, z ∗ = 2.414. (d) b = −1, g = −2, the symbiotic fixed points being x∗ = 0.707, z ∗ = 0.414. Fig. 6. Comparison of the symbiotic solutions x(t) (solid line) and z(t) (dashed-dotted line) with the solutions x(t) = z(t) (dashed line) of the decoupled equations (40) for the same initial conditions x0 = z0 = 1.8 > 1, but for different symbiotic parameters b and g : (a) b = 0.5, g = 0.080 < gc = 0.086; the fixed points of the symbiotic equations being 33 x∗ = 2.823, z ∗ = 1.292. (b) b = 0.5, g = −0.1, the fixed points of the symbiotic equations being x∗ = 1.742, z ∗ = 0.852. (c) b = −0.3, g = 1, the symbiotic fixed points being x∗ = 0.582, z ∗ = 2.393. (d) b = −0.1, g = −0.3, the symbiotic fixed points being x∗ = 0.927, z ∗ = 0.782. Fig. 7. Logarithmic behavior of the exponentially growing solutions x(t) (solid line) and z(t) (dashed-dotted line) for different symbiotic parameters and initial conditions: (a) b = 1, g = 0.1, x0 = 148, z0 = 0.135. (b) b = 0.5, g = 1 > gc = 0.086, x0 = 0.368, z0 = 0.05. (c) b = 1, g = 0.1 > gc = 0, x0 = 0.135, z0 = 2.72. (d) b = −1, g = 0.1, x0 = 148, z0 = 0.05. Fig. 8. Logarithmic behavior of the solutions in the case of finite-time death and singularity, x(t) (solid line) and z(t) (dashed-dotted line), for the parasitic relations with the symbiotic coefficients b = −1, g = −2, under the initial conditions x0 = 1.8, z0 = 0.5. For these parameters, the critical time is tc = 0.87. Fig. 9. Region of stability (shaded) in the parameter plane b − g for the fixed points in the case of symbiosis with asymmetric interactions. Fig. 10. Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed-dotted line), as functions of time, in the case of asymmetric interactions, for the same initial conditions, such that x0 > z0 , with x0 = 1.8 and z0 = 0.1, for different parameters b and g : (a) b = 0.5, g = 0.1 < gc = 0.125; the fixed points being x∗ = 2.764, z ∗ = 1.276. (b) b = 2, g = −0.5; the fixed points being x∗ = 1.618, z ∗ = 0.191. (c) b = −0.25, g = 2; with the fixed points x∗ = 0.638, z ∗ = 2.275. (d) b = −0.3, g = −0.4; the fixed points being x∗ = 0.833, z ∗ = 0.667. Fig. 11. Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed-dotted line), as functions of time, in the case of asymmetric interactions, for the same initial conditions, such that x0 < z0 , with x0 = 0.1 and z0 = 1.8, for different parameters b and g : (a) b = 0.5, g = 0.1 < gc = 0.125; the fixed points being x∗ = 2.764, z ∗ = 1.276. (b) b = 2, g = −0.5; the fixed points being x∗ = 1.618, z ∗ = 0.191. (c) b = −0.25, g = 2; with the fixed points x∗ = 0.638, z ∗ = 2.275. (d) b = −0.3, g = −0.4; the fixed points being x∗ = 0.833, z ∗ = 0.667. Fig. 12. Logarithmic behavior of the solutions in the case of asymmetric interactions, in the presence of the finite-time singularity, x(t) (solid line) and z(t) (dashed-dotted line), for the symbiotic coefficients b = 2, g = −1.5 and the initial conditions x0 = 1.8, z0 = 0.1. The critical time here is tc = 4.43. 34 Fig. 13. Finite-time death in the case of asymmetric interactions. The species x(t) (solid line) kill the species z(t) (dashed-dotted line) at the death time td : (a) b = −2, g = −1.5; the initial conditions are x0 = z0 = 0.1; the death time is td = 3.8. (b) b = 1, g = −1.5; the initial conditions are x0 = 0.1, z0 = 1.8; the death time is td = 2.6. Fig. 14. Finite-time death in the case of asymmetric interactions. The species x(t) (solid line) kill the species z(t) (dashed-dotted line) at the death time td . The symbiotic coefficients are b = −2, g = −1.5; the initial conditions are x0 = 0.1, z0 = 6; the death time is td = 0.559. Fig. 15. Finite-time death in the case of asymmetric interactions. The species z(t) (dashed-dotted line) kill the species x(t) (solid line) at the death time td . The symbiotic coefficients are b = −2, g = −1.5; the initial conditions are x0 = 1.8, z0 = 0.1; the death time is td = 1.016. Fig. 16. Region of stability (shaded) in the parameter plane b − g for the fixed points in the case of symbiosis without direct interactions. Fig. 17. Finite-time death in the case of symbiosis without direct interactions. Temporal behavior of solutions x(t) (solid line) and z(t) (dashed-dotted line) for different symbiosis parameters and initial conditions: (a) b = −0.75, g = −0.5, x0 = 0.8, z0 = 3; the death time being td = 0.204. (b) b = −1.5, g = 1, x0 = 1, z0 = 0.1; with the death time td = 2.412. (c) b = −1, g = −2, x0 = 0.8, z0 = 0.1; with the death time td = 4.279. (d) b = −2, g = −1, x0 = 0.1, z0 = 1; the death time being td = 1.459. 35 Figure 1: Region of stability (shaded) in the parameter plane b − g for the fixed points in the case of symbiosis with mutual interactions. 36 5 4 3 2 1 0 0 2.5 2 1.5 1 0.5 (3) 0 0 (3) (2) (1) (a) z(t) x(t) 1 2 3 4 5 6 (c) z(t) x(t) (1) (2) 0.5 1 1.5 2 2.5 3 7 6 5 4 3 2 1 0 0 3 2 1 0 0 (b) x(t) z(t) 2 4 6 8 10 (2) (1) (d) x(t) z(t) 1 2 3 4 5 Figure 2: Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed- dotted line), as functions of time, in the case of symmetric mutual interactions, when x0 > x∗ and z0 > z ∗ , for different parameters b and g and initial conditions {x0 , z0}: (a) b = −0.1, g = 1, {0.75, 3.8}1 , {1, 3.8}2 , {2.5, 3.8}3 ; the fixed points being x∗ = 0.730, z ∗ = 3.702. (b) b = 0.5, g = 0.01 < gc = 0.086, {2.5, 1.5}1 , {4, 1.5}2 , {7, 1.5}3 ; x∗ = 2.043, z ∗ = 1.021. (c) b = −0.5, g = −0.1, {0.8, 1}1 , {0.8, 1.5}2 , {0.8, 2.499}3 , x∗ = 0.681, z ∗ = 0.936. (d) b = 0.5, g = −0.1 < gc = 0.086, {3, 1}1 , {3, 3}2 , x∗ = 1.742, z ∗ = 0.852. 37 4 3 2 1 0 0 1.5 1 0.5 0 0 (a) z(t) (3) (2) (1) x(t) 2 4 6 8 10 (c) (1), (2), (3) 6 5 4 3 2 1 0 0 4.5 4 3.5 3 2.5 2 1.5 1 (3) (2) (1) (b) z(t) 2 4 6 8 10 (d) z(t) x(t) (2) (1) z(t) x(t) 2 4 6 8 10 0.5 (3) 0 0 2 4 6 8 10 Figure 3: Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed- dotted line), as functions of time, in the case of symmetric mutual interactions, when x0 > x∗ but z0 < z ∗ , for different parameters b and g and initial conditions {x0 , z0}: (a) b = 0.05, g = 0.5 < gc = 0.603, {4, 0.05}1 , {4, 0.5}2 , {4, 2}3 ; the fixed points being x∗ = 1.130, z ∗ = 2.298. (b) b = 0.7, g = 0.02, {6, 0.05}1 , {6, 0.5}2 , {6, 0.9}3 ; x∗ = 4.258, z ∗ = 1.093. (c) b = −1, g = 0.1, {0.45, 0.01}1 , {1, 0.01}2 , {1.5, 0.01}3, x∗ = 0.438, z ∗ = 1.281. (d) b = −0.01, g = 1, {3, 0.05}1 , {3, 2}2 , {3, 3.33}3, x∗ = 0.730, z ∗ = 3.702. 38 (3) (2) (1) 3 2.5 2 1.5 1 0.5 0 0 1.2 1 0.8 0.6 0.4 0.2 0 0 z(t) (a) x(t) 2 4 6 8 10 12 z(t) (c) (1), (2), (3) x(t) 1 2 3 4 5 6 7 5 4 3 2 1 0 0 1.2 1 0.8 0.6 0.4 0.2 0 0 (b) x(t) (1), (2), (3) z(t) 2 4 6 8 10 12 x(t) (d) (1), (2), (3) z(t) 2 4 6 8 10 Figure 4: Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed- dotted line), as functions of time, in the case of symmetric mutual interactions, when x0 < x∗ and z0 < z ∗ , for different parameters b and g and initial conditions {x0 , z0}: (a) b = 0.1, g = 0.45 < gc = 0.468, {0.05, 2}1 , {0.5, 2}2 , {1.1, 2}3 ; the fixed points being x∗ = 1.333, z ∗ = 2.5. (b) b = 0.7, g = 0.025 < gc = 0.027, {0.1, 0.01}1, {1.4, 0.01}2, {4.9, 0.01}3; x∗ = 5, z ∗ = 1.143. (c) b = −1, g = 0.1, {0.01, 0.1}1 , {0.2, 0.1}2 , {0.487, 0.1}3 , x∗ = 0.4874, z ∗ = 1.051. (d) b = −0.1, g = −1, {0.01, 0.001}1 , {0.01, 0.3}2, {0.01, 0.51}3 , x∗ = 0.951, z ∗ = 0.513. 39 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 3 2 1 0 0 (a) x(t) z(t) 2 4 6 8 10 z(t) (c) x(t) 2 4 6 8 10 4 3 2 1 0 0 1.2 1 0.8 0.6 0.4 0.2 0 0 (b) x(t) z(t) 2 4 6 8 10 (d) x(t) z(t) 2 4 6 8 10 Figure 5: Comparison of the symbiotic solutions x(t) (solid line) and z(t) (dashed-dotted line) with the solutions x(t) = z(t) (dashed line) of the decoupled equations (40) for the same initial conditions x0 = z0 = 0.1 < 1, but for different symbiotic parameters b and g : (a) b = 0.25, g = 0.1 < gc = 0.25; the stationary points of the symbiotic equations being x∗ = 1.411, z ∗ = 1.164. (b) b = 2, g = −0.5, the fixed points of the symbiotic equations being x∗ = 3.562, z ∗ = 0.360. (c) b = −1, g = 2, the symbiotic fixed points being x∗ = 0.293, z ∗ = 2.414. (d) b = −1, g = −2, the symbiotic fixed points being x∗ = 0.707, z ∗ = 0.414. 40 4 3 2 1 0 0 3 2.5 2 1.5 1 0.5 0 0 (a) x(t) z(t) 2 4 6 8 10 (c) z(t) x(t) 1 2 3 4 5 2.5 2 1.5 1 0.5 0 0 2 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 (b) x(t) z(t) 1 2 3 4 5 6 7 8 (d) x(t) z(t) 1 2 3 4 5 Figure 6: Comparison of the symbiotic solutions x(t) (solid line) and z(t) (dashed-dotted line) with the solutions x(t) = z(t) (dashed line) of the decoupled equations (40) for the same initial conditions x0 = z0 = 1.8 > 1, but for different symbiotic parameters b and g : (a) b = 0.5, g = 0.080 < gc = 0.086; the fixed points of the symbiotic equations being x∗ = 2.823, z ∗ = 1.292. (b) b = 0.5, g = −0.1, the fixed points of the symbiotic equations being x∗ = 1.742, z ∗ = 0.852. (c) b = −0.3, g = 1, the symbiotic fixed points being x∗ = 0.582, z ∗ = 2.393. (d) b = −0.1, g = −0.3, the symbiotic fixed points being x∗ = 0.927, z ∗ = 0.782. 41 10 8 6 4 2 0 −2 −4 0 10 8 6 4 2 0 −2 −4 0 (a) ln(x) ln(z) 2 4 6 8 10 12 14 (c) ln(x) ln(z) 5 4 3 2 1 0 −1 −2 −3 0 60 50 40 30 20 10 0 (b) ln(x) ln(z) 2 4 6 8 10 (d) ln(x) ln(z) 4 8 12 16 −10 0 5 10 15 20 25 Figure 7: Logarithmic behavior of the exponentially growing solutions x(t) (solid line) and z(t) (dashed-dotted line) for different symbiotic parameters and initial conditions: (a) b = 1, g = 0.1, x0 = 148, z0 = 0.135. (b) b = 0.5, g = 1 > gc = 0.086, x0 = 0.368, z0 = 0.05. (c) b = 1, g = 0.1 > gc = 0, x0 = 0.135, z0 = 2.72. (d) b = −1, g = 0.1, x0 = 148, z0 = 0.05. 42 8 6 4 2 0 −2 −4 −6 −8 0 ln(z) ln(x) ≈ 0.87 t c 0.2 0.4 0.6 0.8 1 Figure 8: Logarithmic behavior of the solutions in the case of finite-time death and sin- gularity, x(t) (solid line) and z(t) (dashed-dotted line), for the parasitic relations with the symbiotic coefficients b = −1, g = −2, under the initial conditions x0 = 1.8, z0 = 0.5. For these parameters, the critical time is tc = 0.87. 43 Figure 9: Region of stability (shaded) in the parameter plane b − g for the fixed points in the case of symbiosis with asymmetric interactions. 44 3 2.5 2 1.5 1 0.5 0 0 3 2.5 2 1.5 1 0.5 0 0 x(t) (a) z(t) 3 6 9 12 15 (c) z(t) x(t) 1 2 3 4 5 6 7 8 2 1.5 1 0.5 0 0 2 1.5 1 0.5 0 0 (b) x(t) z(t) 1 2 3 4 5 (d) x(t) z(t) 1 2 3 4 5 6 7 8 Figure 10: Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed- dotted line), as functions of time, in the case of asymmetric interactions, for the same initial conditions, such that x0 > z0 , with x0 = 1.8 and z0 = 0.1, for different parameters b and g : (a) b = 0.5, g = 0.1 < gc = 0.125; the fixed points being x∗ = 2.764, z ∗ = 1.276. (b) b = 2, g = −0.5; the fixed points being x∗ = 1.618, z ∗ = 0.191. (c) b = −0.25, g = 2; with the fixed points x∗ = 0.638, z ∗ = 2.275. (d) b = −0.3, g = −0.4; the fixed points being x∗ = 0.833, z ∗ = 0.667. 45 3 2 1 0 0 3 2.5 2 1.5 1 0.5 0 0 x(t) (a) z(t) 3 6 9 12 15 (c) z(t) x(t) 1 2 3 4 5 6 2 1.5 1 0.5 0 0 2 1.5 1 0.5 0 0 (b) x(t) z(t) 1 2 3 4 5 6 7 8 (d) z(t) x(t) 1 2 3 4 5 6 Figure 11: Convergence to stationary states of the solutions x(t) (solid line) and z(t) (dashed- dotted line), as functions of time, in the case of asymmetric interactions, for the same initial conditions, such that x0 < z0 , with x0 = 0.1 and z0 = 1.8, for different parameters b and g : (a) b = 0.5, g = 0.1 < gc = 0.125; the fixed points being x∗ = 2.764, z ∗ = 1.276. (b) b = 2, g = −0.5; the fixed points being x∗ = 1.618, z ∗ = 0.191. (c) b = −0.25, g = 2; with the fixed points x∗ = 0.638, z ∗ = 2.275. (d) b = −0.3, g = −0.4; the fixed points being x∗ = 0.833, z ∗ = 0.667. 46 35 30 25 20 15 10 5 0 −5 0 ln(x) ln(z) = 4.43 t c 1 2 3 4 t 5 Figure 12: Logarithmic behavior of the solutions in the case of asymmetric interactions, in the presence of the finite-time singularity, x(t) (solid line) and z(t) (dashed-dotted line), for the symbiotic coefficients b = 2, g = −1.5 and the initial conditions x0 = 1.8, z0 = 0.1. The critical time here is tc = 4.43. 47 1.2 1 0.8 0.6 0.4 0.2 0 0 (a) x(t) z(t) t = 3.815 d 2 4 6 8 10 2 1.5 1 0.5 0 0 (b) z(t) x(t) t = 2.6 d 1 2 3 4 5 6 7 8 Figure 13: Finite-time death in the case of asymmetric interactions. The species x(t) (solid line) kill the species z(t) (dashed-dotted line) at the death time td : (a) b = −2, g = −1.5; the initial conditions are x0 = z0 = 0.1; the death time is td = 3.8. (b) b = 1, g = −1.5; the initial conditions are x0 = 0.1, z0 = 1.8; the death time is td = 2.6. 48 z(t) 6 5 4 3 2 1 0 0 x(t) t = 0.5593 d 0.5 1 1.5 2 Figure 14: Finite-time death in the case of asymmetric interactions. The species x(t) (solid line) kill the species z(t) (dashed-dotted line) at the death time td . The symbiotic coefficients are b = −2, g = −1.5; the initial conditions are x0 = 0.1, z0 = 6; the death time is td = 0.559. 49 2 1.5 1 0.5 0 0 x(t) z(t) t = 1.016 d 1 2 3 4 5 Figure 15: Finite-time death in the case of asymmetric interactions. The species z(t) (dashed- dotted line) kill the species x(t) (solid line) at the death time td . The symbiotic coefficients are b = −2, g = −1.5; the initial conditions are x0 = 1.8, z0 = 0.1; the death time is td = 1.016. 50 Figure 16: Region of stability (shaded) in the parameter plane b − g for the fixed points in the case of symbiosis without direct interactions. 51 3 2.5 2 1.5 1 0.5 0 0 1 0.8 0.6 0.4 0.2 0 0 z(t) (a) x(t) t = 0.2042 d 0.5 1 1.5 2 2.5 (c) x(t) z(t) t = 4.2788 d 2 4 6 8 10 x(t) (b) z(t) t = 2.4123 d 1 2 3 4 5 6 7 z(t) x(t) (d) t = 1.4568 d 1 2 3 4 1 0.8 0.6 0.4 0.2 0 0 1 0.8 0.6 0.4 0.2 0 0 Figure 17: Finite-time death in the case of symbiosis without direct interactions. Temporal behavior of solutions x(t) (solid line) and z(t) (dashed-dotted line) for different symbiosis parameters and initial conditions: (a) b = −0.75, g = −0.5, x0 = 0.8, z0 = 3; the death time being td = 0.204. (b) b = −1.5, g = 1, x0 = 1, z0 = 0.1; with the death time td = 2.412. (c) b = −1, g = −2, x0 = 0.8, z0 = 0.1; with the death time td = 4.279. (d) b = −2, g = −1, x0 = 0.1, z0 = 1; the death time being td = 1.459. 52
physics/0310096
2
0310
2016-03-20T08:49:23
Combined surface acoustic wave and surface plasmon resonance measurement of collagen and fibrinogen layers
[ "physics.bio-ph", "cond-mat.soft", "q-bio.QM" ]
We use an instrument combining optical (surface plasmon resonance) and acoustic (Love mode acoustic wave device) real-time measurements on a same surface for the identification of water content in collagen and fibrinogen protein layers. After calibration of the surface acoustic wave device sensitivity by copper electrodeposition, the bound mass and its physical properties -- density and optical index -- are extracted from the complementary measurement techniques and lead to thickness and water ratio values compatible with the observed signal shifts. Such results are especially usefully for protein layers with a high water content as shown here for collagen on an hydrophobic surface. We obtain the following results: collagen layers include 70+/-20 % water and are 16+/-3 to 19+/-3 nm thick for bulk concentrations ranging from 30 to 300 ug/ml. Fibrinogen layers include 50+/-10 % water for layer thicknesses in the 6+/-1.5 to 13+/-2 nm range when the bulk concentration is in the 46 to 460 ug/ml range.
physics.bio-ph
physics
a Combined surface acoustic wave and surface plasmon resonance measurement of collagen and fibrinogen layer physical properties J.-M. Friedt FEMTO-ST Time & Frequency, and SENSeOR SAS, Besan¸con, France ([email protected]) L.A. Francis Sensors, Microsystems and Actuators Laboratory of Louvain (SMALL), ICTEAM Institute, Universit´e catholique de Louvain (UCL), Belgium Abstract We use an instrument combining optical (surface plasmon resonance) and acous- tic (Love mode surface acoustic wave device) real-time measurements on a same surface for the identification of water content in collagen and fibrinogen protein layers. After calibration of the surface acoustic wave device sensitivity by copper electrodeposition and surfactant adsorption, the bound mass and its physical properties -- density and optical index -- are extracted from the complementary measurement techniques and lead to thickness and water ratio values compatible with the observed signal shifts. Such results are especially usefully for protein layers with a high water content as shown here for collagen on an hydrophobic surface. We obtain the following results: collagen layers include 70±20% water and are 16±3 to 19±3 nm thick for bulk concentrations ranging from 30 to 300 µg/ml. Fibrinogen layers include 50±10% water for layer thicknesses in the 6±1.5 to 13±2 nm range when the bulk concentration is in the 46 to 460 µg/ml range. Keywords: fibrinogen, density, thickness surface acoustic wave, surface plasmon resonance, collagen, 1. Introduction Sorption processes at the solid/liquid interface by which (bio)molecules bind to material surfaces are of interest for biosensors, biomaterials, material and surface science. Understanding the three-dimensional organization (including density, solvent content and thickness) of the resulting sorbed film and its evo- lution during the adsorption process is crucial for many applications in these domains. For biosensors, more specifically, there is a need to monitor the re- sponse in real-time, in order to assess the adsorption kinetics, and to be able to Preprint submitted to Elsevier November 1, 2018 distinguish contributions coming from the dry sorbed mass, which is the physical criterion for estimating sensitivity, and those that should be attributed to effects intimately associated to the layer organization, like sorbent-bound water and hydrodynamic effects for example. While a wide variety of methods can qualita- tively detect the formation of the sorbed film, almost none of them, taken alone, is quantitative and able to reveal the film organization, and only few can moni- tor the process in real-time. Scanning probe microscopies might fulfill all these requirements, mainly for submonolayers sorbed on smooth surfaces [1]. Neutron reflectivity [2], X-ray Photoelectron Spectroscopy (XPS) [3], mass spectroscopy and radiolabeling are quantitative techniques able to directly measure the dry sorbed amount. Of all the direct detection (i.e. label-less) techniques, we have identified acoustic and optical methods as being the only ones fulfilling two fun- damental criteria of our measurements: time resolved and in-situ (liquid phase) measurement of the physical properties of the adsorbed layer. Various methods of direct detection of biochemical layers have been devel- oped, either based on the disturbance of an acoustic wave [4] (quartz crystal microbalance [5] -- QCM -- and surface acoustic wave devices [6] -- SAW) or of an evanescent electromagnetic wave (optical waveguide sensors [7, 8], surface plas- mon resonance [9, 10] -- SPR). While each one of these transducers individually provides reliable qualitative curves during protein adsorption on their function- alized surfaces, extraction of quantitative physical parameters such as optical index, density, viscosity or water content requires modeling of the adsorbed lay- ers [11]. The modeling includes multiple parameters which must be identified simultaneously: hence the need for the combination of (acoustic and optical) detection methods in a single instrument [12, 13, 14, 15] to separate contribu- tions as a same layer is reacting with the surface under investigation. Multiple investigators have identified such a combination of measurement methods as fruitful means of extracting independent physical properties of adsorbed layers, including the challenging combination of QCM and SPR [16, 17, 18, 19, 20] or comparing the results of successive experiments using different instruments [21, 22, 23, 24], combining QCM and reflectometry [25] or measuring separately using the two techniques [26], or SAW and SPR [27]. We here use a combination of Love mode SAW device and SPR to identify values of density, water content and thickness of surfactant films and protein layers (collagen and fibrinogen) adsorbed on methyl-terminated surfaces. This combined measurement is necessary when attempting to convert a raw signal as observed at the output of a transducer (angle shift for SPR, phase and magni- tude shift for SAW or frequency and damping for a QCM) to the actual protein mass bound to the surface, which is the physical criterion for estimating the expected highest possible sensitivity of a biosensor since it provides an esti- mate of the density of active sites on the surface. We furthermore compare the signals obtained from quartz crystal microbalance with dissipation monitoring (QCM-D [28, 29]) measurements to that of the SAW and, based on the results obtained from the analysis of the SAW/SPR combination, show how SAW and QCM interact differently with the layer. The QCM displays a strong sensitivity to viscous interactions with adsorbed layers as was shown previously [30, 31]. 2 SAW devices are sensitive to mass loading, visco-elastic interactions and electri- cal charge accumulation on the sensing area [32], but with different influences due to the different frequencies and hence penetration ratio of the shear acoustic wave with respect to the layer thickness. Love mode surface acoustic waves were chosen for their high mass sensitiv- ity and their compatibility with measurements in liquid media [33, 34, 35, 36]. Being based on the propagation of a shear horizontal acoustic wave, their inter- action with the surrounding liquid is restricted to an evanescent coupling with the viscous liquid. Although bulk liquid viscosity properties affect the acoustic wave propagation [37, 38], including the phase shift of Love mode SAW [39], we will throughout this investigation consider that the phase shift affecting the SAW device is solely related to adsorbed mass and not to viscous effects of the adsorbed layer, in order to reduce the number of unknowns. Such a crude assumption could be removed by exploiting the SAW insertion losses, for intro- ducing the adsorbed layer viscosity. Throughout the analysis proposed here, we consider hydrodynamic interactions of the acoustic wave with the solvent filled adsorbed layer, as well as the equivalent optical index of the protein-solvent mixture, without focusing on the dynamic viscosity of this adsorbed layer but only on the viscosity of the fluid yielding shear wave evanscent coupling with the solvent. The chosen protein layers consist of collagen and fibrinogen, selected as ref- erences both for their interest in engineering biocompatible surfaces [40, 41] but most significantly for their strong solvent content and hence acoustic proper- ties challenging to analyze. Collagen is a fibrillar protein of the extracellular matrix possessing self-assembly properties, involved in biorecognition processes. The collagen macromolecule consists of a triple helix with dimensions about 300 nm in length and about 1.5 nm in diameter, and weights about 300 kilo- daltons [42]. Organization of collagen films adsorbed on hydrophobic surfaces - CH3 -- terminated self-assembled monolayers (SAMs) and polystyrene -- from 30 to 40 µg/ml -- were fully characterized under water or after drying using AFM [43, 44], X-ray photoelectron spectroscopy [45] and radioassays. By AFM scratching experiments, thicknesses of the film adsorbed on methyl-terminated surfaces was estimated to be about 20 nm under water and 7-8 nm after drying. Furthermore, it was found that the measurement was strongly influenced at weak applied forces (<0.5 nN), in relation with the long-range repulsion (∼50 to ∼250 nm) observed by AFM force-distance curves, strongly suggesting that at least some molecules of the film must protrude into the solution [44]. Ad- sorbed amount of (dry) collagen on methyl-terminated surfaces were estimated to be between 0.4 and 0.8 µg/cm2 by combining AFM and XPS measurements. Values near 0.5 µg/cm2 were determined to be adsorbed on polystyrene by radiolabeling. Fibrinogen, on the other hand, is a blood protein that presents three globular domains linked together by fibrilar segments. Similarly to collagen, its molecular weight is about 340 kilodaltons. Fibrinogen adsorbed on various surfaces has been imaged by AFM as well [46], down to molecular resolution [47, 48, 49, 50]. In this presentation, similar films will be investigated, but under the new 3 Figure 1: Picture of the experimental setup combinging a SPR instrument (670 nm-wavelength laser with mechanical sweep of the illumination angle) with a piezoelectric (ST-cut quartz) substrate acting as both SPR sensing area since coated with a 50-nm Au layer and propagating a Love-mode SAW. Bottom-left inset: schematic view of he measurement head, emphasizing the need for integrated microfluidics to prevent liquid from reaching the interdigitated trans- ducers. In this schematic chart, an electrochemical application is considered by showing the (bent) counter electrode and (straight) reference electrode located over the gold-coated sensing area acting as working electrode. perspective of combined acoustic and optical measurements for extracting the thickness, mass, and solvent density in a buffer medium. While the SAW/SPR technique (Fig. 1) and data processing with our proposed formalism was already shown [51] to be appropriate in the identification of some physical properties of a rigid adsorbed layer (S-layer) [52], we here quantitatively analyze the organi- zation of collagen and fibrinogen adsorbed layers which are expected to possess a substantial water content. This condition is expected to lead to the largest differences between acoustic and optical signals since both techniques respond differently to a viscous (solvent containing) layer: acoustic methods tend to overestimate the bound mass due to hydrodynamic interactions, while optical methods provide an estimate of the dry bound mass after appropriate modeling of the response but cannot resolve both parameters, thickness and optical index of the layer which SPR is sensitive to. This optical index can be assumed to be the weighted value of the index of water and that of proteins to the volumic part that these components occupy in the film. SPR response is thus dependent 4 CCDarraylinear+temperaturesensorRF cablesto networkanalyzer670 nmlaser diodeSAWrotatingmirror: +/−2.5oprismmicro−fluidicsLove−modeSAWlaser50 nmAu on the film organization and on the dry adsorbed amount. Unlike the model used in other studies [53], we here assume that the SAW device signal shift is predominantly due to added rigidly bound mass on the electrode. Indeed we have shown that, while the QCM is sensitive via hydro- dynamic interactions with the topography induced by rough copper electrode- position [54], SAW is much less sensitive [55]. The contribution of the wave coupling with the viscous fluid to the phase shift [39] will be neglected through- out this investigation. Indeed, collagen films were analyzed using SAW/SPR after different conditionings, inducing a different film organization and related viscoelastic behavior (as probed by AFM and QCM respectively) for a similar adsorbed amount. These differences in the film properties were found to result in only minor changes in the SAW response. For all these reasons, the SAW phase response is considered in this article to be only mass dependent during adsorption phenomena. Hence, we use a proportionality relationship between the mass of the layer ∆m per unit area A = 5× 5.5 mm2 here (∆m/A) -- includ- ing the rigidly bound water -- and the frequency shift: ∆m where f0 is the frequency at which the phase is monitored in an open-loop configuration, ∆f is the frequency shift obtained after conversion from phase to frequency shift through the experimentally measured phase to frequency linear relation- ship, and S is the mass sensitivity calibrated by copper electrodeposition. Since the mass sensitivity calibration is a central part in extracting quantitative re- sults from the experimental data, we confirm the results obtained with copper electrodeposition by measuring the SAW signal change during adsorption of a surfactant, cetyltrimethylammonium bromide (CTAB). Hence, four sets of ex- periments will be described in the experimental section below, after presenting the instrument setup: A = ∆f S×f0 1. experimental calibration of the gravimetric sensitivity of the SAW trans- ducer in liquid phase is a core step to the quantitative analysis: such a step is performed by reversible copper electrodeposition, allowing for multiple cycles under different deposited mass configurations to be repeated, 2. because the mechanical properties and surface roughness of copper differ from the biofilm we are investigated, the calibration is validated using CTAB monolayer deposition, 3. collagen thick films will be assembled on the surface under different bulk concentrations, 4. fibrinogen thick films will be assembled on the surface under different bulk concentrations. 2. Experimental section Instrumentation The combined SAW/SPR instrument developed for per- forming this experiment has been described previously [52]: a modified Ibis II SPR instrument (IBIS Technologies BV, Netherlands) is used to inject a 670 nm laser in a quartz substrate (Kretschmann configuration) and monitor the reflected intensity vs. angle with an accuracy of ±2.555◦/200 pixels at two 5 locations separated by about 2 mm on the sensing surface. The ST-cut quartz substrate is patterned with double-fingers interdigitated electrode for launching a Love mode acoustic wave at a center frequency of 123.5 MHz. The guiding layer is made of a 1.13 µm thick PECVD silicon dioxide layer. The phase and insertion loss of the acoustic wave device are monitored using an HP 4396A network analyzer while a custom software records the full reflected intensity vs. angle SPR curve. The 5 × 5.5 mm2 sensing area is coated by e-gun evaporation with nominal thicknesses of 10 nm T i and 50±5 nm Au to support surface plas- mon resonance, and is either grounded to minimize varying salt concentration effects on the signal measured by the SAW device during biological experi- ments, or connected as the working electrode during copper electrodeposition calibration through a bias-T circuit. Both sets of data -- SAW phase at a given frequency and SPR reflected intensity curves -- are time stamped with an ac- curacy of 1 second on the common personnal computer internal clock reference for future processing, which for the SPR curve involves polynomial fitting for improving 100-fold the accuracy on the localization of the reflected intensity minimum. The temperature coefficient of the Love mode device has been mea- sured to be 34±1 ppm/K, in agreement with values reported in the literature [36], and the insertion loss is in the -22 dB to -30 dB range when the grounded sensing area is covered by a liquid. Both dual-finger interdigitated transducers (wavelength: 40 µm) are protected from the liquid by a 100 µm-high SU8 wall capped with a piece of quartz forming sealed protective chambers [56]. The sides of the SAW devices are glued to an 7×17 mm opening in a 1.6 mm-thick epoxy (FR4) printed circuit board (PCB) with the bottom area of the SAW tangent to the PCB, allowing optical contact between the bottom surface of the SAW de- vice and a cylindrical prism through optical index matching oil (noil = 1.518). After wire bonding the SAW device to the PCB, the region surrounding the SAW device is coated with 1 mm wide, 1-2 mm thick epoxy walls using H54 two-parts epoxy glue (Epoxy Technology, Billerica, MA, USA), forming a 180 to 200 µl open well in which the solution is injected using a micropipette. While both SPR and SAW sensors display theoretically similar sensitivities and detection limits [57], the acoustic sensor is extremely sensitive to temper- ature variations and care must be taken to properly control the environment around the instrument to keep the temperature stable [12]. Recording all SPR curves of the reflected intensity vs. angle during the experiment for post-processing is required since the shape of the dip is unusual (Fig. 2) and cannot be processed properly by the commercial software provided with the instrument which expects the substrate to be glass. As seen in Fig. 2 in which the reflected intensity vs. angle curve is displayed, the periodic dips observed when the gold coated SAW sensor is placed over the half-cylinder prism in air is attributed to the birefringence of quartz which is the cause of the interference pattern after reflection of the laser on the metallic surface [52]. The orientation of the quartz substrate on which the SAW propagates has been chosen on purpose to separate as much as possible the successive interference minima. The dip due to SPR is identified when the surface is coated with water, and the position of this dip only is monitored over time. A typical raw SPR 6 Figure 2: SPR dip observed on a quartz substrate coated with 1.13 µm SiO2 and 10+50 nm T i + Au. Notice the additional features on both angle extrema of the curve attributed to interference from quartz birefringence. reflected intensity vs. angle curve is displayed in Fig. 2: the mean incidence angle is about 72±2◦ (not calibrated in the Ibis II instrument) and the abscissa is graduated in relative angle shift. The QCM-D measurements were performed using the electronics supplied by Q-Sense AB (Sweden) using AT-quartz plano-plano resonators (Chintele Quartz Technology Co. Ltd, Zhejiang, P.R. China) coated, by evaporation, with 10 nm T i and 50 nm Au, key-hole shaped counter electrode and fully coated sensing electrode, leading to a fundamental resonance frequency around 4.7 MHz and a quality factor in liquid around 5500. These experiments were performed both in a lab-made open cuvette and in the commercially supplied flow cell (with static liquid during the measurement). Both sets of measurements are compatible. However, since the QCM setup is distinct from the SAW/SPR setup, we will not attempt any kinetics comparison between both datasets. Chemicals and Materials The Cu electrodeposition step was performed using a solution of 10−2 M CuSO4 in 10−2 M H2SO4. The counter elec- trode was a 0.25-mm-diameter platinum wire shaped in 3/4 of a circle and the pseudo-reference electrode was made of a 1 mm-diameter 99.98+% copper 7 −2500−1500−50005001500250020253035404550556065∆θ (mo)reflected intensity (a.u.: %)air tris buffer acetate buffer SPR dip interference(quartzbirefringence) wire (Huntingdon, UK), cleaned by briefly dipping in 70% nitric acid before use. The electrochemical reaction was controlled by a PC3-300 Gamry Instruments (Warminster, USA) potentiostat. Collagen was supplied by Roche (Boehringer-Mannheim, Mannheim, Ger- many) as a sterile acidic solution of 3 mg/ml solution of collagen type I from calf skin. It was diluted in PBS buffer (137 mM N aCl, 6.44 mM KH2P O4, 2.7 mM KCl and 8 mM N a2HP O4) to reach the wanted concentration just prior to use. Type I fibrinogen from human plasma was obtained from Sigma- Aldrich. Both QCM and SAW/SPR Au sensing electrodes were coated with an octadecanethiol hydrophobic self-assembled layer, leading to a contact angle above 110◦. Alkanethiols were self-assembled by dipping the QCMs in a 10−3 M octadecanethiol (Sigma-Aldrich) solution in ethanol after a 15 minutes cleaning step in UV-O3 and rinsing with ethanol. The SAW devices were placed in a chamber with an ethanol saturated atmosphere, and 150 to 180 µl of the same thiol solution was deposited on the open well above the sensing area. Adsorp- tion was subsequently allowed for 3 hours, after which the surfaces were rinsed with ethanol and dried under a stream of nitrogen. The thiol solution and later biochemical solutions are only in contact with the chemically inert H54 epoxy glue and quartz slides capping the SAW device and never with the FR4 epoxy and copper of the PCB, thus minimizing risks of contamination of the solutions and of the sensing surface. 3. Results 3.1. SAW sensitivity calibration Figure 3 illustrates the calibration procedure using copper electrodeposition: a cyclic voltametry sweep from +600 mV to -60 to -100 mV (depending on the mass of copper to be deposited) vs a copper wire acting as a pseudo-reference electrode provides an estimate of the deposited mass, assuming a 100% efficiency of the electrochemical process. A current Ij is monitored by the potentiostat at a sampling rate of 5 Hz, j being a sample index. By selecting the voltage range for which the current is negative (copper deposition), we obtain the deposited mass by numerical integration MCu = ΣjIj × δt F × mCu ne where F is the Faraday constant (96440 C), δt = 0.2 s is the time interval between two samples, mCu = 63.5 g/mol is the molar weight of copper and ne = 2 the number of electrons required for copper reduction. Simultaneously the phase shift ∆φ of the SAW device due to the added mass is measured and converted to a frequency variation ∆f via the linear phase to frequency rela- tionship measured on the Bode plot: ∆f /∆φ = 1680± 5 Hz/◦ in our case. Once the deposited mass is estimated, the SAW sensitivity S is computed considering the center oscillation frequency f0 = 123.2 MHz, the working electrode area A and the frequency shift ∆f due to the added mass: S = ∆f A , leading to an f0 × ∆m 8 Figure 3: Copper electrodeposition calibration curve. The resulting experimentally measured sensitivity indicated on the bottom graph is 165 cm2/g. From top to bottom: raw SAW phase shift and converted frequency shift via the phase to frequency linear relationship observed on the Bode plot; SPR angle shift showing fluctuations with the potential sweep due to electroreflectance effects; voltage applied to the potentiostat (versus a copper wire pseudo- reference electrode); current flowing through the potentiostat and, on top, the mass deposited on the sensing area obtained by integrating the current over time and converting to a mass of deposited copper, and bottom the experimentally measured sensitivity calculated knowing the frequency shift (top graph) and the deposited mass (bottom graph). experimentally measured sensitivity of 165±10 cm2/g for one SAW device and 145±15 cm2/g for another similar sensor (data not shown), the difference being related to packaging issues and mostly the influence of the protective SU8 wall on the acoustic path. Both values were confirmed with CTAB adsorption on octadecanethiol coated gold at a bulk concentration of 1.3×10−4 M: the surface density of surfactant molecules upon adsorption on gold is known from the surface area per molecule of 50 A2 [58] and the molar weight of 364 g/mol, a mass per unit area of 120 ng/cm2 is deduced assuming the formation of a monolayer. The monolayer formation lead to a frequency shift of the SAW device of 2.5 kHz from which a surface density of 125 ng/cm2 is obtained. Interestingly, the simultaneously observed SPR angle shift can only be interpreted with a layer of optical index at least 1.5, which leads to a layer thickness of 1.5±0.5 nm fully coated with CTAB 9 5001000150020002500300035004000450080100120140device #38/5, Cu calibrationSAW Df (o)50010001500200025003000350040004500−0.200.20.40.6E (V)50010001500200025003000350040004500−2024x 10−4time (s)I (A)50010001500200025003000350040004500−1000100SPR Dq (mo)76.0 72.9 68.8 67.5 67.1 67.1 68.1 24.5 23.2 23.2 24.2 25.2 Df (kHz) Dm (ng) 1035 996 936 916 903 902 911 367 321 321 326 327 S (cm2/g) 164 163 164 164 166 166 167 149 161 161 166 172 (no solvent) after simulation using the formalism described in [59] and following the procedure described in more details later in this article. The large uncer- tainty on the layer thickness is related to the low molecular weight of CTAB leading to a small mass increase which leads to a phase shift of only about 5 times the noise level of the SAW phase. Such a high optical index is compatible with that provided in the literature for thiol monolayers as an extension of the optical index of bulk alkane solutions [60]. The surface mass density is compati- ble with that observed with QCM-D measurements following the assumption of a rigid mass, as validated by the good overlap of curves obtained at overtones 3, 5 and 7 when normalized by the overtone number, and the low dissipation increase (∆D < 1 × 10−6). 3.2. SAW/SPR measurements Once the mass sensitivity of the SAW device is calibrated, the combined SAW and SPR measurements are used to monitor protein adsorption and ex- tract the physical properties of the resulting films. Collagen and fibrinogen adsorption were performed on methyl-terminated surfaces of the SAW/SPR de- vice at concentrations of 30 or 300 µg/ml and 46 or 460 µg/ml respectively. The sequence of solutions brought in contact with the sensing surface was succes- sively PBS (or PBS-water-PBS), protein in PBS, PBS, water and finally PBS. Figure 4 displays a typical signal observed during collagen adsorption on an hydrophobic surface, for (a) 300 µg/ml and (b) 30 µg/ml bulk concentrations. Immediately after the conditioning with the collagen solution, both SAW and SPR show a clear response, reaching after about one hour a clear plateau or a beginning of a plateau, depending on the conditioning collagen concentration. The SAW and SPR shifts at the end of the adsorption with a collagen con- centration of 30 µg/ml were respectively 17.0±0.7 ◦ and 0.39±0.07 ◦, while a concentration of 300 µg/ml lead to 19.2±0.5 ◦ and 0.65±0.1 ◦ for SAW and SPR respectively. Adding PBS, after collagen conditioning, resulted in very small loss (< 1 ◦ for SAW and < 0.05 ◦ for SPR) of the response, stable with time, and subsequent washing by water and PBS again provoked no signal change. This means that molecules do not desorb with time once adsorbed on the substrate. The SAW phase to frequency conversion factor as observed on the Bode plot is 1680 Hz/◦: the resulting frequency variation is translated into a bound mass thanks to the calibrated sensitivities -- in our case 165 g/cm2 or 145 g/cm2 depending on the SAW device being used. The resulting hydrated collagen ad- sorbed films are estimated to weight 1.75±0.15 and 2.10±0.20 µg/cm2 for the 30 and 300 µg/ml solutions respectively. Fig. 5 displays a similar measurement on 460 µg/ml fibrinogen. A similar sequence of conditioning steps as previously presented for collagen were applied. PBS washing results in a stable loss of response attributable, like for collagen, to a loss of fibrinogen molecules physisorbed to the adsorbed film. There is again no significant desorption of adsorbed molecules. The observed shifts re- sulting from fibrinogen adsorption were 14.0±0.7 ◦ and 0.65±0.1 ◦ for SAW and SPR respectively. For SAW, using the same relations as was presented for col- 10 (a) (b) Figure 4: (a) 300 µg/ml and (b) 30 µg/ml collagen adsorption curves. Top: phase shift monitored at the constant frequency of f0=123.2 MHz, which can then be converted to an equivalent frequency shift as would be observed in a Phase-Locked Loop oscillator configura- tion via the linear phase to frequency relationship. During experiment (a), collagen adsorption lead to an insertion loss increase of 6.5 dB, from 22.9 dB when the sensing area is covered with buffer to 29.4 dB after collagen adsorption. During experiment (b), this insertion loss increased by 3.3 dB, from 24.4 dB prior to collagen adsorption to 27.7 dB after adsorption. Bottom: SPR dip position angle shift. The sequence is PBS buffer (3 solution exchanges in (a), 2 solution exchanges in (b)), collagen in the same PBS buffer, PBS rinsing step, DI water and PBS. lagen, the surface density of the hydrated fibrinogen films were estimated to be 0.75±0.1 and 1.5±0.3 µg/cm2 for 46 and 460 µg/ml concentrations respectively. 3.3. QCM-D independent validation mesurements Collagen and fibrinogen adsorption were performed on methyl-terminated QCM surfaces following the same conditions as were used for the SAW/SPR. Changes in resonance frequency and dissipation of the first four odd overtones were measured as a function of time. An example of a QCM-D measurement is presented in Fig. 6 for the same 300 µg/ml collagen as in Fig. 4 (b) but on a different setup: both experiments (SAW/SPR and QCM) are based on a lab- made open cuvette in which 200 µl of the new solution is manually injected by a micro-pipette. While the kinetics of the SAW and SPR data can be compared, we will not attempt to compare it with the kinetics of the QCM since the latter was included in a different setup -- static well geometry does influence the adsorption kinetic [61]. We will however see that the asymptotic behavior of both experiments per- formed separately lead to the same conclusion, namely the large amount of water trapped in the collagen layer. Frequency shifts are presented after normalizing by the overtone number as classically done when assuming the adsorbed layer to be rigidly bound to the sensing surface, and by the square root of the over- tone number as would be expected for a strong hydrodynamic interaction [54]. The frequency responses of the different overtones do not overlap when normal- ized by the overtone number while they do when normalized by the square root 11 1000200030004000500060007000120125130135140time (s)SAW Df (o)10002000300040005000600070001.21.41.61.822.2SPR Dq (o)time (s)50010001500200025003000350040004500500055006000105110115120125SAW Df (o)time (s)500100015002000250030003500400045005000550060001.41.51.61.71.81.922.1time (s)SPR Dq (o) Figure 5: 460 µg/ml fibrinogen adsorption curves. Top: phase shift monitored at the constant frequency of 123.2 MHz, which can then be converted to an equivalent frequency shift as would be observed in a Phase-Locked Loop oscillator configuration via the linear phase to frequency relationship. Fibrinogen adsorption lead to an insertion loss rise by 0.5 dB, from 23.0 dB prior to adsorption to 23.5 after adsorption. Bottom: SPR dip position angle shift. The sequence is PBS buffer (2 solution exchanges), DI water, PBS, fibrinogen in the same PBS buffer, PBS rinsing step, DI water and PBS. of the overtone number, indicating a predominantly viscous interaction of the QCM with the film and surrounding solvent. Leading to the same conclusion, the dissipation rises to an enormous value above 100×10−6. Quantitatively, if this film was assumed to be a rigid layer, these frequency shifts would corre- spond to wet masses in the range 8 to 12 µg/cm2 as obtained by applying the Sauerbrey relation. These values are far larger than the corresponding values deduced by SAW measurements (2.10±0.20 µg/cm2). This points to the fact that SAW measurements are far less sensitive to viscoelastic interactions with the adsorbed film than the QCM. All measurements for collagen and fibrinogen adsorptions at the different concentrations were performed and the characteris- tic results as explained above are summarized in Table 1. The same conclusion as above can be made in all cases, QCM measurements denoting a behavior of the adsorbed film strongly interacting throug hydrodynamic coupling with the bound and surrounding solvent, and the corresponding wet adsorbed masses be- 12 50010001500200025003000350040004500500095100105110115time (s)SAW Df (o)500100015002000250030003500400045005000100012001400160018002000time (s)SPR Dq (mo)channel2 channel1B B W B B W B=PBS bufferW=DI water 460 mg/ml fibrinogen √ Figure 6: QCM-D 300 µg/ml collagen adsorption experiment. Top: normalized QCM res- onance frequency assuming a predominantly hydrodynamic interaction (scaling law as 1/ n with n being the overtone number) and assuming a rigid layer (scaling law as 1/n). Notice that the latter set of curves do not overlap and hence the rigid mass hypothesis is not cor- rect. Bottom: QCM dissipation monitoring. Notice the very large damping value reached (∆D > 100 × 10−6). ing always superior to those deduced by SAW. Fibrinogen shows less viscoelastic properties than the collagen film and lower protein concentrations result in a less viscoelastic behavior of the adsorbed film. Notice the excellent agreement between our results and those cited in the literature about dry mass of collagen layers adsorbed under similar conditions as measured by AFM and XPS: 20 nm thick layers under water and 7-8 nm after drying [44], and adsorbed amount of (dry) collagen between 0.4 and 0.8 µg/cm2 are consistent with our 19±3 nm thick layers and a dry mass of 2.1 × 0.35 = 0.7 µg/cm2, justifying our initial assumptions when modeling the adsorbed layer interaction with the acoustic and optical waves. 4. Discussion 4.1. Physical properties of the layers The observed SAW phase shift monitored at 123.2 MHz thus translates into a surface mass density of 1.75±0.15 µg/cm2 and 2.10±0.20 µg/cm2 for bulk 13 10002000300040005000600070008000900010000−1200−1000−800−600−400−2000D f (Hz)time (s)10002000300040005000600070008000900010000020406080100120time (s)DD (· 10−6)Dfn/n Dfn/sqrt(n) surface d (nm) x (%) SAW/SPR SAW/SPR Analyte (bulk concentration, µg/ml) Cu S-layer CTAB collagen (30µg/ml) collagen (300µg/ml) fibrinogen (46µg/ml) fibrinogen (460µg/ml) density (ng/cm2) 560±20 125±15 1750±150 2100±200 750±100 1500±500 2 − 12 4.7 ± 0.7 1.0 ± 0.1 16.0 ± 3.0 19.0 ± 3.0 6.0 ± 1.5 13.0 ± 2.0 √ ∆fn/ (Hz) QCM (Hz) QCM ∆fn/n n ? 75 ± 15 100 25 ± 15 35 ± 10 50 ± 10 50 ± 10 ?-1000 NO NO 1000 1200 110±5 NO NO 45=900 8=160 NO NO 55±5(cid:39) 1110 100=1700 ∆D (×10−6) QCM 50 3-5 0.2-0.5 100 >120 4-10 8-10 Table 1: Summary of the results extracted from simultaneous SAW and SPR measurements, and comparison with the results obtained from QCM-D measurements. The "surface density" value is obtained by converting the phase shift observed on the SAW device monitored in an open loop configuration to frequency through the linear relationship monitored on the Bode plot, and the resulting frequency shift to a mass density per unit area by using the mass sensitivity deduced from the copper calibration. The layer thickness d and protein content of the layer x are deduced from the simulation of the SPR shift as a function of layer thickness and optical index (assuming a protein layer optical index of 1.45 and a protein layer density of 1.4 g/cm3) and correspond to the pair of parameters matching both the SPR and SAW data. The last three columns are related to QCM-D measurements and illustrate the fact that the higher the water content in the layer under investigation, the higher the dissipation (∆D) and the more appropriate the hydrodynamic model of the wave interacting with the viscous solvent (with a 1/ n normalization factor) over the rigid mass model (with a 1/n normalization factor). The equality sign in the 6th column (in which the rigid mass assumption is made and in which the Sauerbrey relationship is applied) relates the normalized frequency shift to a surface density in ng/cm2 using the proportionality factor 20.1 ng/Hz for the 4.7 MHz resonance frequency resonator. √ concentrations of collagen of 30 and 300 µg/ml respectively. The SPR angle shift data are modeled following the formalism developed in [59] in which each component of the electrical field is propagated through a stack a planar di- electic interfaces. The optical indices of the metallic layers were taken from the literature [62] considering the 670 nm wavelength: 10 nm titanium with nT i = 2.76 + i3.84 and 45 nm gold layer with nAu = 0.14 + i3.697, while the op- tical indexes of the underlying dielectric substrate were taken as nquartz = 1.518 coated with 1200±100 nm silicon dioxide nSiO2 = 1.45. While only the last dielectric layer (SiO2) below the metallic layer is necessary to estimate the po- sition of the plasmon peak, we include the full stack of planar layers in our simulation in order to include possible interference effects. The optical index of the protein nprotein is assumed to be in the 1.45 to 1.465 range, as is the case for most polymers [63, 64, 9]. The properties of the protein layer, sup- posed to be homogeneous, are deduced from the assumption that a weighted average of its components (water and protein) can be used: if the ratio of protein in the layer is x, then the optical index of the layer is assumed to be nlayer = x · nprotein + (1 − x) · nwater [65] (nwater = 1.33) and its density ρlayer = x · ρprotein + (1 − x) · ρwater (ρwater = 1 g/cm3 and ρprotein is assumed to be equal to 1.4 g/cm3 range [66, 11, 63, 64]). The latter value is used in our simulations. The methodology for extracting the physical parameters of the 14 protein layers is the following: we first simulate the predicted SPR dip angle shift as a function of layer protein content x and layer thickness d, as shown in Fig. 7 where we display the SPR angle shift due to the added dielectric layer formed by the proteins (i.e. dip angle with solvent before adsorption subtracted from the dip angle after adsorption). Figure 7: Simulation of the angle shift (ordinate) as a function of layer thickness (abscissa) and protein content ratio x (right and top margin) for various layers, including or not including the thiol layer and various protein layer optical index (1.45 or 1.465). The mass density in ng/cm2 are indicated within the graph for the points compatible with the observed SPR angle shift (390±70 m◦ for 30 µg/ml bulk collagen concentration and 650±100 m◦ for 300 µg/ml bulk collagen concentration): the only point compatible with the 1750±150 ng/cm2 (2100±200 respectively) observed by the SAW device results in a layer thickness of 16±3 nm (19±3 nm respectively) and a protein content ratio of 25±15% (35±10% respectively). Optics simula- tions were performed either assuming a 2.6 nm thick thiol layer [60] (refractive index: 1.50) on the T i/Au layer followed by a variable thickness of collagen (mixture of protein with optical index 1.465 or 1.45 and water with optical index 1.33: crosses and circles), or assuming that the thiol layer leads to a negligible effect and not including it in the simulation, with a protein optical index of 1.465 (dots). The resulting mass density ranges (in ng/cm2) are calculated for the intersection point of each of these curves with the observed angle shift for each possible protein to water content ratio: all values indicated on the figure are for solvent swollen layers, including the mass of the trapped solvent. We mark on such a plot the experimentally observed angle shift, which leads to a set of possible {x, d} couples compatible with the SPR experiment. For each of these set we compute the surface density of the layer d × ρlayer and 15 051015202500.10.20.30.40.50.60.70.80.91layer thickness (nm)SPR angle shift ∆θ (o)o: with thiols, n=1.45 +: with thiols, n=1.465 300 µg/ml 30 µg/ml thiol: 2.6 nm thick, n=1.50 2565 2216 1972 1824 21281882 1624 1438 2646 2322 1770 1568 1369 1825 1610 1254 1075 x=0.1 x=0.2 x=0.3 x=0.4 x=0.5 x=0.6 x=0.7 x=1.0 .: no thiols, n=1.465 compare it with the value ∆m/A = ∆f /(S × f0) experimentally observed with the SAW measurement. A single set {x, d} is compatible with both the SAW and SPR data. Notice that the resulting mass density d × ρlayer includes the mass of trapped solvent in the layer. The results of both SPR and SAW signals can only be compatible with a water content of the collagen layers in the range 75%±15% for 30 µg/ml concentration (x=0.25±0.15) and 65%±10% for 300 µg/ml (x=0.35±0.1), with a lower water content for higher bulk collagen concentrations, and correspond to the only set of parameters compatible with the SPR angle shift and the mass variation observed by the SAW device, assuming a purely rigid mass effect and negligible viscous interaction. The very high water content obtained here is confirmed by the large dissipation monitored on the QCM-D. As was shown earlier in [54] and by others [31], a large dissipation shift in- dicates strong hydrodynamic interaction of the vibrating QCM surface with the surrounding solvent, and is confirmed by the overtone scaling law ∆fn/ (as opposed to the expected ∆fn/n =constant for a rigid layer). One also sees here the strong overestimate of the bound protein mass if assuming a rigid mass and applying the Sauerbrey proportionality relationship between mass and fre- quency shift. The estimated collagen film thickness in the 15 to 20 nm range (Fig. 7) is in agreement with AFM topography maps obtained ex-situ under similar conditions [43]. √ n =constant Table 1 summarizes our measurement results on collagen and fibrinogen lay- ers following the data processing methodology described in the previous para- graph. We observe that the parameters extracted from the SAW/SPR simulta- neous measurements are in good qualitative agreement with the QCM-D results: a layer leading to a high dissipation increase (∆D > 10× 10−6) and an overtone normalization law with the square root of the overtone number is characterized by a high water content, as seen with collagen. On the other hand a small dissipation increase and a normalization law of the QCM frequencies with the overtone number are indicative of a low water content and a layer behaving as a rigid mass bound to the surface, as observed with CTAB and the S-layer. Fibrinogen is an intermediate case, for which the QCM frequency does not scale well with the overtone number nor with the square root of the overtone number, and for which the dissipation increase is average. We indeed find a rather low water content for such layers. The uncertainty on the collagen layer density and hence thickness is large: as opposed to highly ordered layers such as those formed by self-assembled monolayers of thiols or by CTAB, we do not expect highly disorganized polymer-like films of fibrous proteins to form exactly sim- ilar structures from one experiment to another. The physical property (water content and hence density and optical index) results on the other hand are con- sistent from one experiment to another, hence the need for the simultaneous measurements at the same location by both optical and acoustic methods (and eventually additional methods such as scanning probe microscopy [49]) in order to obtain statistically significant properties of the film. 16 4.2. Kinetics analysis We attempt now to provide some hint at differences in the kinetics as ob- served by acoustic and optical methods. We compare the kinetics as monitored by SAW and SPR by normalizing the response of the sensor once the saturation level is reached, just after the rinsing buffer step. We thus plot in Figs. 8 (a) and (b) the curves S(t) − S(t0) S(tf ) − S(t0) (1) where t0 is the adsorption starting time, tf the time at which a saturation signal has been reached and the protein solution is replaced by buffer, while S(t) is a signal which can either be in our case the SAW phase shift ∆φ or the SPR angle shift ∆θ. We now first analyze the expected behavior of the normalized quantities of the two signals when assuming a rigid mass interaction and the model presented previously here, and will then discuss some of the possible causes of deviation from this expected behavior. (a) (b) Figure 8: Comparison of the kinetics of collagen adsorption as monitored by an acoustic (SAW) method and an optical (SPR) method, for bulk concentrations of (a) 30 µg/ml and (b) 300 µg/ml. Notice that the same hydrophobic-thiols coated gold area of monitored by both techniques. The kinetics as observed by acoustic methods appears faster than that observed by optical methods: as explained in the text, the mass of solvent swollen layer is overestimated by acoustic methods and underestimated by optical methods, leading to differences in the observed kinetics during layer formation. Inset: normalized SAW response as a function of the normalized SPR response. A rigid interaction in which both signal are proportional to the adsorbed protein quantity would lead to all points lying on the line of slope 1 (see text). If we first assume that the SAW signal shift is proportional to the bound mass -- including the trapped solvent -- and that the SPR signal is proportional to the added protein quantity, then ∆φ ∝ d × ρ(x) and ∆θ ∝ d × n(x) where d is the common layer thickness measured by both techniques. By introducing these proportionality relationships in Eq. 1, we obtain normalized quantities expressed as ρ(x)−ρ0 for the SPR angle, where ρf−ρ0 for the SAW phase and n(x)−n0 nf−n0 17 100020003000400050006000700000.20.40.60.81normalized response (a.u.)time (s)SAWphase SPRangle 00.10.20.30.40.50.60.70.80.9100.10.20.30.40.50.60.70.80.91SAW normalized phase shiftSPR normalized angle shift10002000300040005000600000.20.40.60.811.2time (s)normalized response (a.u.)SPRangle SAWphase 00.10.20.30.40.50.60.70.80.9100.10.20.30.40.50.60.70.80.91SAW normalized phase shiftSPR normalized angle shift an f index indicates the value of the quantity at the date at which saturation is reached and index 0 indicated the beginning of the adsorption. If we furthermore assume, as has been done previously in the quantita- tive analysis of the data, that ρlayer = x · ρprotein + (1 − x) · ρwater and nlayer = x · nprotein + (1 − x) · nwater, and considering additionally that ρ0 = ρwater and n0 = nwater, we finally find that both normalized quantities express as x·(ρprotein−ρwater) . Since both techniques, SAW and SPR, are probing the same surface, not only do they measure a common layer thickness, but also a common protein to water ratio, so that a common final value xf can be attributed to both nf and ρf , and finally the two expressions simplify and give x·(ρprotein−ρwater) for the normalized SAW data and x·(nprotein−nwater) xf·(nprotein−nwater) = x xf and x·(nprotein−nwater) xf·(ρprotein−ρwater) = x xf ρf−ρwater nf−nwater . Figure 9: Normalized signal from acoustic (SAW) and optical (SPR) measurements: if both techniques were measuring the same adsorbed mass property, the data would be lying on the y = x line, as seen for the rigid S-layer. For collagen and fibrinogen, all datasets lie above y = x. In conclusion of this calculation, we see that following two assumptions about the proportionality of the SAW and SPR signals with the physical properties of the layer and the evolution of the physical property of the protein layer as a function of water content, we reach the same law for the evolution of the 18 00.10.20.30.40.50.60.70.80.9100.10.20.30.40.50.60.70.80.91SAW normalized phase shiftSPR normalized angle shift30 µg/ml CH1 (collagen) 30 µg/ml CH2 (collagen)300 µg/ml CH1 (collagen)300 µg/ml CH2 (collagen)slope=1 S−layer normalized SAW and SPR signals. We test this result by eliminating the time parameter and plotting the SAW normalized phase shift as a function of SPR normalized angle shift. We predict that all points should lie on the line of slope 1. The S-layer data shown previously [52] do not show any significant difference in the kinetics as detected by the optical and acoustic methods, and hence indeed all normalized points fall close to this line (Fig. 9). On the other hand the solvent swollen fibrinogen and collagen adsorbed layers presented here seem to display a faster kinetics as detected by the acoustic method than the one detected by the optical method. Therefore we see in inset of Fig. 8 (a) and (b) that all points lie above the line of slope 1. Such a graph provides a quantitative relationship between measurements obtained by acoustic and optical methods. From the calculation we just presented, we can attribute the discrepancy of the normalized point distribution to either of the two assumptions: either the SAW signal is not strictly proportional to the bound mass (assumption ∆φ ∝ d × ρlayer) or the simple linear law relating the physical properties of the layer (density and optical index) to the physical properties of the individual components of the layer -- protein and water -- is not accurate. In view of the previously published literature [11, 53], we favor the first possibility and propose a possible contribution of the hydrodynamic interaction of the collagen layer to the SAW signal. The acoustic signal overestimates the bound mass by being affected by the bound solvent as well, while the optical method under-estimates the bound mass due to the lowered optical index of the adsorbed layer associated with the bound solvent. Throughout this analysis, neither the viscoelasticity of the adsorbed layer or of the surrounding solvent, nor the experimentally measured acoustic losses, have been considered. The shortcoming of these simplifications might be solved by adding a set of charts relating the frequency shift with layer thickness and viscosity, and most significantly insertion loss with these parameters. Such an investigation was endeavoured in [39, 67], in the former reference for solvent only (no adsorbed layer) and in the latter for a single layer in air, since the coupled phase/magnitude shift induced by both mass loading and viscoelastic coupling makes the separation of each contribution complex when considering finite thickness layers coated by a viscous solvent. The insertion loss evolution observed between the beginning and end of each experiment has been indicated in the caption of each experimental SAW chart result for the reader to assess the validity of this assumption: while a minute insertion loss change in the case of fibrinogen (0.5 dB) hints at the validity of our assumptions, the large (3.3 -- 6.5 dB) insertion loss observed during collagen adsorption hints at the need for a more complex analysis to grasp all the subtelties of the adsorption process. 5. Conclusion We have shown how a two steps measurement, first using copper electrode- position for instrument calibration followed by a measurement on the actual biological layer under investigation, on a combined SAW/SPR instrument leads to an estimate of the water content and layer thickness. While each individual 19 technique cannot achieve such identification, both common parameters (wa- ter content ratio and layer thickness) can be identified with the combination. Such results are especially useful for protein layers with high water content for which the hydrodynamic interaction can become predominant over rigid ad- sorbed mass. We have verified the calibration process using CTAB, confirming the 120 ng/cm2 surface density due solely to the densely organized monolayer, and illustrated the technique with an estimate of the water content of colla- gen, leading to a 70%-water/30%-protein result, which is in agreement with the large dissipation increase (quality factor decrease) observed by QCM-D. Furthermore, QCM-D overtone normalization factor with such solvent swollen layers is the inverse of the square root of the overtone number (and not the inverse of the overtone number as would be the case for a rigid layer). Apply- ing the proportionality factor between QCM frequency shift and surface mass density as predicted by the Sauerbrey relationship, assuming a rigidly bound mass, leads to a strong overestimate of the adsorbed mass. We have also shown that fibrinogen layers have a water content of 50±10 %, compatible with a lower dissipation factor increase in the QCM-D measurements. Finally, we have dis- cussed the comparison of kinetics as observed by optical and acoustic techniques and propose that the discrepancies observed in the case of solvent swollen layers might be attributed to minor hydrodynamic interactions of the SAW with the protein layer. Introducing in this analysis the SAW insertion losses would allow to resolve one more unknown, namely the adsorbed layer dynamic viscosity. 6. Acknowledgments All literature references were fetched on the Library Genesis at gen.lib. rus.ec, whose service is invaluable to our research activities. Funding was pro- vided by European Comission Framework Program grants FP5-IST PAMELA and FP7-ICT LoveFood. References References [1] D. Keller, Scanning force microscopy in biology, in: A. Baszkin, W. Norde (Eds.), Physical Chemistry of Biological Interfaces, Marcel Dekker Inc., New York, 2000. [2] J. Lu, R. Thomas, The application of neutron and x-ray specular reflec- tion to proteins at interfaces, in: A. Baszkin, W. Norde (Eds.), Physical Chemistry of Biological Interfaces, Marcel Dekker Inc., New York, 2000. [3] M. Plunkett, P. Claesson, M. Ernstsson, M. Rutland, Comparison of the adsorption of different charge density polyelectrolytes: microbalance and X-ray photoelectron spectroscopy study, Langmuir 19 (2003) 4678 -- 4681. 20 [4] B. Cavi´c, G. Hayward, M. Thompson, Acoustic waves and the study of biochemical macromolecules and cells at the sensor-liquid interface, Analyst 124 (1999) 1405 -- 1420. [5] M. Rodahl, P. Dahlqvist, F. Hook, B. Kasemo, The quartz crystal microbal- ance with dissipation monitoring, in: E. G. E, Lowe (Eds.), Biomolecular Sensors, Taylor & Francis, London, 2002, pp. 304 -- 316. [6] E. Gizeli, Acoustic transducers, in: E. G. E, Lowe (Eds.), Biomolecular Sensors, Taylor & Francis, London, 2002, pp. 176 -- 206. [7] R. Cush, J. Cronin, W. Stewart, C. Maule, J. Molloy, N. Goddard, The resonant mirror: a novel optical biosensor for direct sensing of biomolecular interactions, Biosensors and Bioelectronics 8 (1993) 347 -- 353. [8] J. Voros, J. Ramsden, G. Cs´ucs, I. Szendro, S. D. Paul, M. Textor, N. Spencer, Optical grating coupler biosensors, Biomaterials 23 (2002) 3699 -- 3710. [9] B. Liedberg, C. Nylander, I. Lundstrom, Surface plasmon resonance for gas detection and biosensing, Sensors and Actuators 4 (1983) 299 -- 304. [10] B. Ivarsson, M. Malmqvist, Surface plasmon resonance: development and use of BIACORE instruments for biomolecular interaction analysis, in: E. Gizeli, C. Lowe (Eds.), Biomolecular Sensors, Taylor & Francis, Lon- don, 2002, pp. 241 -- 268. [11] F. Hook, B. Kasemo, T. Nylander, C. Fant, K. Sott, H. Elwing, Variations in coupled water, viscoelastic properties, and film thickness of a Mefp-1 pro- tein film during adsorption and cross-linking: a quartz crystal microbalance with dissipation monitoring, ellipsometry, and surface plasmon resonance study, Anal. Chem. 73 (2001) 5796 -- 5804. [12] L. Bailey, D. Kambhampati, K. Kanazawa, W. Knoll, C. Frank, Using surface plasmon resonance and the quartz crystal microbalance to monitor in situ the interfacial behavior of thin organic films, Langmuir 18 (2002) 479 -- 489. [13] A. Laschitsch, B. Menges, D. Johannsmann, Simultaneous determination of optical and acoustic thicknesses of protein layers using surface plasmon resonance spectroscopy and quartz crystal microweighing, Appl. Phys. Lett. 77 (14) (2000) 2252 -- 2254. [14] A. Bund, A. Baba, S. Berg, D. Johannsmann, J. Lubben, Z. Wang, W. Knoll, Combining surface plasmon resonance and quartz crystal mi- crobalance for the in-situ investigation of the electropolymerization and doping/dedoping of poly(pyrrole), J. Phys. Chem. B 107 (2003) 6743 -- 6747. 21 [15] M. Plunkett, Z. Wang, M. Rutland, D. Johannsmann, Adsorption of pNI- PAM layers on hydrophobic gold surfaces measured in situ by qcm and spr, Langmuir 19 (2003) 6837 -- 6844. [16] C. Keller, K. Glasmastar, V. Zhdanov, B. Kasemo, Formation of supported membranes from vesicles, Physical Review Letters 84 (23) (2000) 5443. [17] A. Bund, A. Baba, S. Berg, D. Johannsmann, J. Lubben, Z. Wang, W. Knoll, Combining surface plasmon resonance and quartz crystal mi- crobalance for the in situ investigation of the electropolymerization and doping/dedoping of poly (pyrrole), The Journal of Physical Chemistry B 107 (28) (2003) 6743 -- 6747. [18] E. Reimhult, C. Larsson, B. Kasemo, F. Hook, Simultaneous surface plas- mon resonance and quartz crystal microbalance with dissipation monitoring measurements of biomolecular adsorption events involving structural trans- formations and variations in coupled water, Analytical chemistry 76 (24) (2004) 7211 -- 7220. [19] E. Reimhult, M. Zach, F. Hook, B. Kasemo, A multitechnique study of liposome adsorption on au and lipid bilayer formation on sio2, Langmuir 22 (7) (2006) 3313 -- 3319. [20] Y. Zong, F. Xu, X. Su, W. Knoll, Quartz crystal microbalance with inte- grated surface plasmon grating coupler, Analytical chemistry 80 (13) (2008) 5246 -- 5250. [21] C. Larsson, M. Rodahl, F. Hook, Characterization of dna immobiliza- tion and subsequent hybridization on a 2d arrangement of streptavidin on a biotin-modified lipid bilayer supported on sio2, Analytical Chemistry 75 (19) (2003) 5080 -- 5087. [22] J. Malmstrom, H. Agheli, P. Kingshott, D. S. Sutherland, Viscoelastic mod- eling of highly hydrated laminin layers at homogeneous and nanostructured surfaces: quantification of protein layer properties using QCM-D and SPR, Langmuir 23 (19) (2007) 9760 -- 9768. [23] P. Ansorena, A. Zuzuarregui, E. P´erez-Lorenzo, M. Mujika, S. Arana, Com- parative analysis of qcm and spr techniques for the optimization of immo- bilization sequences, Sensors and Actuators B: Chemical 155 (2) (2011) 667 -- 672. [24] R. Konradi, M. Textor, E. Reimhult, Using complementary acoustic and optical techniques for quantitative monitoring of biomolecular adsorption at interfaces, Biosensors 2 (4) (2012) 341 -- 376. [25] M. Edvardsson, S. Svedhem, G. Wang, R. Richter, M. Rodahl, B. Kasemo, Qcm-d and reflectometry instrument: applications to supported lipid struc- tures and their biomolecular interactions, Analytical chemistry 81 (1) (2008) 349 -- 361. 22 [26] R. P. Richter, A. R. Brisson, Following the formation of supported lipid bilayers on mica: a study combining AFM, QCM-D, and ellipsometry, Biophysical journal 88 (5) (2005) 3422 -- 3433. [27] F. Bender, P. Roach, A. Tsortos, G. Papadakis, M. I. Newton, G. McHale, E. Gizeli, Development of a combined surface plasmon resonance/surface acoustic wave device for the characterization of biomolecules, Measurement Science and Technology 20 (12) (2009) 124011. [28] M. Rodahl, B. Kasemo, A simple setup to simultaneously measure the resonant frequency and the absolute dissipation factor of a quartz crystal microbalance, Rev. Sci. Instrum. 67 (1996) 3238 -- 3241. [29] M. Rodahl, F. Hook, C. Frederiksson, C. Keller, A. Krozer, P. Brzezinski, M. Voinova, B. Kasemo, Simultaneous frequency and dissipation factor QCM measurements of biomolecular adsorption and cell adhesion, Faraday Discuss. 107 (1997) 229 -- 246. [30] M. Voinova, M. Rodahl, M. Jonson, B. Kasemo, Viscoelastic acoustic re- sponse of layered polymer films at fluid-solid interfaces: continuum me- chanics approach, Physica Scripta 59 (1999) 391 -- 396. [31] M. Rodahl, B. Kasemo, On the measurement of thin liquid overlayers with the quartz-crystal micobalance, Sensors and Actuators A 54 (1996) 448 -- 456. [32] J. Ricco, S. Martin, T. Zipperian, surface acoustic wave gas sensor based on film conductivity changes, Sensors and Actuators 18 (1985) 319 -- 333. [33] G. McHale, F. Martin, M. Newton, Mass sensitivity of acoustic wave devices from group and phase velocity measurements, J. Appl. Phys. 92 (6) (2002) 3368 -- 3373. [34] E. Gizeli, Design considerations for the acoustic waveguide biosensor, Smart. Mater. Struct. 6 (1997) 700 -- 706. [35] A. Wang, J. Cheeke, C. Jen, Sensitivity analysis for Love mode acoustic gravimetric sensors, Appl. Phys. Lett. 64 (22) (1994) 2940 -- 2942. [36] J. Du, G. Harding, J. Ogilvy, P. Dencher, M. Lake, A study of love-wave acoustic sensors, Sensors and Actuators A 56 (1996) 211 -- 219. [37] B. Jakoby, M. Vellekoop, Viscosity sensing using a love-wave device, Sensors and Actuators A 68 (1998) 275 -- 281. [38] F. Herrmann, D. Hahn, S. Buttgenbach, Separate determination of liquid density and viscosity with sagitally corrugated love-mode sensors, Sensors and Actuators 78 (1999) 99 -- 107. 23 [39] L. E. Fissi, J.-M. Friedt, S. Ballandras, Modeling the rf acoustic behav- ior of love-wave sensors loaded with organic layers, in: IEEE Ultrasonic Symposium, New York, USA, 2007, pp. 484 -- 487. [40] L. di Marzo, W. Hunter, R. Schultz, J. Risavy, In vivo study of collagen- impregnated double velour prosthesis, Vascular and Endovascular Surgery 23 (3). [41] N. Mar´ın-Pareja, M. Cantini, C. Gonz´alez-Garc´ıa, E. Salvagni, M. Salmer´on-S´anchez, M.-P. Ginebra, Different organization of type I col- lagen immobilized on silanized and nonsilanized titanium surfaces affects fibroblast adhesion and fibronectin secretion, Applied Materials & Inter- faces. [42] H. Lodish, A. Berk, S. Zipursky, P. Matsudaira, D. Baltimore, J. Darnell, Molecular Cell Biology. 4th edition, W. H. Freeman, New York, 2000. [43] F. Denis, P. Hanarp, D. Sutherland, J. Gold, C. Mustin, P. Rouxhet, Y. Dufrene, Protein adsorption on model surfaces with controlled nanoto- pography and chemistry, Langmuir 18 (3) (2002) 819 -- 828. [44] V. D. Cupere, J. V. Wetter, P. Rouxhet, Nanoscale organization of collagen and mixed collagen-pluronic adsorbed layers, Langmuir 19 (2003) 6957 -- 6967. [45] Y. Dufrene, T. Marchal, P. Rouxhet, Probing the organization of adsorbed protein layers: complementarity of atomic force microscopy, X-ay photo- electron spectroscopy and radiolabeling, Applied Surface Science 144 -- 145 (1999) 638 -- 643. [46] B. Drake, C. Prater, A. Weisenhorn, S. Gould, T. Albrecht, C. Quate, D. Cannell, H. Hansma, P. Hansma, Imaging crystals, polymers, and pro- cesses in water with an atomic force microscope, Science 243 (1989) 1586 -- 1589. [47] R. Wigren, H. Elwing, R. Erlandsson, S. Welin, I. Lundstrom, Structure of adsorbed fibrinogen obtained by scanning force microscopy, FEBS Letters 280 (1991) 225 -- 228. [48] P. Cacciafesta, A. Humphris, K. Jandt, M. Mile, Human plasma fibrinogen adsorption on ultraflat titanium oxide surfaces studied with atomic force microscopy, Langmuir 16 (2000) 8167 -- 8175. [49] K. Choi, J.-M. Friedt, F. Frederix, A. Campitelli, G. Borghs, Simultaneous atomic force microscope and quartz crystal microbalance measurements: Investigation of human plasma fibrinogen adsorption, Appl. Phys. Lett. 81 (2002) 1335 -- 1337. 24 [50] B. Snopok, K. Kostyukevych, O. Rengevych, Y. Shirshov, E. Venger, I. Kolesnikova, E. Lugovskoi, A biosensor approach to probe the struc- ture and function of the adsorbed proteins: fibrinogen at the gold surface, Semiconductor Physics, Quantum Electronics & Optoelectronics (National Academy of Sciences of Ukraine) 1 (1998) 121 -- 134. [51] L. Francis, J.-M. Friedt, C. Zhou, P. Bertrand, In situ evaluation of density, viscosity and thickness of adsorbed soft layers by combined surface acoustic wave and surface plasmon resonance, Analytical Chemistry 78 (12) (2006) 4200 -- 4209. [52] J.-M. Friedt, L. Francis, G. Reekmans, R. D. Palma, A. Campitelli, U. Sleytr, Simultaneous surface acoustic wave and surface plasmon res- onance measurements: electrodeposition and biological interactions moni- toring, J. Appl. Phys. 95 (4) (2004) 1677 -- 1680. [53] K. Saha, F. Bender, A. Rasmusson, E. Gizeli, Probing the viscoelastic- ity and mass of a surface-bound protein layer with an acoustic waveguide device, Langmuir 19 (4) (2003) 1304 -- 1311. [54] J.-M. Friedt, K. Choi, F. Frederix, A. Campitelli, Simultaneous atomic force microscope and quartz crystal microbalance measurements: Methodology validation using electrodeposition, J. Electrochem. Soc. 150 (2003) H229 -- H234. [55] J.-M. Friedt, L. Francis, K. Choi, A. Campitelli, Combined atomic force microscope and acoustic wave devices: Application to electrodeposition, J. Vac. Sci. Technol. A 21 (2003) 1500 -- 1505. [56] L. Francis, J.-M. Friedt, C. Bartic, A. Campitelli, A SU8 liquid cell for surface acoustic waves biosensors, in: SPIE Photonics Europe, Strasbourg (France), 2004, pp. 353 -- 363. [57] M. Puiu, A.-M. Gurban, L. Rotariu, S. Brajnicov, C. Viespe, C. Bala, Enhanced sensitive love wave surface acoustic wave sensor designed for immunoassay formats, Sensors 15 (5) (2015) 10511 -- 10525. [58] K. Eskilsson, V. Yaminsky, Deposition of monolayers by retraction from so- lution: ellipsometric study of cetyltrimethylammonium bromide adsorption at silica-air and silica-water interfaces, Langmuir 14 (1998) 2444 -- 2450. [59] P. Grossel, J.-M. Vigoureux, F. Baıda, Nonlocal approach to scattering in a one-dimensional problem, Phys. Rev. A 50 (5) (1994) 3627 -- 3637. [60] A. Ulman, An introduction to ultrathin organic films (from Langmuir- Blodgett to Self Assembly), Academic Press Inc.: San Diego, 1991. 25 [61] F. Kardous, L. E. Fissi, J.-M. Friedt, F. Bastien, W. Boireau, R. Yahiaoui, J.-F. Manceau, S. Ballandras, Integrated active mixing and biosensing us- ing low frequency vibrating mixer and Love-wave sensor for real time de- tection of antibody binding event, Journal of Applied Physics 109 (2011) 094701. [62] E. Palik, Handbook of optical constants of solids, Academic Press Inc.: New York, 1997. [63] F. Hook, J. Voros, M. Rodahl, R. Kurrat, P. Boni, J. Ramsden, M. Textor, N. Spencer, P. Tengvall, J. Gold, B. Kasemo, A comparative study of pro- tein adsorption on titanium oxide using in situ ellipsometry, optical waveg- uide lightmode spectroscopy, and quartz crystal microbalance/dissipation, Colloids and Surfaces B 24 (2002) 155 -- 170. [64] R. Marsch, R. Jones, M. Sferrazza, Adsorption and displacement of a glob- ular protein on hydrophilic and hydrophobic surfaces, Colloids and Surfaces B 23 (2002) 31 -- 42. [65] U. Kreibig, M. Vollmer, Optical Properties of Metal Clusters, Springer Series in Material Science (Springer-Verlag: New York), 1995. [66] F. Caruso, D. Furlong, K. Ariga, I. Ichinose, T. Kunitake, Characteriza- tion of polyelectrolyte-protein multilayer films by atomic force microscopy, scanning electron microscopy, and fourier transform infrared reflection- absorption spectroscopy, Langmuir 14 (1998) 4559 -- 4565. [67] L. E. Fissi, J.-M. Friedt, S. Ballandras, L. Robert, F. Ch´erioux, Acous- tic characterization of thin polymer layers for love mode surface acoustic waveguide, in: IEEE International Frequency Control Symposium, Hon- olulu, USA, 2008, pp. 711 -- 716. 26
1903.08529
2
1903
2019-07-03T13:40:55
Computational modeling of active deformable membranes embedded in 3D flows
[ "physics.bio-ph", "physics.comp-ph", "physics.flu-dyn" ]
Active gel theory has recently been very successful in describing biologically active materials such as actin filaments or moving bacteria in temporally fixed and simple geometries such as cubes or spheres. Here we develop a computational algorithm to compute the dynamic evolution of an arbitrarily shaped, deformable thin membrane of active material embedded in a 3D flowing liquid. For this, our algorithm combines active gel theory with the classical theory of thin elastic shells. To compute the actual forces resulting from active stresses, we apply a parabolic fitting procedure to the triangulated membrane surface. Active forces are then dynamically coupled via an Immersed-Boundary method to the surrounding fluid whose dynamics can be solved by any standard, e.g. Lattice-Boltzmann, flow solver. We validate our algorithm using the Green's functions of [Berthoumieux et al. New J. Phys. 16, 065005 (2014)] for an active cylindrical membrane subjected (i) to a locally increased active stress and (ii) to a homogeneous active stress. For the latter scenario, we predict in addition a so far unobserved non-axisymmetric instability. We highlight the versatility of our method by analyzing the flow field inside an actively deforming cell embedded in external shear flow. Further applications may be cytoplasmic streaming or active membranes in blood flows.
physics.bio-ph
physics
Computational modeling of active deformable membranes embedded in three-dimensional flows Christian Bächer *1 and Stephan Gekle †1 1Biofluid Simulation and Modeling, Theoretische Physik VI, Universität Bayreuth, Bayreuth, Germany 9 1 0 2 l u J 3 ] h p - o i b . s c i s y h p [ 2 v 9 2 5 8 0 . 3 0 9 1 : v i X r a Abstract Active gel theory has recently been very successful in describing biologically active materials such as actin fila- ments or moving bacteria in temporally fixed and simple geometries such as cubes or spheres. Here we develop a computational algorithm to compute the dynamic evolution of an arbitrarily shaped, deformable thin membrane of active material embedded in a 3D flowing liquid. For this, our algorithm combines active gel theory with the classical theory of thin elastic shells. To compute the actual forces resulting from active stresses, we apply a parabolic fitting procedure to the triangulated membrane surface. Active forces are then dynamically coupled via an Immersed-Boundary method to the surrounding fluid whose dynamics can be solved by any standard, e.g. Lattice-Boltzmann, flow solver. We validate our algorithm using the Green's functions of [Berthoumieux et al. New J. Phys. 16, 065005 (2014)] for an active cylindrical membrane subjected (i) to a locally increased active stress and (ii) to a homogeneous active stress. For the latter scenario, we predict in addition a so far unobserved non-axisymmetric instability. We highlight the versatility of our method by analyzing the flow field inside an actively deforming cell embedded in external shear flow. Further applications may be cytoplasmic streaming or active membranes in blood flows. Published as: Phys. Rev. E 99(6), 062418 (2019) 1 Introduction Many biological processes such as cell division or locomotion depend on the ability of living cells to convert chemical energy into mechanical work [1]. A prominent mechanism to achieve such a conversion are motor proteins which perform work through a relative movement of cross-linked cytoskeletal filaments. This movement induces active stresses in the cell cortex [2 -- 7] which are transmitted via anchor proteins to the plasma membrane separating the interior of the cell from its surroundings. Active stresses have an inherent non-equilibrium character and are the basis of physically unique processes in active fluid layers such as instabilities [8, 9], the emergence of spontaneous flows [10] or the creation of geometrical structures [11, 12]. They are furthermore responsible for large shape deformations during cell morphogenesis [13 -- 15], cell division [16 -- 21], cell locomotion [22 -- 26], cell rheology [27, 28] or spike formation on artificial vesicles [29]. In recent years a theoretical framework has been developed describing cytoskeleton and motor proteins together as an active continuous gel [30 -- 39]. This active gel theory in general treats the cell cortex as a viscoelastic medium with additional active contributions. On time scales short compared to the viscoelastic relaxation time active gel theory can be formulated in the elastic limit [3]. For thin active 2D membranes embedded in 3D space, active gel *[email protected][email protected] 1 theory can be reformulated into force balance equations using the formalism of differential geometry [3, 6]. Any active stress in the membrane is then balanced by a counter-stress, usually due to viscous friction, from the external medium. The force balance equations can be applied on fixed membrane geometries such as spheres, cylinders or flat layers. This often results in the prediction of regions in parameter space where the prescribed shape is expected to become unstable [11, 16, 29, 40 -- 42]. Despite being a powerful qualitative tool, such calculations cannot make statements about the precise shape of the active membrane after the instability has set in. To obtain actual shape predictions, a number of works start instead from a parametrized, free membrane shape which is adjusted so as to fulfill the force balance equations for a given set of parameters and boundary conditions. This procedure can be carried out either analytically [3, 43] or numerically [15, 20, 21, 44 -- 46]. For example, Berthoumieux et al. [3] derived Green's functions to predict the deformation of an infinitely long cylindrical ac- tive, elastic membrane resulting from the application of a point active stress. For certain parameter ranges, these Green's functions exhibit divergences which have been interpreted as mechanical (buckling or Rayleigh-Plateau- like) instabilities of the cylindrical membrane. Sain et al. [21] investigated the axisymmetric dynamics of the furrow in cytokinesis. Turlier et al. [20] computed the time evolution of an axisymmetric membrane undergoing cytokinesis by advecting tracer points discretizing the membrane. Callan-Jones et al. [15] predicted a transition to a polarized cell shape because of an instability of the cell cortex. Reymann et al. [45] matched axisymmetric, theoretical results to observed cell shapes during cytokinesis using measured velocity and order parameter fields as input for the theory. Heer et al. [47] determined the equilibrium shape of an elastic tissue layer folded into a de- formable ellipsoidal shell, where myosin activity is incorporated as a preferred curvature of the shell. Klughammer et al. [48] analytically calculated the flow inside a sphere that slightly deforms due to a traveling band of surface tension, which mimics cortical active tension, under the assumption of rotational symmetry. In all of these works the full dynamics of the external fluid was neglected and nearly all of them restricted themselves to deformations from simple rest shapes in the steady state. There currently exists no analytical or numerical method to compute the dynamical deformation of an arbitrarily shaped active membrane immersed in a 3D moving Newtonian fluid. In this work, we develop a computational algorithm to predict the dynamic deformation of arbitrarily shaped thin membranes discretized by a set of nodes connected via flat triangles. Starting from prescribed active stresses on the discretized membrane our algorithm computes the corresponding forces on every membrane node via a parabolic fitting procedure taking into account the full deformed surface geometry. Knowledge of the nodal forces then enables the dynamic coupling to a surrounding 3D fluid via the Immersed-Boundary Method (IBM). The Navier-Stokes dynamics for the surrounding fluid is solved here by the Lattice-Boltzmann method (LBM), but other flow solvers can straightforwardly be incorporated. With this, our method allows to study the dynamic evolution of active membranes in general external flows. It thus builds a bridge between the extensive literature on active fluids and the similarly extensive work on elastic cells, vesicles and capsules in flows [49 -- 56]. In biological situations often a flowing environment is present rendering the dynamic coupling of external fluid and membrane deformations necessary. Our proposed method allows such a coupling and thus the computation of dynamically evolving non-equilibrium shapes. Possible applications of our method include the study of active cell membranes inside the blood stream or active cellular compartments in cytoplasmic streaming flows. First in section 2, we extensively describe the LBM and IBM for a dynamic coupling of an elastic membrane and a suspending fluid. In section 3 we start with the problem formulation in the framework of thin shell theory using differential geometry. Next, we describe our algorithm for three dimensional active force calculation in section 4. In section 5 we provide an in-depth validation of our algorithm based on the analytical results for a cylindrical active membrane in case of infinitesimal deformations by Berthoumieux et al. [3]. As an application, section 6.1 presents the flow field inside a dividing elastic cell, where active stresses trigger membrane deformations that in turn lead to fluid flow. In section 6.2 we analyze the same system in externally driven shear flow. Eventually, we conclude our work in section 7. 2 Dynamic coupling of membrane and fluid 2.1 Lattice-Boltzmann method for fluid dynamics The Lattice-Boltzmann method (LBM) is an efficient and accurate flow solver which is well described in the literature [57 -- 60]. In the following we therefore summarize only the basic concepts. In contrast to macroscopic, 2 e.g. finite element, methods based directly on the discretized Navier-Stokes equation (NSE), LBM starts from the Boltzmann equation which is a common tool in statistical mechanics. Using Chapman-Enskog analysis the NSE are recovered from the Boltzmann equation [60, 61]. The Boltzmann equation provides insight into a system on mesoscopic length scales by means of the continuous particle distribution function f(r, p, t) where r, p and t refer to position of the particles, momentum of the particles and time, respectively. The expression f(r, p, t) dr dp dt gives the probability to find a particle (fluid molecule) in the phase-space volume dr dp at (r, p) in the time interval t to t + dt. The dynamic evolution of the particle distribution function f(r, p, t) is given by df dt + v · ∇rf + F · ∇pf = Ω = ∂f ∂t (1) with Ω being the collision operator accounting for re-distribution of molecules due to collisions. Discretization of space, momentum space and time leads from the Boltzmann equation (1) to the Lattice- Boltzmann equation. The discretization of the spatial domain is carried out using a cubic Eulerian grid. The distance between the fluid nodes is ∆x = 1 in simulation units. In contrast to other methods, LBM also discretizes the momentum (velocity) space, i.e., f = f(xj, pi, t) such that only a discrete set of velocities is allowed at each node. Here, we use the common D3Q19 scheme with 19 discrete velocity vectors, which is illustrated in figure 1 a). Thus, each node contains one population f(xj, pi, t) for each momentum pi, which moves with the corre- sponding velocity ci away from the node xj within one time step. As an abbreviation the distribution functions are labeled by an index according to their discrete momentum/velocity, i.e., f(xj, pi, t) = fi(xj, t). Finally, dis- crete time steps from t to t + ∆t are considered. Under discretization the Boltzmann equation in (1) becomes the Lattice-Boltzmann equation [60] fi(xj + ci∆t, t + ∆t) = fi(xj, t) + Ωi(xj, t)∆t. (2) The numerical integration of the Lattice-Boltzmann equation in time is split into two steps, collision and propaga- tion. Collision is done by an approximation of the collision operator in equation (2). The idea for the approximation of the collision operator is that the populations should relax towards the Maxwellian equilibrium distribution func- tion in absence of driving forces. Here, we use the multiple relaxation time scheme. In the framework of the multiple relaxation time scheme the relaxation is done in moment space of the populations with individual relax- ation rates for the different moments [58, 60]. Moments corresponding to conserved density and momentum do not have a relaxation time, while two relaxation rates related to bulk and shear viscosity, respectively, are chosen for moments corresponding to the stress tensor. For further discussion on the relaxation rates and moments used in our LBM implementation we refer to refs. [58, 62]. In the second step, the streaming, the populations after collision propagate according to the associated velocities. We note that our LBM implementation supports adding thermal fluctuations to the fluid dynamics mimicking a given thermal energy kBT corresponding to fluctuating hydrodynamics [62]. Thermal fluctuations are taken into account by adding a random noise to those relaxed moments of the multiple relaxation time scheme which correspond to elements of the stress tensor [58, 62]. We note that our approach using thermal fluctuations is different from considering a separate temperature field [63 -- 65] as e.g. required for convection flows [63]. Although they are not per se necessary, we employ thermal fluctuations in the present manuscript to speed up the onset of instabilities. Although we do not consider solid walls in the present manuscript, we note that they can be realized by bounce back boundary conditions [60], where populations streaming towards nodes beyond a boundary are simply re- flected. In case of moving elastic objects (2.2) having close contact to solid walls as it is considered e.g. in ref. [66], at least one fluid node between the object and the solid boundary is required in our method. Typical LBM grids used in this work have dimensions of 126x72x72 or 100x100x100. Except in the case of the shear flow with moving boundaries in section 6.2, our LBM simulations use periodic boundary conditions in figure 6 consists of 107 time steps. Onset of Rayleigh-Plateau for the fluid. A typical simulation run, e.g. like instabilities in figure 7 typically appears after approximately 8 × 106 time steps. Simulations are performed with the simulation package ESPResSo [67 -- 69] which has been extended to include thin membranes using the Immersed Boundary Method described next. 3 a) b) Fig. 1: a) D3Q19 LBM scheme with the discrete velocity set (solid arrows) connecting nearest and next nearest neighboring fluid nodes (dots). b) 3D illustration of the Immersed Boundary Method. A continuous membrane is discretized by Lagrangian nodes (red and orange dots) that are connected by triangles. The membrane is immersed in an Eulerian grid representing the fluid (blue dots). The velocity at a Lagrangian node is obtained by interpolation from the eight closest fluid nodes (illustrated for the two orange membrane nodes in the middle by the blue shaded cubes). The same stencil is used to transmit the forces from the membrane to the fluid. 2.2 Immersed-Boundary method for membrane dynamics The framework of the Immersed-Boundary Method (IBM) [70 -- 72] allows for the coupling of a cellular membrane to the suspending fluid, where the fluid is simulated using LBM as described in the previous section 2.1. The IBM consists of two central steps: elastic forces acting on the membrane are transmitted to the fluid and due to no-slip boundary condition the mass-less membrane is advected with the local fluid velocity. The membrane of a cell is represented by an infinitely thin elastic sheet in the framework of thin shell theory [73 -- 75]. In our numerical simulations the membranes is discretized by a set of nodes that are connected by flat triangles [51, 76 -- 78]. These represent the elastic membrane as a Lagrangian mesh immersed into the Eulerian fluid mesh as illustrated in figure 1 b). Physical behavior of the membrane is characterized by appropriate constitutive laws which are described in sec- tion 3. The resulting forces are calculated on the deformed Lagrangian membrane mesh as described in section 4. To transmit these forces into the fluid, the incompressible NSE for the fluid velocity u(x, t) becomes [77] ∂u ∂t + (u · ∇) u = − 1 ρ∇p + ν∆u + 1 ρZ f(cid:0)X0, t(cid:1) δ(cid:0)X0 − x(cid:1) d2X0, with p the pressure, ν the kinematic viscosity, ρ the density, and f the membrane force per area acting on the membrane, which is parametrized by X0. The Dirac δ function ensures that the forces only act at the position of the actual membrane. Correspondingly, the Lattice-Boltzmann equation (2) changes to (3) (6) Fj =Z f(cid:0)X0, t(cid:1) δ(cid:0)X0 − xj(cid:1) d2X0 fi(xj + ci∆t, t + ∆t) = fi(xj, t) + Ωi(xj, Fj, t)∆t. (4) (5) After an update of the fluid dynamics (LBM algorithm), which now include the forces from the membrane Fj, the mass-less membrane nodes are advected with the local fluid velocity thus satisfying exactly the no-slip boundary condition [71]. Moving with the local fluid velocity is expressed for a membrane node xn ∈ X0 by [70, 71] dxn dt = u (xn(t), t) =Z u(x, t)δ (x − xn) d3x. with u (xn, t) being the fluid velocity at the position of the membrane node. In simulations equation (6) is inte- grated numerically using an Euler scheme in order to move each membrane node from time t to t + ∆t. 4 xyz The core of both steps, force transmission and movement with local fluid velocity, is the interpolation between the Eulerian fluid grid and the Lagrangian membrane grid. Considering the discretization and the resulting spatial mismatch of membrane nodes and fluid nodes, as illustrated in figure 1 b), it becomes clear that an interpolation is necessary. On the one hand the ideal point force at the site of a membrane node must be spread to the adjacent fluid nodes. On the other hand to obtain the local fluid velocity at the site of a membrane node the velocity of the adjacent fluid nodes is interpolated. This means the δ distributions in equation (4) and (6) must be discretized. The interpolation between fluid and membrane is carried out by an eight-point stencil. As illustrated for two membrane nodes by the blue shaded cubes in figure 1 b), for each membrane node a cube containing the eight nearest fluid nodes is considered. A linear interpolation between the eight fluid points is performed [79]. A typical mesh for a cylinder in the present study contains 17100 Lagrangian membrane nodes and 34200 triangles. Simulating a cylindrical membrane, as it is partly shown in figure 4, is done using periodic boundary conditions for the fluid in all directions and for the membrane in the direction of the cylinder axis. The latter is achieved by connecting membrane nodes at the end of the box to those at the beginning of the box. 2.3 Validation of LBM/IBM for passive, elastic membranes Our implementation of the force calculation and the LBM/IBM have extensively been tested for passive, elastic membranes in previous publications [52, 56, 78, 80 -- 82]. The algorithm for the passive, elastic force calculation has been validated in references [52, 78] by comparison with exact results in static situations as well as for a capsule in shear flow. For red blood cells flowing through a rectangular channel very good agreement with in vitro experiments has been found [56, 80]. The mixed LBM/IBM has been extensively validated for suspensions of red blood cells and rigid particles in complex geometries. In [81] the concentration profile of cells across the channel diameter (affected by cross streamline migration, which is triggered by the passive elasticity of the cells) has been successfully compared to well-established literature results. In ref. [82] we have performed a validation based on the Zweifach-Fung effect for red blood cell suspension in branching channels. Furthermore, in the Supplementary Information of ref. [82] we have shown red blood cell shapes for a single cell in tube flow and in a rectangular channel, respectively, that are in very good agreement with previous studies using dissipative particle dynamics [83] and boundary integral method [56], respectively. The stability and behavior of stiff particles realized by IBM has been shown and validated in ref. [81] and [82] for a spherical particle based on the Stokes relation (sphere pulled through a quiescent fluid as well as a fixed sphere in homogeneous flow) and an ellipsoidal particle rotating in shear flow. In addition to the well-established passive elasticity, in this paper we introduce active elastic forces into sim- ulations of membranes, see section 4. We refer to section 5 for the validation in case of an active, elastic cell membrane. 3 Force balance in thin shell formulation including active stresses 3.1 Physical model We consider the plasma membrane and the cytoskeletal filaments within the cell cortex, sketched in figure 2, as a single physical entity which for simplicity we denote in the following as "the membrane". Since this membrane is small in height compared to the cell diameter it is described as an infinitely thin shell. Thin shell theory treats the membrane as a two dimensional manifold embedded in the three dimensional environment, e.g., the intra- and extra-cellular fluid, thereby accounting for membrane curvature. For a condensed introduction into the required differential geometry on such manifolds as well as present conventions we refer the reader to ref. [84] and ap- pendix A1 of this work, respectively. Here, we only note that Greek indices refer to coordinates on the membrane, i.e., α, β = 1, 2, and Einstein summation convention is used. In this work, we consider the short-time limit of a purely elastic membrane noting that our algorithm can, without substantial difficulty, be extended to viscous or viscoelastic membranes. The key quantity in the framework of thin shell theory are the in-plane surface stresses tαβ and moments mαβ [73, 74, 85]. The in-plane surface stress tensor or stress resultant [74] tαβ is the 3D elastic stress tensor for the material forming the membrane projected onto the membrane and integrated over the membrane thickness h [74] 5 Fig. 2: (left) The cell cortex underlying the plasma membrane consists of cytoskeletal filaments and motor proteins. The latter constantly convert energy into mechanical work. (right) Membrane and cortex are condensed into a thin shell with normal vector n and in-plane coordinates e1, e2. Mechanical work of motor proteins results in an active in-plane surface stress t β aα. Forces from the cytoplasm inside the cell and from the extra-cellular medium onto the membrane can be treated by the 3D fluid stress tensors σin ij and σout ij . (dimensions of a force per length, i.e., N/m) 1. It is a function of in-plane coordinates α, β and contains only in-plane components such that its matrix representation has dimensions 2x2. Besides the in-plane surface stresses, we introduce the normal surface stress which is sometimes denoted a shearing force [73] or transverse shear surface stress tα n [76, 78]. The in-place surface stresses tαβ contain a contribution from passive, elastic stresses as well as from active, force-generating mechanisms. Both contributions superpose linearly and thus can be treated separately [3, 6]. Passive elastic stresses can be further split into different contributions such as, e.g., shear and bending resistances [50, 78, 88]. The moment tensor or stress couples [74] mαβ accounts for stress distribution across the membrane [74] (dimensions of a torque per unit length, i.e., N). We denote the corresponding normal surface moment by mn. For explicit materials building up the membrane corresponding constitutive laws are required to derive explicit forms for in-plane surface stresses and moments. Considering the passive, elastic properties of membranes in most cases constitutive laws are formulated in terms of a strain energy functional. For many biological membranes such as red blood cells the passive, elastic forces arise from the resistance to shear deformation of the cortical, cytoskeletal network and the resistance to bending deformation and area dilatation of the plasma membrane. For shear resistance and area dilatation Skalak et al. [89] proposed an appropriate energy density functional wSK(I1, I2) = κS (7) 12(cid:16)(I2 2(cid:17) . 1 + 2I1 − 2I2) + CI2 The energy density depends on the invariants I1, I2 of the transformation between undeformed and deformed membrane [50, 89]. Both invariants are defined in equation (A2.3) and (A2.4) of the appendix. Resistance to shear is characterized by the shear modulus κS while resistance to area dilatation is characterized by CκS with C much larger than unity for nearly area-incompressible membranes. For bending resistance the Helfrich model is used [78, 90] wHF = 2κB(H − H0)2 + κKK (8) with κB being the bending modulus, H denoting the local mean curvature, H0 the spontaneous curvature, κK the Gaussian modulus and K the Gaussian curvature. On the one hand, from a given energy functional, in-plane surface stresses and moments can be deduced by functional derivatives with respect to the strain tensor [74, 87] and curvature tensor, respectively [78, 91, 92]. On the other hand, in numerical algorithms the explicit introduction of stresses is typically bypassed and forces on the nodes discretizing the membrane are often computed directly by deriving a discretized version of the energy functional, here equations (7) and (8), with respect to the node positions [50, 78, 93]. This approach is used here as well for the passive elastic forces. For bending force calculation we use the method denoted by B in ref. [52]. 1The term in-plane surface stresses mainly follows the terminology of Deserno [84]. Guven [86] and Deserno [84] denote the correspond- ing force vector with its in-plane components being tαβ as surface stress. For the same quantity Berthoumieux et al. [3] and Salbreux and Jülicher [6] as well as Guckenberger and Gekle [78] use the term in-plane tension tensor. Daddi-Moussa-Ider et al. [87] call this quantity stress tensor. Finally, Green and Adkins [74] call it stress resultant. 6 σoutσintβaαne1e2 For actively generated forces, however, this approach is not applicable since, due to their inherent non-equilibrium character, an energy functional cannot be defined in a strict sense. Instead, active contributions are usually given in terms of active stresses whose strength and direction can either be temporally constant or depend on additional quantities such a local concentration of motor proteins [6]. We will in the following construct a numerical method which explicitly applies these active stresses and derives the corresponding active forces on the discretized mem- brane nodes. These forces are then used to introduce a two-way coupling between active membrane dynamics and a surrounding hydrodynamic flow. For simplicity of demonstration and in order to connect to the analytical axisymmetric solutions of [3], we restrict ourselves to temporally constant active in-plane surface stresses. Time- dependent active stresses or active moments can be included without substantial modification of our numerical framework. 3.2 Force balance for deformable active membranes embedded in a 3D fluid As sketched in figure 2 we consider a membrane immersed in an external fluid and enclosing an internal fluid, the cytosol. Coupling of the membrane to the internal and external fluid with the stress tensor σin ij and σout ij , respectively, is described by the traction jump [51, 76, 88] − fj =(cid:16)σout ij(cid:17) ni ij − σin with Latin subscripts denoting 3D coordinates and the components ni of the local unit normal vector onto the membrane pointing towards the external fluid. Transforming to the in-plane coordinate system the traction jump can - as a vector on the membrane in general - be decomposed into tangential and normal component − f = −f αeα − f nn. (10) Neglecting membrane inertia the traction jump between internal and external fluid is balanced by membrane forces (per unit area). In the present situation, these arise from elastic and active contributions. Looking ahead, the numerical method which we will construct in the following section will compute elastic forces f α e via the classical discretized energy functional route, bypassing for simplicity the introduction of elastic stresses, while active forces need to be computed explicitly from active stresses and moments. It is thus convenient to write the force balance for a thin shell [3, 6, 73, 87, 88] as e and f n (9) (15) (16) ∇0αtαβ ∇0αtα na + f β a + C0β α tα a + f n na − C0αβtαβ a + C0β ∇0αmαβ α mα na − C0αβmαβ ∇0αmα e = f β e = f n na = 0β a = −0αβtαβ , tα α tα na a (11) (12) (13) (14) , , mα with active contributions to the in-plane and normal surface stress tαβ na and active contributions to the moments a mαβ na. The geometrical quantities are the curvature tensor C0αβ, the Levi-Civita tensor 0αβ and the co-variant a derivative ∇0α taken on the deformed membrane as defined in the appendix. On the basis of equation (11) and (12) it becomes clear that the negative traction jump is the force exerted from the membrane on the fluid. We now na = 0) consider the active contributions to the traction jump in the case of vanishing active moments (mαβ and vanishing active transverse shear surface stress (tα a = tαβ na = 0) a ,α + Γ0α a = mα a + Γ0β αγtαγ a αγtγβ a = ∇0αtαβ f β a = −C0αβtαβ f n a , where we have used the definition of the co-variant derivative on the membrane (see equation (A1.20)) in the first line. To simulate the temporal dynamics and coupling to the external fluid of our discretized active membrane, we need to compute the forces on each membrane node corresponding to active in-plane surface stresses. According to equations (15) and (16) the curvature tensor, the Christoffel symbols, and the active in-plane surface stresses together with their derivatives have to be known locally on each node on a deformed surface. In the next section we will develop an algorithm to compute these quantities numerically for the discretized thin shell embedded in a 3D environment. 7 Fig. 3: The central node at rc is surrounded by N neighbors. The discretization is chosen such that all nodes on the cylindrical mesh in figure 4 have 6 neighbors. For the ellipsoid in figure 10, topology requires at least 12 triangles with N = 5 while all others have N = 6. From the surrounding triangles a local normal vector n at rc can be computed. Together with the position of one neighbor local in-plane coordinate vectors eξ, eη can be constructed. Applying the first step of Gram-Schmidt orthonormalization leads to eξ and the cross product of n and eξ to eη. 4 Algorithm for active force calculation on arbitrarily shaped discretized membranes Our algorithm starts from active in-plane surface stresses and computes the corresponding active forces on the discretized membrane. The key ingredient of the algorithm is the discrete computation of geometrical properties on the discretized, deformed membrane, such as the curvature tensor or Christoffel symbols. This is achieved by a parabolic fitting procedure and using the force balance equations (15) and (16) as described in subsection 4.2 based on a local coordinate system which we introduce in subsection 4.1. With these forces the active membrane dynamics can be bidirectionally coupled to a surrounding fluid flow which is computed separately using either an overdamped dynamics or a Lattice-Boltzmann method. 4.1 Local coordinate system As sketched in figure 3 each membrane node rc has a neighborhood consisting of N nodes, which are labeled in an ordered fashion around the central node. The choice of the starting node is arbitrary initially, but has to be retained through the simulation. In figure A1 we provide evidence that the choice of the starting node does not affect simulation results. At the site of the central node a local coordinate system (eξ, eη, n) can be defined, where we denote the in-plane coordinate vectors (eα (α = ξ, η) in the appendix) by eξ and eη. We determine the local unit normal vector n by the mean weight by angle of the normal vectors on the surrounding triangles [94]. The first in-plane vector eξ is calculated using the vector from the central node to the first neighbor x1 = r1 − rc. The first step of the Gram-Schmidt orthogonalization is applied and the vector is normalized The second in-plane coordinate vector is calculated by the cross-product eξ = x1 − (x1 · n) n x1 − (x1 · n) n . eη = n × eξ n × eξ . (17) (18) Using this method we can assign to every node at every time step a unique coordinate system which is adapted to the local deformed membrane geometry. We denote coordinates along eξ and eη by ξ and η, respectively. In a figurative sense, we co-move with a cytoskeletal filament initially positioned at rc pointing towards x1. 4.2 Parabolic fitting To obtain local geometrical quantities such as the curvature tensor on the deformed membrane we perform a parabolic fitting procedure based on the local coordinate system derived in the previous section. An arbitrary point 8 c132456neξeη ¯r in the neighborhood of the central node rc can be expressed by a Taylor expansion around rc up to the second order. For the i-th component of the vector ¯r with i = x, y, z we obtain ¯ri(ξ, η) ≈ rc i + Aiξ + Biη + 1 2(cid:16)Ciξ2 + Diη2 + 2Eiξη(cid:17) , with ξ, η being the coordinates along eξ and eη, respectively. Using this expression we can apply a parabolic fitting procedure (see also [51, 52] where a similar procedure has been used to compute passive bending forces) considering all N neighboring nodes with a squared deviation from the fitted surface χ2 i = NXν=1 i − ¯rν (rν i )2 , (20) with i = x, y, z. By minimizing the χ2 i we obtain the coefficients Ai − Ei. Using equation (19) we are able to analytically calculate the derivative of χ2 i with respect to the coefficients Ai − Ei. The derivative of χ2 i being zero in case of minimization then leads to a linear equation system for the Ai − Ei. This linear equation system is solved numerically in the simulation using LU decomposition. The paraboloid fitted to the neighborhood around rc provides a good approximation of the local curvature for typical cell shapes [52]. By construction, the fitting coefficients equal the derivatives of the membrane parametrization vector ¯r with respect to local coordinates at the site of the central node Ai = ¯ri,ξ Bi = ¯ri,η Ci = ¯ri,ξξ Di = ¯ri,ηη Ei = ¯ri,ξη or, equivalently, A = ¯r,ξrc B = ¯r,ηrc C = ¯r,ξξrc D = ¯r,ηηrc E = ¯r,ξηrc. Thus, we are able to calculate geometrical quantities, as defined in the appendix, at the site of the central node in local coordinates with α, β, γ = ξ, η, such as the metric tensor (19) (21) (22) (23) (24) (25) (26) (27) the curvature tensor and the derivatives of the metric tensor, e.g., gα β = A · A A · B B · A B · B! , −E · n −D · n.! , Cα β = −C · n −E · n gξξ,ξ = (A · A),ξ = 2A · C, which are necessary for the calculation of the Christoffel symbols. The tensor g α β is obtained by inverting the β metric gα β since gαγgγ β = δ γ . We note that this procedure can be used without any restriction on the deformed membrane as well. Correspondingly, we obtain the metric tensor g0α β, the curvature tensor C0α β, and the Christoffel symbols on the deformed surface. β α. Thus, we can further calculate C β α = Cαγgγ β and C α β = g αγC Equation (15) for the in-plane, active force includes a partial derivative tαβ a ,α, which is calculated using another parabolic fitting procedure. As done in equation (19) for the position in local coordinates we can expand the active in-plane surface stress tαβ around the components of the central node tcαβ a ¯tαβ a (ξ, η) = tcαβ a + Aαβ a ξ + Bαβ a ξ2 + Dαβ a η2 + 2Eαβ a η + 1 2(cid:16)Cαβ a ξη(cid:17) α β Corresponding to equation (20) we consider here the squared deviation from the expanded active stress ¯t a χ2 a = (tα βν a − ¯tα βν a NXν=1 )2 9 α βν a being one component of the active in-plane surface stress of the ν-th neighboring node in local coordi- with t nates. Thus, the corresponding fitting coefficients Aa and Ba represent the partial derivative of the active in-plane surface stress component with respect to the coordinate ξ and η, respectively, whereas the coefficients Ca − Ea represent the second derivatives. We note that the matrix for the fit is the same as for the position. We perform one fitting procedure for each component, i.e., α = ξ, η and β = ξ, η. With the method proposed here, we are able to calculate the co-variant derivative, which consists of a partial derivative and Christoffel symbols, and perform the contraction of active in-plane surface stress tensor and curva- ture tensor in the local coordinate system to determine the traction jump given by equations (15) and (16), which we repeat here for the deformed membrane in the local coordinate system α β β αγtγ β a ,α + Γ0α a = t f a = −C0α βtα β f n , a a + Γ0 β αγtαγ a (28) (29) in case of vanishing active moments and vanishing transverse shear stress. We note that the in-plane surface stress is the analogue of the Cauchy stress tensor in 3D [3, 88] and thus acting on the deformed membrane [95]. Equation (28) and (29) give the discrete forces (per area) on one membrane node. These forces then enter the fluid solver as illustrated in section 2.2 or are used for the relaxation dynamics in the overdamped limit (see appendix A3). To convert the traction jump from equations (28) and (29) into a nodal force we use Meyer's mixed area approach [78, 94]: in the case of non-obtuse triangles the area is calculated using Voronoi area AVoronoi defined by AVoronoi = 1 8 NXν=1 (cot(αν) + cot(βν))xν (30) with the angles α and β opposite to the edge connecting the central node and the ν-th neighbor node within the adjacent triangles and in the case of obtuse triangles the midpoint of xν that is opposite of the obtuse angle is chosen instead of the circumcenter point for each triangle. 4.3 Specification of active in-plane surface stress Active stresses in membranes are often generated by ATP-fueled molecular motors "walking" along cross-linked rod-like structures such as actin filaments or microtubules. As a direct consequence, active stresses are often anisotropic with the direction of contractile/extensile stresses specified in a material frame moving and deforming along with the membrane itself. This naturally leads to a convenient specification of the active in-plane surface stress tensor tαβ in the local coordinate system introduced in section 4.1. Since the labeling of the neighbors a around each node remains unchanged throughout the simulation, the distance vector x1 = r1 − rc represents a material vector. Its normalized in-plane counterpart eξ, given in equation (17), together with eη, given in equa- tion (18), constitute the associated material frame. Imagining one cytoskeletal filament for an illustrative picture, the anchoring position of the filament is tracked by the node position rc, while the orientation of the filament is tracked by the fixed choice of the first neighbor r1. We again note that the choice of the starting node for the labeling, here 1, does not affect simulation results, as shown in figure A1. The active in-plane surface stress tensor tαβ a itself is not computed by our method but needs to be specified as an input quantity according to a corresponding constitutive law for the surface stress [3, 6]. Our algorithm allows for an arbitrary choice of active in-plane surface stress, subject to condition (14) for vanishing active mo- ments, including spatially heterogeneous, anisotropic or time-dependent stresses. For the latter, coupling to a convection-diffusion model of active substances such as ATP or myosin (with the magnitude of tαβ a proportional to local ATP/myosin concentration) is methodologically possible. Thus, if the concentration field of ATP/myosin is solved/prescribed on the membrane, e.g. by discretizing the convection-diffusion equation, the active stress can be calculated from the local concentration. The corresponding active forces and their coupling to the surrounding fluid dynamics are then straightforwardly achieved by the present algorithm, which allows for a spatially varying active stress. In the present work, we consider only temporally constant active stresses. Active stresses can thus conveniently be specified in the initial configuration of the membrane. For this, we first choose an intuitive coordi- nate system (e1,e2) appropriate for the initial shape of the undeformed cell membrane. The active in-plane surface 10 stresses in this coordinate system are of the form aα = t 1 a1 t 1 a2 t β a2! t 2 a1 t 2 (31) where the mixed form with upper and lower index is chosen such that the right-hand side contains physical material-specific constants with dimensions of N/m [3]. Because of the dynamical deformation of the membrane, the active in-plane surface stress of equation (31) needs to be mapped into the local coordinate system at each node. This can be achieved by mapping the coordinate system (e1,e2) to the local coordinate system (eξ,eη). In many situations, the intuitive coordinate system (e1,e2) will correspond to cylindrical coordinates or spherical coordinates, e.g., for a rounded cell during mitosis [14]. By construction of the intuitive in-plane coordinate system and the local coordinate system both normal vectors, e.g., er for a cylinder or a sphere and the local n are equal initially. Thus, the initial in-plane coordinate system (e1,e2) can be converted directly into the local coordinate system (eξ,eη) on every node individually tα β a = D tαβ a D−1, (32) with D being a rotation matrix around the local unit normal vector n, with α = 1, 2 and α = ξ, η. We thus obtain the active in-plane surface stress along the local coordinate vectors eξ and eη. The rotation in equation (32) is performed once at the beginning of a simulation. Since the local coordinate system eξ, eη is co-moving with the membrane material, the active in-plane surface stress tensor t , expressed in these coordinates, does not change over time. α β a The parabolic fitting procedure in equation (27) requires the difference in active stress between neighboring nodes. The differences in active stress are calculated in the intuitive coordinate system, in which the active stresses are prescribed. Because active stresses do not change in time, calculation of the differences can be done also once at the beginning of the simulation. Projection into the local coordinate system for each node is done in the same way as for active stresses in equation (32). Along a certain direction, the cytoskeletal filaments may tend to contract or to expand, respectively, resulting in a contractile or extensile active in-plane surface stress [37]. The contractile or extensile nature of the cytoskeletal filaments manifests itself in the sign of the active stress, namely a positive (negative) stress corresponds to a contractile (extensile) nature. In active matter consisting of cytoskeletal filaments, active stresses often possess different signs in different directions. Imagining two anti-parallel polar filaments that are cross-linked by a motor protein walking in opposite directions on both filaments, a relative extensile shift of both filaments together with a lateral contraction occurs. 5 Validation In this section we provide an in-depth validation of our algorithm to compute the dynamics of active membranes embedded in a 3D fluid. For this, we first compute the deformation obtained when an initially unstressed cylindrical membrane is subjected to a localized perturbation due to active stresses. Our numerical results are in excellent agreement with analytical predictions by Berthoumieux et al. [3] which were obtained using a Green's function approach in the limit of small deformations. Next, we apply homogeneous active stresses again to an initially cylindrical membrane. In agreement with the analytical predictions of [3], we observe two kinds of axisymmetric instabilities. Going beyond the axisymmetric calculations of [3], our numerical method then predicts the existence of a third non-axisymmetric instability which, to the best of our knowledge, has not been observed thus far. To account for the dynamics of the surrounding fluid, we employ two qualitatively different approaches. In the first approach, we use simple overdamped dynamics such that the surrounding fluid acts purely as a viscous frictional damping, see appendix A3. In the second approach, we consider the full fluid dynamics of the surround- ing liquid by solving the Navier-Stokes equations using a Lattice-Boltzmann method, see section 2.1. Two-way coupling to the active membrane dynamics is provided by the Immersed-Boundary Method, as detailed in section 2.2. 11 5.1 Green's function formalism We here briefly recall the central analytical results of [3] which will be used to validate our numerical computations in the subsequent paragraphs. In [3] the active in-plane surface stresses have the following form aα = Ta + T z a δ(z) 0 t β 0 Ta + T φ a δ(z)! (33) on the initially unperturbed cylinder surface where the local coordinates α and β correspond to polar coordinates z and φ, respectively (cf. 4.3). We note that the component t z az being positive represents a contractile stress along az being negative an extensile stress, as seen on the basis of the buckling instability reported the cylinder axis and t z for negative stress in ref. [3]. A positive t φ aφ represents a contractile stress in azimuthal direction, which causes a contraction of the cylinder. The latter becomes clear by considering the deformation caused by T φ a according to the Green's function. In equation (33) Ta represents an isotropic homogeneous background active stress while T z a a are the amplitudes of point active stresses along each of the two coordinate axes. δ(z) is the Dirac delta and T φ distribution. As in ref. [3], we focus on the radial deformation ur(z) of a cylinder with initial radius R resulting from an azimuthal in-plane surface stress T φ a . At the end of section 5.3 we perform a validation for an axial in-plane sur- face stress T z a . Berthoumieux et al. [3] consider a 3D elastic material and perform a projection onto the membrane resulting in the surface stretching modulus S and the bending modulus B as surface elastic parameters, together with the 3D Poisson ratio ν. The strength of the homogeneous active in-plane stress is measured by the dimen- sionless number g = Ta SR2 . The radial deformation of a cylindrical shell with radius R is then given by the Green's function Grφ(z) as S and bending forces are quantified by the relative bending modulus b = B An expression for G(z) is given in Fourier space [3] by ur(z) R = −Grφ(z)T φ a = − T φ a RS G(z). (34) (35) G(q) = 2b(Rq)4 + (g − 2νb)(Rq)2 + 2(1 − ν2) − g 1 with G(z) being the inverse Fourier transform of RG(q). 3 κS and B = 1 In our numerical method, elastic forces are derived from the Skalak and Helfrich energy functionals as defined in equation (7) and (8), respectively, which represent an accurate and widely used description of the elastic properties of biological cell membranes. In the limit of small deformations we show in section A2 that the Skalak and Helfrich model, which we use in this study, and the elastic model used by Berthoumieux et al. [3] are related by S = 2 2 κB. In the following these relations are used to calculate the Green's function in equation (35) and the corresponding deformation in equation (34) for comparison to our simulations. We furthermore set the bulk Poisson ratio ν = 1 2, corresponding to a 3D incompressible material, and the Skalak parameter C = 1. Then the Skalak law becomes equivalent to the Neo-Hookean membrane law which is built to model a membrane made of a 3D incompressible material [88]. For the shear modulus S we show in appendix A2 that the shear related in-plane surface stresses determining the Green's function agree between the Skalak law and the model used by ref. [3] in the limit of small deformation. Although we can relate the parameter of Helfrich model to the model used in Berthoumieux et al. [3] by B = 1 2 κB as illustrated in the appendix, the Helfrich model alters the in-plane surface stresses, which in turn alter the Green's function. However, since we consider the limit B → 0 this has no effect for the Green's function used for validation of our simulations. 5.2 Singular active perturbation We start by simulations of overdamped dynamics of an initially stress-free cylindrical shell subjected to a local increase in active in-plane surface stress. The active in-plane surface stress from equation (33) simplifies to aα = 0 t β 0 0 T φ a! δ(z), 12 (36) a) b) c) Fig. 4: a) We consider a local increase in active in-plane surface stress of a cylindrical membrane. b) and c) Membrane meshes as in main text with two different resolutions a) ∆z = 0.2, ∆φ = 0.2 and b) ∆z = 0.1, ∆φ = 0.15, respectively. Fig. 5: The final deformation resulting from three dimensional simulations of a singular active stress shows the same shape as the analytical prediction of the Green's function. However, peak height deviates from the theory. This can be attributed to the singular nature of a = −0.01, κS = 1, the perturbation which is difficult to resolve numerically. Deformations are obtained for the parameter set T φ C = 1, and κB = 10−3. as sketched in figure 4 a). Simulations are carried out with R = 1, κS = 1, C = 1, κB = 0.001, and T φ a = −0.01 in simulation units. We employ two different triangulations of the cylindrical shape which are shown in 4 b) and c). In the coarse mesh, the axial distance between rings of nodes is ∆z = 0.2 and the azimuthal spacing is about ∆φ = 0.2 radians. The finer mesh uses an axial spacing of ∆z = 0.1 and an azimuthal spacing of about ∆φ = 0.15 radians. In figure 5 we compare the final shape of the shell as observed in 3D simulations for the two different resolutions and compare it to the prediction of the Green's function in equation (34). The analytical Green's function shows a peak in deformation of finite width centered at the site of active in-plane surface stress perturbation (z = 0) that decays with increasing distance, reaches a shallow minimum at around z ≈ ±0.7 and then approaches zero for z → ±∞. Although for both resolutions the resulting amplitude of the peak in deformation from the simulation is close to the Green's function and both show a similar shape, we observe a significant deviation of the 3D simulation results from the theoretical expectation. Especially, the simulations cannot reproduce correctly the predicted shallow minima next to the main peak. We note that the deviations for singular perturbation also appear using Lattice Boltzmann/Immersed Boundary method instead of overdamped dynamics (results not shown). This observed deviation can be explained by the idealized singular nature of the active in-plane surface stress which is impossible to accurately reproduce on a discretized membrane shape. Along the cylinder axis, i.e., along z-direction, only one ring of nodes with z = 0 is attributed with finite active in-plane surface stress. Correspond- ingly, only these nodes are subjected to active forces and cause the neighboring nodes to move due to the elastic nature of the cylinder. We note that the central nodes - with active in-plane surface stress - experience a signif- icantly larger deformation than their direct, adjacent neighbor nodes, as seen in figure 5. This steep gradient in deformation resulting in locally very large curvature cannot be completely resolved by the parabolic fit. Thus, the procedure fails to resolve the deformation caused by a point perturbation, although obtained deformations are similar in shape and nearly match the amplitude of the Green's function. Given that in real applications, all pertur- bations can be expected to be non-singular, we proceed to investigate the behavior of our algorithm for spatially smooth active stresses. 13 Tφaz-0.005 0 0.005 0.01 0.015 0.02 0.025-4-3-2-1 0 1 2 3 4ur / Rz / RGreen's functionΔz = 0.2, Δφ = 0.2Δz = 0.1, Δφ = 0.15 5.3 Localized smooth perturbation As a prototypical smooth distribution, we choose a Gaussian distributed active in-plane surface stress of the form aα = 0 t β 0 0 T φ a! exp − z2 R2! , (37) where T φ a superposing a distribution of Green's functions (equation (34)) leading to the deformation is again a constant amplitude. For this distribution, the predicted deformation can be obtained by ur(z) R ∞Z−∞ = − Grφ(z − z0)T φ a exp(−z02/R2) dz0. (38) We use the Green's function in Fourier space given in equation (35) and the Fourier transform of the Gaussian ∞Z−∞ exp(cid:16)−z2/R2(cid:17) exp(−iqz)d(cid:18) z R(cid:19) = √π exp(− 1 4(Rq)2). Using the convolution theorem in Fourier space and transforming back to real space leads to ur(z) R 1 2π R S = − ∞Z−∞ √πT φ 2b(Rq)4 − 2νb(Rq)2 + 2(1 − ν2) exp(iqz)dq. a exp(− 1 4(Rq)2) The integral can be solved analytically in the limit of small bending rigidity b (cid:28) 1 (which corresponds well with the chosen simulation parameters) to obtain ur(z) R = − T φ a S 1 2(1 − ν2) exp(− z2 R2 ). (39) Alternatively, we can solve the integral in equation (38) numerically, which does not lead to any differences for small bending modulus (results not shown). In figure 6 a) we compare the 3D simulation results using overdamped dynamics for parameters R = 1, κS = 1, C = 1 , κB = 10−5, and T φ a = −0.01 to the analytical prediction of equation (39). The Gaussian distribution of active in-plane surface stress leads to a much smoother and broader peak of deformation than the singular perturbation of the previous subsection. Our simulation results are now in very good agreement with the theoretical prediction. To go one step further, with applications in mind such as an active membrane in a flowing liquid, we now replace the simple overdamped fluid dynamics with a full Navier-Stokes dynamics solved by the Lattice-Boltzmann method and coupled to the active membrane via the Immersed-Boundary method as described in section 2. In figure 6 b) we compare simulation results obtained by LBM/IBM to the theoretical Green's function. Again, our simulations are in very good agreement with the analytical theory. In figure 6 b) we include three sets of active in-plane surface stress and shear modulus with constant κB = 0.00018 which are chosen such that the ratio of active in-plane surface stress and shear modulus g remains constant. Thus, the Green's function predicts identical deformation in all three cases which is indeed observed in our simulations. We now proceed to the validation for a perturbation in z-stress, i.e., the active in-plane surface stress takes the form aα = T z a 0 t β 0 0! exp(− z2 R2 ), with the constant amplitude T z a . Corresponding to equation (38) the deformation can be obtained by [3] ur(z) R = ∞Z−∞ Grz(z − z0)T z a exp(−z02/R2) dz0, 14 (40) (41) with Grz(s) = ν RS G(z) and takes in analogy to equation (39) the form z2 R2 ). 2(1 − ν2) exp(− = T z a S ur(z) R ν (42) In figure 6 c) we compare LBM/IBM simulation results for three different perturbations and shear moduli for R = 1, C = 1, and κB = 0.00018 to the theory. Our simulation results are in very good agreement with the theory and by comparing figure 6 b) and c) we observe half the maximum deformation for T φ a which is indeed predicted by the theory for ν = 1 2. As a further test of algorithm accuracy we now perform a convergence study based on the setup in figure 6 b) by as well as in c). We compare the deformation obtained by simulation usim calculating the relative error per node defined as and by Green's function uGreen a = T z r r NzsXzi  = 1 ((usim r (zi) − uGreen r (zi)) /uGreen r (0))2 , (43) where Nz denotes the number of nodes along the cylinder axis and the difference is in relation to the maximal deformation given by the Green's function. In figure 6 d) we show the relative error per node in dependency of the number of membrane nodes. With increasing resolution the error per node steadily decreases in both cases. The converge rate, however, is different for T φ a which decreases more quickly with N−2 . This may be due to the fact that the Green's function is derived from linearized equations of motion [3], whereas simulations are also valid for larger deformations. A perturbation in T z a shows half the maximal deformation as a perturbation in T φ a as predicted in equation (42) and thus is in better agreement with the Green's function. The steady decrease in error demonstrates the accuracy of the presented algorithm for active force calculation. a which decreases with slope N−1 and T z z z From figure 6 we conclude that our algorithm presented in sec. 4 together with elastic force calculations gives reliable results for a reasonably smooth distribution of active in-plane surface stress. The very good accuracy of the predictions is achieved for the simple overdamped dynamics as well as for the substantially more complex and flexible combination of an active membrane with the lattice-Boltzmann/Immersed-Boundary method. 5.4 Homogeneous perturbation: instability diagram To provide another test of our algorithm we perform simulations of a cylindrical membrane which now is subjected to a homogeneous active in-plane surface stress aα = Ta 0 Ta! , 0 t β (44) with Ta = const. This situation corresponds to a membrane with constant motor protein activity and isotropic cortex architecture. We note again that positive Ta represents contractile and negative Ta extensile stress in z- or φ-direction. Although they did not explicitly compute the deformation for this situation, ref. [3] predicts two unstable regions in g-b-parameter space for which the Green's function could be shown to diverge. The predicted instability thresholds serve us as a further validation of our simulation method. By varying both the relative active in-plane surface stress g = Ta/S = 3Ta/(2κS) and the relative bending modulus b = B/(SR2) = 3κB/(4κSR2) we obtain the phase diagram in figure 7 and compare it to the predicted instability thresholds given by Berthoumieux et al. [3]. On the right, for large g, Berthoumieux et al. [3] predict an instability occurring for Ta > 4 3 κS(1 − ν2) shown by the vertical orange line in figure 7. Indeed, for simulations in this range we observe an instability with local contraction of the cylinder (see inset of figure 7 for a shape illustration). The threshold obtained by our simulations closely matches the analytically predicted threshold. This instability is analogous to a Rayleigh-Plateau instability of a liquid jet with the positive, contractile active in-plane surface stress playing the role of the surface tension. To the left of the threshold a fairly large region is observed in which the initial cylindrical shape remains stable. For negative g (extensile stress), [3] predicts a buckling R2 . To compare to our simulations, we prescribe an active stress only along the instability when Ta < −2q κBκS 15 a) c) b) d) Fig. 6: a) Comparison of the deformation obtained by the 3D overdamped dynamics with the theoretical expectation for a cylindrical shell with 160 nodes along z-direction subjected to a Gaussian distributed active in-plane surface stress with κS = 1, κB = 10−5, a = −0.01. 3D simulations are in very good agreement with the theory. b) Comparison of the deformation obtained with and T φ LBM/IBM simulations for C = 1 and κB ≈ 10−4 with the theoretical prediction for the same setup with perturbation in φ-stress T φ a and c) with perturbation in z-stress T z a . d) Increasing resolution of both membrane and fluid mesh show convergence of the relative error per node for the parameter set of b) for a perturbation in T φ (red dots) as well as for the parameter set of c) for a a −1 perturbation in T z (blue triangles). The error decreases proportional to N , respectively. In b) -- c) a z simulations are done for 300 nodes in z-direction. −2 and proportional to N z Fig. 7: Phase diagram of a cylindrical shell with relative, active in-plane surface stress g = Ta/S = 3Ta/(2κS) and relative bending modulus b = B/(SR2) = 3κB/(4κSR2) from 3D LBM/IBM simulations in comparison to the theoretically predicted thresholds [3]. On the left of the dotted line we apply a negative active stress, on the right of the dotted line a positive active stress. For small negative or positive active stress (around the dotted line) the cylindrical membrane is stable. For large negative stresses, a buckling instability is observed. For large positive active in-plane surface stress a Rayleigh-Plateau like instability occurs. The theoretically predicted instability thresholds [3] and our 3D LBM/IBM simulations are in excellent agreement in both cases. Insets illustrate the shape of the shell corresponding to different values of active in-plane surface stress. The labels a, b, c refer to figure 8. 16 0 0.002 0.004 0.006 0.008 0.01-3-2-1 0 1 2 3Overdamped dynamicsur / Rz / RGreen's functionTaφ = -0.01, κS = 1.0 0 0.002 0.004 0.006 0.008 0.01-6-4-2 0 2 4 6LBM/IBMur / Rz / RGreen's functionTaφ = -0.01, κS = 1.0Taφ = -0.005, κS = 0.5Taφ = -0.02, κS = 2.0 0 0.001 0.002 0.003 0.004 0.005-6-4-2 0 2 4 6LBM/IBMur / Rz / RGreen's functionTaz = 0.01, κS = 1.0Taz = 0.005, κS = 0.5Taz = 0.02, κS = 2.010-410-3101102103Nz-2Nz-1Relative error per node εNumber of nodes NzSimulation results TaφSimulation results Taz 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-1.5-1-0.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5◯⬤b = B / ( S R2 ) = 3 κB / (4 κS R2)g = Ta / S = 3 Ta / (2κS )buckling threshold from [3]◯⬤⬤◯◯⬤⬤⬤◯◯✷✷⬤◯◯✷✷✷◯◯✷⬤◯⬤⬤◯⬤⬤⬤⬤⬤⬤⬤⬤⬤⬤◯◯⬤⬤⬤◯◯◯⬤⬤⬤◯◯⬤⬤⬤⬤⬤⬤◯◯◯⬤◯◯◯✷⬤⬤RP-like threshold from [3]✷stable cylinderbuckling instabilityRP-like instabilitycbabucklingRP-likestabletβa,α=(cid:18)Ta<0000(cid:19)tβa,α=(cid:18)Ta>000Ta>0(cid:19) az = Ta and t φ aφ = 0. This corresponds to a contracting, cylindrical membrane and the resulting cylinder axis, i.e., t z shape beyond the threshold is illustrated in figure 7 at the bottom left. Our simulations agree very well with the predicted instability onset depending on the relative active in-plane surface stress and the relative bending modulus. In addition, we investigate the transition to buckling in more detail and carry out simulations imposing an active aφ = Ta, which corresponds exactly to the scenario considered tension in azimuthal and axial direction t z by [3]. In figure 8 we compare these simulations (top row) with the ones in figure 7 (bottom row), respectively. At large (negative) active stress in a) we observe an instability in both simulations, however only the simulation with a purely axial stress clearly corresponds to a buckling instability. The instability in the top row exhibits a non-axisymmetric character. For slightly smaller active stress in 8 b) the non-axisymmetric instability remains for the isotropic stress, but the membrane becomes stable for z-stress only. Decreasing the active stress further, we observe a stable cylindrical membrane in both cases, as shown in c). az = t φ From figure 8 we conclude that an additional instability is present (not related to buckling) caused by a finite azimuthal stress t φ aφ. This additional instability induces non-axisymmetric deformations and thus is not observed in the axisymmetric treatment of ref. [3] nor in our axisymmetric simulations for which also the onset of the buckling instability is in exact agreement with the analytical prediction as shown in figure S2 of the Supporting Information. Fig. 8: Membrane shape for different values of negative active in-plane surface stress for fixed bending modulus b = 0.01125. In the upper row the membrane is subjected to both z- and φ-stress, while in the lower row the membrane is subjected to z-stress only. For z-stress only we observe a buckling instability in a). However, for isotropic stress we observe an instability introducing non- axisymmetric deformations at intermediate stresses in b) where no buckling instability occurs. c) For smaller active stress the cylindrical membrane remains stable in both cases. Corresponding points in the phase diagram in figure 7 are labeled with a to c. So far, we have validated our algorithm to agree with theoretical predictions on the basis of the Green's function. To obtain a validation in the non-linear regime beyond the Green's function, we compare our 3D simulations to simulations of an axisymmetric membrane, as detailed in the Supplemental Material [96]. We do this by comparing the membrane shape in the case of a buckling instability. In figure 9 we show the deformation obtained from the axisymmetric simulation and compare it to three simulations using 3D LBM/IBM. The three 3D simulations are done for different resolutions ∆z/R. We use the non-dimensional parameters g = −0.75 and b = 0.01125. On the one hand all three 3D simulations show the same deformation and wavelength and on the other hand they agree in the wavelength with the axisymmetric simulation method. The wavelength of the buckling instability is about 1.75R for both very different simulation techniques. In case of small bending elasticity Berthoumieux et al. [3] predict a wavelength at the threshold, where the denominator of the Green's function in eq. (35) becomes zero for finite wave vector q, of λ = 1.9R which is reasonably close to the value observed with simulations. The difference may arise from the periodicity of the shell in our simulations and/or from finite bending together with being beyond the threshold. Nevertheless, the excellent agreement between the axisymmetric and 3D simulations provides strong evidence for the reliability of our algorithm also in the range of large deformations. 6 Model application: cell division in suspending fluid In the following we present a first application of our method including fluid flow. This illustrates the versatility and applicability of our combined LBM/IBM method for active cell membranes. For this, we consider a dividing ellipsoidal cell. Except the fact that we employ an elastic rather than a viscous cortex, our setup resembles the situation of cell cytokinesis [16, 17, 19 -- 21]. Cell cytokinesis as part of cell division is a prominent subject of active matter research in biological physics [13, 18, 97]. Most previous studies, e.g. [20] or [21], investigated the dynamics of cell cytokinesis for an axisymmetric membrane without considering internal fluid flow. Refs. [98] and [99] analyzed the flow field inside a dividing cell where the contractile ring is modeled as an additional force using 17 a)g=−0.9b)g=−0.3c)g=−0.075 Fig. 9: Comparison of the deformation obtained for the buckling instability of an initially cylindrical membrane subjected to a homogeneous active in-plane surface stress with g = −0.75 and b = 0.01125. We compare the wavelength of the deformation obtained in 3D LBM/IBM simulations with different resolutions ∆z/R to one obtained by axisymmetric simulation. For a sample illustration of the 3D shape we refer to the inset on the left hand side of reference 7. All simulations show the same wavelength for the buckling instability. the immersed boundary method and phase field model, respectively. Ref. [100] analyzed the flow field by means of the phase field model as well, but considered the cortical ring as shrinking elastic loop. Here, we consider cell division triggered by active stresses including an external flow leading to a fully 3D asymmetric membrane shape during the division. We first analyze the flow field dynamically evolving inside the dividing cell surrounded by a quiescent medium in section 6.1 and then extend this setup by considering a dividing cell in an external shear flow in section 6.2. 6.1 Flow field inside a dividing cell We consider a prolate ellipsoidal cell of diameter 7 µm and length 14 µm which is endowed with an isotropic active stress Ta = 8 × 10−5 N/m. In addition, in an interval of ∆θ ≈ π 12 around the equator the active stress in azimuthal direction is increased by a factor of six according to a step function. The membrane is endowed with shear elasticity κS = 5 · 106 N/m, C = 1, and bending elasticity κB = 2 · 10−19 Nm which are in the range of typical cell membranes. We present the initial 3D membrane shape in figure 10 a) and a deformed 3D shape in figure 10 b). In both cases we illustrate the region with increased active stress along φ-direction by the red shaded area. The shape and the developing flow field in the central plane, which includes the long axis of the ellipsoid, are shown over time in figure 10 c) to h). Here, the active stress triggers active deformation of the membrane which in turn triggers fluid flow inside the cell. We note that the fluid velocity at the position of the membrane corresponds to membrane motion, which moves with local fluid velocity due to no-slip condition as described in section 2.2. At the beginning the membrane contracts around the poles (left and right in figure 10 c)) due to the isotropic contractile active stress. The contraction at the poles causes a rounding which triggers a flow field pointing away from the poles at the beginning. Simultaneously, the membrane starts contracting around the equator, see figure 10 c) and d). In figure 10 d) four vortices are present around the equator and four at the corners of the figure. After some time the contraction at the poles stops, see figure 10 e) and f) and only the contraction at the equator proceeds. With progressive contraction a flow away from the midplane towards the poles develops. As it is visible in figure 10 g) and h), the site of maximal velocity towards the poles is located at x ≈ ±10. Thus, it does not coincide with the center of the two spheroids pinching off at x ≈ ±18, but is rather shifted towards the equatorial plane at x = 0. Approaching the poles of the ellipsoid the velocity decreases. 18 -0.15-0.1-0.05 0 0.05 0.1 0.15 0 1 2 3 4 5ur in arbitrary unitsz / RΔz / R = 0.090Δz / R = 0.075Δz / R = 0.064axisymmetric simulation Fig. 10: a) Similar to cell division an elastic cell membrane, which is subjected to homogeneous active stress and a local increase in azimuthal active stress in the red shaded area, contracts locally as shown in b). (c) - (h) show the outline of the deforming membrane (red nodes) in the central plane and the developing flow field (arrows) inside the cell over time. Eventually, a flow away from the contracting region towards the poles of the cell is observed. Arrows indicate the flow direction while the color indicates the flow velocity. Relative time is t/tmax = a) 0, b) 1, c) 4×10−4, d) 0.08, e) 0.16, f) 0.51, g) 0.86, h) 1.0 . 6.2 Dividing cell in shear flow In the previous section we considered an initially quiescent fluid. Here, we go one step further and include an externally driven flow interacting with the membrane. We apply a shear flow with a shear rate γ ≈ 1400 s−1. All other parameters are the same as in the previous section. The cell membrane as illustrated in figure 11 a) deforms now due to the local increase in active stress (red shaded area in figure a) and b)) but also due to the external shear flow. This becomes visible by the non-symmetrically deformed membrane in figure 11 b). Figures 11 c) to h) show the flow field relative to the shear flow, i.e., from each velocity vector the corresponding background flow is subtracted. The time evolution of the cell shape and the flow field is also illustrated in the Supplemental Video 1. In contrast to the previous section, the shear flow triggers an asymmetric deformation and in turn an asymmetric flow inside the cell. The rounding together with the flow from the poles towards the equator is less pronounced (compare figure 11 d) to figure 10 d)). 19 a)b)c)402002040x in sim. units201001020y in sim. unitsd)402002040x in sim. units201001020y in sim. unitszerointermediatelargee)402002040x in sim. units201001020y in sim. unitsf)402002040x in sim. units201001020y in sim. unitsg)402002040x in sim. units201001020y in sim. unitsh)402002040x in sim. units201001020y in sim. units Fig. 11: An elastic dividing cell membrane in shear flow. a) An increase in azimuthal active stress in the red shaded area of an ellipsoidal cell membrane triggers a local contraction as shown in b) while the externally driven shear flow also deforms the membrane. (c) - (h) show the outline of the deforming membrane (red nodes) in the central plane and the developing flow field (arrows) as perturbation to the shear flow over time. We here refer also to the Supplemental Video 1. The presence of the shear flow renders both the cell shape and the flow field asymmetric. Relative time is t/tmax = a) 0, b) 1, c) 1.3×10−3, d) 0.07, e) 0.17, f) 0.52, g) 0.88, h) 1.0 . The flow field inside a dividing cell suspended in a shear flow shows how the actively deforming membrane couples to a background flow and imposes perturbation on the shear flow. Both the active stress present in the cell cortex as well as the external flow trigger membrane deformation. 20 a)b)c)402002040x in sim. units3020100102030y in sim. unitsd)402002040x in sim. units3020100102030y in sim. unitszerointermediatelargee)402002040x in sim. units3020100102030y in sim. unitsf)402002040x in sim. units3020100102030y in sim. unitsg)402002040x in sim. units3020100102030y in sim. unitsh)402002040x in sim. units3020100102030y in sim. units 7 Conclusion We presented a computational algorithm to compute the dynamical deformation of arbitrarily shaped active bio- logical (cell) membranes embedded in a 3D fluid. Active stresses in cells typically arise from the activity of motor proteins. Constitutive equations for active stresses in membranes have been developed recently [6, 30] in the framework of differential geometry and form the theoretical basis for our computational method. The membranes are discretized by a set of nodes connected via flat triangles. The key ingredient of our algorithm is the computa- tion of the active force acting on each node starting from prescribed active stresses via a parabolic fitting procedure on the deformed membrane. Besides active forces, the method also includes passive elastic forces derived from the well-established Skalak and Helfrich models for cell membranes. In simple cases, the surrounding fluid can be considered a purely frictional medium, such that the membrane nodes follow simple overdamped dynamics in time. For more realistic situations, we introduced a powerful and versatile coupling between the active membrane and the surrounding fluid via the Immersed-Boundary method. This technique incorporates the full Navier-Stokes dynamics for the surrounding liquid, solved here via the Lattice-Boltzmann method. Thus, our method allows us to go beyond the determination of equilibrium shapes of active elastic membranes and allows for simulations of dynamically deforming biological cells immersed in an external flow. We successfully validated our algorithm for two distinct situations of an elastic, initially cylindrical membrane: (i) a local, Gaussian distributed active stress and (ii) a homogeneous active stress. For (i) our numerical results are in excellent agreement with the analytically predicted deformation of Berthoumieux et al. [3] and show conver- gence with increasing resolution. Overdamped dynamics and IBM/LBM dynamics are in good agreement. For (ii) we recovered both the buckling as well as the Rayleigh-Plateau-like instability predicted by [3]. Comparison to our own numerical solutions of the axisymmetric problem shows very good agreement with the full 3D algorithm also in the regime of large deformation not covered by [3]. In addition, our computations reveal the existence of a thus far unobserved non-axisymmetric instability in the case of extensile axial and azimuthal stresses. In order to illustrate the versatility of our method, we analyzed the flow field inside an elastic, dividing cell membrane in shear flow. This represents the first investigation of the dynamic two-way coupling between active deformations and externally driven fluid flow. In this work, we considered temporally constant active stresses, but the inclu- sion of time-dependent active stresses computed, e.g., by a convection-diffusion model of active substances, is straightforwardly possible. Our computational method significantly extends the range of physical problems to which existing active mem- branes theories can be applied. First, it is not restricted to simple shapes such as cylinders or spheres (or small deformations thereof) allowing efficient and accurate treatment of arbitrary membrane shapes and deformations. Second, our method couples the active membrane dynamics to the full Navier-Stokes dynamics of the surrounding fluid. In the described LBM/IBM scheme a viscosity contrast of the inner and outer fluid - as it is well known in case of red blood cells [50] - can furthermore be incorporated, which allows for a even more realistic model of living cells. This opens up a wide range of applications in external flows such as active cells in the blood stream or active cellular compartments in cytoplasmic streaming flows which currently remain largely unexplored. A particularly interesting application could be the formation of platelets from megakaryocytes which, according to a set of recent experiments [101, 102], crucially depends on the interplay between active processes and external flows. 8 Acknowledgments C. B. thanks the Studienstiftung des deutschen Volkes for financial support and acknowledges support by the study program "Biological Physics" of the Elite Network of Bavaria. This project is part of the collaborative research centre TRR225 (subproject B07) funded by the German Research Foundation (DFG). We gratefully acknowledge computing time provided by the SuperMUC system of the Leibniz Rechenzentrum, Garching, as well as by the Bavarian Polymer Institute and financial support from the Volkswagen Foundation. 21 Appendix A1 Membrane in thin shell formulation In this appendix, we summarize the necessary basics and conventions of differential geometry on thin shells used in this work. For a more detailed description we refer the reader to refs. [73, 84]. The 2D manifold in general is parametrized by two coordinates s1, s2. We denote vector components on the manifold by Greek letters α, β = 1, 2 and vector components in Euclidean space by Latin letters i, j, k = x, y, z. Moreover, we use the Einstein summation convention, i.e., double occurrence of an index in sub and superscript implies a sum over this index. A partial derivative with respect to sβ is denoted by a comma, i.e., for an arbitrary vector v this implies vα,β = ∂vα/∂sβ. The membrane in the undeformed state is parametrized by the vector From the local in-plane coordinates the metric tensor is defined by with g = det(gαβ). The inverse metric gαβ can be obtained by the expression X(s1, s2, t). eα = X,α gαβ = eα · eβ, gαγgγβ = δβ α (A1.1) (A1.2) (A1.3) (A1.4) (A1.8) (A1.9) (A1.10) (A1.11) (A1.12) α denoting the Kronecker delta being 1 for α = β and 0 otherwise. For a general vector vα, vα or tensor β, tαβ subscript indices denote co-variant components and superscript indices denote contra-variant α , tα with δβ tαβ, t β components. An index can be raised by vα = vβgβα , α = tαγgγβ t β or lowered by vα = vβgβα , tαβ = t γ α gγβ. A surface element is defined by dS = √gds1ds2. The Christoffel symbols are given by αβ = 1 Γγ 2 gγδ [gαδ,β + gβδ,α − gαβ,δ] . (A1.5) (A1.6) (A1.7) The in-plane coordinate vectors eα provide a local coordinate system together with the unit normal vector on the membrane n = e1 × e2 ke1 × e2k . Considering the local unit normal vector allows for the definition of the curvature tensor A co-variant derivative of an arbitrary vector vα or tensor tαβ defined on the membrane is given by Cαβ = −eβ,α · n. ∇αvβ = vβ ∇αtβγ = tβγ On the membrane the Levi-Civita tensor is given by ,α + Γβ ,α + Γβ αγvγ αδtδγ + Γγ αδtβδ. αβ = √g 0 −1 0! , αβ = 1 √g 0 −1 0! . 1 1 22 An arbitrary vector vα defined on the manifold with respect to eα with α = 1, 2 can be decomposed into an in-plane and a normal contribution vα = tαβeβ + tα n the component of a vector. nn (A1.13) with tαβ being the component of a tensor and tα A connection from the in-plane coordinates to the Euclidean coordinates can be drawn by the expression eα = ei αEi (A1.14) α being the i-th component of eα and Ei being the i-th Euclidean unit vector. A three dimensional tensor with ei tij can be projected onto the membrane via tαβ = tijei αej β. (A1.15) External or internal forces may lead to a deformation of the membrane characterized by the deformation field u. The membrane in the deformed state is parametrized by X0(s1, s2) = X(s1, s2) + u(s1, s2). (A1.16) We denote all vectors or tensors that are evaluated on the deformed surface by a prime. Corresponding to the change of local coordinate vectors e0α = X0,α and normal vector n0 both the metric tensor and the curvature tensor changes to g0αβ = e0α · e0β C0αβ = −e0β,α · n0. The Christoffel symbols have to be computed using g0αβ and the co-variant derivative becomes ∇0αvβ = ∂0αvβ + Γ0β αγvγ ∇0αtβγ = ∂0αtβγ + Γ0β αδtδγ + Γ0γ αδtβδ. A2 Elastic in-plane surface stresses (A1.17) (A1.18) (A1.19) (A1.20) In the following we compare the elastic in-plane surface stresses used in ref. [3] to those obtained for Skalak energy density in equation 7. We consider the displacement vector in equation A1.16 decomposed into axial and normal deformation u = uzez + urn. (A2.1) (A2.2) For the given energy density in equation 7 the in-plane surface stresses are obtained by [74, 103, 104] SK = 2 tαβ J ∂wSK ∂I1 gαβ + 2J ∂wSK ∂I2 g0αβ with the invariants (A2.3) (A2.4) and with J = √I2 + 1. Using equations A1.16, A1.17 and the deformation in equation A2.1 we obtain for the metric on the deformed membrane in the limit of small deformations 0 I1 = gαβg0αβ − 2 I2 = det(gαβ) det(g0αβ) − 1, (A2.5) g0αβ = 1 + 2∂zuz 0 R! 1 + 2 ur 23 and for the in-plane surface stresses SK = 2 tzz SK = 2 tφφ R(cid:21) R3(cid:21) . R2 C∂zuz + (1 + C) ur These equations can be compared to the tensions in equation (11) of ref. [3] for the elastic model used by Berthoumieux et al. [3]. The latter is based on the Hooke's law in 3D which is projected onto the membrane. By comparing the stresses we find agreement in the limit of small deformations for the relation of the stretching modulus S to the shear modulus used in equation (7) of 3 κS(cid:20)(1 + C)∂zuz + C 3 κS(cid:20) 1 ur (A2.6) (A2.7) S = 2 3 κS (A2.8) and C = 1. This relation is used for the Green's function to match both elastic models. We furthermore can compare the bending modulus used in Berthoumieux et al. [3] to the one appearing in the Helfrich energy in equation (8) [78]. Berthoumieux et al. [3] defines B = Eh3/(24(1 − ν2)) with the Youngs modulus E and the thin shell height h. Comparing this to the expression κB = Eh3/(12(1 − ν2)) of Pozrikidis [105] we obtain the relation (A2.9) B = 1 2 κB. A3 Ovedamped dynamics method and simulation analysis A3.1 Overdamped dynamics As a different approach than the LBM/IBM, we use a model program based on overdamped dynamics to solve for the final, equilibrium shape of the membrane in case of validation. The resulting active and elastic forces F calculated for every node enter the equation of motion of the corresponding node rc which is given by F = γ rc (A3.1) where γ is a friction coefficient. We solve the equations for all nodes using Euler integration scheme. We fix the nodes at the boundaries of the cylinder by harmonic springs of strength 1000κS. This results in nearly vanishing deformation at the boundary of the cylinder. A3.2 Simulation analysis To obtain the shape shown for example in figure 6 a) or b) we average the final, radial deformation over all nodes at certain z-position. Due to the averaging and inherent errors in the bending algorithm [52] the deformation ur does not reach exactly zero far away from the perturbation in active stress, but show a constant offset of 5 × 10−4, which we eliminate in the figures. A4 Choice of reference neighbor does not influence simulation results In section 4.1 we explain how the local coordinate system is built using the first neighboring node as reference for the first in-plane coordinate vector eξ. This fact is also important for the projection of the active stress into the local coordinate system, as mentioned in section 4.3. In the following we show that the choice of the reference neighboring node is arbitrary and does not influence simulation results. Therefore, we first consider the cylindrical membrane subjected to a homogeneous perturbation in active in- a = 0.01 and plane surface stress along z-direction T z κS = 1.0. We systematically change the reference neighbor node serving for construction of the local coordinate system as illustrated in figure A1 a). Figure A1 b) shows that the deformation obtained in simulations is the same a . We use exactly the setup analyzed in figure 6 c) with T z 24 for every choice of reference neighbor node and agrees very well with the theoretically obtained Green's function. In case of T z a perturbation the in-plane derivatives of the active in-plane surface stress are crucial and trigger the deformation. Thus, figure A1 provides evidence for the correct calculation of derivatives regardless of the choice of reference neighbor node and furthermore shows the choice is arbitrary. Furthermore, in order to take into account a more complex membrane geometry, we use the setup in figure 10 and compare the dynamics for five different choices of reference neighboring node. Choosing the sixth neighbor is not possible for the ellipsoidal geometry as the surface tiling requires at least 12 nodes with five neighbors only. We consider an initially ellipsoidal cell membrane endowed with active stress, which in azimuthal direction increases around the equator. The increased active stress triggers the cell membrane to contract. In figure 10 we show the dynamically evolving flow field inside the cell. Here, we redo the simulation and for each simulation we choose a different neighbor node to build the local coordinate system, which moves with the deforming membrane in time. In figure A1 c) we show the five different choices. In figure A1 d) we consider the point in time corresponding to figure 10 g) and show the radial position of all membrane nodes as function of the position along the axis. All simulations with different reference neighbor show the same membrane shape and are in very good agreement. We note that slight deviations occur at the site of strongest indentation, which results also in the strongest curvature. At this position, the parabolic fit is not capable of covering the strongly deformed membrane shape completely and thus slight deviations occur. In figure A1 e) we do the same for a dividing cell in shear flow corresponding to figure 11 g). We show the node positions within the plane containing the long axis of the cell and the shear axis of the external flow. All simulations with different reference neighbor show the same membrane shape in shear flow and are in very good agreement. Thus, figure A1 d) and e) provide evidence that the choice of reference neighbor is indeed arbitrary and does not alter the simulation in case of a dynamically deforming membrane coupled to a suspending fluid. 25 Fig. A1: a) Color code for the neighbor node serving as reference to construct the local coordinate system on the cylindrical membrane at the site of the central node. b) Deformation obtained for a cylindrical membrane subjected to a Gaussian perturbation in active in-plane surface stress along z-direction for different reference neighbor nodes. The setup is identical to figure 6 c) with a = 0.01 and κS = 1.0. Obtained deformations are in very good agreement regardless of the choice of reference neighbor T z node. c) Neighbor node serving as reference to construct the local coordinate system on the ellipsoidal membrane at the site of the central node. d) Cell membrane shape for different reference neighboring node νref at a given time corresponding to figure 10 g). Except for slight deviations in the region of largest curvature in case of neighboring node 2 and 5, all membrane shapes are in very good agreement. e) Membrane shapes in shear flow for different reference neighboring node νref at a given time corresponding to figure 11 g) are in very good agreement. Thus, we prove evidence that the choice of the reference neighboring node does not affect simulation results. 26 c621345a)b) 0 0.001 0.002 0.003 0.004 0.005-4-2 0 2 4ur / Rz / RGreen's functionνref = 1νref = 2νref = 3νref = 4◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃◃νref = 5▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹▹νref = 6c)c13524d)40302010010203040axial position0510152025radial positionνref=1νref=2νref=3νref=4νref=5e)40302010010203040axial position201510505101520lateral positionνref=1νref=2νref=3νref=4νref=5 References [1] Bruce Alberts, Alexander Johnson, Julian Lewis, Martin Raff, Keith Roberts, and Peter Walter. Molecular Biology of the Cell with CD. Garland Science, 5 edition, 2007. [2] Andrew G. Clark, Ortrud Wartlick, Guillaume Salbreux, and Ewa K. Paluch. the Cell Surface during Animal Cell Morphogenesis. Current Biology, 24(10):R484 -- R494, May 2014. doi:10.1016/j.cub.2014.03.059. Stresses at [3] Hélène Berthoumieux, Jean-Léon Maître, Carl-Philipp Heisenberg, Ewa K Paluch, Frank Jülicher, and Guillaume Salbreux. Active elastic thin shell theory for cellular deformations. New Journal of Physics, 16(6):065005, June 2014. doi:10.1088/1367-2630/16/6/065005. [4] Arnab Saha, Masatoshi Nishikawa, Martin Behrndt, Carl-Philipp Heisenberg, Frank Jülicher, and Stephan W. Grill. Determining Physical Properties of the Cell Cortex. Biophysical Journal, 110(6):1421 -- 1429, March 2016. doi:10.1016/j.bpj.2016.02.013. [5] Tetsuya Hiraiwa and Guillaume Salbreux. Role of Turnover in Active Stress Generation in a Filament Network. Physical Review Letters, 116(18), May 2016. doi:10.1103/PhysRevLett.116.188101. [6] Guillaume Salbreux and Frank Jülicher. Mechanics of active surfaces. Physical Review E, 96(3), September 2017. doi:10.1103/PhysRevE.96.032404. [7] Arvind Ravichandran, Gerrit A. Vliegenthart, Guglielmo Saggiorato, Thorsten Auth, and Gerhard Gomp- per. Enhanced Dynamics of Confined Cytoskeletal Filaments Driven by Asymmetric Motors. Biophysical Journal, 113(5):1121 -- 1132, September 2017. doi:10.1016/j.bpj.2017.07.016. [8] Tanniemola B. Liverpool and M. Cristina Marchetti. Instabilities of Isotropic Solutions of Active Polar Filaments. Physical Review Letters, 90(13), April 2003. doi:10.1103/PhysRevLett.90.138102. [9] Niladri Sarkar and Abhik Basu. Role of interfacial friction for flow instabilities in a thin polar-ordered active fluid layer. Physical Review E, 92(5), November 2015. doi:10.1103/PhysRevE.92.052306. [10] Rajesh Ramaswamy and Frank Jülicher. Activity induces traveling waves, vortices and spatiotemporal chaos in a model actomyosin layer. Scientific Reports, 6:20838, February 2016. doi:10.1038/srep20838. [11] Edouard Hannezo, Bo Dong, Pierre Recho, Jean-François Joanny, and Shigeo Hayashi. Cortical instability drives periodic supracellular actin pattern formation in epithelial tubes. Proc. Nat. Acad. Sci. (USA), 112 (28):8620 -- 8625, July 2015. [12] Kripa Gowrishankar and Madan Rao. Nonequilibrium phase transitions, fluctuations and correlations in an active contractile polar fluid. Soft Matter, 12(7):2040 -- 2046, 2016. doi:10.1039/C5SM02527C. [13] Stephan W Grill. Growing up is stressful: Biophysical laws of morphogenesis. Current Opinion in Genetics & Development, 21(5):647 -- 652, October 2011. doi:10.1016/j.gde.2011.09.005. [14] Guillaume Salbreux, Guillaume Charras, and Ewa Paluch. Actin cortex mechanics and cellular morphogen- esis. Trends in Cell Biology, 22(10):536 -- 545, October 2012. doi:10.1016/j.tcb.2012.07.001. [15] A. C. Callan-Jones, V. Ruprecht, S. Wieser, C. P. Heisenberg, and R. Voituriez. Cortical Flow- January 2016. Physical Review Letters, 116(2):028102, Driven Shapes of Nonadherent Cells. doi:10.1103/PhysRevLett.116.028102. [16] G. Salbreux, J. Prost, and J. F. Joanny. Hydrodynamics of Cellular Cortical Flows and the Formation of Contractile Rings. Physical Review Letters, 103(5), July 2009. doi:10.1103/PhysRevLett.103.058102. 27 [17] Jakub Sedzinski, Maté Biro, Annelie Oswald, Jean-Yves Tinevez, Guillaume Salbreux, and Ewa Paluch. Polar actomyosin contractility destabilizes the position of the cytokinetic furrow. Nature, 476(7361):462 -- 466, August 2011. doi:10.1038/nature10286. [18] Rebecca A. Green, Ewa Paluch, and Karen Oegema. Cytokinesis in Animal Cells. Annual Review of Cell and Developmental Biology, 28(1):29 -- 58, November 2012. doi:10.1146/annurev-cellbio-101011-155718. [19] Inês Mendes Pinto, Boris Rubinstein, and Rong Li. Force to Divide: Structural and Mechanical Re- quirements for Actomyosin Ring Contraction. Biophysical Journal, 105(3):547 -- 554, August 2013. doi:10.1016/j.bpj.2013.06.033. [20] Hervé Turlier, Basile Audoly, Jacques Prost, and Jean-François Joanny. Furrow Constriction in Animal Cell Cytokinesis. Biophysical Journal, 106(1):114 -- 123, January 2014. doi:10.1016/j.bpj.2013.11.014. [21] Anirban Sain, Mandar M. Inamdar, and Frank Jülicher. Shape Changes during Cytokinesis. doi:10.1103/PhysRevLett.114.048102. Physical Review Letters, 114(4):048102, Dynamic Force Balances and Cell January 2015. [22] A E Carlsson. Mechanisms of cell propulsion by active stresses. New Journal of Physics, 13(7):073009, July 2011. doi:10.1088/1367-2630/13/7/073009. [23] Danying Shao, Herbert Levine, and Wouter-Jan Rappel. Coupling actin flow, adhesion, and morphology in a computational cell motility model. Proceedings of the National Academy of Sciences, 109(18):6851 -- 6856, 2012. [24] Wieland Marth, Simon Praetorius, and Axel Voigt. A mechanism for cell motility by active polar gels. Journal of The Royal Society Interface, 12(107):20150161, 2015. [25] Andrew C Callan-Jones and Raphaël Voituriez. Actin flows in cell migration: From locomotion and polarity to trajectories. Current Opinion in Cell Biology, 38:12 -- 17, February 2016. doi:10.1016/j.ceb.2016.01.003. [26] Eric J. Campbell and Prosenjit Bagchi. A computational model of amoeboid cell swimming. Physics of Fluids, 29(10):101902, October 2017. doi:10.1063/1.4990543. [27] Elisabeth Fischer-Friedrich, Yusuke Toyoda, Cedric J. Cattin, Daniel J. Müller, Anthony A. Hyman, and Frank Jülicher. Rheology of the Active Cell Cortex in Mitosis. Biophysical Journal, 111(3):589 -- 600, August 2016. doi:10.1016/j.bpj.2016.06.008. [28] Elisabeth Fischer-Friedrich. Active Prestress Leads to an Apparent Stiffening of Cells through Geometrical Effects. Biophysical Journal, 114(2):419 -- 424, January 2018. doi:10.1016/j.bpj.2017.11.014. [29] Felix C. Keber, Etienne Loiseau, Tim Sanchez, Stephen J. DeCamp, Luca Giomi, Mark J. Bowick, M. Cristina Marchetti, Zvonimir Dogic, and Andreas R. Bausch. Topology and dynamics of active nematic vesicles. Science, 345(6201):1135 -- 1139, 2014. [30] K. Kruse, J. F. Joanny, F. Jülicher, J. Prost, and K. Sekimoto. Generic theory of active polar gels: A paradigm for cytoskeletal dynamics. The European Physical Journal E, 16(1):5 -- 16, January 2005. doi:10.1140/epje/e2005-00002-5. [31] F Jülicher, K Kruse, J Prost, and J Joanny. Active behavior of the Cytoskeleton. Physics Reports, 449(1-3): 3 -- 28, September 2007. doi:10.1016/j.physrep.2007.02.018. [32] Sriram Ramaswamy. The Mechanics and Statistics of Active Matter. Annual Review of Condensed Matter Physics, 1(1):323 -- 345, August 2010. doi:10.1146/annurev-conmatphys-070909-104101. [33] M. C. Marchetti, J. F. Joanny, S. Ramaswamy, T. B. Liverpool, J. Prost, Madan Rao, and R. Aditi Simha. Hydrodynamics of soft active matter. Reviews of Modern Physics, 85(3):1143 -- 1189, July 2013. doi:10.1103/RevModPhys.85.1143. 28 [34] J. Prost, F. Jülicher, and J-F. Joanny. Active gel physics. Nature Physics, 11(2):111 -- 117, February 2015. doi:10.1038/nphys3224. [35] Wylie W. Ahmed, Etienne Fodor, and Timo Betz. Active cell mechanics: Measurement and the- ory. Biochimica et Biophysica Acta (BBA) - Molecular Cell Research, 1853(11):3083 -- 3094, 2015. doi:10.1016/j.bbamcr.2015.05.022. [36] Étienne Fodor, Cesare Nardini, Michael E. Cates, Julien Tailleur, Paolo Visco, and Frédéric van Wi- Physical Review Letters, 117(3), July 2016. jland. How Far from Equilibrium Is Active Matter? doi:10.1103/PhysRevLett.117.038103. [37] Daniel Needleman and Zvonimir Dogic. Active matter at the interface between materials science and cell biology. Nature Reviews Materials, 2(9):17048, July 2017. doi:10.1038/natrevmats.2017.48. [38] Julia M. Yeomans. Nature's engines: Active matter. Europhysics News, 48(2):21 -- 25, March 2017. doi:10.1051/epn/2017204. [39] Frank Jülicher, Stephan W Grill, and Guillaume Salbreux. Hydrodynamic theory of active matter. Reports on Progress in Physics, 81(7):076601, July 2018. doi:10.1088/1361-6633/aab6bb. [40] Edouard Hannezo, Jacques Prost, and Jean-François Joanny. Mechanical Instabilities of Biological Tubes. Physical Review Letters, 109(1), July 2012. doi:10.1103/PhysRevLett.109.018101. [41] Rui Zhang, Ye Zhou, Mohammad Rahimi, and Juan J de Pablo. Dynamic structure of active nematic shells. Nat. Commun., 7:1 -- 9, November 2016. [42] Thuan Beng Saw, Amin Doostmohammadi, Vincent Nier, Leyla Kocgozlu, Sumesh Thampi, Yusuke Toyama, Philippe Marcq, Chwee Teck Lim, Julia M Yeomans, and Benoit Ladoux. Topological defects in epithelia govern cell death and extrusion. Nature Publishing Group, 544(7649):212 -- 216, April 2017. [43] Ananyo Maitra, Pragya Srivastava, Madan Rao, and Sriram Ramaswamy. Activating Membranes. Physical Review Letters, 112(25), June 2014. doi:10.1103/PhysRevLett.112.258101. [44] Verena Ruprecht, Stefan Wieser, Andrew Callan-Jones, Michael Smutny, Hitoshi Morita, Keisuke Sako, Vanessa Barone, Monika Ritsch-Marte, Michael Sixt, Raphaël Voituriez, and Carl-Philipp Heisenberg. Cor- tical Contractility Triggers a Stochastic Switch to Fast Amoeboid Cell Motility. Cell, 160(4):673 -- 685, February 2015. doi:10.1016/j.cell.2015.01.008. [45] Anne-Cecile Reymann, Fabio Staniscia, Anna Erzberger, Guillaume Salbreux, and Stephan W Grill. Corti- cal flow aligns actin filaments to form a furrow. eLife, 5, October 2016. doi:10.7554/eLife.17807. [46] Carl A. Whitfield and Rhoda J. Hawkins. Immersed Boundary Simulations of Active Fluid Droplets. PLOS ONE, 11(9):e0162474, September 2016. doi:10.1371/journal.pone.0162474. [47] Natalie C. Heer, Pearson W. Miller, Soline Chanet, Norbert Stoop, Jörn Dunkel, and Adam C. Martin. Actomyosin-based tissue folding requires a multicellular myosin gradient. Development, 144(10):1876 -- 1886, May 2017. doi:10.1242/dev.146761. [48] Nils Klughammer, Johanna Bischof, Nikolas D. Schnellbächer, Andrea Callegari, Péter Lénárt, and Ul- rich S. Schwarz. Cytoplasmic flows in starfish oocytes are fully determined by cortical contractions. PLOS Computational Biology, 14(11):e1006588, November 2018. doi:10.1371/journal.pcbi.1006588. [49] Dominique Barthès-Biesel. Modeling the motion of capsules in flow. Current Opinion in Colloid & Inter- face Science, 16(1):3 -- 12, February 2011. doi:10.1016/j.cocis.2010.07.001. [50] Jonathan B. Freund. Numerical Simulation of Flowing Blood Cells. Annual Review of Fluid Mechanics, 46 (1):67 -- 95, January 2014. doi:10.1146/annurev-fluid-010313-141349. 29 [51] Alexander Farutin, Thierry Biben, and Chaouqi Misbah. 3D numerical simulations of vesicle and Journal of Computational Physics, 275:539 -- 568, October 2014. inextensible capsule dynamics. doi:10.1016/j.jcp.2014.07.008. [52] Achim Guckenberger, Marcel P. Schraml, Paul G. Chen, Marc Leonetti, and Stephan Gekle. On the bend- ing algorithms for soft objects in flows. Computer Physics Communications, 207:1 -- 23, October 2016. doi:10.1016/j.cpc.2016.04.018. [53] Timothy W. Secomb. Blood Flow in the Microcirculation. Annual Review of Fluid Mechanics, 49(1): 443 -- 461, January 2017. doi:10.1146/annurev-fluid-010816-060302. [54] Peter Balogh and Prosenjit Bagchi. flow in complex geometry. doi:10.1016/j.jcp.2017.01.007. A computational approach to modeling cellular-scale blood Journal of Computational Physics, 334:280 -- 307, April 2017. [55] Christian Bächer, Lukas Schrack, and Stephan Gekle. Clustering of microscopic particles in constricted blood flow. Physical Review Fluids, 2(1):013102, January 2017. doi:10.1103/PhysRevFluids.2.013102. [56] Achim Guckenberger, Alexander Kihm, Thomas John, Christian Wagner, and Stephan Gekle. Numeri- cal -- experimental observation of shape bistability of red blood cells flowing in a microchannel. Soft Matter, 14(11):2032 -- 2043, 2018. doi:10.1039/C7SM02272G. [57] Sauro Succi. The Lattice Boltzmann Equation: For Fluid Dynamics and Beyond. Oxford university press, 2001. [58] Burkhard Dünweg and Anthony JC Ladd. Lattice Boltzmann simulations of soft matter systems. In Ad- vanced Computer Simulation Approaches for Soft Matter Sciences III, pages 89 -- 166. Springer, 2009. [59] Cyrus K. Aidun and Jonathan R. Clausen. Lattice-Boltzmann Method for Complex Flows. Annual Review of Fluid Mechanics, 42(1):439 -- 472, January 2010. doi:10.1146/annurev-fluid-121108-145519. [60] Timm Krüger, Halim Kusumaatmaja, Alexandr Kuzmin, Orest Shardt, Goncalo Silva, and Erlend Magnus Viggen. The Lattice Boltzmann Method: Principles and Practice. Springer, December 2016. [61] Stewart Harris. An Introduction to the Theory of the Boltzmann Equation. Dover, Mineola, NY, 2004. OCLC: 845367147. [62] Burkhard Dünweg, Ulf D. Schiller, and Anthony J. C. Ladd. the Physical Review E, 76(3):036704, September 2007. Statistical mechanics of fluctuating lattice Boltzmann equation. doi:10.1103/PhysRevE.76.036704. [63] Xiaoyi He, Shiyi Chen, and Gary D. Doolen. A Novel Thermal Model for the Lattice Boltzmann Journal of Computational Physics, 146(1):282 -- 300, October 1998. Method in Incompressible Limit. doi:10.1006/jcph.1998.6057. [64] Rodrigo C. V. Coelho, Anderson Ilha, Mauro M. Doria, R. M. Pereira, and Valter Yoshihiko Aibe. Lattice Boltzmann method for bosons and fermions and the fourth-order Hermite polynomial expansion. Physical Review E, 89(4):043302, April 2014. doi:10.1103/PhysRevE.89.043302. [65] Rodrigo C.V. Coelho and Mauro M. Doria. Lattice Boltzmann method for semiclassical fluids. Computers & Fluids, 165:144 -- 159, March 2018. doi:10.1016/j.compfluid.2018.01.019. [66] Zunmin Zhang, Wei Chien, Ewan Henry, Dmitry A. Fedosov, and Gerhard Gompper. Sharp-edged geomet- ric obstacles in microfluidics promote deformability-based sorting of cells. Physical Review Fluids, 4(2): 024201, February 2019. doi:10.1103/PhysRevFluids.4.024201. 30 [67] H.J. Limbach, A. Arnold, B.A. Mann, and C. Holm. ESPResSo -- an extensible simulation package for research on soft matter systems. Computer Physics Communications, 174(9):704 -- 727, May 2006. doi:10.1016/j.cpc.2005.10.005. [68] D. Roehm and A. Arnold. Lattice Boltzmann simulations on GPUs with ESPResSo. The European Physical Journal Special Topics, 210(1):89 -- 100, August 2012. doi:10.1140/epjst/e2012-01639-6. [69] Axel Arnold, Olaf Lenz, Stefan Kesselheim, Rudolf Weeber, Florian Fahrenberger, Dominic Roehm, Peter Košovan, and Christian Holm. ESPResSo 3.1: Molecular Dynamics Software for Coarse-Grained Models. In Meshfree Methods for Partial Differential Equations VI, pages 1 -- 23. Springer, Berlin, Heidelberg, 2013. doi:10.1007/978-3-642-32979-1_1. [70] Charles S. Peskin. The immersed boundary method. Acta Numerica, 11, January 2002. doi:10.1017/S0962492902000077. [71] Rajat Mittal and Gianluca Iaccarino. Immersed Boundary Methods. Annual Review of Fluid Mechanics, 37 (1):239 -- 261, January 2005. doi:10.1146/annurev.fluid.37.061903.175743. [72] L. Mountrakis, E. Lorenz, and A. G. Hoekstra. Revisiting the use of the immersed-boundary lattice- Physical Review E, 96(1), July 2017. Boltzmann method for simulations of suspended particles. doi:10.1103/PhysRevE.96.013302. [73] Albert Edward Green and Wolfgang Zerna. Theoretical Elasticity. Clarendon Press, Oxford, 1954. [74] Albert Edward Green and John Edward Adkins. Large Elastic Deformations and Non-Linear Continuum Mechanics. Clarendon Press, 1960. [75] W. T. Koiter. On the mathematical foundation of shell theory. In Proc. Int. Congr. of Mathematics, Nice, volume 3, pages 123 -- 130, 1970. [76] C. Pozrikidis. Interfacial Dynamics for Stokes Flow. Journal of Computational Physics, 169(2):250 -- 301, May 2001. doi:10.1006/jcph.2000.6582. [77] Yunchang Seol, Wei-Fan Hu, Yongsam Kim, and Ming-Chih Lai. An immersed boundary method for simulating vesicle dynamics in three dimensions. Journal of Computational Physics, 322:125 -- 141, October 2016. doi:10.1016/j.jcp.2016.06.035. [78] Achim Guckenberger and Stephan Gekle. Theory and algorithms to compute Helfrich bending forces: A review. Journal of Physics: Condensed Matter, 29(20):203001, May 2017. doi:10.1088/1361-648X/aa6313. [79] Patrick Ahlrichs and Burkhard Dünweg. Simulation of a single polymer chain in solution by combining lattice Boltzmann and molecular dynamics. The Journal of Chemical Physics, 111(17):8225 -- 8239, October 1999. doi:10.1063/1.480156. [80] S. Quint, A. F. Christ, A. Guckenberger, S. Himbert, L. Kaestner, S. Gekle, and C. Wagner. 3D tomography of cells in micro-channels. Applied Physics Letters, 111(10):103701, September 2017. doi:10.1063/1.4986392. [81] Stephan Gekle. Strongly Accelerated Margination of Active Particles in Blood Flow. Biophysical Journal, 110(2):514 -- 520, January 2016. doi:10.1016/j.bpj.2015.12.005. [82] Christian Bächer, Alexander Kihm, Lukas Schrack, Lars Kaestner, Matthias W. Laschke, Christian Wagner, and Stephan Gekle. Antimargination of Microparticles and Platelets in the Vicinity of Branching Vessels. Biophysical Journal, 115(2):411 -- 425, July 2018. doi:10.1016/j.bpj.2018.06.013. [83] Dmitry A. Fedosov, Matti Peltomäki, and Gerhard Gompper. Deformation and dynamics of red Soft Matter, 10(24):4258 -- 4267, May 2014. blood cells in flow through cylindrical microchannels. doi:10.1039/C4SM00248B. 31 [84] Markus Deserno. Fluid lipid membranes: From differential geometry to curvature stresses. Chemistry and Physics of Lipids, 185:11 -- 45, January 2015. doi:10.1016/j.chemphyslip.2014.05.001. [85] Roman Vetter, Norbert Stoop, Thomas Jenni, Falk K. Wittel, and Hans J. Herrmann. Subdivision shell International Journal for Numerical Methods in Engineering, 95(9): elements with anisotropic growth. 791 -- 810, August 2013. doi:10.1002/nme.4536. [86] Jemal Guven. Membrane geometry with auxiliary variables and quadratic constraints. Journal of Physics A: Mathematical and General, 37(28):L313 -- L319, July 2004. doi:10.1088/0305-4470/37/28/L02. [87] Abdallah Daddi-Moussa-Ider, Achim Guckenberger, and Stephan Gekle. Long-lived anomalous thermal diffusion induced by elastic cell membranes on nearby particles. Physical Review E, 93(1), January 2016. doi:10.1103/PhysRevE.93.012612. [88] Dominique Barthès-Biesel. Motion and Deformation of Elastic Capsules and Vesicles in Flow. Annual Review of Fluid Mechanics, 48(1):25 -- 52, January 2016. doi:10.1146/annurev-fluid-122414-034345. [89] R. Skalak, A. Tozeren, R. P. Zarda, and S. Chien. Strain energy function of red blood cell membranes. Biophysical Journal, 13(3):245, 1973. [90] W Helfrich. Elastic Properties of Lipid Bilayers: Theory and Possible Experiments. Zeitschrift für Natur- forschung C, 28(11-12):693 -- 703, 1973. [91] D. J. Steigmann. Fluid Films with Curvature Elasticity. Archive for Rational Mechanics and Analysis, 150 (2):127 -- 152, December 1999. doi:10.1007/s002050050183. [92] Roger A. Sauer, Thang X. Duong, Kranthi K. Mandadapu, and David J. Steigmann. A stabilized finite element formulation for liquid shells and its application to lipid bilayers. Journal of Computational Physics, 330:436 -- 466, 2016. doi:10.1016/j.jcp.2016.11.004. [93] Martin Kraus, Wolfgang Wintz, Udo Seifert, and Reinhard Lipowsky. Fluid vesicles in shear flow. Physical review letters, 77(17):3685, 1996. [94] Mark Meyer, Mathieu Desbrun, Peter Schröder, and Alan H. Barr. Discrete differential-geometry operators for triangulated 2-manifolds. In Visualization and Mathematics III, pages 35 -- 57. Springer, 2003. [95] DS Chandrasekharaiah and Lokenath Debnath. Continuum Mechanics. Elsevier, 2014. [96] See supplemental material at [url will be inserted by publisher] for an axisymmetric algorithm simulating an acitve cylindrical membrane. [97] Thomas D. Pollard. Nine unanswered questions about cytokinesis. The Journal of Cell Biology, 216(10): 3007 -- 3016, October 2017. doi:10.1083/jcb.201612068. [98] Yibao Li, Ana Yun, and Junseok Kim. An immersed boundary method for simulating a single ax- isymmetric cell growth and division. Journal of Mathematical Biology, 65(4):653 -- 675, October 2012. doi:10.1007/s00285-011-0476-7. [99] Jia Zhao and Qi Wang. Modeling cytokinesis of eukaryotic cells driven by the actomyosin contrac- tile ring: MODELING AND SIMULATIONS OF CYTOKINESIS OF EUKARYOTIC CELLS. Inter- national Journal for Numerical Methods in Biomedical Engineering, 32(12):e02774, December 2016. doi:10.1002/cnm.2774. [100] Seunggyu Lee. Mathematical Model of Contractile Ring-Driven Cytokinesis in a Three-Dimensional Do- main. Bulletin of Mathematical Biology, 80(3):583 -- 597, March 2018. doi:10.1007/s11538-018-0390-x. [101] Markus Bender, Jonathan N. Thon, Allen J. Ehrlicher, Stephen Wu, Linas Mazutis, Emoke Deschmann, Martha Sola-Visner, Joseph E. Italiano, and John H. Hartwig. Microtubule sliding drives proplatelet elon- gation and is dependent on cytoplasmic dynein. Blood, 125(5):860 -- 868, January 2015. 32 [102] Antoine Blin, Anne Le Goff, Aurélie Magniez, Sonia Poirault-Chassac, Bruno Teste, Géraldine Sicot, Kim Anh Nguyen, Feriel S. Hamdi, Mathilde Reyssat, and Dominique Baruch. Microfluidic model of the platelet-generating organ: Beyond bone marrow biomimetics. Scientific Reports, 6:21700, February 2016. doi:10.1038/srep21700. [103] E. Lac, D. Barthès-Biesel, N. A. Pelekasis, and J. Tsamopoulos. Spherical capsules in three-dimensional unbounded Stokes flows: Effect of the membrane constitutive law and onset of buckling. Journal of Fluid Mechanics, 516:303 -- 334, October 2004. doi:10.1017/S002211200400062X. [104] Abdallah Daddi-Moussa-Ider, Maciej Lisicki, and Stephan Gekle. Hydrodynamic mobility of a sphere Physics of Fluids, 29(11):111901, November 2017. moving on the centerline of an elastic tube. doi:10.1063/1.5002192. [105] C. Pozrikidis. Effect of membrane bending stiffness on the deformation of capsules in simple shear flow. Journal of Fluid Mechanics, 440, August 2001. doi:10.1017/S0022112001004657. 33 List of Figures 1 2 3 4 5 6 7 . . ij and σout ij . . a) D3Q19 LBM scheme with the discrete velocity set (solid arrows) connecting nearest and next nearest neighboring fluid nodes (dots). b) 3D illustration of the Immersed Boundary Method. A continuous membrane is discretized by Lagrangian nodes (red and orange dots) that are connected by triangles. The membrane is immersed in an Eulerian grid representing the fluid (blue dots). The velocity at a Lagrangian node is obtained by interpolation from the eight closest fluid nodes (illustrated for the two orange membrane nodes in the middle by the blue shaded cubes). The same stencil is used to transmit the forces from the membrane to the fluid. . . . . . . . . . . . . . . . (left) The cell cortex underlying the plasma membrane consists of cytoskeletal filaments and motor proteins. The latter constantly convert energy into mechanical work. (right) Membrane and cortex are condensed into a thin shell with normal vector n and in-plane coordinates e1, e2. Mechanical work of motor proteins results in an active in-plane surface stress t β aα. Forces from the cytoplasm inside the cell and from the extra-cellular medium onto the membrane can be treated by the 3D fluid stress tensors σin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The central node at rc is surrounded by N neighbors. The discretization is chosen such that all nodes on the cylindrical mesh in figure 4 have 6 neighbors. For the ellipsoid in figure 10, topology requires at least 12 triangles with N = 5 while all others have N = 6. From the surrounding triangles a local normal vector n at rc can be computed. Together with the position of one neighbor local in-plane coordinate vectors eξ, eη can be constructed. Applying the first step of Gram-Schmidt orthonormalization leads to eξ and the cross product of n and eξ to eη. . . . . a) We consider a local increase in active in-plane surface stress of a cylindrical membrane. b) and c) Membrane meshes as in main text with two different resolutions a) ∆z = 0.2, ∆φ = 0.2 and b) ∆z = 0.1, ∆φ = 0.15, respectively. . . . . . . . . . . . . . . . . . . . . . . . . The final deformation resulting from three dimensional simulations of a singular active stress shows the same shape as the analytical prediction of the Green's function. However, peak height deviates from the theory. This can be attributed to the singular nature of the perturbation which a = −0.01, is difficult to resolve numerically. Deformations are obtained for the parameter set T φ κS = 1, C = 1, and κB = 10−3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . a) Comparison of the deformation obtained by the 3D overdamped dynamics with the theoreti- cal expectation for a cylindrical shell with 160 nodes along z-direction subjected to a Gaussian distributed active in-plane surface stress with κS = 1, κB = 10−5, and T φ a = −0.01. 3D sim- ulations are in very good agreement with the theory. b) Comparison of the deformation obtained with LBM/IBM simulations for C = 1 and κB ≈ 10−4 with the theoretical prediction for the same setup with perturbation in φ-stress T φ a . d) Increasing resolution of both membrane and fluid mesh show convergence of the relative error per node for the parameter set of b) for a perturbation in T φ a (red dots) as well as for the parameter set of c) for and proportional to a perturbation in T z N−2 . . . . . . . . . . Phase diagram of a cylindrical shell with relative, active in-plane surface stress g = Ta/S = 3Ta/(2κS) and relative bending modulus b = B/(SR2) = 3κB/(4κSR2) from 3D LBM/IBM simulations in comparison to the theoretically predicted thresholds [3]. On the left of the dotted line we apply a negative active stress, on the right of the dotted line a positive active stress. For small negative or positive active stress (around the dotted line) the cylindrical membrane is stable. For large negative stresses, a buckling instability is observed. For large positive active in-plane surface stress a Rayleigh-Plateau like instability occurs. The theoretically predicted instability thresholds [3] and our 3D LBM/IBM simulations are in excellent agreement in both cases. Insets illustrate the shape of the shell corresponding to different values of active in-plane surface stress. The labels a, b, c refer to figure 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . a (blue triangles). The error decreases proportional to N−1 , respectively. In b) -- c) simulations are done for 300 nodes in z-direction. a and c) with perturbation in z-stress T z . . . . . . . . . . . z z 4 6 8 13 13 16 16 34 9 8 10 Membrane shape for different values of negative active in-plane surface stress for fixed bending modulus b = 0.01125. In the upper row the membrane is subjected to both z- and φ-stress, while in the lower row the membrane is subjected to z-stress only. For z-stress only we observe a buckling instability in a). However, for isotropic stress we observe an instability introducing non- axisymmetric deformations at intermediate stresses in b) where no buckling instability occurs. c) For smaller active stress the cylindrical membrane remains stable in both cases. Corresponding points in the phase diagram in figure 7 are labeled with a to c. . . . . . . . . . . . . . . . . . . Comparison of the deformation obtained for the buckling instability of an initially cylindrical mem- brane subjected to a homogeneous active in-plane surface stress with g = −0.75 and b = 0.01125. We compare the wavelength of the deformation obtained in 3D LBM/IBM simulations with differ- ent resolutions ∆z/R to one obtained by axisymmetric simulation. For a sample illustration of the 3D shape we refer to the inset on the left hand side of reference 7. All simulations show the same wavelength for the buckling instability. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . a) Similar to cell division an elastic cell membrane, which is subjected to homogeneous active stress and a local increase in azimuthal active stress in the red shaded area, contracts locally as shown in b). (c) - (h) show the outline of the deforming membrane (red nodes) in the central plane and the developing flow field (arrows) inside the cell over time. Eventually, a flow away from the contracting region towards the poles of the cell is observed. Arrows indicate the flow direction while the color indicates the flow velocity. Relative time is t/tmax = a) 0, b) 1, c) 4×10−4, d) 0.08, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . e) 0.16, f) 0.51, g) 0.86, h) 1.0 . 11 An elastic dividing cell membrane in shear flow. a) An increase in azimuthal active stress in the red shaded area of an ellipsoidal cell membrane triggers a local contraction as shown in b) while the externally driven shear flow also deforms the membrane. (c) - (h) show the outline of the deforming membrane (red nodes) in the central plane and the developing flow field (arrows) as perturbation to the shear flow over time. We here refer also to the Supplemental Video 1. The presence of the shear flow renders both the cell shape and the flow field asymmetric. Relative time is t/tmax = a) 0, b) 1, c) 1.3×10−3, d) 0.07, e) 0.17, f) 0.52, g) 0.88, h) 1.0 . . . . . . . . . . . . a) Color code for the neighbor node serving as reference to construct the local coordinate system on the cylindrical membrane at the site of the central node. b) Deformation obtained for a cylindrical membrane subjected to a Gaussian perturbation in active in-plane surface stress along z-direction a = 0.01 for different reference neighbor nodes. The setup is identical to figure 6 c) with T z and κS = 1.0. Obtained deformations are in very good agreement regardless of the choice of reference neighbor node. c) Neighbor node serving as reference to construct the local coordinate system on the ellipsoidal membrane at the site of the central node. d) Cell membrane shape for different reference neighboring node νref at a given time corresponding to figure 10 g). Except for slight deviations in the region of largest curvature in case of neighboring node 2 and 5, all membrane shapes are in very good agreement. e) Membrane shapes in shear flow for different reference neighboring node νref at a given time corresponding to figure 11 g) are in very good agreement. Thus, we prove evidence that the choice of the reference neighboring node does not affect simulation results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A1 . . . . . . . 17 . 18 . 19 . 20 . 26 35 SupportingInformation-Computationalmodellingofactivedeformablemembranesembeddedin3DflowsC.Bächer1andS.Gekle11BiofluidSimulationandModeling,TheoretischePhysikVI,UniversitätBayreuth,Bayreuth,GermanyS1NumericsforanaxisymmetricactiveelasticcylinderHereweprovideadetailedderivationofthenumericalmethodusedtoobtaintheaxisymmetricshapesshowninfigure9ofthemaintext.S1.1Axisymmetric,cylindricalshellWeconsideracylindricalshellundertheassumptionofaxisymmetry.Inthefollowingwecalculatethetractionjumpfortheaxisymmetricmembraneforlinearelasticity.Theshellisparametrizedusingcylindricalcoordianteswiths1=zands2=φsothatX(z,φ)=Rcosφex+Rsinφey+zez.(S1)Weobtainforthein-planecoordinatesez=(0,0,1),eφ=(−Rsinφ,Rcosφ,0),andn=(−cosφ,−sinφ,0).WefollowBerthoumieuxetal.[1]regardingconventions,thusthenormalvectorpointsinwardsintothecylinder.Themetricandcurvaturetensorbecomegαβ= 100R2!Cαβ= 000R!.(S2)Duetoaxisymmetrythedeformationcanbewrittenasu=uz(z)ez+un(z)n.(S3)andfollowingequation(S3)weobtainthemetriconthedeformedsurfaceusingg0αβ=e0α·e0βg0αβ= 1+2∂zuz00R2−2Run!.(S4)Accordingly,thein-planestraintensorcanbecalculateduαβ= ∂zuz00−Run!.(S5)WiththeexpressionsabovewecancalculatethecurvaturetensorC0αβ=(∂α∂βX0)·n0(S6)1 onthedeformedmembranetobeC0αβ= ∂2zun00R−un!.(S7)TheLevi-Civitatensoronthedeformedsurfaceforsmalldeformationsis0αβ=(R+R∂zuz−un) 01−10!.(S8)TheChristoffelsymbolsonthedeformedmembraneareΓz0zz'∂2zuz,Γz0φφ'R∂zun,Γφ0zφ'−1R∂zun,Γφ0φz'−1R∂zun,Γφ0φφ=0,Γφ0zz=0,Γz0φz=0,Γz0zφ=0.S1.2ForcebalanceinaxisymmetricformulationInthissectionweconsidertheforcebalanceequationsinthecaseofanaxisymmetric,cylindricalshell.Theforcebalance(eqs.(11)-(12)inthemaintext)takesthefollowingformincylindricalcoordinates∇0ztzza+∇0φtφza−C0zztzna−C0zφtφna+fze=fz(S9)∇0ztzφa+∇0φtφφa−C0φztzna−C0φφtφna+fφe=fφ(S10)∇0ztzna+∇0φtφna+C0zztzza+C0zφtzφa+C0φztφza+C0φφtφφa+fne=fn.(S11)ElasticforcesIntheframeworkofour3Dsimulationspresentedinthemaintext,elasticforcesarecomputedbydirectderivationofadiscretizedenergyfunctionalwithrespecttonodalpositions.Here,wefollowadifferentrouteanduseelasticin-planesurfacestressesinordertofollowcloselyref.[1].ElasticpropertiesoftheshellareconsideredintheframeworkofHooke'slawforathreedimensional,elastic,isotropicsolidwiththestresstensorσij=E1+ν(cid:18)eij+ν1−2νekkδij(cid:19)(S12)dependinginalinearfashiononthestraintensoreij=ui,j+uj,iwithEbeingtheYoung'smodulusandνthePoissonratio.Fromthebulkequation(S12)expressionsfortheintriniscin-planesurfacestresstensorandmomentsonthethinshellcanbederived[1]¯teαβ=2S ∂zuz−νunR00νR2∂zuz−unR!(S13)¯meαβ=2BR 0−R2∂2zun−νunν∂2zun+unR20!,(S14)withSthestretchingandBthebendingmodulus.FollowingSalbreuxandJülicher[2]byusingeq.(16)and(17)of[2]wecalculatethein-planesurfacestressandmomenttensorviatαβe=¯tαβe+12(cid:16)mγαCβγ+mγβCαγ(cid:17)(S15)mαβe=−¯mαγeβγ(S16)andobtaintzze=¯tzze(S17)tzφe=¯tzφe(S18)tφze=¯tφze(S19)tφφe=¯tφφe−2BνR3∂2zun−2BR5un(S20)2 andmzze=−2B∂2zun+2BνR2un(S21)mzφe=0(S22)mφze=0(S23)mφφe=−2BνR2∂2zun−2BR4un.(S24)Weobtainforthenormalsurfacestresstzne=∂zmzz+2(∂2zuz)mzz+(−1R∂zun)mzz+(R∂zun)mφφ(S25)tφne=0,(S26)andbyinsertingthein-planemomentsweeventuallygettzne=−2B∂3zun+2BνR2∂zun.(S27)Thusfollowingfβe=∇0αtαβe+C0βαtαn,e(S28)fne=∇0αtαn,e−C0αβtαβe(S29)wegetfortheelasticforcesfze=2S∂2zuz−2SνR∂zun(S30)fφe=0(S31)fne=−2B∂4zun−(cid:18)2BR4+2SR2(cid:19)un+2SνR∂zuz(S32)ActiveforcesWeneglectthecontributionsofactivemomentsandnormalstressandusetheactivein-planesurfacestressinthegeneralformtβaα= tza(z)00tφa(z)!.(S33)Fromtheforcebalanceincylindricalcoordinatesweobtaintheactiveforcewithcomponentsfza=∂ztza−1Rtza∂zun+1Rtφa∂zun(S34)fφa=0(S35)fna=∂2zuntza+(cid:18)1R+1R2un(cid:19)tφa.(S36)Bothelasticforceandactiveforcetogetherdeterminethetractionjump,whichisrequiredforsimulations.S1.3NumericalmethodInordertodeterminetheshapeofthemembraneinthesteadystatefortheaxisymmetricproblemweperformanoverdampedrelaxation.Wediscretizethecontourofthemembrane(whichisalineincaseoftheaxisymmetricformulation)byaseriesofmarkerpoints.Accordingtothetractionjumpinequations(S9)to(S11)wecalculatetheforcesofthemembraneusingquinticsplinesforthederivativesofthedeformation.Oncetheforcesaredetermined,weconsidertheequationsofmotionintheoverdampedlimitforeachmembranenodeF=γr.(S37)Thismeansweintroduceafrictiontermintheequationsofmotionandneglectinertia.EquationsofmotionaresolvedbytheEulerintegrationscheme.Asboundaryconditionforthedeformationwechoosethefirstandsecondderivativetovanishonbothends.3 -0.002 0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016-4-2 0 2 4un / Rz / RGreen's functionsimulation Δz = 0.10simulation Δz = 0.05simulation Δz = 0.025Fig.S1:ComparisonofthenumericalsolutionoftheaxisymmetricproblemforthreedifferentresolutionswiththeGreen'sfunctionana-lyticallypredictedbyBerthoumieuxetal.[1].Numericsandanalyticsagreeverywellinthecaseofasmall,local,activein-planesurfacestressTφa=−0.025.ShapesareobtainedforR=1,S=1,b=0.02,andg=0.S1.4ResultsInordertovalidateouraxisymmetricnumericalmethod,wefirstcomparethenumericalresultstothepredictedGreen'sfunctionofBerthoumieuxetal.[1].FortheintroductionoftheGreen'sfunctionwerefertothesectionV.Aofthemaintext.AsinsectionV.Bofthemaintextweapplyanactivein-planesurfacestresstensoroftheformtβaα= 000Tφa!δ(z).(S38)WecompareaxisymmetricsimulationstotheanalyticalGreen'sfunctioninfigureS1fortheparametersetTφa=−0.025,Tza=0,b=0.02,g=0forthreedifferentresolutions∆z.FigureS1showsthatournumericalmethodontheonehanddoesnotdependonspatialdiscretizationandontheotherhandisinverygoodagreementwiththetheoreticalexpecteddeformationduetotheGreen'sfunctiongivenbyref.[1]includingtheshallowminimanexttothemainpeak.AsinsectionV.Dofthemaintextforthefull3Dmethod,wenextconsiderafinite,homogeneous,activein-planesurfacestressta6=0withoutanysingularperturbation,i.e.,Tza,Tφa=0.InthiscaseBerthoumieuxetal.[1]predictstwokindsofinstabilitiesdependingonthemagnitudeofactivein-planesurfacestress:fornegativeactivein-planesurfacestressbeyondta<−2q3BSR2abucklinginstabilitytakesplaceandforpositiveactivein-planesurfacestresstheypredictaninstabilityforta>2S(1−ν2).Thisallowsustoperformsimulationsfordifferentparametersets(g,b)andtoconstructaphasediagramclassifyingthefinalshape.TheresultingphasediagramisshowninfigureS2.Fornegative,activein-planesurfacestressg<0ouraxisymmetricnumericsareinverygoodagreementwiththepredictedinstabilitythresholdforabroadrangeofrelativebendingmodulib.Inthecaseofpositive,activein-planesurfacestressg>0thecylindricalshellcontracts.Whereasthemembranein3DshowsaninstabilitysimilartoaRayleigh-Plateauinstabilityofaliquidjet,in2Dthecylindricalmembranecontractshomogeneously,sincevolumeconservationisnotenforcedintheaxisymmetriccase.Beyondacertainthreshold,thedeformationbecomeslargerthanthecylinderradiusR.Thisservesusasacriterionforaninstability.TheGreen'sfunctionfoundbyBerthoumieuxetal.[1]divergesforg=1.5,whichisbeyondourinstabilitythreshold.ThediscrepancymaybeexplainedbythefactthatourcriterionismorerealisticthanthatofBerthoumieuxetal.whichisbasedontheGreen'sfunctiononly.Nevertheless,wenotethatourinstabilitythresholddoesnotdependonthebendingmodulus,aspredictedbyref.[1].InordertodiscussthelatterissueinfurtherdetailweshowinfigureS3thefinaldeformationufinaln/Rdependingontherelative,activein-planesurfacestressg.Wedefinetheinstabilitythresholdbyufinaln=R.Beyondthisthresholdthedeformationincreasesstrongly,butournumericsdonotcoverthedivergencepredicted.However,4 0 0.01 0.02 0.03 0.04 0.05-3-2.5-2-1.5-1-0.5 0 0.5 1 1.5b = B / S R2g = ta / S◯⬤✷stablebuckling instabilityinstability⬤⬤⬤⬤⬤◯◯◯◯◯◯◯◯◯◯⬤⬤⬤⬤⬤⬤⬤◯◯◯⬤⬤⬤⬤⬤◯✷◯◯◯◯◯◯◯✷✷✷✷✷◯◯◯◯◯⬤⬤⬤⬤◯◯✷✷✷✷✷✷✷◯◯◯◯◯✷◯◯◯◯◯◯◯⬤⬤⬤✷◯◯◯◯◯✷✷◯◯◯⬤◯◯◯◯◯◯◯✷✷Fig.S2:Phasediagramofacylindricalshellwithhomogeneous,relative,activein-planesurfacestressg=ta/Sandrelativebendingmodulusb=B/(SR2)foraxisymmetricsimulations.ThePoissonratioisν=12.Fornegativeactivein-planesurfacestressabucklinginstabilityoccurs.Thesimulationsmatchthepredicted[1]criticalactivein-planesurfacestressverywell.Forpositiveactivein-planesurfacestressesaninstabilityoccurscharacterizedbyadeformationlargerthanthecylinderradiuswhichhappensevenbeforethethresholdpredictedby[1](orangedottedline).Insetsshowmembraneshapescorrespondingtothedifferentphases. 0 0.5 1 1.5 2 2.5 3 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 unfinal / Rg = ta / SFig.S3:Finaldeformationofacylinderwithhomogeneousactivein-planesurfacestressintheaxisymmetricsimulationsforabendingmodulusb≈0.015.Nearthethresholdg=1.5thedeformationincreasesstronglyandnearlydiverges(divergencenotcompletelycoveredbynumerics).5 wehavetonotethatthebehaviorbeyondthethresholdisnotphysicalinastrictsense.Inconclusion,thissectionprovesthatouraxisymmetricnumericalsimulationsareinverygoodagreementwiththepredictedtheoryandconsequentlycanbeusedforvalidationofthethreedimensionalmethodasdoneinfigure8ofthemaintext.References[1]H.Berthoumieux,J.-L.Maître,C.-P.Heisenberg,E.K.Paluch,F.Jülicher,andG.Salbreux.Activeelasticthinshelltheoryforcellulardeformations.NewJournalofPhysics,16(6):065005,June2014.[2]G.SalbreuxandF.Jülicher.Mechanicsofactivesurfaces.PhysicalReviewE,96(3),Sept.2017.6
1012.4798
1
1012
2010-12-21T20:57:49
The collective motion of nematodes in a thin liquid layer
[ "physics.bio-ph", "physics.flu-dyn" ]
Many organisms live in confined fluidic environments such as the thin liquid layers on the skin of host organisms or in partially- saturated soil. We investigate the collective behaviour of nematodes in a thin liquid layer, which was first observed by Gray and Lissmann, [J. Exp. Biol. 41, 135 (1964)]. We show experimentally that nematodes confined by a thin liquid film come into contact and only separate again after some intervention. We attribute this collective motion to an attractive force between them arising from the surface tension of the layer and show that for nearby nematodes this force is typically stronger than the force that may be exerted by the nematodes' muscles. We believe this to be the first demonstration of the "Cheerios effect" acting on a living organism. However, we find that being grouped together does not significantly alter the body stroke and kinematic performance of the nematode: there are no statistically significant changes of the Strouhal number and the ratio of amplitude to wavelength when aggregated. This result implies that nematodes gain neither a mechanical advantage nor disadvantage by being grouped together; the capillary force merely draws and keeps them together.
physics.bio-ph
physics
The collective motion of nematodes in a thin liquid layer‡ Sean Gart,∗ Dominic Vella,† and Sunghwan Jung∗ Received Xth XXXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XX First published on the web Xth XXXXXXXXXX 200X DOI: 10.1039/b000000x 0 1 0 2 c e D 1 2 ] h p - o i b . s c i s y h p [ 1 v 8 9 7 4 . 2 1 0 1 : v i X r a Many organisms live in confined fluidic environments such as the thin liquid layers on the skin of host organisms or in partially- saturated soil. We investigate the collective behaviour of nematodes in a thin liquid layer, which was first observed by Gray and Lissmann, [J. Exp. Biol. 41, 135 (1964)]. We show experimentally that nematodes confined by a thin liquid film come into contact and only separate again after some intervention. We attribute this collective motion to an attractive force between them arising from the surface tension of the layer and show that for nearby nematodes this force is typically stronger than the force that may be exerted by the nematodes' muscles. We believe this to be the first demonstration of the "Cheerios effect" acting on a living organism. However, we find that being grouped together does not significantly alter the body stroke and kinematic performance of the nematode: there are no statistically significant changes of the Strouhal number and the ratio of amplitude to wavelength when aggregated. This result implies that nematodes gain neither a mechanical advantage nor disadvantage by being grouped together; the capillary force merely draws and keeps them together. 1 Introduction Organisms need to move for a variety of reasons including to avoid predators, to find nutrients or to find a mate. This mo- tion often takes place in fluid environments and hence may in- volve hydrodynamic interactions, which have been studied ex- tensively in recent years. At the microscopic scale, hydrody- namic interactions between organisms can result in the genera- tion of large-scale circulations 1,2, a highly concentrated layer in a shear field 3 or the non-Brownian motion of a test parti- cle 4. This collective behaviour can also enhance mixing and transport of chemical nutrients 5 -- 7. In the majority of recent studies, collective movements take place in the bulk of flu- ids and hence owe their origin to bulk hydrodynamic interac- tions. Here we report the temporary aggregation of nematodes caused by the presence of a liquid -- gas interface. Ostensibly, this is similar to the dancing of Volvox (algal colonies) at the surface of water 8, but the physics is very different. Nematodes locomote by creating undulatory movement similar to other invertebrates. The propagating body-wave is generated by contracting and relaxing muscles under the con- trol of the neuromuscular system. The maximum force that nematodes are capable of generating has been measured in micro-fabricated devices, and shown to be on the order of 1 µN 9. This limiting force suggests that the application of a ∗ Department of Engineering Science and Mechanics, Virginia Polytechnic Institute and State University, Blacksburg, VA 24061, USA. Fax: 01 540-231- 4574; Tel: 01 540-231-5146; E-mail: [email protected] † ITG, Department of Applied Mathematics and Theoretical Physics, Univer- sity of Cambridge, Wilberforce Road, Cambridge, CB3 0WA, UK. ‡ Electronic Supplementary Information (ESI) available: Videos 1 and 2]. See DOI: 10.1039/b000000x/ [Supplementary force with magnitude greater than 1 µN may change or con- strain the organism's behaviour. It is often observed that nematodes migrate up the side of their culture jars and move into a thin liquid film region. When this occurs they are also often observed to aggregate together 10. Qualitatively, such collective behaviour has been found in many types of nematodes (such as the hookworm larvae and parasitic nematodes 11) and has been shown not to originate in chemical interactions 12. In this article we shall suggest that this aggregation is caused by lateral capillary forces -- similar to the "the Cheerios effect" 13 observed with floating objects. Capillary forces arise from the interfacial tension of fluid phase boundaries, and act along the interfaces. They can, therefore, be important for small organisms living close to fluid interfaces; they allow, for example, the pond skater to walk on the surface of water 14. When a liquid film is thinner than the diameter of a nematode body the nematode experi- ences both vertical and lateral forces because of the surface tension of the interface. The vertical component of surface tension pulls its body down against the substrate, lowering the lateral slipping during locomotion and allowing them to crawl on a surface. The horizontal component of surface tension, meanwhile, will vanish when the nematode is alone but may, in principle, lead to an interaction with a nearby nematode. In this paper, we shall investigate the collective behaviour of nematodes in a thin liquid layer. We characterize the locomo- tory performance of nematodes by measuring the geometric parameters of their body stroke and forward moving veloc- ity. We compare this locomotory performance for the cases of aggregated and individual nematodes. We then present a sim- 1 -- 6 1 Fig. 1 Two nematodes (Panagrellus redivivus) merging: (a) two long nematodes and (b) two small nematodes. Typically, it takes about 1 sec for them to merge completely from the initial point of touching. Over several hours some nematodes migrate up the side of the glass. We characterize the motion of nematodes a few cen- timetres above the bulk of the culture medium. At this height, only a thin liquid layer remains on the side glass wall, which is quite different from the slurry bulk culture below. Videos of nematodes are taken using a Sony Handycam with a macro lens and Virtualdub software was used to format video. Data are taken from the same chamber with the same ambient air temperature to avoid any thermotatic behaviours. The MAT- LAB image processing toolbox is used to analyze footage and process data. Peak-to-peak amplitude A and wavelength λ of the nematodes are measured at each stage of the aggregation process. Nematode speed V is recorded by tracking the cen- troid of the body typically over a few periods of undulation. The frequency f of nematode undulation and the length L of each nematode are also measured. Figure 1 shows two nematodes coming together as they crawl in proximity to each other. Typically, their heads touch and their forward motions draw the rest of the bodies together. Once they are grouped, the capillary force makes it difficult for the nematodes to separate themselves. We observe that ne- matodes stay together from as little as a few seconds to as long as a minute depending on whether they are hit by others (see Fig. 2). We often observed events in which two nematodes of dif- ferent sizes aggregate as in Fig. 1(b). Before two nema- todes merge, the individuals have different body strokes and speed. Typically, smaller nematodes have shorter wavelength and amplitude and lower speed. However, after they aggre- gate they form a single body form -- the smaller nematode adopting the body form of the larger nematode. To characterize the aggregation behaviour, the peak-to-peak amplitude over wavelength and non-dimensional Strouhal number (the ratio of lateral to forward velocities) were calcu- lated for individual and paired nematodes. The ratio of wave- form amplitude to wavelength is a primary indicator of chang- ing body stroke in swimming organisms 16 while the Strouhal number gives an indication of the efficiency of the locomo- Fig. 2 Various stages of aggregating, joint, and separating nematodes. Two nematodes are aggregated at t = 6.7 sec, and stay together for more than 10 sec. However, they are hit by a third nematode at t = 17.7 sec ultimately leading to the separation of all three nematodes. (Supplementary Video 1‡) plified theoretical model to estimate the typical lateral forces due to surface tension and compare it with the typical forces generated by nematodes during locomotion. Our study ratio- nalizes the physical basis of the aggregation but demonstrates ultimately that it changes little about the locomotion stroke. 2 Experimental methods 2.1 Nematodes crawling on a thin fluid Panagrellus redivivus was used as the model organism in our experiments. P. redivivus, often called the micro-worm, is a free living, bacteriophagous nematode. It can be found in soil, vegetation, and dung 15, demonstrating its ability to move in various fluidic conditions. Cultures of P. redivivus were main- tained at a temperature of 18 ◦C in a potato medium. P. redi- vivus grows to around 1 mm in length L and 50 µm in diameter 2rc. Nematode cultures were placed in sealed glass chambers. 2 1 -- 6 t = 0 s1 mm0.27 s0.54 s0.81 s1.08 s1.35 s1.62 st = 0 s0.47 s0.94 s1.41 s1.88 s2.35 s2.82 s1 mm(a)(b)t = 0 sect = 2.7 sect = 6.7 sect = 17.7 sect = 21.0 sect = 23.0 sec1 mm122133212131221 Fig. 3 (a) Amplitude versus wavelength, and (b) f A versus forward velocity, V , for individual and grouped nematodes. In each case, the lines correspond to the mean ratios for individual (dashed line) and grouped (solid line) nematodes. tion stroke 17. Data are taken from 24 individual nematodes, and 23 cases of grouped nematodes and the 95% confidence intervals for mean ratio and mean Strouhal number are con- structed for each condition (individual or grouped). The 95% confidence interval for the mean ratio of amplitude to wave- length was (0.79,0.95) amongst individuals whilst amongst grouped nematodes the corresponding confidence interval was (0.76,0.90). Since these confidence intervals overlap, we con- clude that there is no significant difference between the mean ratio of stroke amplitude to wavelength. However, we note that these values are higher than the value reported in crawl- ing nematodes on a soft agar gel (0.63±0.03) 18. The 95% confidence interval of mean Strouhal number for individu- als was (0.54,0.76) while that for grouped nematodes was (0.57,0.84). Since these confidence intervals overlap, we again conclude that there is no statistically significant differ- ence between the mean Strouhal number for grouped and in- dividual nematodes. In conclusion, the features of their loco- motion are not changed by being aggregated. Figure 3 demon- strates this similarity in the locmotion characteristics for indi- vidual and grouped nematodes graphically. To characterize the dominant mechanism for nematode ag- Fig. 4 Collision time vs. crawling time. Collision time is measured as the time interval from point-wise touch to grouped two-body and crawling time is defined as the ratio of body length to crawling velocity. gregation, we consider two possible time-scales associated with the aggregation. One is the capillary time-scale account- ing for the surface tension force that draws them together, de- fined as Tcapillary = 6πµrc/γ where µ is the fluid viscosity (4.0 cP for a filtered potato medium) and γ is the surface tension coefficient (note that this holds even for zipping bodies 19). The other is the crawling time defined as Tcrawl = L/V . Experimentally, we define the collision time as the time in- terval from the initial contact to the grouped two-body. The grouped two-body is referred as a configuration when the two nematodes are touching along their body length but do not necessarily have their heads and tails in the same phase. For example, the second column in Fig. 1(a) is when two nema- todes make the initial contact and the sixth column represents the grouped two-body. Hence, in Fig. 1(a) the collision time is 1.88 sec. A comparison between crawling and observed col- lision time scales is shown in Fig. 4. We observe that the capillary time scale is expected to be on the order of 10−5 sec and hence is significantly smaller than the measured col- lision time scales. Furthermore, the scaling for Tcapillary does not explain the observed relationship between size, speed and collision time. However, the crawling time depends on the body size and the crawling velocity, and is on the order of the measured collision times. From this, it seems clear that the initial aggregation is driven by crawling, rather than by cap- illary forces. We therefore hypothesize that aggregation ini- tially occurs because of collisions but that, once aggregated, the capillary force keeps the two nematodes in one body form. 1 -- 6 3 00.20.40.60.811.200.20.40.60.81(cid:47)A (mm)(cid:104) (mm) IndividualGrouped00.10.20.30.40.50.600.10.20.30.40.5fA (mm/s)V (mm/s) IndividualGrouped02468100123456L/V (sec)Collision Time (sec) 3 Model We now seek to rationalize the observation that nematodes seem to be bound together after having come close to one an- other. We begin by considering the relative importance of the various physical forces in the system. Firstly, we note that the Bond number, Bo = ρgL2/γ ∼ O(0.1): thus surface ten- sion dominates gravity for nematodes. In what follows, there- fore, we shall neglect gravity entirely. We also note that the Reynolds number, which measures the relative importance of inertia to viscous effects, Re = ρV L/µ ∼ O(10−3) and so liq- uid inertia is entirely negligible here. The insignificance of inertia means that velocities are proportional to forces and it is enough to calculate the typical interaction force between two nearby nematodes. In principle, it would then be possible to calculate the perturbation caused to their motion by this force, along the lines of capillary zippering 19, though we have al- ready seen that this perturbation must be small. To estimate the typical interaction force between two nema- todes we calculate the typical force that 'immersion forces' can exert on two horizontal cylinders. This has not been studied before although the force between floating cylinders has been calculated 13,20 as has the force between immersed spheres 21. (The case of immersed objects is different from that of floating objects because in the former case the substrate provides whatever vertical force is required to satisfy the ver- tical force balance condition. The vertical force balance must be considered separately for floating objects.) We consider the idealised setup illustrated in Fig. 5 (a), which is qualitatively similar to that observed for an actively moving nematode (see Fig. 5 (b)). Fig. 5 (a) Schematic of cross-sectional view of two worms. (b) Experimental observations of meniscus around a moving nematode. (Supplementary Video 2‡) 4 1 -- 6 Both between the cylinders and outside, the angle of the meniscus to the horizontal satisfies φ = π − (θc + ψ) at the contact line, where θc is the contact angle on the cylinder. El- ementary geometry then shows that the radii of curvature of the menisci in the inside and outsideregions are Ri = Ro = , d − 2rc sinψi 2sin(θc + ψi) cosθg + cos(cid:0)θc + ψo rc(1− cosψo) (1) (cid:1) . The total horizontal force consists of two components: a force due to surface tension acting at the contact line and a pressure force acting along the wetted perimeter. The horizontal force per unit length due to surface tension acting at the contact line is given by Fγ = γ(cosφi − cosφo) . (2) Here, a positive force corresponds to an attractive interaction. There is also a contribution to the force from the difference in pressures between the inner and outer menisci. The pressure in the fluid in these regions is given by p − pa = − γ R. The inward pressure force per unit length is then (cid:90) ψo (cid:21) Fp = γrc sinϕ Ri dϕ − 0 sinϕ Ro dϕ . (3) (cid:20)(cid:90) ψi 0 Therefore, the total horizontal force per unit length on the cylinder is rc Ri (1−cosψo). (4) (1−cosψi)− rc Ro Ftot γ = cosφi−cosφo + In eqn. (4) we have the force for given values of ψi and ψo. However, we need to estimate these angles to estimate the total horizontal force. To obtain an equation relating ψo and ψi we shall assume equal capillary pressures in the inside and outside regions, since then Ri = Ro. This may be justified physically since it is unlikely that nematodes will be in contact with the solid substrate all along their length. Hence any difference in capillary pressure would be quickly equalized. We shall also make the simplifying assumption that the nematodes and the glass that they sit on are both perfectly wetting so that θc = 0. With these simplifications, we find that the nonlinear terms cancel and we have simply cosψo = sinψi − 1 2 δ (5) where δ = d/2rc > 1. Therefore, the total horizontal force per unit length simplifies to (cid:21)(cid:18) (cid:19) sinψi − cosψi − 1 1 + rc Ri . (6) (cid:20) 2 δ Ftot γ = ψiψooiPiPo2rc50 μmWorm'scross-sectionMeniscus(b)(a)dRiRo nematodes. This result indicates that the body form is not sig- nificantly changed by the capillary force and, further, that no mechanical advantage is achieved. Nevertheless, grouped ne- matodes are observed to stay together for a long time (more than a few minutes in some cases). It is known that many other organisms move and live in circumstances where capil- lary forces are important and so it seems likely that, in some of these instances at least, the capillary force may be exploited to the organism's benefit, as has already been shown for indi- vidual water treaders, Mesovelia 22. 5 Acknowledgements D.V. is supported by an Oppenheimer Early Career Fellow- ship. References 1 T.J. Pedley, and J.O. Kessler, Annu. Rev. Fluid Mech., 24, 313 (1992). 2 A. Be ´Er, H.P. Zhang, E.L. Florin, S.M. Payne, E. Ben- Jacob, and H.L. Swinney, Proc. Nat. Acad. Sci, 106 428 (2009). 3 W. Durham, J. Kessler, and R. Stocker, Science, 323 1067 (2009). (2000). 4 X.L. Wu and A. Libchaber, Phys. Rev. Lett., 84 3017 5 I. Tuval, L. Cisneros, C. Dombrowski, C.W. Wolgemuth, J. Kessler, and R.E. Goldstein, Proc. Natl Acad. Sci. USA, 102 2277 (2005). 6 A. Sokolov, I.S. Aranson, J.O. Kessler, and R.E. Gold- stein, Phys. Rev. Lett., 98 158102 (2007). 7 M.J. Kim, and K.S. Breuer, Phys. Fluids, 16 L78 (2004). 8 K. Drescher, K. C. Leptos, I. Tuval, T. Ishikawa, T. J. Pedley, and R.E. Goldstein, Phys. Rev. Lett., 102 168101 (2009). 9 J.C. Doll, H. Nahid, N. Klejwa, R. Kwon, S.M. Coulthard, B. Petzold, M.B. Goodman, and B.L. Pruitt, Lab Chip, 21, 9, 1449 (2009). 10 J. Gray and H.W. Lissmann, J. Exp. Bio. 41 135 (1964). 11 N.A. Croll, "The behaviour of Nematodes", Edward Arnold, London (1970). 12 N.A. Croll, Nematologica, 16, 382 (1970). 13 D. Vella and L. Mahadevan, Am. J. Phys. 73, 817 (2005). 14 J.W.M. Bush and D.L. Hu, Annu. Rev. Fluid Mech., 38, 339 (2006). 15 N.F. Gray, Ann. App. Biol. 102, 501 (1983). 16 M.E.J. Holwill, "Low Reynolds number undulatory propulsion in organisms of different sizes" in Scale ef- 1 -- 6 5 Fig. 6 Force-displacement curves for different values of ψi: ψi = 1 (red), ψi = 1.5 (green) and ψi = 2 (blue). Here δ = d/2rc and we have assumed that θ = 0, as is likely to be the case for the worms. For this force to be attractive, we require that 2 δ sinψi − where tanα = δ /2. For a given displacement, i.e. given δ , we have that the force is attractive provided that cosψi > 1, that is, sin(cid:0)ψi − α(cid:1) > δ√ (cid:19) (cid:18) δ 2+4 δ√ δ 2 + 4 + arctan δ 2 (7) ψi > arcsin For simplicity, we shall assume that the contact line is pinned so that ψi is constant as the cylinders move closer to- gether. By plotting the force-displacement curves we find that, as expected, the attractive force increases as the two worms get closer together. Some typical results are shown in Fig. 6. These plots also demonstrate that the force only acts over a lengthscale comparable to the radius of the worms; the force is relatively short ranged. This is consistent with the observation that zipping is driven by collision, rather than being driven by surface tension. Finally, the total horizontal force on a worm of length 1 mm could easily reach the order of 50 µN, and consequently overwhelm the typical muscle force that can be exerted by a single nematode ∼ O(1) µN. 4 Discussion We have studied the collective behaviour of nematodes in thin liquid layers, thereby uncovering the possibility of interactions between locomoting organisms without hydrodynamic effects. This effect is observed even when the nematodes are of very different sizes. We have shown that this interaction is not ini- tially driven by surface tension forces but that once an inter- action happens surface tension forces ensure that the nema- todes remain aggregated. However, we have also shown that the characteristics of locomotion (namely the Strouhal number and body form) do not differ between individual and grouped 1.01.52.02.53.001234F/γδ fects in animal locomotion, edited by T.J. Pedley, London: Acadmic Press (1977). 17 G.K. Taylor, R.L. Nudds, and A.L.R. Thomas, Nature, 425, 707 (2003). 18 J. Karbowski, C.J. Cronin, A. Seah, J.E. Medel, D. Cleary, and P.W. Sternberg, J. Theor. Bio. 242, 652 (2006). 19 D. Vella, H.-Y. Kim and L. Mahadevan, J. Fluid Mech 502 89 -- 98 (2004). 20 D.Y.C. Chan, J.D. Henry, and L.R. White, J. Colloid Inter- face Sci. 79 410 (1981). 21 P.A. Kralchevsky and K. Nagayama "Particles at Fluid in- terfaces and Membranes", Elsevier, Amsterdam (2001). 22 D.L. Hu and J.W.M. Bush, Nature, 437, 733-736 (2005). 6 1 -- 6
1305.2340
2
1305
2013-11-01T05:17:04
Radial propagation in population dynamics with density-dependent diffusion
[ "physics.bio-ph", "q-bio.PE" ]
The population dynamics that evolves in the radial symmetric geometry is investigated. The nonlinear reaction-diffusion model, which depends on population density, is employed as the governing equation for this system. The approximate analytical solution to this equation has been found. It shows that the population density evolves from initial state and propagates as the traveling wave-like for the large time scale. One can be mentioned that, if the distance is insufficient large, the curvature has ineluctable influence on density profile and front speed. In comparison, the analytical solution is in agreement with the numerical solution.
physics.bio-ph
physics
Radial propagation in population dynamics with density-dependent diffusion Division of Physics, School of Science, University of Phayao, Mueang Phayao, Phayao 56000, Thailand (Dated: September 21, 2018) Waipot Ngamsaad∗ The population dynamics that evolves in the radial symmetric geometry is investigated. The non- linear reaction-diffusion model, which depends on population density, is employed as the governing equation for this system. The approximate analytical solution to this equation has been found. It shows that the population density evolves from initial state and propagates as the traveling wave- like for the large time scale. One can be mentioned that, if the distance is insufficient large, the curvature has ineluctable influence on density profile and front speed. In comparison, the analytical solution is in agreement with the numerical solution. PACS numbers: 82.40.Ck, 87.23.Cc, 87.18.Hf, 05.45.-a The growth and dispersal of species in populations un- dergo the density spreading as the traveling wave front [1]. This phenomenon becomes an active research topic for many decades. In theoretical framework, the dynam- ics of population can be modeled as diffusion with re- action processes. The paradigmatic model is known as the Fisher equation [2], which has been originated as a model for the population genetics [1, 3 -- 6]. The solution to this equation has demonstrated the propagating as the traveling wave front in population dynamics [1, 2, 7]. This equation and its variant have also appeared in vari- ous systems, including chemical dynamics [1], nerve pulse propagation [3], flow in porous media [8], combustion the- ory [3, 5, 6], wound healing [9, 10], tissue engineering [11, 12] and bacterial pattern formation [1, 13, 14]. In the original Fisher model [2], the evolution of pop- ulation density u(r, t), at spatial position r and time t, is governed by the simplest nonlinear reaction-diffusion equation [1, 2]. The reaction term is modeled as logistic law and it describes the growth of population with lim- ited supply. The movement of individual is modeled as random walk [15], where the diffusion coefficient is con- stant. However, the motion of the biological population is not purely random but they move with sense. To remedy this issue, the directed motion model, in which individ- uals tend to move in the direction of decreasing popu- lations as fast as increasing density, has been proposed [16, 17]. The diffusion coefficient in this model depends on the population density [5, 6, 16 -- 18]. Later, a general form of the logistic law has been found [6]. With these modifications, the density-dependent reaction-diffusion equation or the extended Fisher model has been pre- sented [1, 5, 6, 19, 20], ∂u ∂t uM(cid:19)p = ∇ ·(cid:20)D(cid:18) u ∇u(cid:21) + αu(cid:20)1 −(cid:18) u uM(cid:19)p(cid:21) , (1) where p > 0, D is diffusion constant, α is rate constant and uM is maximum population density. ∗ [email protected] The solution of Eq. (1) in one dimension (1D) is known as sharp traveling wave, propagating with constant front speed [1, 5, 6, 21]. In our previous work, a general form of solution to Eq. (1) in one-dimensional space have been found [20]. This solution shows that the population den- sity evolves, from a specific initial condition, as a self- similar pattern that converges to the traveling wave at large time scale [20]. Although the solution of Eq. (1) in one-dimensional space has been known [1, 5, 6, 20, 21], its behavior in higher dimensions has not well understood. Typically, the population dynamics takes place in two dimensions (2D), sometimes in three dimensions (3D). Therefore, the solution of Eq. (1) in dimension higher than one could provide better insight into the dynamics of population. In this work, we study the population dynamics that evolves with radial symmetric geometry. In this form, the system is governed by the extended Fisher equation (1) in axisymmetric coordinate system. Before describing further, we change following quantities to be dimension- less: u′ = u/uM , t′ = αt and r the radial symmetric extended Fisher equation in dimen- sionless form is given by ′ =p(p + 1)α/Dr. Then, ∂u ∂t = ∂2um ∂r2 + γ r ∂um ∂r + u − um, (2) where m = p + 1, r = r, 0 ≤ r < ∞, γ = N − 1 and N is dimension. Here, the prime symbols are dropped for convenience. Eq. (2) recovers dynamics in 1D when γ = 0. Since the exact solution of Eq. (2) in 1D has been found [20], we focus on its solution in 2D, as well as in 3D. Eq. (2) does not support the traveling wave solution because the presence of the gradient term (γ/r)∂um/∂r [1]. It reduces to 1D problem at r → ∞, which has the planar traveling wave as solution. Nevertheless, the be- havior of this system at the distance that is not so large has been unclear. It has been mentioned that the effects of curvature can make the front propagation in reaction- diffusion system somewhat to be different from the pla- nar case [7, 22]. Previously, Eq. (2) in cylindrical co- ordinate has been analyzed by the perturbation method [19]. However, the solution is shown in the large distance that yields the usual traveling wave. More recent, the Lie symmetry method has been employed to solve Eq. (2) for m = 2 in cylindrical coordinate (γ = 1) [23]. Although the exact solution has been found, it is another class and does not reflect the density distribution of population. In this work, we adapt the technique similar to our previous studies [20, 24] to solve for the solution of Eq. (2). We have found the approximate radial symmetric solution for Eq. (2) in the intermediate regime that the distance is not so large. The solution reveals the curvature effect on the spreading of the population, both of density profile and front speed, in this regime explicitly. To verify the analytical solution, we have solved Eq. (2) by a numer- ical method. The numerical result seems to agree with this approximate solution. It confirms that our approx- imate solution is plausible to describe the intermediate behavior of system. Before finding the solution in general case, we study the asymptotic behavior of Eq. (2), as time goes to in- finity, first. Eq. (2) can be viewed as a one-dimensional nonlinear convection-reaction-diffusion equation with the varying drift coefficient γ/r, similar to [25]. The gradient term (γ/r)∂um/∂r in Eq. (2) is large in the vicinity of front position R(t), otherwise it becomes small [7]. Thus, we change the gradient term to (γ/R)∂um/∂r. The drift coefficient becomes small as R → ∞. In another hand, the nonlinear convection-reaction-diffusion equation with constant drift coefficient ν has been analyzed previously [24, 26, 27]. It has found that the solution in this case, at large time, converges to the sharp traveling front at speed c =q1 + (ν/2)2 − ν/2 [24, 26, 27]. For the small varying drift coefficient, we assume that the approximate front speed for Eq. (2) can be obtained by setting ν = γ/R, c =r1 +(cid:16) γ 2R(cid:17)2 − γ 2R . (3) The front speed (3) approaches to 1 as R → ∞, which is equal to the constant front speed of planar wave [1, 5, 6, 20, 21]. If the solution of Eq. (2) exists, it should result the front speed (3) as the asymptotic behavior. We now perform the analysis to find the general form of density profile that propagates at the front speed of Eq. (3). In approximation, we rewrite Eq. (2) in the form of ∂u ∂t =(cid:18) ∂ ∂r where ∂r + κ∗(cid:19)(cid:18) ∂ − κ(cid:19) um + u + κ(r) =r1 +(cid:16) γ 2r(cid:17)2 κ∗(r) =r1 +(cid:16) γ 2r(cid:17)2 γ 2r γ 2r + − , . ∂κ ∂r um, (4) (5) (6) We note that κ∗ − κ = γ/r and κ∗κ = 1. The correction term in Eq. (4), ∂κ/∂r =(cid:26)1 −h1 + (2r/γ)2i−1/2(cid:27) γ 2r2 , 2 approaches to O(1/r2) for r ≫ γ/2 and to 1/γ+O(r2) for r ≪ γ/2. Fortunately, this correction term well behaves because it decays from 1/γ to zero as r ≫ 0. In addition, at r ≪ γ/2, the correction term does not much affect the initial state while u ≪ 1. Next, we introduce the transformation dη = dr/κ, which can be evaluated to η(r) = κr + γ 2 ln (κr) . (7) With the transformation (7), Eq. (4), by dropping the correction term, becomes ∂u ∂t = κ−1(cid:18) ∂ ∂η + 1(cid:19) κ−1(cid:18) ∂ ∂η − κ2(cid:19) um + u. (8) For r ≫ γ/2, we approximate that κ ≈ 1 + O(1/r). Applying this approximation to Eq. (8), we obtain ∂u ∂t ≈ ∂2um ∂η2 + u − um. (9) Eq. (9) is equivalent to Eq. (2) in 1D, but evolving with η as the spatial coordinate. By adapting the result from Ref. [20], we obtain the solution to Eq. (9), u(r, t) = ×(1 −(cid:20) ρet [ρp (ept − 1) + 1] 1 p ep(η(r)−η0) ρp (ept − 1) + 1(cid:21) 1 p p+1) 1 , (10) where η0 = η(r0), r0 is initial front position and ρ is initial density amplitude. By setting γ = 0, Eq. (10) recovers the solution in 1D [20]. We note that u(r, t) vanishes after front position for r ≥ R(t), which will be determined later. As r → 0, we have η → −∞. This causes the density profile at the origin approaches to u(0, t) = ρet/ [ρp (ept − 1) + 1] p , which is actually the solution of 1 ∂u(0, t) ∂t = u(0, t) − um(0, t). (11) This implies no diffusion at the origin. At a sufficient large time that ept′ ≫ 1 and conse- quently ρpept′ ≫ 1, we estimate the transition point t′ ≈ − ln ρ. (12) For t ≫ t′, the solution (10) emerges a pattern form of the traveling wave-like p+1(cid:27) 1 eu(r, t) =(cid:26)1 −hρ−1e(η(r)−t−η0)i p p For r ≫ γ/2, we approximate that η(r) ≈ r + γ 2 ln r. It is seen that the logarithmic term does not vanish even at large distance, unless γ = 0. Therefore, Eq. (13) con- tains ineluctable curvature term, which can be called the curved traveling wave-like. . (13) The front position can be calculated from Eq. (13) by determining the first position R(t) that density falls to for front position zero oreu(R, t) = 0. After evaluating, we obtain equation η(R) − η0 = t − t′. (14) From Eq. (14), we see that the front position does not simply linearly depend on time as in 1D case [20]. By differentiating, respected to time, both sides of Eq. (14), we obtain ∂η(R) dR dt = 1. This allow us to calculate the ∂R front speed c = dR dt , that is c =(cid:18) ∂η(R) ∂R (cid:19)−1 =r1 +(cid:16) γ 2R(cid:17)2 − γ 2R . (15) The front speed (15) obtained from this analysis recovers Eq. (3) as expected. Once again, we have seen that the front speed is altered by the curvature as found in other similar systems [7, 22]. At sufficient large distance that R ≫ γ/2, the front speed can be approximated as a constant c ≈ 1. This is equal to the front speed in 1D case (γ = 0) [1, 5, 6, 20, 21]. If we define the following quantities: φ(r) = e(p+2)(η(r)−η0)/(p+1), τ (t) = ρp (ept − 1) + 1, and u(r, t) = ρete(η(r)−η0)/(p+1)w(r, t), we can rewrite Eq. (10) as the scaling function w(φ, τ ) = 1 τ β(cid:19) , τ β F(cid:18) φ (16) p(p+1) , F (ξ) = (cid:2)ξ−p/(p+2) − 1(cid:3)1/p where β = p+2 and ξ = φ/τ β. In the term of transformed density w(φ, τ ), as a function of transformed space φ and time τ , evolv- ing of the population density in the radial symmetric geometry still holds the self-similarity with the scaling law of Eq. (16). Moreover, this self-similar pattern con- verges to the traveling wave-like (13) as time becomes large. The connection between self-similar solution and traveling wave solution can be described as intermediate asymptotics of the system [28]. To compare with the analytical solution, we employ the standard explicit finite difference scheme [29] to solve the radial symmetric extended Fisher equation (2) numeri- cally. Eq. (11) is imposed as the boundary condition at origin (r = 0). The boundary condition at the edge of computational domain is free, as the front never reaches to this position. The initial density profile for the numer- ical calculation is set to the same value of the analytical one, u(r, 0). The initial front position is chosen such that r0 ≫ γ/2, since the analytical solution is expected to be accurate at large distance. The evolution of population density profiles, obtained from the analytical solution (10) and the numerical solu- tion, are demonstrated in Fig. (1). The density initiates from a sharp profile then grows locally to the saturated value, while it spreads out to unoccupied region. At the early state, while density is small, the correction term 3 (the last term in Eq. (4)) does not interfere the analytical density, as mentioned above. Our approximation is not accurate as the density grows to unity at early regime. However, it is seen that both of the analytical solution and the numerical solution seem to be in agreement as time and distance become large. The front position R(t) is also measured directly from the density profiles in Fig. (1). We notice that the small numerical deviation in density can make the front posi- tion to be shifted from the actual value. Therefore, the density that is less than 10−6 can be considered as zero in measuring the front position. The plot of analytical front position versus numerical front position is shown in Fig. (2). By calculating t for given measured R, the front position obtained from the density profiles satisfies Eq. (14). Once again, both of the analytical front posi- tion and the numerical front position are in agreement. Noting that, although it looks similar to, the data cannot be well fitted with the solution from 1D [20], for not so large distance. 1 0.8 0.6 0.4 0.2 ) t , r ( u t=0 0 0 time t=102 10 20 30 40 50 r 60 70 80 90 100 110 FIG. 1. (Color online) Demonstration of evolution of the ra- dially symmetric population density profile u(r, t) (10) in 2D (γ = 1) by comparing with the numerical solution. The solid lines represent the exact solutions and the circle markers rep- resent the numerical solutions. The parameters are as follows: p = 2, ρ = 0.05 and r0 = 5. The density profiles are initiated at t = 0 and evolve until t = 102. ) t ( R 100 80 60 40 20 0 0 10 20 30 40 50 t 60 70 80 90 100 110 FIG. 2. (Color online) The corresponding front position R(t) extracted from the density profiles in Fig. (1). The solid lines represent the exact solutions and the circle markers represent numerical solutions. In summary, we study the population dynamics that is described by the density-dependent reaction-diffusion equation, so called the extended Fisher model. We have found the approximate solution in the radial symmetric form in two- and three-dimensional space to this equa- tion. The analytical result shows that the evolution of population density is self-similar. At large time scale, the population density propagates as the curved traveling wave-like. The analytical solutions seem to be in agree- ment with the numerical solutions. Finally, it is revealed that the density profile and the propagating speed of the evolving population are influenced by the ineluctable cur- vature at distance that is insufficient large. 4 [1] J. Murray, Mathematical Biology I: An Introduction and N. Shigesada, J. Theor. Biol. 188, 177 (1997). (Springer-Verlag, New York Berlin Heidelberg, 2002). [14] E. Ben-Jacob, I. Cohen, and H. Levine, Adv. Phys. 49, [2] R. Fisher, Ann. Eugenics 7, 355 (1937). [3] D. Aronson and H. Weinberger, in Partial Differential Equations and Related Topics, Vol. 446, edited by J. A. Goldstein (Springer, Berlin Heidelberg, 1975) pp. 5 -- 49. [4] D. G. Aronson and H. F. Weinberger, Adv. Math. 30, 33 (1978). [5] W. Newman, J. Theor. Biol. 85, 325 (1980). [6] W. Newman, J. Theor. Biol. 104, 473 (1983). [7] V. Volpert and S. Petrovskii, Phys. Life Rev. 6, 267 (2009). [8] D. Aronson, in Nonlinear Diffusion Problems, Vol. 1224, edited by A. Fasano and M. Primicerio (Springer, Berlin Heidelberg, 1986) pp. 1 -- 46. 395 (2000). [15] J. Skellam, Biometrika 38, 196 (1951). [16] W. Gurney and R. Nisbet, J. Theor. Biol. 52, 441 (1975). [17] W. Gurney and R. Nisbet, J. Theor. Biol. 56, 249 (1976). [18] M. Gurtin and R. MacCamy, Math. Biosci. 33, 35 (1977). [19] T. P. Witelski, Appl. Math. Lett. 8, 57 (1995). [20] W. Ngamsaad and K. Khompurngson, Phys. Rev. E 85, 066120 (2012). [21] P. Rosenau, Phys. Rev. Lett. 88, 194501 (2002). [22] T. P. Witelski, K. Ono, and T. J. Kaper, Nat. Resour. Model. 13, 339 (2000). [23] A. H. Bokhari, M. Mustafa, and F. Zaman, Nonlinear Anal. 69, 4803 (2008). [9] J. A. Sherratt and J. Murray, Proc. R. Soc. Lond. B 241, [24] W. Ngamsaad and K. Khompurngson, 29 (1990). [10] P. K. Maini, D. S. McElwain, and D. I. Leavesley, Tissue Eng. 10, 475 (2004). Phys. Rev. E 86, 062901 (2012). [25] K. Uchiyama, Arch. Rational Mech. Anal. 90, 291 (1985). [26] B. Gilding and R. Kersner, J. Phys. A: Math. Gen. 38, [11] B. G. Sengers, C. P. Please, and R. O. Oreffo, J. R. Soc. 3367 (2005). Interface 4, 1107 (2007). [12] M. J. Simpson, K. K. Treloar, B. J. Binder, P. Haridas, K. J. Manton, D. I. Leavesley, D. S. McElwain, and R. E. Baker, J. R. Soc. Interface 10 (2013). [13] K. Kawasaki, A. Mochizuki, M. Matsushita, T. Umeda, [27] M. Mansour, Rep. Math. Phys. 66, 375 (2010). [28] G. Barenblatt and Y. Zel'dovich, Annu. Rev. Fluid Mech. 4, 285 (1972). [29] W. H. Press, B. P. Flannery, S. A. Teukolsky, and W. T. Vetterling, Numerical Recipes in C: The Art of Scientific Computing (1988).
1802.05075
1
1802
2018-02-14T12:49:29
Nonlinear elasticity of the extracellular matrix fibers facilitates efficient inter-cellular mechanical communication
[ "physics.bio-ph" ]
Biological cells embedded in fibrous matrices have been observed to form inter-cellular bands of dense and aligned fibers, through which they mechanically interact over long distances. Such matrix-mediated cellular interactions have been shown to regulate a variety of biological processes. The current study was aimed at exploring the effects of elastic nonlinearity of the fibers contained in the extracellular matrix (ECM) on the transmission of mechanical loads between contracting cells. Based on our biological experiments, we developed a finite-element model of two contracting cells embedded within a fibrous network. The individual fibers were modeled as showing either linear elasticity, compression-microbuckling, tension-stiffening or both of the latter. Compression-buckling resulted in smaller loads occurring in the ECM, but these were more directed toward the neighboring cell. The latter decreased with increasing cell-to-cell distance; when cells were >15 cell-diameters apart, no such inter-cellular interaction was observed. Tension-stiffening further contributed to directing the loads toward the neighboring cell, though to a smaller extent. The contraction of two neighboring cells resulted in mutual attraction forces, which were considerably increased by tension-stiffening, and decayed with increasing cell-to-cell distances. Nonlinear elasticity contributed also to the onset of force polarity on the cell boundary. The density and alignment of the fibers within the inter-cellular band were considerably greater when fibers buckled under compression, with tension-stiffening further contributing to this structural remodeling. Our model demonstrates the contribution of nonlinear elasticity of biological gels to directionality and efficiency of mechanical-signal transfer between distant cells.
physics.bio-ph
physics
Nonlinear elasticity of the extracellular matrix fibers facilitates efficient inter-cellular mechanical communication Ran S Sopher, Hanan Tokash, Sari Natan, Mirit Sharabi, Ortal Shelah, Oren Tchaicheeyan, Ayelet Lesman* School of Mechanical Engineering, Faculty of Engineering, Tel-Aviv University, Israel * Corresponding author: Dr. Ayelet Lesman Wolfson Mechanical Engineering Building, Room 331 School of Mechanical Engineering, Faculty of Engineering Tel-Aviv University Tel Aviv 69978 Israel Tel: +972-3-6408233 Email: [email protected] Abstract Biological cells embedded in fibrous matrices have been observed to form inter-cellular bands of dense and aligned fibers, through which they mechanically interact over long distances. Such matrix-mediated cellular interactions have been shown to regulate a variety of biological processes. The current study was aimed at exploring the effects of elastic nonlinearity of the fibers contained in the extracellular matrix (ECM) on the transmission of mechanical loads between contracting cells via the ECM. Based on our biological experiments with fibroblasts in fibrin gels shortly after seeding, we developed a finite- element model of two contracting cells embedded within a fibrous network. The individual fibers were modeled as showing either linear elasticity, compression-microbuckling, tension- stiffening or both of the latter. Fiber compression-buckling resulted in smaller loads occurring in the ECM, but these were more directed toward the neighboring cell. The latter decreased with increasing cell-to-cell distance; when cells were >15 cell-diameters apart, no such inter-cellular interaction was observed. Tension-stiffening further contributed to directing the loads toward the neighboring cell, though to a smaller extent. The contraction of two neighboring cells resulted in mutual attraction forces, which were considerably increased by tension-stiffening, and decayed with increasing cell-to-cell distances. Nonlinear elasticity contributed also to the onset of force polarity on the cell boundary, manifested in larger contractile forces on the part of the cell boundary pointing toward the neighboring cell. The density and alignment of the fibers within the inter-cellular band were considerably greater when fibers buckled under compression, with tension-stiffening further contributing to this structural remodeling. While previous studies have established the role of the ECM nonlinear mechanical behavior in increasing the range of inter-cellular force transmission, our model demonstrates the contribution of nonlinear elasticity of biological gels to directionality and efficiency of mechanical-signal transfer between distant cells. 1. Introduction The cellular actomyosin machinery actively generates forces that are transmitted to the cell surroundings to induce loads (displacements, strains and stresses) within the extracellular matrix (ECM); these can persist hundreds of microns away (1), and are known to influence cell morphology, migration and differentiation (2). Long-range loads have been thus proposed as a means for cells to mechanically communicate with each other, and demonstrated to play a key role in various biological, physiological and pathological processes, as diverse as capillary sprouting (3), cancel invasion (4), heart-beat synchronization (5) and morphogenesis (6). The fibrous ECM demonstrates nonlinear elastic behavior that is manifested in compressive-softening and tension-strain-stiffening (7, 8). These are attributed to the mechanical behavior of the individual fibers contained in the matrix, showing strain stiffening in tension (8–12) and microbuckling under compression (9, 13, 14). The fibrous structure of the ECM also contributes to its macroscale elastic nonlinearity as a result of fiber reorganization under applied loading, manifested in fiber realignment and densification (15– 17). Previous experimental studies have demonstrated that the nonlinear elasticity of the matrix act as a mechanism facilitating long-range transmission of loads, enabling cells to sense and respond to mechanical signals sent by other cells located at far distances. For example, Notbohm et al. (18) found that the contraction of fibroblasts within a fibrin matrix of nonlinear elastic behavior induced displacements that travelled considerably further than predicted when assuming linear elasticity. Vanni et al. (19) similarly found that the contraction of a single fibroblast within collagen gel induces strains that can propagate up to 800 μm through the substrate. Winer et al. (20) found that local strain-stiffening of fibrin gels facilitates the transmission of forces between fibroblasts or mesenchymal stem cells up to ~500-μm (~30-cell-diameters) apart. This is in a striking contrast to cells cultured on linear elastic gels, which sense and respond to loads induced by other cells in a distance limited to ~25 μm (21). These experimental findings have been elucidated by numerous analytical procedures, finite-element (FE) simulations and other computational models developed to study the effects of ECM elastic nonlinearity on the transmission of cell-contraction-induced loads through the network in which the cell(s) are embedded. Safran and colleagues (22–24) presented an analytical model demonstrating the long-range decay of displacements induced by the contraction of a single circular cell embedded in a medium showing nonlinear-elastic behavior (manifested in tension-stiffening and compression-softening). These predications were later supported by computer simulations in fibrous networks demonstrating that fiber buckling results in displacements and stresses travelling considerably farther than they would in a linear-elastic medium (25, 26). Also important, the fibrous structure of the ECM, and particularly the load-induced geometric rearrangement of the network, have been found to contribute to the nonlinear elastic behavior (especially stiffening) of the ECM and to act as a mechanism considerably increasing the range of inter-cellular force transmission and sensing (1, 15–17, 27–30). These studies distinguish the ECM nonlinear mechanical behavior, and particularly fiber microbuckling, as a mechanism supporting long-range force transmission through the ECM. Less is known about the effect of the mechanical behavior of the fibers constituting the ECM on force transmission between cells. Previous experimental studies (3, 4, 27, 31, 32) have revealed that contraction of multiple cells embedded in fibrous biological gels induces structural remodeling of the fibers in the inter-cellular medium manifested in the formation of aligned and densely packed fiber 'bands' connecting the contracting cells. This observation demonstrates not only the long- range nature of the force transmission through the ECM – which is both facilitated by and contributing to such structural remodeling – but also reflects the tendency of cell-contraction- induced loads to be delivered in a highly directional manner toward neighboring cells. Harris et al. (31) was the first to show that the forces exerted by tissue explants placed millimeters apart in collagen gel stretched and aligned the fibers between the explants, thereby forming inter-cellular 'bands'. More recent studies have demonstrated the ability of cells to respond to such inter-cellular bands. Korff and Augustin (3) found that forces applied by two endothelial-cell spheroids embedded in collagen gel and placed up to 700 m apart, resulted in directional sprouting of capillaries along aligned fibrils between the spheroids. Shi et al. (4) further showed that the directional remodeling of collagen fibers accelerated the transition of gel-cultured mammary-acini cells to an invasive phenotype. Such biological observations highlight the importance of the structural remodeling of the inter-cellular matrix in supporting long-range inter-cellular mechanical interactions, thereby regulating various biological processes. Still, the physical mechanisms facilitating the formation of such ECM 'bands' and the manner in which they mediate force transmission between cells are poorly understood. Particularly, despite the fact that several computational models were able to capture the tendency of loads to concentrate within the inter-cellular medium and align the fibers contained in this region (16, 27, 29, 30, 32), a quantitative exploration of the influence of ECM elastic nonlinearity on regulating matrix-mediated mechanical interaction between cells is warranted. In this work, we explore the contribution of the nonlinear elastic properties of the ECM fibers to the structural remodeling of the inter-cellular ECM and to the transfer of mechanical loads between neighboring cells. Experimentally, we show the ability of fibroblast cells to structurally align and densify the fibers between them even shortly after seeding in fibrin gels, when they are still primarily spherical. Based on our experimental setting, we developed FE simulations of two contracting cells (separated by 1.5-19 cell diameters) embedded within fibrous nonlinear elastic networks. The simulation outcomes indicate that cell-contraction- induced loads are highly directed toward neighboring cells owing to the nonlinear elastic behavior of the matrix and its constituting fibers; this observation is coupled with elevated structural remodeling of the inter-cellular region of the ECM. We link these observations with efficient transfer of mechanical loads between cells. We also show that inter-cellular interactions manifest in attraction forces occurring between neighboring cells, and also lead to the onset of force polarity on the cell boundaries. The model presented herein contributes to the understanding of biological processes involving ECM-mediated interactions that can influence cell differentiation, migration, and morphogenesis; it can also expand the knowledge basis required for designing biomaterials that support efficient inter-cellular mechanical interactions. 2. Methods 2.1. Biological experiments Approximately 5,000 NIH3T3 - GFP-Actin cells were seeded in 20 μl of fibrin gel (5 mg/ml fibrinogen) labelled with Alexa Fluor 456 as described in previous studies (e.g. (18)). The gel was scanned at several time points post seeding using confocal laser-scanning microscope (Zeiss LSM 880 lens x40, water immersion) to capture distant cells forming 'bands' between them, as visual evidence of mechanical interaction (18). Image analysis was followed using ImageJ (NIH, Bethesda, MD, USA; https://imagej.nih.gov/ij/) and the OrientationJ plugin (EPFL, Switzerland, 2017; http://bigwww.epfl.ch/demo/orientation/#soft) to examine the orientation and intensity (indicative of the fibrin density, and normalized to the mean intensity of the entire image) of several regions contained in the ECM between two interacting cells or far away from the cells. 2.2. Computational modeling In order to explore the effects of the mechanical behavior of the ECM on the mechanical loads occurring within the medium between contracting cells, a FE model was developed. This simulated two identical contracting cells embedded in a fibrous network, while the cell- to-cell distance and the mechanical properties of the fibers contained in the network were adjusted to create 558 model variants (31 cell-to-cell distances, 6 types of fiber mechanical properties, 3 levels of cell contraction). The model was greatly based on computational models described in previous studies (mainly Notbohm et al. (26) and Liang et al. (33)) with some major modifications, which are described below. 2.2.1. Network geometry and architecture A two-dimensional (2D) array of multiple identical square box-X units was initially created. Each unit contained four corners acting as nodes and two horizontal, two vertical and two diagonal sides acting as elements. Each of the four horizontal/vertical sides was shared with another X-box unit; similarly, each of the four corners was shared between four identical units meeting at this corner, resulting in each corner/node being the edge of eight sides/elements (referred to by Notbohm et al. (26) and others as connectivity=8; Figure 1a). The locations of all nodal positions were modified by relocating each node to a randomly selected location contained within a circular region of radius equal to the length of the horizontal/vertical sides of the aforementioned square box-X. This created an array of elements of different lengths, where each linear element connecting a pair of nodes represented a fiber segment stretched between two cross-linking points, thereby simulating the fibrous network contained in the ECM. Elements were removed from the network to create two circular void regions (15, 16, 25, 26, 33) representing cells embedded within the ECM; the centers of both circles were coincident with a horizontal line passing through the center of the network (Figure 1b). Equivalent model variants with a single cell each (the center of which was coincident with the center of the network) were also generated. Cell diameter and mean fiber-segment length were assumed to be 15.2 μm and 3.5 μm, respectively, according to our biological experiments with fibroblast cells embedded in 5 mg/ml fibrin gel (Figure 2). Cell-to-cell distance ranged between 1.5 to 19 cell diameters (i.e., approximately 23-289 μm, based on our laboratory experiments; full list of cell-to-cell distances: 1.5, 1.8, 2.1, 2.9, 3.4, 4.0, 4.6, 5.3, 5.9, 6.5, 7.1, 7.8, 8.4, 9.0, 9.6, 10.3, 10.9, 11.5, 12.1, 12.8, 13.4, 14.0, 14.6, 15.2, 15.9, 16.5, 17.1, 17.8, 18.4, 19.0). The diameter of the network was 100 times larger than that of the cell (i.e., 1520 μm), so that the distance between each cell and the boundaries of the network was at least twice larger than the cell-to- cell distance. (a) (b) ] a P k [ s s e r t S 5 4 3 2 1 Compression 0 Tension -10% -5% 0% 5% Strain 10% Linear Strain stiffening Buckling Buckling + strain stiffening -1 -2 (c) Figure 1: Details of the finite-element (FE) model: (a) The elementary unit of the model (before randomization). Nine such units are shown. (b) A small portion of a two-dimensional FE model of two cells contracting within the fibrous network. The contracting cells are modeled as void regions (blue circles) to which a boundary condition of radial contractile displacement (red arrows) is applied. (c) Stress-strain curves representing the mechanical properties assigned to each of the fiber elements consisting the FE model. Four material models were used to simulate the mechanical behavior of the ECM fibers, as listed in Section 2.2.2. Lines are slightly shifted for visualization purposes. 2.2.2. Mechanical properties Four material models were used to simulate the mechanical behavior of the individual ECM fibers (Equation 1, Figure 1c; (26, 33)): i) 'linear': linear elastic material with tensile and compressive elastic moduli (Young's moduli, E) of 11.5 kPa (ρ=1); ii) 'buckling': elastic material with tensile E of 11.5 kPa and compressive E 2- (buckling 1/2, ρ=0.5), 5- (buckling 1/5, ρ=0.2) or 10- (buckling 1/10, ρ=0.1) times smaller, which simulates fiber buckling; iii) 'strain stiffening': hyper-elastic material with compressive E of 11.5 kPa and tensile E of 11.5 kPa within the logarithmic-strain (i.e. true strain) range of 0-2% (ρ=1); for tensile strains larger than 2%, stress increases exponentially to simulate strain-stiffening behavior; iv) 'buckling+stiffening': hyper-elastic material with tensile E of 11.5 kPa within the logarithmic-strain range of 0-2%, which increases exponentially for strains larger than 2%; compressive E is 10 times smaller than 11.5 kPa (ρ=0.1). Further information about the assignment of mechanical properties to the ECM fibers is included in the Supplementary Materials. 𝜌 ∙ 𝐸𝑟𝑒𝑓, 𝜆 < 0 𝐸 = { 𝐸𝑟𝑒𝑓, 0 ≤ 𝜆 < 𝜆𝑠 𝐸𝑟𝑒𝑓 ∙ 𝑒 𝜆−𝜆𝑠 𝜆0 , 𝜆 ≥ 𝜆𝑠 Equation 1: Values of elastic modulus (Young's modulus, E) assigned to the truss elements of the FE model. Eref=11.5 kPa is the reference elastic modulus, λ=Δl/l is the engineering strain occurring within an element of length l, λs=2% is the strain above which strain stiffening occurs, λ0=0.05 is the strain stiffening coefficient and ρ is the buckling ratio (9, 33). 2.2.3. Boundary conditions Cell contraction was modeled by applying a boundary condition of radial contractile displacement (equals to 10%, 25% and 50% of the cell radius) to all nodes constituting the cell boundaries (Figure 1b) (in line with (26–30, 33)). The circular boundary of the entire network was fixed for translations and rotations in all directions (15, 29, 30, 33). In all simulations, the possibility of the boundary fixation affecting the model outcome measures was eliminated by ensuring that the strains, stresses and strain energy densities (SEDs) occurring along the network boundaries were negligible compared with those occurring in the cell vicinity (15). 2.2.4. Numerical method Linear truss elements (i.e. supporting uniaxial tension and compression and unrestrained rotation about the nodes, which represent fiber cross-links, but with infinite resistance to bending) were used to model the fiber segments, following our preliminary study showing marginal differences in the model outcomes between truss- and beam- element networks (in line with (15, 16, 33)). The cross-sectional area of all elements was assumed to be constant at 0.031416 μm2 (see Supplementary Materials). The model included approximately 590,000 elements and 148,000 nodes. The ABAQUS® Standard/Implicit FE solver (ABAQUS, Version 2017, Dassault Syst`emes Simulia Corp., Providence, RI, 2017) in its nonlinear analysis mode was used to process all model variants, with running time of approximately 5 minutes per simulation (on an i7-, 3.60GHz- CPU and 32 GB-RAM station). 2.2.5. Outcome measures Values of displacement, strain, stress (logarithmic/true tensile and compressive strains and stresses), SED and reaction force were calculated at the nodes and/or centroids of all elements. A script coded in MATLAB® (version R2016b, MathWorks Inc., Natick, MA, USA) was used to determine the following outcome measures for all model variants: (a) Total contraction force: the sum of all reaction forces occurring in the nodes constituting the cell perimeter and pulling them along the radial axis toward the cell center (arrows at the top panel of Figure 7a). (b) Net cell-interaction force: the sum of projections of the radial contraction forces on the line connecting the centers of the two cells. This outcome measure is indicative of the level of inter-cellular attraction (negative) or repulsion (positive). (c) Contraction-force front-to-rear polarity ratio: the fraction of the total contraction force occurring within a 60° arc of the cell boundary pointing toward the neighboring cell (orange arc at the top panel of Figure 7b) to that occurring in the arc pointing toward the exactly opposite direction (sky-blue arc). This outcome measure is used to evaluate the relative amount of cell-contraction force directed into sending signals to the neighboring cell; it can also be used as an indication of the direction (toward/away from the neighboring cell) where the cell is likely to spread or migrate (force polarity is known to affect cell motility and morphology; (34)). (d) The mean loads (strains, stresses and SED) occurring within a disc surrounding an individual cell, of radius equals to half of the cell-to-cell distance (orange disc at the top panel of Figure S6a); (e) Directionality ratio: the fraction of the sum of loads occurring in all elements falling within a 60° sector pointing toward the neighboring cell (orange sector at the top panel of Figure 5a) to the sum of loads occurring in all elements falling within the entire aforementioned disc. (f) Asymmetry ratio: the fraction of the sum of loads occurring in all elements falling within the aforementioned 60° sector to the sum of loads occurring in all elements falling within a 60° sector pointing toward the opposite direction (sky-blue sector at the top panel of Figure 5b; similar to the 'signal parameter' described elsewhere (15)). The two latter outcome measures are used to evaluate the relative amount of the load (strain, stress or energy) caused by cell contraction directed toward the neighboring cell (in very simplistic words, how much of the effort that the cell is putting while contracting is potentially delivered to the neighboring cell, or how efficient the cell contraction is in terms of inter-cellular mechanical signal transfer); directionality and polarity/asymmetry ratios of 0.17 and 1.0, respectively, indicate no preferred orientation of loads toward the neighboring cell. (g) The relative change (in %) in density of elements contained in the 60° sector pointing toward the neighboring cell (orange sector at the top panel of Figure 8a) as a result of cell contraction; (h) Change in the mean angle of the orientation of the fibers contained in aforementioned 60° sector as a result of cell contraction. The latter two outcome measures were calculated in view of previous experimental studies showing that contraction of neighboring cells causes densification and realignment of the ECM fibers stretched between the cells; cells may respond to such structural changes, for example by migrating along the aligned fibers (34). All aforementioned outcomes were derived for both cells contained in each model variant; the mean of the outcomes of the two cells was then calculated, and is regarded as the model outcome (as presented below). 3. Results 3.1. Biological experiments of inter-cellular mechanical interactions Actin-GFP fibroblast cells were embedded in fluorescently labelled fibrin gels at a low cellular density such that cells were well separated from each other. Within two hours from cell seeding, while most of the cells were still of a spherical shape, cells were observed to deform the fibrous matrix, creating highly remodeled matrix 'bands' between neighboring cell pairs (Figure 2a). Image analysis of the fibrous structure of the ECM revealed elevated fiber density and alignment within the inter-cellular medium compared with other regions of the ECM (Figure 2b,c). This observation demonstrates the ability of cells, still in their spherical shape, to deform the surrounding matrix in a highly directional manner toward neighboring cells. We used these biological experiments as the basis of a computational study aimed at examining the contribution of the nonlinear elastic behavior of the fibrous matrix to mechanical interaction between cells. Our model was designed to test the experimentally- observed directionality of deformations occurring in the ECM, when isolated from other factors potentially creating load anisotropy (e.g. elongated cell geometry or non-uniform contraction). We therefore developed a FE model of two uniformly contracting circular cells situated at increasing distances from each other within a structural fibrous mesh, and explored the effects of fiber elastic nonlinearity on the distribution and propagation of loads (forces, displacements, strains, stresses and SEDs) through the matrix, and particularly in the region between neighboring cells. (a) y t i s n e t n i n a e m f o % 400 350 300 250 200 150 100 50 0 Inter-cellular medium Away from the inter-cellular medium Inter-cellular medium Away from the inter- cellular medium (c) ) s e e r g e d ( n o i t a t n e i r o r e b i F 70 60 50 40 30 20 10 0 (b) Figure 2: (a) Two NIH3T3 - GFP-Actin cells (green) embedded in fibrin gel labelled with Alexa Fluor 456 (grey). The images show the overlay of cells and fibrin matrix (left) and only-cells image (right). (b) Orientation of the fibrin fibers contained in the ECM in the inter-cellular matrix and away from the cells; 0° is defined as the line connecting the centers of the two cells. (c) Mean intensity as measured both within the inter-cellular medium and away from the cells. Data in (b) and (c) are derived from six pairs of cells as demonstrated in (a). 3.2. Finite-element analysis of the transmission of loads induced by cell contraction through a fibrous matrix Our FE simulations of a single cell contracting within a fibrous matrix show that the distribution of loads within the matrix was not homogeneous. Particularly, the majority of loads were carried through a small number of fiber segments constituting 'force chains' (mostly apparent in the buckling model variants; Figure 3 and Figure S3). Despite this non- homogeneity, the load distribution around a single cell was nearly isotropic about the cell center and did not demonstrate any preferred orientation. Tensile loads were most elevated in the fibers aligned approximately perpendicularly to the cell edges (thereby forming 'tethers' or 'force chains' propagating away from the contracting cell; Figure 3a,b and Figure S3a-d). The compressive strains, which were generally of magnitude approximately twice-larger than the tensile strains, were the largest within the fibers aligned approximately tangentially to the cell perimeter (thus forming 'rings' around the cell; Figure 3c,d and Figure S3e-h). The model variants simulating the contraction of a pair of neighboring cells similarly demonstrated propagation of loads through distinct fibrillar paths, with tensile-strain tethers forming between the cells and compressive-strain rings encircling the two cells (Figure 4a-d and Figure S4a-h). These observations are attributable to the contraction of the cell pulling mainly the fibers perpendicular to the cell edges and squeezing primarily the tangential ones. The tendency of contraction-induced loads to concentrate in such tethers and rings was greatly augmented by fiber compression-buckling, with fiber tension-stiffening marginally contributing to this effect (Figure S3 and Figure S4). Fiber compression-buckling expectedly resulted in the compressive strains occurring in the ECM as a result of cell contraction being larger, as opposed to the tensile loads and SEDs being considerably smaller; tension- stiffening again only marginally affected the magnitude of loads (Figure S3 and Figure S4). It can be therefore concluded that the contraction of isolated single cells induces substrate deformations that are symmetrically distributed about the cell center, yet propagate through distinct fibrillar arrangements, with magnitudes that are predominantly dictated by fiber microbuckling; the propagation of loads through tethers or rings and their dependency on fiber buckling is also manifested in the contraction of two neighboring cells. Tensile strain Compressive strain Strain energy density (a) (c) (e) r a e n i L g n i n e f f i t s + g n i l k c u B (b) 0 [%] 2.3 4.5 (d) 0 [%] 5.3 10.5 (f) 0 [Pa] 3 6 Figure 3: Contour plots showing the tensile (logarithmic) strains (left column), compressive (logarithmic) strains (middle column) and strain energy densities (SEDs, right column) occurring in the fiber segments within the vicinity of a single, isolated contracting cell, for 25% contraction. Plots were produced for two of the material models used to simulate the mechanical behavior of the individual fibers (linear, buckling + strain stiffening; Figure 1c). Equivalent plots produced for all four material models used to simulate the mechanical behavior of the individual fibers (Figure 1c) are contained as supplementary materials. Tensile strain Compressive strain Strain energy density (a) (c) (e) r a e n i L g n i n e f f i t s + g n i l k c u B (b) 0 [%] (d) (f) 2.3 4.5 5.3 10.5 0 [%] 3 6 0 [Pa] Figure 4: Contour plots showing the tensile strains (left column), compressive strains (middle column) and SEDs (right column) occurring in the fiber segments within the vicinity of two neighboring contracting cells (here the cell-to-cell distance is equivalent to 3.4 cell diameters as an example), and particularly within the inter- cellular medium, for 25% contraction. Plots were produced for two of the material models used to simulate the mechanical behavior of the individual fibers (linear, buckling + strain stiffening; Figure 1c). Equivalent plots produced for all four material models used to simulate the mechanical behavior of the individual fibers (Figure 1c) are contained as supplementary materials. 3.3. The effect of fiber compression-buckling on load transfer between cells We next analyzed the distribution of loads occurring in the vicinity of two neighboring contracting cells in more quantitative terms. The contour plots in Figure 4 and Figure S4 show that fiber microbuckling resulted in the tensile strains and SEDs occurring within the cell vicinity being generally smaller, yet more concentrated within the inter-cellular medium; fiber linear elasticity produced more even distribution of loads around each cell. The mean tensile strain and SED occurring in a disk surrounding each of the cells revealed considerably lower values for the compression-buckling model variants compared with those where fibers resisted compression (Figure S6a and Figure S6c). This indicates that it is 'easier' for cells to contract within a bucklable matrix, which is ascribable to its lower resistance. To quantify the tendency of loads to concentrate in the inter-cellular medium, which arises from the contraction of two cells practically stretching this region of the ECM, we calculated the fraction of deformations and SEDs falling within the matrix area pointing toward the neighboring cell (outcome measure e in Section 2.2.5, referred to as 'directionality ratio'). We further evaluated these loads against those falling within the ECM area pointing toward the opposite direction (outcome measure f in Section 2.2.5, referred to as 'asymmetry ratio'). We found that fiber buckling under compression resulted in the distribution of SEDs, tensile strains and compressive strains being more directed toward the neighboring cell (i.e., concentrating in the inter-cellular region), as reflected in higher directionality and asymmetry ratios (for 10%, 25% and 50% cell contraction; Figure 5a,b, Figure S7, Figure S8 and Figure S9). The most distinguished effect was found for SEDs at the shortest cell-to-cell distances: at cell contraction of 25% and cell-to-cell distance of 2.1 cell diameters, for example, fiber buckling resulted in approximately 0.40 of the total SEDs being directed toward the neighboring cell, compared with 0.24 for linear-elastic fibers and 0.17 for a single, isolated contracting cell (Figure 5a). 'Stronger' buckling ratios (i.e. smaller ρ in Equation 1) normally resulted in a more considerable fraction of loads falling within the inter-cellular medium (small panel at the top right of Figure 5a), which further emphasizes the critical role of fiber buckling in directing the ECM cell-contraction-induced loads toward the neighboring cell. Asymmetry ratios, which reflect the asymmetric distribution of the loads about the cell center, were generally more affected by fiber buckling than the directionality ratios. In particular, in the compression-buckling model variant, the total SED falling within the fraction of matrix pointing toward the neighboring cell was approximately 2.8-fold larger than on the opposite side (i.e. asymmetry ratio of 2.8), as compared with only 1.4 in the equivalent linear-elastic model variant, and 1.0 for a single, isolated cell (example values are for cell contraction of 25% and cell-to-cell distance of 2.1 cell diameters, Figure 5c). The effect of microbuckling was noticeable up to a distance of 9 cell diameters, above which the directionality and asymmetry ratios approached the values for a single, isolated cell (0.17 and 1.0, respectively, which indicate no preferred orientation of loads toward the neighboring cell), implying that mechanical inter-cellular signaling no longer occurred (Figure 5a,b, Figure S7, Figure S8 and Figure S9). In general, loads were smaller and all the above- mentioned trends were weaker for smaller cell contraction (25% versus 10% cell contraction; Figure S7 and Figure S8). Overall, the results indicate that fiber microbuckling results in lower-magnitude loads that are more efficiently directed to the neighboring cell. o i t a r y t i l a n o i t c e r i d D E S o i t a r y r t e m m y s a D E S 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 15 10 5 0 cell 1 cell 2 0.5 0.4 0.3 0.2 0.1 0.0 2.1 3.4 9.0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) cell 1 cell 2 3.0 2.0 1.0 0.0 2.1 3.4 9.0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (b) Figure 5: Directionality (a) and asymmetry (b) ratios of SED. The model variants shown include several cell-to- cell distances (in terms of cell diameter, D), and four of the material models used to simulate the mechanical behavior of the individual fibers (Figure 1c), for 25% cell contraction. Directionality and asymmetry ratios of 0.17 and 1.0, respectively, indicate no preferential orientation of loads toward the neighboring cell. The values calculated for the single-cell model variants (assuming a disc of radius equals to half of the cell-to-cell distance of 2 D) are shown for comparison. On the top right, an equivalent bar chart with fewer cell-to-cell distances and three material models used to simulate the mechanical behavior of the individual fibers, representing different buckling ratios (ρ=1/2, ρ=1/5 and ρ=1/10). 3.4. Effect of fiber tension-strain-stiffening on load transfer between cells Fiber strain-stiffening in tension further contributed to concentration of loads in the inter- cellular medium as reflected by higher directionality and asymmetry ratios (Figure 5b,c, Figure S7, Figure S8 and Figure S9). The effect of stiffening was most evident in smaller cell-to-cell distances and larger cell contractions (50% as opposed to 25% and 10%), which is attributable to the tensile strains exceeding the critical stiffening threshold of 2% tensile strain (Equation 1). For cell contraction of 25% and cell-to-cell distance of 1.8 cell diameters, fiber stiffening resulted in directionality ratio of 0.62 and asymmetry ratio of 14, compared with 0.27 and 1.6 for the equivalent linear-elastic model variant, and 0.17 and 1.0, respectively for a single, isolated cell (Figure 5b,c). At that short distance, the matrix between cells experiences high level of tensile strain, thereby resulting in higher SEDs (cells need to stretch more to achieve a certain level of contraction; discussed in Section 4). When fibers were modeled as both buckling under compression and stiffening in tension, the directionality and asymmetry ratios were substantially elevated, reaching values of up to 0.52 and 4.3, respectively, at cell contraction of 25% and cell-to-cell distance of 2.1 cell diameters (Figure 5a,b). For 50% cell contraction, fiber nonlinear elasticity contributed to directionality ratio of up to 0.80 at the cell-to-cell distance of 2.1 cell diameters (Figure S9). For cell contraction of 10%, the differences between equivalent strain-stiffening and non-strain-stiffening model variants were marginal (Figure S7), which is ascribable to the tensile strains rarely exceeding 2.5% (strain stiffening was assumed to occur above 2%; Section 2.2.2). The influence of fiber tension-stiffening was noticeable up to approximately 3-4 cell diameters, a relatively smaller distance than that found for the effect of compression-buckling (9 cell-diameters, Section 3.3). Overall, in larger cell contractions and smaller cell-to-cell distances, fiber strain stiffening in tension – alone or in combination with compression microbuckling – further contributes to concentration of loads within the inter-cellular region of the matrix, as demonstrated in the amplification of the directionality and asymmetry ratios of the load distributions. 3.5. The effect of fiber nonlinear elasticity on cell-induced forces Analysis of the force balance acting on the cell boundaries (Figure 6 and Figure S10) and its dependency on the cell-to-cell distance can serve as an indication of how the ECM mediates the propagation of forces between contractile cells. A key question is whether the contraction of two neighboring cells results in mutual attraction or repulsion. We thus calculated the projection of the net force pulling the cell boundary toward the cell center (contraction force) on the line connecting the centers of the two cells (outcome measure b in Section 2.2.5, referred to as 'net cell-interaction force') and plotted it against the cell-to-cell distance (Figure 7a, Figure S11a and Figure S12a). We found that the direction of the net force that the cell is applying to the ECM is consistently the direction opposite the neighboring cell. It can thus be inferred that in our simulations the cells always attract each other. The cell- interaction force was only slightly influenced by fiber compression-buckling, but considerably affected by tension-stiffening (Figure 7a, Figure S11a and Figure S12a). For example, at cell contraction of 25% and cell-to-cell distance of 2.1 cell diameters, fiber stiffening resulted in the interaction force exceeding 0.4 nN, compared with 0.07 nN (i.e. 5.7- fold) for the equivalent linear-elastic model variant (Figure 7a). This effect of strain stiffening was most evident for the larger levels of cell contraction; for 50% contraction, there was an almost 30-fold difference in the interaction force between the strain-stiffening and linear- elastic cases (Figure S12a). The cell-interaction force decreased with increasing cell-to-cell distance, and for all levels of cell contraction reached a plateau at distances larger than 6-7 cell diameters (Figure 7a, Figure S11a and Figure S12a). Also of interest is the force polarity that develops on the cell boundary. This can potentially trigger cell morphological changes and migratory preference along the direction of force polarization (34). Therefore, we also computed the front-to-rear force polarity ratio (with 'front' referring to the direction pointing toward the neighboring cell; outcome measure c in Section 2.2.5), as indicative of the direction toward which the cell(s) are likely to spread or migrate (Figure 7b, Figure S11b and Figure S12b). We found that force polarity was greatly elevated as the cells were closer together, particularly when the matrix fibers exhibited both buckling and stiffening (Figure 7b). At cell contraction of 25% and cell-to-cell distance of 2.1 cell diameters, for example, the force applied in the direction pointing toward the neighboring cell was approximately 1.2- (linear-elastic fibers), 2.2- (compression- buckling), 2.0- (tension-stiffening) and 4.0- (buckling and stiffening) fold larger than the force applied in the opposite direction (Figure 7b). The force polarity considerably increased with the level of cell contraction; at 50% contraction this reached as much as 6.2 (tension- stiffening) and 22.7 (buckling and stiffening) (cell-to-cell distance of 2.1 cell diameters, Figure S12b); for a smaller cell-to-cell distance of 1.8 cell diameters, polarity exceeded 100. This is attributable to the large tensile deformations occurring in the fibers contained in the inter-cellular region of the matrix (Section 3.4). These findings indicate that nonlinear elasticity contributes to attraction forces and the onset of force polarity between neighboring contracting cell(s); cells are thus more likely to adapt their shape or migrate toward their neighbours when embedded within a nonlinear elastic matrix compared with linear-elastic ones. r a e n i L g n i n e f f i t s + g n i l k c u B (a) (b) (c) (d) 2.7 0 ×10-2 [nN] 5.4 Figure 6: Contour plots showing the reaction forces occurring on the cell boundaries for a single (left column) or two (right column; here distance between the neighboring cells is equivalent to 3.4 cell diameters as an example) contracting cells, for 25% contraction. Plots were produced for two of the material models used to simulate the mechanical behavior of the individual fibers (linear, buckling + strain stiffening; Figure 1c). Equivalent plots produced for all four material models used to simulate the mechanical behavior of the individual fibers (Figure 1c) are contained as supplementary materials. ] N n [ e c r o f n o i t c a r e t n i - l l e C -1.0 -0.8 -0.6 -0.4 -0.2 0.0 10.0 l y t i r a o p e c r o F 8.0 6.0 4.0 2.0 0.0 cell 1 cell 2 -0.2 -0.1 -0.1 0.0 2.1 3.4 9.0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) cell 1 cell 2 2.5 2.0 1.5 1.0 0.5 0.0 2.1 3.4 9.0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (b) Figure 7: (a) Net cell-interaction force occurring on the cell boundary (as projected on the line connecting the cell centers), for 25% contraction. The interaction force is defined as positive for repulsion and negative for attraction. (b) The polarity ratio of the contraction force occurring on the cell boundary. Polarity ratio of 1.0 indicates no preferential orientation of loads toward the neighboring cell. Equivalent bar charts showing the net cell-interaction force and the contraction force polarity ratio for 10% and 50% contraction are contained as supplementary materials. 3.6. Effect of nonlinear elasticity on structural remodeling of the inter-cellular matrix Structural remodeling of the ECM, including fiber densification and alignment, can influence the biological activity of cells, including morphology and migration (34), and is therefore of particular interest. Accordingly, we examined how the above-described preferred directionality of loads and their tendency to concentrate in the inter-cellular matrix influence fiber density and alignment. We found that the inter-cellular region of the matrix contained more fibers as a result of contraction. The level of increase in fiber density was considerably greater when the fibers buckled under compression, with strain stiffening under tension further contributing to this effect. The level of increase was greater for higher levels of cell contraction (Figure 8a, Figure S13a and Figure S14a). The fibers contained in the inter- cellular medium were also more aligned along the line connecting the cell centers, which is the direction of maximum tensile loading imposed by such cell contractions. The alignment was more considerable when the fibers buckled under compression and stiffened under tension, with buckling again showing a more considerable effect. Higher levels of cell contraction resulted in a more considerable alignment (Figure 8b, Figure S13b and Figure S14b). Similarly to the model outcomes discussed above, these trends – along with the changes in fiber density and orientation themselves – were more noticeable when the two neighboring cells were closer together. Overall, nonlinear elasticity of the fibers forming the ECM contributes to the increase in the density and alignment of the fibers contained in the inter-cellular region of the matrix, facilitating the formation of the 'bands' visible in our experimental findings (Figure 2) as well as in other biological contexts (3, 4, 34). cell 1 cell 2 15% 10% 5% 0% 2.1 3.4 9.0 20% 15% 10% 5% 0% l ] g e d [ e g n a n o i t a t n e i r o Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) -6 -4 -2 0 2.1 3.4 9.0 -8 -7 -6 -5 -4 -3 -2 -1 0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (b) y t i s n e d r e b i f n i e g n a h c % r e b i f n a e m n i e g n a h c Figure 8: (a) Relative change in density of fiber segments contained in the inter-cellular medium, and (b) mean change in the angles of the orientation of the fibers contained in the inter-cellular medium (orange sector at the top right panel), as a result of 25% cell contraction. 0° is defined as the horizontal line pointing toward the center of the neighboring cell; accordingly, negative change indicates realignment of the fiber to point toward the center of neighboring cell (i.e. to be aligned more horizontally). 4. Discussion The computational model presented in this study suggests that nonlinear elastic behavior of the fibers contained in the ECM – manifested in compression-microbuckling and tension- stiffening – results in the loads occurring in the fibers in the vicinity of contracting cells being more concentrated within the inter-cellular regions of the matrix and more directed toward neighboring contracting cells. Efficiency can be regarded as the quality of a process to occur with a minimum amount of wasted efforts; in more quantitative terms, it is the ratio of useful output to the total input. In the context of the current model, efficiency can be quantified as the fraction of cell-induced loads applied in the direction pointing toward the neighboring cell (and thus potentially result in mechanical signaling) to the total load induced by the cell contraction. Accordingly, the directionality-ratio outcomes presented herein (Sections 3.3 and 3.4) indicate that nonlinear elastic behavior of the individual fibers constituting biological gels contributes to the efficiency of the transfer of mechanical signals between neighboring cells, which may trigger biological response. The model outcomes further show that cells exert less energy when contracting within a matrix consisting of bucklable fibers, which is attributable to the smaller resistance of the matrix to deformations (Section 3.3); this further highlights the contribution of fiber buckling to efficient mechanical communication between distant cells via the ECM. While previous models have revealed the important contribution of elastic nonlinearity of biological gels to increasing the range of propagation of cell- contraction-induced loads through the ECM and specifically between neighboring cells (22, 26, 27, 29), our model is the first to demonstrate and quantify the direct effects of the mechanical behavior of the individual fibers on the directionality of mechanical loads and the efficiency of load transfer between cells. The concentration of loads and the rearrangement of fibers within the inter-cellular medium (experimental: Section 3; computational: Section 3.6), which were observed in all model variants (though to different levels, depending on the modeled fiber elastic behavior, level of cell contraction and cell-to-cell distance), are ascribable to the simultaneous contraction of two cells that is practically applying tension to the inter-cellular band, while other regions of the network are not subjected to such elevated loading. The increased tension in this region of the ECM obliges the fibers to align, and also results in compression occurring perpendicularly – which is augmented by fiber compression-buckling – leading to fiber densification in the inter-cellular medium (the level of increase in fiber density within the inter-cellular region was found to be larger when fibers were bucklable; Section 3.6). This may induce 'compressive stiffening' of the inter-cellular ECM band (35, 36), particularly when the fibers are modeled as bucklable (36). The increased effective stiffness of this region of the matrix, which is subjected to elevated tension, can lead to the observed preferential orientation of loads toward the neighboring cell, which is more evident in fiber-buckling networks. The further increase in the directionality of loads (mainly SEDs) owing to tension- stiffening (Section 3.4) is attributable to the greater resistance of the strain-stiffened fibers (including these consisting the inter-cellular ECM band) to tensile deformation (i.e. stiffening). While previous studies (15, 26, 27, 29) have captured the tendency of cell-contraction- induced loads to concentrate in the inter-cellular regions of the matrix, our study provides a systematic quantitative evaluation of the transfer of mechanical signals between neighboring cells, and the dependence of which on the nonlinear elastic mechanical properties of the matrix fibers and the distance between the cells. For example, Notbohm et al. (26) and Abhilash et al. (29) showed that cell-contraction-induced tensile strains occur almost entirely in the band between the two cells, which was ascribed to fiber microbuckling. Ma et al. (27) further suggested that 24 hours post cell seeding in collagen gel, the preferential orientation of cell-contraction-induced stresses toward neighboring cells is facilitated by increased fiber alignment. In our simulations we show that preferential orientation of cell-contraction- induced loads toward neighboring cells can occur without a priori fiber alignment, but both the load orientation and fiber alignment occurred simultaneously with cell contraction. The increased fiber density and alignment within the inter-cellular band – which was demonstrated in our simulations to be most considerable when the individual fibers contained in the ECM were modeled as of nonlinear mechanical behavior (Section 3.6) – is corroborated by our experimental findings (Section 3), and is in agreement with the cell- contraction-induced fiber densification and alignment previously demonstrated both in vitro (1, 3, 26, 27, 32) and in silico (16, 28–30, 32). Previous biological experiments have further suggested that the formation of inter-cellular bands of dense and aligned fibers regulate cell function in various biological processes, including directing the growth of capillary sprouts originated in spheroids in collagen gels (3), controlling cancer invasion in mechanically interacting acini (4, 32) and directing fibroblast patterning (20). This diversity of biological contexts implies that long-range inter-cellular mechanical interactions via fibrous gels has a universal role in cell and tissue function. Our model highlights the important role played by the nonlinear elastic behavior of the ECM fibers in the formation of these inter-cellular bands. Our FE model allowed us to directly infer the interaction forces acting between neighboring contracting cells and examine their dependency on the cell-to-cell distance, as opposed to previous analytical models that analyzed the interaction based on energy considerations (37, 38). In our simulations, we found that neighboring contracting cells always attract each other up to a cell-cell distance of 6-7 cell diameters, with the attraction forces being larger when the cells are closer together. Previous experimental study found that endothelial cells cultured on linear synthetic gels apply both repulsive and attractive forces to each other (21). The different behaviour may be due to the fibrillar structure of the ECM, which is absence in synthetic gels. Our simulations predict inter-cellular mechanical signal transmission to occur up to a distance of 15 cell-diameters apart. Specifically, we observe structural changes within the inter-cellular matrix to occur up to a distance of 15.3 cell-diameters (alignment and densification), tensile and SED's directionality ratios predict preferential load transmission to occur up to 9 cell-diameters, and force interaction predicts sensing up to 6.5 cell-diameters. The influence of fiber tension-stiffening on inter-cellular loads is notable up to a cell-to-cell distance of only 4 cell diameter (Section 3.4), while the effect of compression-buckling spans an over-twice longer distance (9-15 cell diameters). This is supported by the recent analytical predications of Xu et al. (22), which imply that tension-stiffening dominates the propagation of matrix displacements in the near vicinity of the contracting cell, whereas compression- buckling dictates the distribution of displacements in the further regions of the matrix. The range of inter-cellular mechanical force transmission predicted by our simulations (approximately 12 cell diameters or 175 μm when assuming that the fibers are bucklable; Section 3.3) is larger than that of Humphries et al. (15) (5 cell diameters), but similar to that of Wang et al. (30); it is also somewhat larger and considerably smaller than predicted in the experimental studies by Ma et al. (27) (120 μm) and Winer et al. (20) (500 μm), respectively. Like in many previous models (15, 16, 22, 25, 33), cells were modeled here as circles. This simplified representation captures the morphology of cells in the first hours after seeding in biological gels (Section 3, Figure 2; (27)). This simplification is in line with the purpose of our model, designed to allow investigation of the mechanical signaling occurring between neighboring contracting cells embedded in biological gels shortly after seeding, prior to cell morphogenesis, and while being isolated from any modifications potentially occurring in the ECM at later stages. Similarly, we assumed uniformly distributed (i.e. isotropic) contractile displacements of the cell membrane. While this simplification is likely to considerably affect the simulation-predicted loads occurring in the very near vicinity of the contracting cell, it has less influence farther away from the cell (27). Furthermore, like in most previous models, prior to cell contraction the fiber segments were assumed to be unstressed, which is again in line with the aim of this simulation – capturing cell-ECM and inter-cellular mechanical interactions at 'time zero'. Overall, our computational model mimics the biological scenario of the early stages of cell-ECM and cell-cell interactions and predicts the ability of cells to initiate mechanical cues and communicate through mechanical pathways shortly after their seeding. These early mechanical cues can guide cell growth and expansion toward nearby cells, as previously described in fibroblasts spreading toward each other already at 5-7 hours post seeding (26). Our model has notable limitations that should be considered when interpreting the results. First, similarly to most previous models (15, 16, 27, 29), the model we developed to simulate an inherently 3D system, is 2D. Such simplification, however, possesses a considerable computational advantage, which allowed us to run a large number of model variants in order to explore the effects of several factors on the outcome measures of interest, while still capturing all aspects of network mechanics, including nonlinear mechanical loading response in the macroscale level and loading-induced fiber realignment (29). Additionally, the mechanical properties assigned to the individual fibers were not directly derived from the literature, but from the simulated macroscale properties of the network, which were juxtaposed against previous experimental findings of the bulk response of collagen gels to uniaxial loading (Supplementary Materials). This is because most of our understanding of the mechanical behavior of fibrous gels comes from bulk measurements, while data on the elastic behavior of the individual fibers contained in the gels are sparse and inconsistent (16). Furthermore, we modeled the fiber segments as truss elements, which encompasses the assumption that these are subjected to uniaxial tension and compression (without bending) and are able to rotated freely one with respect to the other (Section 2.2.4). Some of the previous models (16, 25, 27, 29, 30, 33), however, did account for bending potentially occurring in the fibers and/or between two fibers meeting at a crosslink. Nevertheless, a preliminary study we conducted showed that using beam elements had affected the model outcomes only slightly, which concurs with previous findings (15, 16, 26, 33, 39). Specifically, Humphries et al. (15) and Heussinger and Frey (39) found that even when bending dominated the stretching energy of the loaded fibers, the forces acting in them were still predominantly axial, while Abhilash et al. (29) further claimed that the loading response of 2D models is less dominated by bending of the network segments. Considering these and other potential limitations of the model, it is recommended that the data presented in this study be interpreted mostly as trends of effects, rather than as absolute values (40). Particularly, the objective of the study was not to reproduce specific numerical results reached through laboratory experiments, but to compare trends of effects of the mechanical behavior of the ECM fibers on the mechanical signaling between neighboring cells, and we do not expect the limitations discussed above to affect such comparative findings. To conclude, the FE model presented herein predicts that nonlinear elastic behavior of the individual fibers constituting the ECM contributes to a highly directional and efficient inter- cellular mechanical signal transmission. Such mechanism regulating long-range cell-cell communication can elucidate previous experimental observations of biological processes involving inter-cellular mechanical signaling; it further highlights the importance of utilizing fibrous biological gels for facilitating long-range inter-cellular communication, in comparison with linear-elastic synthetic gels (such as polyethylene glycol and polyacrylamide gels), which are commonly used for cell culturing (41). Major challenges for future work include a more realistic description of the physiological scenario, including: 3D modeling; a more accurate representation of the network architecture in terms of connectivity and fiber arrangement; dynamic modeling accounting for the time-dependent cell morphology and loading, as well as the viscoelastic and plastic nature of ECM. Acknowledgments: We thank Prof. Samuel Safran and Prof. Yair Shokef for their recommendations while preparing the manuscript. This study was supported by the Israel Science Foundation number 1474/16 and the Israel Science Foundation- I-CORE number 1902/12. References 1. Rudnicki, M.S., H.A. Cirka, M. Aghvami, E.A. Sander, Q. Wen, and K.L. Billiar. 2013. Nonlinear strain stiffening is not sufficient to explain how far cells can feel on fibrous protein gels. Biophys. J. 105: 11–20. 2. Murrell, M., P.W. Oakes, M. Lenz, and M.L. Gardel. 2015. Forcing cells into shape: The mechanics of actomyosin contractility. Nat. Rev. Mol. Cell Biol. 16: 486–498. 3. Korff, T., and H.G. Augustin. 1999. Tensional forces in fibrillar extracellular matrices control directional capillary sprouting. J. Cell Sci. 112: 3249–3258. 4. Shi, Q., R.P. Ghosh, H. Engelke, C.H. Rycroft, L. Cassereau, J.A. Sethian, V.M. Weaver, and J.T. Liphardt. 2014. Rapid disorganization of mechanically interacting systems of mammary acini. Proc. Natl. Acad. Sci. 111: 658–663. 5. Nitsan, I., S. Drori, Y.E. Lewis, S. Cohen, and S. Tzlil. 2016. Mechanical communication in cardiac cell synchronized beating. Nat. Phys. 12: 472–477. 6. Stopak, D., and A.K. Harris. 1982. Connective tissue morphogenesis by fibroblast traction. I. Tissue culture observations. Dev. Biol. 90: 383–398. 7. Wen, Q., and P.A. Janmey. 2013. Effects of nonlinearity on cell-ECM interactions. Exp. Cell Res. 319: 2481–2489. 8. 9. Storm, C., J.J. Pastore, F.C. MacKintosh, T.C. Lubensky, and P.A. Janmey. 2005. Nonlinear elasticity in biological gels. Nature. 435: 191–194. Steinwachs, J., C. Metzner, K. Skodzek, N. Lang, I. Thievessen, C. Mark, S. Münster, K.E. Aifantis, and B. Fabry. 2015. Three-dimensional force microscopy of cells in biopolymer networks. Nat. Methods. 13: 171–176. 10. Piechocka, I.K., R.G. Bacabac, M. Potters, F.C. Mackintosh, and G.H. Koenderink. 2010. Structural hierarchy governs fibrin gel mechanics. Biophys. J. 98: 2281–2289. 11. Gentleman, E., A.N. Lay, D.A. Dickerson, E.A. Nauman, G.A. Livesay, and K.C. Dee. 2003. Mechanical characterization of collagen fibers and scaffolds for tissue engineering. Biomaterials. 24: 3805–3813. 12. Rijt, J.A.J. Van Der, K.O. Van Der Werf, M.L. Bennink, P.J. Dijkstra, and J. Feijen. 2006. Micromechanical Testing of Individual Collagen Fibrils. Mol. Biosci. 6: 697– 702. 13. Conti, E., and F.C. MacKintosh. 2009. Cross-linked networks of stiff filaments exhibit negative normal stress. Phys. Rev. Lett. 102: 88102. 14. Münster, S., L.M. Jawerth, B.A. Leslie, J.I. Weitz, B. Fabry, and D.A. Weitz. 2013. Strain history dependence of the nonlinear stress response of fi brin and collagen networks. Proc. Natl. Acad. Sci. 110: 12197–12202. 15. Humphries, D.L., J.A. Grogan, and E.A. Gaffney. 2017. Mechanical Cell–Cell Communication in Fibrous Networks: The Importance of Network Geometry. Bull. Math. Biol. 79: 498–524. 16. Reinhardt, J.W., and K. Gooch. 2017. An Agent-Based Discrete Collagen Fiber Network Model of Dynamic Traction Force-Induced Remodeling. J. Biomech. Eng. . 17. Stein, A.M., D.A. Vader, D.A. Weitz, and L.M. Sander. 2011. The Micromechanics of Three-Dimensional Collagen-I Gels. Complexity. 16: 22–28. 18. Notbohm, J., A. Lesman, D.A. Tirrell, and G. Ravichandran. 2015. Quantifying cell- induced matrix deformation in three dimensions based on imaging matrix fibers. Integr. Biol. 7: 1186–1195. 19. Vanni, S., B. Christoffer Lagerholm, C. Otey, L.L. Taylor, and F. Lanni. 2003. Internet-based image analysis quantifies contractile behavior of individual fibroblasts inside model tissue. Biophys. J. 84: 2715–2727. 20. Winer, J.P., S. Oake, and P.A. Janmey. 2009. Non-linear elasticity of extracellular matrices enables contractile cells to communicate local position and orientation. PLoS One. 4: e6382. 21. Reinhart-King, C.A., M. Dembo, and D.A. Hammer. 2008. Cell-Cell Mechanical Communication through Compliant Substrates. Biophys. J. 95: 6044–6051. 22. Xu, X., and S.A. Safran. 2015. Nonlinearities of biopolymer gels increase the range of force transmission. Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys. 92: 32728. 23. Shokef, Y., and S.A. Safran. 2012. Scaling laws for the response of nonlinear elastic media with implications for cell mechanics. Phys. Rev. Lett. 108: 178103. 24. Shokef, Y., and S.A. Safran. 2012. Erratum: Scaling Laws for the Response of Nonlinear Elastic Media with Implications for Cell Mechanics [Phys. Rev. Lett. 108, 178103 (2012)]. Phys. Rev. Lett. 109: 169901. 25. Ronceray, P., C.P. Broedersz, and M. Lenz. 2016. Fiber networks amplify active stress. Proc. Natl. Acad. Sci. 113: 2827–2832. 26. Notbohm, J., A. Lesman, P. Rosakis, D.A. Tirrell, and G. Ravichandran. 2015. Microbuckling of fibrin provides a mechanism for cell mechanosensing. J. R. Soc. Interface. 12: 20150320. 27. Ma, X., M.E. Schickel, M.D. Stevenson, A.L. Sarang-Sieminski, K.J. Gooch, S.N. Ghadiali, and R.T. Hart. 2013. Fibers in the extracellular matrix enable long-range stress transmission between cells. Biophys. J. 104: 1410–1418. 28. Hall, M.S., F. Alisafaei, E. Ban, X. Feng, C.-Y. Hui, V.B. Shenoy, and M. Wu. 2016. Fibrous nonlinear elasticity enables positive mechanical feedback between cells and ECMs. Proc. Natl. Acad. Sci. 113: 14043–14048. 29. Abhilash, A.S., B.M. Baker, B. Trappmann, C.S. Chen, and V.B. Shenoy. 2014. Remodeling of fibrous extracellular matrices by contractile cells: Predictions from discrete fiber network simulations. Biophys. J. 107: 1829–1840. 30. Wang, H., A.S. Abhilash, C.S. Chen, R.G. Wells, and V.B. Shenoy. 2015. Long-range force transmission in fibrous matrices enabled by tension-driven alignment of fibers. Biophys. J. 107: 2592–2603. 31. Harris, A.K., D. Stopak, and P. Wild. 1981. Fibroblast Traction as a Mechanism for Collagen Morphogenesis. Nature. 290: 249–251. 32. Ban, E., J.M. Franklin, S. Nam, L.R. Smith, H. Wang, R.G. Wells, O. Chaudhuri, J.T. Liphardt, and V.B. Shenoy. 2018. Article Mechanisms of Plastic Deformation in Collagen Networks Induced by Cellular Forces. Biophys. J. 114: 450–461. 33. Liang, L., C. Jones, S. Chen, B. Sun, and Y. Jiao. 2016. Heterogeneous force network in 3D cellularized collagen networks. Phys. Biol. 13: 66001. 34. Kurniawan, N.A., P.K. Chaudhuri, and C.T. Lim. 2016. Mechanobiology of cell migration in the context of dynamic two-way cell-matrix interactions. J. Biomech. 49: 1355–1368. 35. Vos, B.E., L.C. Liebrand, M. Vahabi, A. Biebricher, G.J.L. Wuite, E.J.G. Peterman, N.A. Kurniawan, F.C. MacKintosh, and G.H. Koenderink. 2016. Programming filamentous network mechanics by compression. arXiv. arXiv ID: 1612.08601. 36. Xu, X., and S.A. Safran. 2017. Compressive elasticity of polydisperse biopolymer gels. Phys. Rev. E. 95: 52415. 37. Golkov, R., and Y. Shokef. 2017. Shape regulation generates elastic interaction between living cells. New J. Phys. 19. 38. Ben-Yaakov, D., R. Golkov, Y. Shokef, and S.A. Safran. 2015. Response of adherent cells to mechanical perturbations of the surrounding matrix. Soft Matter. 11: 1412– 1424. 39. Heussinger, C., and E. Frey. 2007. Force distributions and force chains in random stiff fiber networks. Eur. Phys. J. E. 24: 47–53. 40. Sopher, R., J. Nixon, E. McGinnis, and A. Gefen. 2011. The influence of foot posture, support stiffness, heel pad loading and tissue mechanical properties on biomechanical factors associated with a risk of heel ulceration. J. Mech. Behav. Biomed. Mater. 4: 572–582. 41. Tibbitt, M., and K.S. Anseth. 2010. Hydrogels as Extracellular Matrix Mimics for 3D Cell Culture. Biotechnol. Bioeng. 103: 655–663. Supplementary Materials Derivation and validation of the mechanical properties assigned to the network fibers The mechanical properties assigned to the fibers contained in the ECM were derived from the simulated macroscale properties of the network, which were juxtaposed with previous experimental findings of the bulk response of collagen gels to uniaxial loading. In detail, a network consisting of truss elements of diameter of 200 nm (which is within the range reported by (17) for the diameter of collagen fibers), and of rectangular shape (length: 535 μm, width: 229 μm), was created as described in Section 2.2.1. The individual fibers were modeled as demonstrating all types of mechanical behavior listed in Section 2.2.2, while the reference elastic modulus (Eref) was initially set at a random value. The bottom edge of this virtual specimen was fixed for all translations and rotations. Uniaxial tension was introduced by applying maximum displacement of 100-μm to the top edge of the specimen. The nominal strain applied to the specimen was continuously calculated by dividing the length change of the rectangle in the vertical axis (i.e., the vertical displacement occurring at the top edge) by the reference, undeformed length (i.e., the initial length of the rectangle). The nominal stress applied to the specimen was calculated as the sum of the vertical components of all reaction forces occurring at the upper edge of the rectangle, divided by the axial cross-sectional area of the specimen (the width of the rectangle multiplied by the 'depth' of the specimen, i.e. the fiber diameter of 200 nm). The calculated stresses were plotted against the calculated strains, and the resulted curve, which demonstrated the macroscale stiffness of the simulated fibrous material, was juxtaposed with the reported mechanical behavior of collagen gel 2.4 mg/ml subjected to uniaxial tension, as measured using a rheometer (9) (Figure S1). The value of Eref assigned to the individual fibers was iteratively adjusted until reaching satisfactory resemblance between the curves (a process similar to that described by (1)), and was ultimately set at 11.5 kPa. Validation of bulk mechanical behavior Buckling + strain stiffening Collagen gel 2.4 mg/ml (Steinwachs et al., 2016) ] a P k [ s s e r t s l i a n m o N 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0% 5% 10% 15% Nominal axial strain Figure S1: Stress-strain curve demonstrating the bulk mechanical behavior of the simulated fibrous network (when assuming fiber diameter of 200 nm, tension-stiffening of the individual fibers and Eref=11.5 kPa) when subjected to uniaxial tension, juxtaposed with an equivalent curve derived experimentally for collagen gel 2.4 mg/ml (9). The stress-strain relationships demonstrated by the simulated bulk material, as derived from the aforementioned analysis when assuming each of the four types of fiber mechanical behavior listed in Section 2.2.2 and implementing the aforementioned value of Eref, are shown in Figure S2a. When assuming tension-stiffening behavior of the individual fibers (material model iii in Section 2.2.2), the bulk materials was the stiffest, followed by fibers demonstrating both tension-stiffening and compression-buckling (material model iv). In other words, the elevated resistance of the fibers contained in the network to tension resulted in an increased tension-stiffness of the bulk material. When assuming fiber compression-buckling alone (material model ii), the bulk material was slightly softer than when assuming linear elasticity (material model i in Section 2.2.2) (Figure S2a). This is attributable to the decreased resistance of the fibers subjected to compression loading (particularly those aligned horizontally, i.e. along the width of the specimen) to such compression. The Poisson's ratio of the simulated material was also calculated when assuming all types of fiber mechanical behavior listed in Section 2.2.2, by dividing the nominal strain along the width of the rectangular specimen (transverse strain) by the nominal strain along the length of the material (axial strain). When assuming linear-elastic behavior of the individual fibers, Poisson's ratio was nearly constant at 0.4-0.5 (Figure S2b). When modeling the fibers as bucklable, increasing axial strain resulted in a gradual increase in the Poisson's ratio, which is in agreement with previous findings (26). Tension-stiffening of the individual fibers resulted in the Poisson's ratio of the simulated bulk material exceeding 1 (Figure S2b), which is in agreement with a previous computational model (32). Pure shear was also exerted to the virtual specimen by applying horizontal displacement to the top edge of the rectangle. The engineering shear strain applied to the specimen was continuously calculated as the change in angle between the horizontal and vertical edges of the rectangle. The engineering shear stress was calculated as the sum of the horizontal components of all reaction forces occurring at the upper edge of the rectangle, divided by the axial cross-sectional area of the specimen. The shear modulus of the bulk material was subsequently estimated as the fraction of the shear stress to the shear strain applied to the specimen. The shear modulus was then plotted against the engineering shear strain (Figure S2c). Tension-stiffening of the individual fibers resulted in the shear modulus increasing with the shear strain. Fiber compression-buckling resulted in the bulk shear modulus being considerably smaller (Figure S2c), which is attributable to a decreased resistance of the fibers to compression. Bulk stress-strain curve under uniaxial tension 0% 2% 4% 6% 8% 10% 12% 14% Nominal axial strain (a) Bulk Poisson's ratio under uniaxial tension 0.5 0.4 0.3 0.2 0.1 0 ] a P k [ s s e r t s l i a n m o N o i t a r s ' n o s s i o P 2 1.5 1 0.5 0 0% 5% 10% 15% Nominal axial strain (b) Bulk shear modulus againt shear strain ] a P k [ s u u d o m l r a e h S 0.25 0.2 0.15 0.1 0.05 0 0% 5% 10% 15% 20% Engineering shear strain (c) Figure S2: The bulk mechanical behavior of the simulated fibrous network (when assuming fiber diameter of 200 nm and Eref=11.5 kPa) for four types of mechanical behavior of the individual fibers: (a) stress-strain relationship under uniaxial tension; (b) Poisson's ratio under uniaxial tension; (c) shear modulus against engineering shear strain. Tensile strain Compressive strain Strain energy density r a e n i L n o i s n e t n i g n i n e f f i t S n o i s s e r p m o c n i g n i l k c u B g n i n e f f i t s + g n i l k c u B (a) (b) (c) (d) 0 [%] (e) (f) (g) (i) (j) (k) 2.3 4.5 (h) 0 [%] 5.3 10.5 (l) 0 [Pa] 3 6 Figure S3: Contour plots showing the tensile strains (left column), compressive strains (middle column) and SEDs (right column) occurring in the fiber segments within the vicinity of a single, isolated contracting cell, for 25% contraction. Plots were produced for all four material models used to simulate the mechanical behavior of the individual fibers (Figure 1c). Tensile strain Compressive strain Strain energy density r a e n i L n o i s n e t n i g n i n e f f i t S n o i s s e r p m o c n i g n i l k c u B g n i n e f f i t s + g n i l k c u B (a) (b) (c) (d) 0 [%] (e) (f) (g) (h) 0 [%] 2.3 4.5 3 6 (i) (j) (k) (l) 0 [Pa] 5.3 10.5 Figure S4: Contour plots showing the tensile strains (left column), compressive strains (middle column) and SEDs (right column) occurring in the fiber segments within the vicinity of two neighboring contracting cells (here the cell-to-cell distance is equivalent to 3.4 cell diameters as an example), for 25% contraction. Plots were produced for all four material models used to simulate the mechanical behavior of the individual fibers (Figure 1c). 2.1 D 3.4 D 9.0 D r a e n i L n o i s n e t n i g n i n e f f i t S n o i s s e r p m o c n i g n i l k c u B g n i n e f f i t s + g n i l k c u B (a) (b) (c) (d) (e) (i) (f) (j) (g) (h) (k) (l) 0 [Pa] 3 6 Figure S5: Contour plots showing the SEDs occurring in the ECM fiber segments surrounding two contracting cells, for 25% contraction. Cell-to-cell distances shown here are equivalent to 2.1 (left column), 3.4 (middle column) and 9.0 (right column) cell diameters. Plots were produced for all four material models used to simulate the mechanical behavior of the individual fibers (Figure 1c). cell 1 cell 2 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (b) i n a r t s e l i s n e t n a e M i n a r t s e v i s s e r p m o c n a e M ] a P [ D E S n a e M 2.5% 2.0% 1.5% 1.0% 0.5% 0.0% 5% 4% 3% 2% 1% 0% 120 100 80 60 40 20 0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (c) Figure S6: Mean tensile (logarithmic) strain (a), compressive (logarithmic) strain (b) and SED (c), occurring within a disc surrounding an individual cell, of radius equals to half of the cell-to-cell distance, for 25% contraction. The model variants shown include several cell-to-cell distances (in terms of cell diameter, D), as well as the equivalent single-cell model variants, and four of the material models used to simulate the mechanical behavior of the individual fibers (Figure 1c). i n a r t s e l i s n e T i n a r t s e v i s s e r p m o C o i t a r y t i l a n o i t c e r i d o i t a r y t i l a n o i t c e r i d 0.4 0.3 0.2 0.1 0.0 0.3 0.2 0.1 0.0 cell 1 cell 2 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) 3.4 (b) Single 1.8 2.1 2.8 4.0 6.5 9.0 11.5 15.3 19.0 o i t a r y t i l a n o i t c e r i d D E S 0.6 0.5 0.4 0.3 0.2 0.1 0.0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (c) Figure S7: Directionality ratios for tensile strains (a), compressive strains (b) and SEDs (c), occurring within a disc surrounding an individual cell, of radius equals to half of the cell-to-cell distance, for 10% contraction. i n a r t s e l i s n e T i n a r t s e v i s s e r p m o C o i t a r y t i l a n o i t c e r i d o i t a r y t i l a n o i t c e r i d 0.5 0.4 0.3 0.2 0.1 0.0 0.4 0.3 0.2 0.1 0.0 Single 1.8 2.1 2.8 Single 1.8 2.1 2.8 3.4 (a) 3.4 (b) cell 1 cell 2 4.0 6.5 9.0 11.5 15.3 19.0 4.0 6.5 9.0 11.5 15.3 19.0 o i t a r y t i l a n o i t c e r i d D E S 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (c) Figure S8: Directionality ratios for tensile strains (a), compressive strains (b) and SEDs (c), occurring within a disc surrounding an individual cell, of radius equals to half of the cell-to-cell distance, for 25% contraction. i n a r t s e l i s n e T i n a r t s e v i s s e r p m o C o i t a r y t i l a n o i t c e r i d o i t a r y t i l a n o i t c e r i d 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0.4 0.3 0.2 0.1 0.0 cell 1 cell 2 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) 3.4 (b) Single 1.8 2.1 2.8 4.0 6.5 9.0 11.5 15.3 19.0 o i t a r y t i l a n o i t c e r i d D E S 1.0 0.8 0.6 0.4 0.2 0.0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (c) Figure S9: Directionality ratios for tensile strains (a), compressive strains (b) and SEDs (c), occurring within a disc surrounding an individual cell, of radius equals to half of the cell-to-cell distance, for 50% contraction. r a e n i L n o i s n e t n i g n i n e f f i t S n o i s s e r p m o c n i g n i l k c u B g n i n e f f i t s + g n i l k c u B (a) (b) (c) (e) (f) (g) (d) (h) 2.7 0 ×10-2 [nN] 5.4 Figure S10: Contour plots showing the reaction forces occurring on the cell boundaries for a single (left column) or two (right column; here distance between the neighboring cells is equivalent to 3.4 cell diameters as an example) contracting cells, for 25% contraction. Plots were produced for all four material models used to simulate the mechanical behavior of the individual fibers (Figure 1c). cell 1 cell 2 ] N n [ e c r o f n o i t c a r e t n i - l l e C -0.10 -0.08 -0.06 -0.04 -0.02 0.00 3.0 2.5 2.0 1.5 1.0 0.5 0.0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) cell 1 cell 2 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (b) l y t i r a o p e c r o F Figure S11: (a) Net cell-interaction force occurring on the cell boundary (as projected on the line connecting the cell centers), for 10% contraction. (b) The polarity ratio of the contraction force occurring on the cell boundary. cell 1 cell 2 ] N n [ e c r o f n o i t c a r e t n i - l l e C -25 -20 -15 -10 -5 0 25 20 15 10 5 0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) * cell 1 cell 2 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (b) l y t i r a o p e c r o F Figure S12: (a) Net cell-interaction force occurring on the cell boundary (as projected on the line connecting the cell centers), for 50% contraction. (b) The polarity ratio of the contraction force occurring on the cell boundary. * The value should actually be approximately 109, but is not shown on the graph. y t i s n e d r e b i f n i e g n a h c % r e b i f n a e m n i e g n a h c cell 1 cell 2 10% 5% 0% Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) l ] g e d [ e g n a n o i t a t n e i r o -3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (b) Figure S13: (a) Relative change in density of fiber segments contained in the inter-cellular medium, and (b) mean change in the angles of the orientation of the fibers contained in the inter-cellular medium (orange sector at the top right panel), as a result of 10% cell contraction. cell 1 cell 2 45% 40% 35% 30% 25% 20% 15% 10% 5% 0% Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 19.0 (a) l ] g e d [ e g n a n o i t a t n e i r o -20 -15 -10 -5 0 Single 1.8 2.1 2.8 3.4 4.0 6.5 9.0 11.5 15.3 18.4 19.0 (b) y t i s n e d r e b i f n i e g n a h c % r e b i f n a e m n i e g n a h c Figure S14: (a) Relative change in density of fiber segments contained in the inter-cellular medium, and (b) mean change in the angles of the orientation of the fibers contained in the inter-cellular medium (orange sector at the top right panel), as a result of 50% cell contraction.
1705.04047
1
1705
2017-05-11T07:34:46
Controlling the morphology and outgrowth of nerve and neuroglial cells: The effect of surface topography
[ "physics.bio-ph" ]
Unlike other tissue types, like epithelial tissue, which consist of cells with a much more homogeneous structure and function, the nervous tissue spans in a complex multilayer environment whose topographical features display a large spectrum of morphologies and size scales. Traditional cell cultures, which are based on two-dimensional cell-adhesive culture dishes or coverslips, are lacking topographical cues and mainly simulate the biochemical microenvironment of the cells. With the emergence of micro- and nano-fabrication techniques new types of cell culture platforms are developed, where the effect of various topographical cues on cellular morphology, proliferation and differentiation, can be studied. Different approaches (regarding the material, fabrication technique, topographical charactertistics, etc.) have been implemented. The present review paper aims at reviewing the existing body of literature on the use of artificial micro- and nano-topographical features to control neuronal morphology, outgrowth and neural network topology. The cell responses from phenomenology to investigation of the underlying mechanisms- on the different topographies, including both deterministic and random ones, are summarized.
physics.bio-ph
physics
1 Controlling the morphology and outgrowth of nerve and neuroglial cells: The effect of surface topography C. Simitzi, A. Ranella and E. Stratakis Institute of Electronic Structure and Laser (IESL), Foundation for Research and Technology- Hellas (FORTH), Heraklion, 71003, Greece Abstract Unlike other tissue types, like epithelial tissue, which consist of cells with a much more homogeneous structure and function, the nervous tissue spans in a complex multilayer environment whose topographical features display a large spectrum of morphologies and size scales. Traditional cell cultures, which are based on two-dimensional cell-adhesive culture dishes or coverslips, are lacking topographical cues and mainly simulate the biochemical microenvironment of the cells. With the emergence of micro- and nano-fabrication techniques new types of cell culture platforms are developed, where the effect of various topographical cues on cellular morphology, proliferation and differentiation, can be studied. Different approaches (regarding the material, fabrication technique, topographical charactertistics, etc.) have been implemented. The present review paper aims at reviewing the existing body of literature on the use of artificial micro- and nano-topographical features to control neuronal morphology, outgrowth and neural network topology. The cell responses from phenomenology to investigation of the underlying mechanisms- on the different topographies, including both deterministic and random ones, are summarized. Keywords Topography, nerve cells, neuroglial cells, micro-/nano-fabrication Statement of significance With the aid of micro-and nanofabrication techniques, new types of cell culture platforms are developed and the effect of surface topography on the cells has been studied. The present review article aims at reviewing the existing body of literature reporting on the use of various topographies to study and control the morphology and functions of cells from nervous tissue, i.e. the neuronal and the neuroglial cells. The cell responses from phenomenology to 1 investigation of the underlying mechanisms- on the different topographies, including both deterministic and random ones, are summarized. 2 Table of contents 1. Introduction 1.1 The in vitro study of the effect of topography on nerve cells 1.2 Neurons and neuroglial cells: the basic components of the nervous tissue 1.3 3D micro/nano surface texturing of biomaterials as a means of manipulation of neuron cell fate 1.4 Different topographies used as cell culture platforms for nerve cell studies 1.5 Interfacing nerve cells with biomaterials: The effect of surface topography 1.6 Cell types, assays and materials 2. The in vitro effects of topography on nerve cell morphology, functions and network topology 2.1 The effect of continuous topographies on nerve cells 2.2 The effect of discontinuous topographies on nerve cells 2.3 The effect of random topographies on nerve cells 2.4 The effect of topography on nerve cells -An insight into the mechanisms 3. The in vitro effects of topography on neuroglial cell morphology and functions 3.1 The effect of continuous topographies on neuroglial cells 3.2 The effect of discontinuous topographies on neuroglial cells 3.3 The effect of random topographies on neuroglial cells 3.4 The effect of topography on neuroglial cells-An insight into the mechanisms 4. Discussion & Conclusions 4.1 Effect on neurons 4.2 Effect on neuroglial cells 4.3 Other issues 2 3 1. Introduction 1.1 The in vitro study of the effect of topography on nerve cells The role of soluble (bio)chemical signals in cell shape, cell adhesion, differentiation and axon guidance, is well established [1]. In addition to the biochemical signals, there is increasing evidence that the physical parameters (e.g. topography and stiffness) of the complex extracellular milieu, constituted by the extracellular matrix (ECM) components and the surrounding cells, are also important [2]. Recent in vitro studies suggest that many neuronal and neuroglial cells respond to stiffness and to other mechanical cues during development (reviewed in [3]). Topography, which can be described as the arrangement of the spatial and structural features (e.g. geometrical architecture, discontinuities motif, contours, etc.) of the extracellular environment, has been also strongly correlated with specific functional characteristics of the cells and the tissues at both physiological (e.g. during development) and pathological states (e.g. wound healing) of nervous tissue [2,4]. The importance of ECM architecture in cells and tissue organization is apparent already from the early developmental stages. Embryonic cells produce their own extracellular scaffolds by secreting many types of molecules in the surrounding space, following a well tiation [5,6]. The different spatial organization of these secreted molecules gives rise to a great variety of natural scaffolds where cells continue to proliferate and organize themselves in order to build up tissues and accomplish all their natural functions [7]. The role of ECM organization and distribution on directing neural crest migration has been early understood [8,9]. During glioma progression, i.e. a brain cancer, individual cancer cells have the tendency to migrate along myelinated fibers of white matter tracts [10]. The role of topography on cellular outgrowth in vitro has been addressed very early; already in 1914 R. G. Harrison cultured embryonic frog spinal neurons in a meshwork of P. Weiss made similar observations with embryonic chicken spinal neurons on grooves generated by brushing clotting blood and established the an attempt to describe the tendency of the cells to orient themselves along anisotropic topographical features of the surface (such as fibers or ridges) [2,11]. The observations that certain physical properties of the substrate can in vitro influence cellular outgrowth and functions opened a new promising research field. However, many of these early experimental situations in which contact guidance was demonstrated were quite complex. Indeed, it was difficult to discriminate between the effects resulting from 3 4 the chemical cues and those resulting from the topographical ones. More specifically, the substrates used, i.e. plasma clots, fish scales, and various grooved surfaces, were anisotropic not only in shape but also in chemistry [12]. Accordingly, there was a need to carefully study the effect of topographical cues vis- a-vis (bio)chemical cues in a reproducible way. With the emergence of micro- and nanofabrication techniques, a plethora of approaches to engineering or tailoring surfaces in a controllable manner are now available and specific topographical patterns, at micro and sub- micron scale can be designed and fabricated at a plethora of different materials [13 16]. Patterning of surfaces has triggered the development of new types of cell culture platforms, where the effect of topographical cues on cellular responses can be investigated and/or manipulated, depending on the field of interest. 1.2 Neurons and neuroglial cells: the basic components of the nervous tissue Unlike other tissues, for example the epithelial tissue, where cells exhibit simple shapes, the nervous tissue is a complex three-dimensional environment whose topographical features span a large spectrum of morphologies and size scales. Nerve cells are the functional units of the nervous tissue which are responsible for transmitting information from and towards the environment through electrical and chemical signals. Different neuronal cell subtypes exist in the central and peripheral nervous tissue, namely the pyramidal cells and Purkinje cells in the central nervous system (CNS) and the sensory, motor and sympathetic neurons in the peripheral nervous system (PNS). Nerve cell shape exhibits a highly polarized pattern, comprising two types of neurite extensions from the cell body, namely the dendrites and the axon. This polarization, termed as neuronal polarity, establishes the signal transmission which underlies neural function. Cell bodies vary significantly in size among the different cell types; however, in the vertebrate nervous system, they typically exhibit diameters in the range of 10 micrometer range (typically 0.2 [17]. he The growth cone, the highly motile leading edge of axon, navigates along specific pathways to reach the correct target, by recognizing and translating an ensemble of bio(chemical) and physical cues, which are mediated either by diffusion or by contact. Growth cone uses its cytoskeletal elements, including actin filaments and microtubules, to move forward in a process that involves the aid of a plethora of regulator proteins, such as the motor protein myosin II [18]. In this way, the growth cone is advanced and the axon is 4 5 elongated (Axon outgrowth). Furthermore, growth cone uses the elements of the navigation system, including signal transduction molecules (e.g. the Rho family of GTPases that control many downstream signalling pathways) which translate environmental attractive or repulsive guidance cues into localized cytoskeletal remodelling (Axon guidance) [19]. Axon guidance can share similar mechanisms with the axon branching, i.e. the formation of axonal branches arisen by splitting of the terminal growth cone, which is another important axonal process in CNS development [20,21]. Extending axons may also grow in tight bundles (fasciculation). Fasciculation is driven by a balance of attractive and repulsive forces on the axons relative to their surrounding environment and by cues provided by other axons [22]. All these neuronal processes briefly described above, including the neuronal polarity establishment, axonal outgrowth, guidance and fasciculation, shape the neuronal morphology and are critical for the formation of the intricate neural networks of the functional neural tissue during development and regeneration. Neuroglial cells present the other main cell type of nervous tissue and provide a rich and supportive environment for neurite outgrowth during development and nerve regeneration. Astrocytes, oligodendrocytes, and microglial cells are three types of glial cells in the mature CNS and Schwann cells are the cells that elaborate myelin in the PNS. Owing to their capability to release neurotrophic factors, to express cell surface ligands and synthesize extracellular matrix (ECM) but also to their oriented shape and structural organization, neuroglial cells present both molecular and topographical guidance stimuli for the development and outgrowth of neurons. A characteristic example includes radial glia, which are organized into regular arrays spanning the walls of the developing brain and guide the migrating embryonic neurons [23]. 1.3 3D micro/nano surface texturing of biomaterials as a means of manipulation of neuron cell fate Numerous techniques are available for the development of patterned biomaterial surfaces, ranging from simple manual scratching to more controlled fabrication methods [24]. Examples of techniques to engineer surfaces in a controllable manner include [13,15,16,25]. Using these techniques, 3D topographical features of tailored geometry, roughness and orientation, complemented by the desired spatial resolution at micron- and submicron-scales, can be realized on material surfaces of different chemical and mechanical properties [25,26]. The main characteristics of the 3D micro/submicron surface texturing 5 6 techniques of biomaterials are listed in the Table 1 [16,25,27 33]. With the aid of these techniques mainly the well-defined geometries, also coined as deterministic can be fabricated. Furthermore, techniques and strategies to fabricate more random surface morphologies at nanoscale have been implemented. Usually, such substrates are fabricated via chemical or physical etching and surface roughness can be controlled by variable exposure time to chemicals or physical agents, like plasma. Another approach for the development of surfaces exhibiting random roughness is the fabrication of monodispersed colloids (e.g. the fabrication of monodisperse silica particles via the Stoeber process). There are few studies using nanoporous gold, which is typically produced by de-alloying of silver- gold alloys, i.e. by selective dissolution of silver from a silver-rich gold alloy [34]. 6 ] 6 1 [ o t d e i l p p a y l i s a e t o N - n o i t a c i r b a f o r c i m r e h t o r o f s i s a b e h T - : y l l a u s U n o c i l i s y l l a u s U d e t a r e n e g - r e s u a f o r e f s n a r t e h T 7 . f e R s e g a t n a v d a s i D s e g a t n a v d A e p y T n r e t t a P l a i r e t a M e l p i c n i r P e u q i n h c e T s l a i r e t a m o i b f o s e u q i n h c e t g n i r u t x e t e c a f r u s n o r c i m b u s / o r c i m D 3 f o s c i t s i r e t c a r a h c n i a M : a 1 e l b a T d n a s e v o o r g g n i t a n r e t l a s e g d i r l a i r e t a m a o t n o n r e t t a p l a c i r t e m o e g a f o e r u s o p x e e v i t c e l e s e h t h g u o r h t r e m y l o p e v i t i s n e s t h g i l y h p a r g o h t i l ] 9 2 [ s e p y t e m o s n i t s o c h g i H - l l a w e d i s d n a t h g i e h l a r u t c u r t s e g r a L - o i t a r t c e p s a h g i H : y t e i r a v e d i W e u q i n h c e t n o i t a c i r b a f d i r b y h A n o i t c u d o r p o t h c r a e s e R - s s e c o r p d e t a c i l p m o C - t l u c i f f i d s i n o i t i s n a r t s s e c o r p w o l S - m o r f g n i g n a r s s e n k c i h T - - 0 0 1 n o i t u l o s e r l a i t a p s h g i H - s o i t a r t c e p s a h g i H - ) A G I L y a r - X . e . i ( s e i t r e p o r p s e r u t c u r t s o r c i m , s l a t e m , s r e m y l o p d n a s y o l l a s c i m a r e c d n a g n i t a l p o r t c e l e . 2 , y h p a r g o h t i l . 1 g n i d l o m . 3 y l l a i t n e u q e s f o g n i t s i s n o c ] 3 3 , 5 2 [ e l b i s s o p t o n n a h t s a e r a r e g r a l n r e t t a p o t r e i s a E - r a l u g e r - o d u e s p e l b i s s e c c a d n a e l p m S - i m o r f y r a v s n r e t t a P n o i t u l o s e r l a i t a p s h g i H - o t m o d n a r r o r a l u g e r r i L B E h t i w f o e g n a r e d i W s k s a m s a s l a t s y r c l a d i o l l o c f o e s U . g . e ( s l a i r e t a m e h t r o f n o i t i s o p e d d n a g n i h c t e r o f d n a s r e m y l o p s u o i r a v f o n o i t a c i r b a f . ) s l a t e m - n o n d n a r a n a l p n o s e r u t c u r t s o n a n s e t a r t s b u s r a n a l p n i t l u s e r s t s i s e r e v i t a g e N - e g a r e v o c e c a f r u s l l a m S - n o i t u l o s e r e r u t a e f r e w o l g n i m u s n o c e m T - i g n i n r e t t a p d e l l o r t n o c - r e t u p m o C - e l a c s o n a n t a n a d n a s s a l g e h t s s o r c a m a e b n o r t c e l e n a s n a c s ) A M M P . g . e ( t s i s e r d n a e v i t i s o p h t o B . e c a f r u s e h t e v i t i s n e s - n o r t c e l e n o d e t a o c t s i s e r e v i t i s n e s - n o r t c e l e e l b a l i a v a e r a s t s i s e r e p y t e v i t a g e n ] 5 2 [ t s o c h g i H - . k s a m r o f d e e n o N - s n r e t t a p n o i s i c e r p h g i H r o n o c i l i s y l l a u s U n u g n o r t c e l e d e l l o r t n o c - r e t u p m o c A 7 s l a i r e t a m o i b f o s e u q i n h c e t g n i r u t x e t e c a f r u s n o r c i m b u s / o r c i m D 3 f o s c i t s i r e t c a r a h c n i a M : b 1 e l b a T ] 1 3 [ e l b i s s o p t o -n d e r i u q e r e c a f r u s n o l o r t n o c d e t i i m L - y r t s i m e h c e u q i n h c e t p e t s e l p i t l u M - s e i t i l i c a f m o o r - n a e l C - t n e m p i u q e e v i s n e p x e & e c a f r u s d e v r u c - X , m a e b n o r t c e l e s a h c u s , s e u q i n h c e t l l a r o f s s e c o r p g n i y l r e d n u e h T - n i s e s s e c o r p g n i r u t c u r t s o r c i m . c t e , y h p a r g o h t i l y a r s c i n o r t c e l e o r c i m e m u l o v d e t c e f f a e h t f o e z i s d e t i i m L - s e i r t e m o e g x e l p m o C f o e g n a r d a o r B h g i h f o m a e b r e s a l d e s l u p A n o i t c a r e t n i t c a t n o c - n o N - e t a r n o i t a c i r b a f h g i H - y t i l i b i c u d o r p e R - r o s e p a h s D 3 g n i d u l c n i g n i y r a v h t i w s e r u t c u r t s h c t e d n a s e p a h s l l a w s u o i r a v d n a e h t n o s o i t a r , s h t p e d t c e p s a e t a r t s b u s e m a s s l a t e m s l a i r e t a m e l a c s e m i t t r o h s t a y t i s n e t n i , s c i m a r e c ( s i l a i r e t a M . e t a r t s b u s n o s e g n i p m i ) s r e m y l o p d n a d e t c e j e e g r e m e d n a n a c d e t l e m r o d e z i r u o p a v n r e t t a P . e c a f r u s m o r f , e l p m a s e h t f o t n e m e v o m y - x y b h t o b f o n o i t a n i b m o c a r o m a e b e h t : g n i r u t c u r t s - o t o h P - o t o h P : g n i r u t c u r t s - t r o h s a r t l U d e s l u p - o t o h P n o i t a l b a r e s a L : g n i r u t c u r t s - o t o h P A G I L y h p a r g o h t i l m a e b - E ) L B E ( y h p a r g o h t i l l a d i o l l o C ] 2 3 , 0 3 [ ] 2 3 , 0 3 [ n o i t r o t s i d p m a t S ) s e i t i l i c a f m o o r n a e l c r o f d e e n o n ( d e s u r e m y l o p n o i t a c i l p e r d n a e r u d e c o r p y s a e d n a e l p m S - i r e t s a m e h t n o e c n e d n e p e D 8 . f e R s e g a t n a v d a s i D s e g a t n a v d A e p y T n r e t t a P l a i r e t a M e l p i c n i r P e u q i n h c e T p m a t s e h t f o g n i l l e w s / g n i k n i r h S - d e t a e r c e b n a c s p m a t s e l p i t l u M - n o i t a n i m a t n o c e t a r t s b u S - e r u d e c o r p y s a e d n a e l p m S - i s r e t s a m e h t f o y t i l i b a s u e R - r e t s a m e l g n i s a m o r f ) m n 0 0 1 ~ ( n o i t u l o s e r h g i H - d e t a e r c e b n a c s p m a t s e l p i t l u M - e h t n o s d n e p e d , s e r u t a e f l a m i n i m h t i w s e m i t l a r e v e s s l o i h t e n a k l a e h t f o n o i t a r t n e c n o c e c n a m r o f r e p f o n o i t a d a r g e d s n o i t i d n o c g n i t n i r p e h t d n a s e c a f r u s d e v r u c d n a t a l f h t o B - d e p m a t s e b n a c d e t n i r p e h t f o n o i t u l o s e r e h T - d e s u e b n a c s p m a t s l a u d i v i d n I - y t i l i b o m k n I - r e t s a m e l g n i s a m o r f s e c a f r u s o t n o . c t e d e t a o c ) s e c a f r u s - 2 O S i , s l o i h t e n a k l a f o g n i n r e t t a P s e t a r t s b u s l a t e M , s r e m y l o p , s e n a l i s , s n i e t o r p r o u A y l n i a m ( c i r e m y l o P s l a i r e t a m g n i t s a c a i v n o i t a c i l p e r e h T - y h p a r g o h t i l t f o S D 3 e h t f o g n i r u c d n a g n i d l o M a c i l p e R , d e n r e t t a p a f o y h p a r g o p o t e n a x o l i s l y h t e m i d y l o p n a n i e c a f r u s d i l o s y l l a u s u , r e m o t s a l e ) S M D P ( - k n i f o ) s M A S ( s r e y a l o n o m o t n o p m a t s S M D P a m o r f . e c a f r u s e t a r t s b u s e h t - t s e r e t n i f o l a i r e t a m e h t d e l b m e s s a f l e s y l l a u s u t c a t n o c o r c i M g n i t n i r p f o r e f s n a r t n r e t t a p e h T - y h p a r g o h t i l t f o S ] 8 2 , 5 2 [ ] 7 2 [ l a t e m e h t m o r f n o i t a n i m a t n o C - t s y l a t a c e l b i s s o p - s a e r a r e g r a l n r e t t a p o t r e i s a E - d e s a b - y h p a r g o h t i l n a h t s o i t a r t c e p s a h g i H - s e u q i n h c e t s e r u t c u r t s y t i l a u q h g i H - s r e k s i h w d n a s d o r , s e r i w o n a N s e i t r e p o r p l a c i n a h c e m s r e t e m a i d r e b i f d n a s r e b i f n e v o w n o n a o t n i n u p s d l o f f a c s a y b d e t s i s s a m s i n a h c e m . t s y l a t a c l a t e m s l a t e m n o i t i s n a r t d n a e l b o N S L V e h t a i v n o i t a m r o f h t w o r g l a t s y r c l a i x A r u o p a V l a c i m e h C n o i t i s o p e d d e t i m L i s e z i s e r o p r e v o l o r t n o c d o o G d e t n e i r o y l m o d n a r r o d e n g i l A n o i t u l o s r e m y l o P y l l a c i t a t s o r t c e l e e r a s r e b i F g n i n n i p s o r t c e l E ] 8 2 , 6 1 [ n o i s n e m i d e r u t a e f d e r i s e d e h T n o i t a n i m a l e d t s i s e r o t o h P : t e W s e t a r h c t e e h t n o s d n e p e d s e s a g c i x o T : y r D c i p o r t o s i n A : y r D s t l u s e r g n i h c t e c i p o r t o s i n a t e w d e n i l c n i h t i w s e i t i v a c n i t o n s r u c c o g n i h c t e c i p o r t o s I , h t p e d f o n o i t c e r i d e h t n i y l n o s t l u s e r d n a , y l l a r e t a l o s l a t u b . s l l a w e d i s s c i t s a l p a n o s e r u t a e f l a c i h p a r g o p o t l a v o m e r e v i t c e l e s y b e c a f r u s l a c i s y h p h g u o r h t l a i r e t a m f o l a c i m e h c r o ) g n i h c t e y r d ( . s n a e m ) g n i h c t e t e w ( e l p m i s d n a n a e l C : t e W g n i h c t e c i p o r t o s i n a y r D , s s a l g , n o c i l i S f o n o i t a e r c e h T g n i h c t E : S L V ; g n u m r o f b A , g n u m r o f o n a v l a G , e i h p a r g o h t i L r o f m y n o r c a n a m r e G : A G I L ; n g i s e d d e d i a - r e t u p m o C : D A C ; l a n o i s n e m i d e e r h t : D 3 : s n o i t a i v e r b b A 8 d i l o S - d i u q i L - r o p a V 9 1.4. Different topographies used as cell culture platforms for nerve cell studies In an attempt to establish a link between the topographical cues with specific cell responses, the patterned surfaces must be well defined and analyzed in terms of geometry. In the case of the random surface patterns, the surface topography cannot be well defined and is usually described in terms of statistical roughness parameters, such as the arithmetic average, Ra or the root mean square, Rq roughness values [35]. For the geometrical description of the nanoporous surfaces, as described in 1.3, parameters such as void coverage or average pore diameter have been used. However, the majority of the studies describe models of deterministic topographies. Such topographies will be presented in this review based on a two-stage classification. Primary classification categorizes the various pattern geometries into continuous and discontinuous. Figure 1A illustrates examples of the cross-sections of the respective topographies with the third dimension being omitted for clarity reasons. Continuous geometries can be further classified into anisotropic and isotropic topographies. Anisotropic topographies are directionally dependent, providing cues preferentially along a single axis. Examples of continuous anisotropic topographies are photolithographically fabricated grooved silicon substrates or electrospun polymer fibers at parallel orientation. Isotropic topographies provide cues along multiple axes, for example randomly-oriented fibers. In the continuous topographies the experimental parameters include feature width and depth and fiber parameters (e.g. orientation, diameter), respectively. Discontinuous geometries are described here based on the anisotropy or asymmetry of the feature. Based on this, pillars of circular (e.g. silicon or gold pillars or post [36]) or squared cross- section are examples of isotropic or symmetric features, while cones of elliptical cross-section is an example of asymmetric or anisotropic ones [37]. In these topographies the experimental parameters include the interfeature spacing or pitch, feature height, etc. Furthermore, discontinuous topographies can further be classified based on the distribution of the features and described as periodic or random. Figure 1: Classification of the deterministic topographies used as cell culture platforms for nerve cell studies. (A) Schematic illustration of the cross-sections of the various topographies, including the continuous, such as grooves and fibers (Fig. 2-4 and 7), and the discontinuous ones, such as pillars and cones (Fig. 5 and 8).The cross-section can exhibit either the same (e.g. as in the case of pillars) or differential (e.g. as in the case of cones) surface area along the feature height. (B) Classification of the deterministic topographies based on the uniformity of the features; example of continuous (top row) and discontinuous (bottom row) topographies, like grooves and pillars, respectively. 9 10 Both continuous and discontinuous topographies can be classified into uniform and graded, depending on whether they provide cues through uniform and gradual changes in the main physical feature (e.g. groove spacing) along a particular direction, respectively (Figure 1B). However, apart from the type of the topographical feature, it has to be emphasized that the feature dimension will also play a critical role on the cell outgrowth [38]. Accordingly, feature dimensions must be correlated with the cell size or the size of the cellular extensions (such as neurites). This can provide a broader insight into the mechanisms of cell shape alteration and cellular growth in response to topography. For example in the case of grooved topographies, a groove size much larger than the cell size may promote cellular growth within the grooves. On the contrary, the same type of cells grown on top of narrow grooves could exhibit an entirely different response. 1.5 Interfacing nerve cells with biomaterials: The effect of surface topography Control of the outgrowth and patterning of nerve cells and neuroglial cells via surface topography is very important in a wide spectrum of neuroscience subfields, ranging from basic research to clinical applications. First, development of cell culture platforms with topographical cues can aid in simulating the cellular environment within a tissue in a simplified manner. In vitro studies with these platforms will help to address questions on how the nerve cells and the neuroglial cells respond to the cues of their environment, beyond the soluble biochemical ones. In other words, how do these cells respond either to the contact-mediated or haptotactic cues presented by the environment (e.g. ECM molecules) or to the physical characteristics of the environment per se (e.g. discontinuities, contours, structural patterns, morphological changes, etc.)? Topographical guidance of axons and neuroglial cells has been shown to be important during development and regeneration. Axons are guided by the morphological changes of the surrounding cells or extracellular space. For instance, during development axons can travel along pre-existing axon tracts (or fascicles) [22]. In peripheral nerve regeneration, following lesion-induced Wallerian degeneration, Schwann cells proliferate, migrate and organize themselves into longitudinal issive substrate for the growth of regenerating peripheral nerves [39,40]. Accordingly, oriented neuroglial cells and their ECM provide indispensable pathways for guided neuronal growth [41]. Another example implicates the 10 11 morphological plasticity of the astrocytes which have been reported to alter the neuronal environment and influence the astrocyte-neurons relationships [42]. Thus, studies focusing on the effect of topographical cues can provide important information on the complex processes of the axonal outgrowth and guidance and nerve-neuroglial interactions during development and regeneration. Second, topographical guidance cues can be implemented/exploited in introducing new conceptual approaches to address old questions in neurobiology. Such an example is the design of microfluidic platforms for the study of separate nerve cell compartments and their biochemical function in vitro have provided important information on the localization of the trophic effects of the ECM molecules to induce axon elongation [43 45]. Another example of cell patterning on interfaces with topography- induced guidance cues includes the design of (multi)electrode arrays platforms for the study of the dynamics of functional neuronal networks [46]. Topographically-guided patterning of neurons on a substrate with integrated electrical contacts for each neurons can ensure a better immobilisation compared to the chemically-guided patterning [47]. Last but not least, the body of knowledge obtained from the in vitro studies can then be translated into the clinical neurosciences, where implantable scaffolds with incorporated topographical guidance cues can be designed in order to promote -in synergy with the biochemical cues- the in vivo tissue regeneration. Parallel aligned but not randomly oriented polymeric fibers -stacked into 3D configuration to into polymeric constructs- have been shown to promote axonal regeneration in a model of peripheral nerve injury, suggesting that topographical cues can influence endogenous nerve repair mechanisms [48]. Moreover, topographical guidance cues, e.g. in the form of aligned nanofibers, could be used to guide neural processes in peripheral nerve electrodes [49]. The present review article aims at summarizing the existing body of literature on the use of artificial micro- and nanoscale surface topographical features to control neuronal and neuroglial cell morphology, outgrowth and neural network topology. A number of questions arise and will be addressed: Which are the geometrical parameters and dimensions of the topographies that influence the cellular responses ? How are neuronal cells influenced by artificial topographical cues? How are neuroglial cells influenced by artificial topographical cues? Which are the mechanisms for neuronal contact guidance by topographical cues? 11 In an attempt to address these issues we present the different studies based on their topographical model according to the classification of section 1.4, i.e. into continuous, discontinuous and random topographies. The effect of the various topographical cues from phenomenology to investigation of the underlying mechanisms - on the nerve cells and the neuroglial cells is presented in the chapter 2. and the chapter 3., respectively. Each chapter summarizes the studies in respective tables. The cell types, assays and materials that have been used in the presented studies are introduced in the next 12 section. Although the passive mechanical properties, such as the stiffness of the substrate are also very important in cell-biomaterials interactions, these are beyond the scope of the present review. There are already remarkable reviews on the effect of mechanical cues on cell processes, in general, and on neuronal development and functioning [3,50] and the reader is advised to these. 1.6 Cell types, assays and materials Cell types The effect of topography on neural phenotype cells has been studied on different cell types, ranging from the PC12 cell line of neuronal phenotype to the terminally differentiated primary nerve cells. PC12 cell line, which is a clonal derivative from a rat pheochromocytome, represents a neuronal cell model of terminally differentiating into neuron-like cells upon stimulation with nerve growth factor (NGF) [51]. When stimulated with NGF, PC12 cells recapitulate several steps of neuronal differentiation as they block proliferation, obtain the phenotype of sympathetic neurons -i.e. they begin to develop processes which are fine in caliber, varicose and fascicles-and become electrically excitable. Primary neurons isolated from both CNS and PNS, depending on the neuronal phenotype under study, have been cultured on different substrates. Brain neurons have been isolated either from hippocampus or the cortex in general, while sources for peripheral neurons can be dorsal root ganglia (DRG; for sensory neurons) and superior cervical ganglia (SCG; for sympathetic neurons). In particular, embryonic hippocampal cells have been extensively studied in vitro due to their ability to spontaneously establish a single axon from equally growing neurites enabling the study of axon initiation mechanisms with respect to topographical cues [36,52 56]. It is the most widely used system for examination of neuronal polarity [57]. Depending on the type of neurons or the developmental age, neurons need to be treated with proper growth factors that are critical for their survival (e.g. NGF for sensory neurons). Usually culture and growth of neurons on different substrates requires a specific protein coating of the surface under study, including collagen, laminin and Poly-D-lysine (PDL). 12 The effect of topography on cells of neuroglial phenotype has been studied mainly on dissociated primary Schwann cells [58 62], but there are few studies using Schwann cell lines, like RT4-D6P2T [63] and CRL-2765 [64]. Studies on the cellular response of astrocytes on micropatterned substrates remain limited [65,66]. Also, there is only one study reporting the culture of oligodendrocyes onto 13 micropatterned substrates [67]. Furthermore, complex models of coculture of neurons with neuroglial cells [58,59] and organotypic models of whole ganglion explants [48,58,59,68 71] on different topographical models have been reported. Such culture systems provide a more reliable representation of the in vivo environment than do single cell culture ones and have been used to assess the spatial interrelations of the different cell population with respect to the underlying topography [58,59]. Finally, some studies use topographical models in order to investigate the e ect of the topographical cues on astrocytes- neurons interactions in an attempt to control astrogliosis. For that reason, models of co-culture [35] or cortical cells containing both populations (e.g. hippocampal neurons and astrocytes) [34] have been used. Assays A great plethora of assays, ranging from cell number and orientation assays to more complex functional and differentiation assays, can be implemented to target various biological questions regarding the effect of topography on nerve cellular functions, as have been very thoroughly reviewed by D. Hoffman -Kim et al. [11]. Assays include immunofluorescence targeting of specific proteins, e.g. neurofilament, Tuj1, etc., or quantitatively evaluating the expression of them by analytical techniques, like western blotting. Morphological analysis of the cell shape and outgrowth but also of the exact position of the cells with respect to the patterned substrates can be assessed via Scanning Electron Microscopy (SEM). This will be very useful in order to establish a link between the topographical cues and the specific cell responses. Differentiation and neurite outgrowth of dissociated cells with respect to the specific topographies can be evaluated via the morphological changes of the cellular shape and the polarity of the cells, i.e. multiple-branched, bipolar or unipolar. Quantitative assays include evaluation of the number of cells forming neurites [72], the neurite density (i.e. the number of neurites per cell) [73 75], the percent of cells forming neurites [37], the length of the longest neurite [54,76] and the total neurite extension length [77]. In studies with whole ganglion explants, neurite outgrowth and Schwann cell migration have been evaluated via measuring the longest distance relative to the explant edge at which 13 the neurite has been outgrown and the Schwann cells has migrated, respectively [59,70,78]. In order to evaluate the various maturation and activation states of the Schwann cells, the expression of specific markers have been monitored, including myelin-specific genes, like MAG, P0, MBP which are significantly upregulated during Schwann cell myelination and NCAM-1 which is 14 associated with the immature Schwann cells [62]. Materials Different materials ranging from polymeric to metallic have been used to develop the various topographical models. Filaments or fibers have been made of synthetic polymers -like polypropylene (PP) [79], polycaprolactone (PCL) [80], poly-L-lactide (PLLA) [71], poly(acrylonitrile-co- methylacrylate (PAN-MA) [48,70] - and natural polymers (e.g. gelatin [63] and chitosan [81]). Polydimethylsiloxane (PDMS) [54,82], polystyrene (PS) [83] and poly(lactic-co-glycolic acid) (PLGA) [84] have been used for grooved patterns. Silicon has been used for the development of grooved substrates [85], pillars [36] and cones [37,58]. Gold has been used mostly for the fabrication of pillars [74]. Patterned substrates with random nanoroughness have been developed using silica nanoparticles [34] or nanoporous gold [34,86]. Moreover, the guidance effect of Schwann cells on neurons has been also exploited with the development of polymeric replicas of Schwann cell monolayer in random or aligned orientations [61]. However, these approaches are beyond the scope of this review paper and the reader is suggested in the review of Hoffman-Kim et al. [11]. 2 The in vitro effects of topography on nerve cell morphology, functions and network topology Topographic guidance of neurite outgrowth and axonal guidance in response to topographical cues has been in vitro investigated with different topographical models, in an attempt to shed light on the mechanisms of axon initiation, axonal branching and neuronal network formation. These studies are reviewed in the current chapter. Nerve cell responses including morphology, neurite outgrowth and differentiation- to various continuous, discontinuous and random topographies reported in the literature are summarized in the Tables 2, 3 and 4, respectively, where the various studies are presented with increasing topographical feature size. 2.1 The effect of continuous topographies on nerve cells 14 15 Topographic guidance of neurite outgrowth and axonal guidance Topographic guidance of neurite outgrowth and axonal guidance has been investigated in vitro using different culture substrates comprising micropatterned features at the submicron-to micron scale. Two types of geometries have mainly been reported. The first includes the anisotropic continuous geometry of alternating grooves. According to these studies, groove depth and width seem to be the critical parameters for axonal guidance and oriented neurite growth. The majority of studies have used microgrooves with a depth size ranging from 0.2 to 69 m. It was shown that as depth increased (i.e. from 400 to 800 nm), the percentage of hippocampal neurons growing parallel to the microchannels increased (Figure 2A) [54]. Cerebral embryonic neurite outgrowth on patterns of 2 m groove and 8 m depth were unaffected by the underlying topography. On the contrary, when the cells were grown on the 2 m deep patterns and the same groove/ridge width, neurites followed the orientation of the groove axis [87]. Remarkably, when the groove depth was shallow (less than 1100 nm), hippocampal neurons lost their parallel orientation tendency and grew perpendicular to the grooves [88]. The study of Rajnicek et al. emphasized also the effect of cell origin (i.e. developmental age and species) on the morphogenetic events of contact guidance. In particular, whereas Xenopus spinal cord neurites grew parallel to grooves independent of feature dimensions, rat hippocampal neurites regulated their direction of neurite growth depending on groove dimensions and developmental age (Figure 2B) [88]. In the studies where groove width was quite large (50-350 m), an interesting effect of groove depth was reported: axons turned at the edges of deep grooves (22-69 m) but not at the edges of shallow grooves (2.5-11 m) on PDL-coated PDMS micropatterned substrates. Specifically, in response to steps of depth h =22 69 m, the vast majority of axons appeared to be guided by surface topography by turning and either remaining inside the grooves or staying on the top surface. In contrast to this response, neurons on shallow grooves (2.5 and 4.6 m) disregarded the topographical steps and extended axons freely into and out of the grooves (Figure 2C). An intermediate response was observed for the neurites grown on substrates with groove depth of 11 m [82] was lost when neurons were grown on matrigel-covered substrates containing deep grooves, emphasizing the complexity of the contact guidance phenomenon; particularly when cells were challenged with competing growth options, stimulated by biochemical and topographical cues [82]. Figure 2: Effect of groove depth on neurite outgrowth. (A) SEM images of hippocampal cells on 15 16 cultured on 2 m-wide and 800 nm-deep (Aa) and 1 m wide and 400 nm deep (Ab) PDMS microchannels with immobilized NGF (0.11 ng/mm2), scale bar=10 m); ( c-d) Graph showing the angle distribution for axons growing on 400 nm-deep (Ac) and 800 nm-deep (Ad) microchannels with immobilized NGF (0.11 ng/mm2); Inset: schematic illustration of the analysis of axon length and angle (Reprinted with permission from [54]). (B) Phase contrast micrographs of E16 rat hippocampal neurons (Ba and Bb) and Xenopus spinal cord neurons (Bc and Bd), grown for 24 and 4 hours, respectively. Cells we nd Bc) or on grooves of 1 µm wide and 130 nm deep (Bb) or 320 nm deep (Bd); Scale bars: 50 µm (Ba-b) and 100µm (Bc-d). (Be-f) Graph representing the orientation responses of Xenopus spinal cord (Be) and rat hippocampal (Bf) neurites on grooved substrata. The direction of neurite growth is presented as the mean percentage of total neurites parallel (white bars) or perpendicular (black bars) to the groove direction ± sd. The groove width is indicated in the upper left of each graph (Reprinted with permission from [88]). (Ca-c) Cross-sectional schematic representation (top row), and corresponding a-tubulin immunofluorescence micrographs (bottom row) illustrating typical neurons cultured on 4.6 (Ca), 11 (Cb) and 69 (Cc) m-deep PDL-coated PDMS grooves; (Cd) Graph showing the percentage of axons that cross a step as a function of step height h. The inset shows a Hoffman DIC image of an axon from a murine cortical neuron (E13, 3 DIV); (Ce) Percentage of axons that overcome an 11 m-high step as a function of angle of approach. The graph shows that as the angle of approach, , increases axons are more likely to choose to bend the angle and disregard the step, bridging onto the next plane (Reprinted with permission from [82]). Feature, i.e. groove and ridge width is also critical. Most of the studies have used microgrooved substrates with a (groove) width ranging from 0.1 to 350 m. Based on the feature size scale, the studies can be classified into two categories: 1) subcellular (nanometer and single-micron) and 2) cellular (tens of microns) to supracellular (up to hundreds of microns) scale. Having usually the same size scale, the one of the two features (i.e. the groove or ridge) is kept constant and the effect of the other on the cell outgrowth is being investigated. In the case of subcellular groove width, oriented growth was observed on all types of patterns, although the 100 nm wide grooves seemed to be less efficient as compared to the patterns with larger widths, suggesting that submicron topographies- although too small to restrict growth- can also orient cellular alignment [83,89]. Ferrari et al studied the neuronal polarity of PC12 cells after NGF stimulation on cyclic- -copolymer grooves of 350 nm depth and 500 nm groove width with varying ridge width. It was shown that cells exhibited a shift from monopolar and bipolar to multipolar morphology, as the ridge width shifted from 500 nm to 2000 nm, with 1000 nm critical dimension [89]. In another study, differentiated PC12 cells exhibited an anisotropic growth on gratings of 200 nm depth and linewidth of 500 nm and 750 nm widths, although the alignment was more striking in the narrower grooves [83]. 16 17 In the cases of cellular to supracellular scale, as the width decreased, the growth orientation of the neurites was promoted in a direction parallel to channel walls and the complexity of neuronal architecture was reduced. Yao et al. used grooves and ridges of 5 and 10 m width (2-3 m deep) to study neuritogenesis in terms of neurite angle with respect to the groove axis [84]. While the PC12 cell somata were grown on top of the ridges, their neurites exhibited a preferential growth in the grooves rather on the ridges, having a signi cantly more parallel growth on small (5 m) than on larger groove sizes (10 m) (Figure 3B). Mahoney et al. studied the effect of grooved topography to study the effect on neuritogenesis with groove widths large enough for the PC12 cells to sit inside [75]. Using substrates of 11 m depth and ridge/groove width ranging from 20-60 m, it was shown that several properties of neurite growth (i.e. neurite polarity, the direction of neurite growth and the length, number and angle of neurites emerging from the cell soma) depended on microchannel width. Cells in narrower channel (20 30 m) extended one or two neurites parallel to the channel walls. Cells in wider channels (40 60 m) exhibited differential response, with i) the cells contacting the wall exhibiting bipolar shape and aligned, while ii) the ones not in contact with a wall exhibiting a perpendicular growth and a shift to multipolar morphology (Figure 3A) [75]. In the high-throughput study of Li et al. a great plethora of different topographical features of 1 m height replica-molded with PDMS have been developed and further used as cell culture platforms for hippocampal neuronal cells. These culture platforms enabled the study of multiple comparisons among the different geometrical parameters of the patterns [90]. Regarding the continuous topographical features, anisotropic features in the form of linear and circular gratings with constant groove width (i.e. 2 and 5 m, respectively) and varying ridge width (i.e. from 2 to 15 m) have been shown to promote axonal growth with a strong guidance response (Figure 3C), which was not the case for the cell outgrowth on the discontinuous topographical features (see Section 2.2) [90]. More specifically, the guided axons were following the transition edge between ridges and grooves. Furthermore, axonal length after 2DIV was 237 m and 214 m in the case of the linear-based and circular-based topographies, respectively (i.e. almost 60% and 48% longer, respectively than the neurons growing on flat substrate). Another parameter that has been evaluated was branching, which was limited in the neurons grown on the gratings. A combination of increased axonal length and reduced branching supports the strong guidance effect and the directional growth of the neurons on the gratings, compared to the other topographical patterns. Moreover, discriminating between dendrites (stained for MAP2) and axons (for Tau1), it was shown that, unlike axons, dendritic growth was found to be insensitive on almost all topographical features tested in this study [90]. 17 18 Though the vast majority of studies using grooved substrates report parallel alignment of neurites, other types of guidance have been reported as well [11]. When dissociated postnatal DRG sensory neurons were grown on grooved PDMS substrates of 50 m depth and groove/ridge width ranging between 30-100 m/30-100 m, respectively, the majority of DRG extended neurites parallel to the groove axis, and a subset of them extended neurites perpendicular to/bridging the groove pattern. The phenomenon termed as seems to be influenced by many parameters, including cell density, groove or ridge width (Figure 3D). Study of the dynamics of bridge formation via time- lapse microscopy revealed that bridges were formed as neurites extended from a neuron in a groove, contacted adjacent plateaus, pulled the neuron up to become suspended over the groove, and the soma translocated to the ridge [53]. Figure 3: The effect of feature width of cellular to supracellular size scale on neurite growth: (Aa- c) SEM images of microchannels; (Ad-f) PC12 cell neurite growth in microchannels of 20 (Ad), 40 (Ae) and 60 (Af) m width; (Ag) Table showing the average values for neurite growth of PC12 cells cultured in microchannels of different width (Reprinted with permission from [75]). (B) Orientation of neurite growth on laminin peptide- ; Parallel neurite growth on 5 m (Bb) and 10 m (Bc) grooves (Reprinted with permission from [84]). (C) Axonal outgrowth of hippocampal neuron cells on gratings with 10 m ridge and 2 m groove (Ca), angles with 5 m ridge and 5 m groove (Cb), circles with 10 m ridge and 5 m groove (Cc) after 2DIV, scale bar: 100 m; (Cd) Average axonal length of neurons growing on gratings, circles, dots and flat surface; (Ce) Dendrite branching index for neurons on patterns with gratings, circles, dots, and flat surface (Reprinted with permission from [90]). (Da) SEM image of DRG after 24 h culture on laminin-coated, micropatterned PDMS substrate with cell density of 125,000 cells/cm2, groove/ridge width of 50/70 m, respectively; (Db-c) Number of neurite bridges as a function of groove width (Db) and ridge width (Dc) (Reprinted with permission from [53]). The second model used to study the topographical guidance of neurite outgrowth and axonal guidance is that of anisotropic continuous geometry of parallel oriented fibers. Research in this area is focused mainly on the effect of parallel oriented submicron-to micron elongation and outgrowth (i.e. neuron. It was shown that sensory and respectively and sometimes crossed [59,91]. Studies with PC12 cells enable the study of neurite guidance in terms of neurite orientation, length and branching. Anisotropic of submicron made of PCL and PLLA, 18 19 bers of submicron diameters (e.g. in the range of 200-400 nm) have been shown to promote neurite guidance in terms of neurite length (compared to the random fibers) of PC12 differentiated cells. In the study of Genchi et al. PC12 differentiated cells exhibited higher neurite length on the parallel compared to the random collagen-coated pHB fibers, although cells on both cases exhibited the same number of neurites/cells (i.e. they were bipolar) [92]. Neurite outgrowth length was increased when the fibers were bound with protein, either physically adsorbed or covalently attached [59,77]. The majority of the studies with fibers, though, investigate the neurite outgrowth in response to the electrospun parallel fibers with a whole DRG explant, as presented in the next section. Directional growth of neurites from whole explants The effect of fiber orientation on neurite outgrowth guidance has been strongly emphasized by many studies using the model of whole DRG explant, which includes the outgrowth of axons and the migration of Schwann and other non-neuronal cells away from it. These studies investigate mainly the effect of the orientation of submicron-to micron (diameter in the range of 400-600 nm) electrospun bers made of different polymers on neurite elongation and outgrowth (i.e.length, etc.) from the ganglion. It was reported that neurite outgrowth on parallel aligned fibers preferentially extended along the direction of the fibers, while it was randomly distributed on the randomly oriented fibers. In order to investigate the n , Corey et al. have used a Fourier method (Fast Fourier Transform-FFT) to quantify the alignment of the PLLA fibers of ~500 ±300 nm diameter with three different orientations and the alignment of the outgrowing neurites. It was shown that neurite alignment was superior on the h compared to that on the intermediate and random alignment, suggesting that fiber alignment has an important effect on neurite alignment (Figure 4A) [69]. This effect of the topography was shown to be amplified with the incorporation of biochemical factors (e.g. ECM proteins or growth factors), either immobilized on the fibers or added in the solution (Figure 4B) [93]. Although the majority of the studies use fibers of subcellular size -and more specifically in the range of submicron-to-micron size- and report neurite alignment along the direction of the fibers, the study of Wen and Tresco used larger fibers in order to investigate the effect of fiber diameter on neurite outgrowth. Specifically, the study investigated the neurite outgrowth using a whole DRG explant model on parallel aligned fibers of supracellular (ie. 500-1000 m), cellular (30 m) and subcellular size (5 m) made of PP. They showed that neurite outgrowth (in terms of alignment, length and density) increased with decreasing diameter, suggesting that the effect of fiber diameter is very critical 19 for neurite outgrowth and alignment (Figure 4D) [79]. Interestingly, this increase in axial alignment was accompanied by a decrease in neurite fasciculation as the filament diameter decreased. More 20 specifically, while on 500 m- outgrowth on the smallest values diameters (i.e. on 5 and 30 m) was dense, spread out and uniformly distributed. This defasciculation as the filament diameter decreases was attributed to the increase of surface area giving more space for cells to adhere and grow. Precoating with ECM molecule amplified this response [79]. Figure 4: Effect of fiber alignment on neurite orientation. (Aa c): (Aa), intermediate (Ab) and low (Ac) alignment; (Ad f): FFT images from 256 x 256 pixel selections from the images in (Aa c), respectively. Yellow depicts greatest intensity, blue depicts least intensity. Note that narrower areas of higher intensity (yellow-orange) in the FFT images cor (Ag i): Neurites ical alignments as depicted in (Aa c); (Aj-l): FFT images from 512 x 512 pixel selections from the images in Ag-Ai (Reprinted with permission from [69]). (Ba-b) High- magnification confocal microscopy images of neurite morphology on random (Ba) and aligned (Bb) PLLA nanofibers; (Bc) Quantitative measurement of neurite outgrowth on the various nanofibers (Reprinted with permission from [93]). (C) The effect of various parameters on neurite outgrowth orientation: Fluorescence -b) DRG cultured on sca glass coverslips and then coated with di and 15(Cb) min; (Cc- sca olds were coated with PLL (Cc) and PLL and then laminin (Cd) from solutions with concentration 516 (Ce-f) DRG cultured on aligned nanofibers that were deposited on glass coverslips that have been precoated with PE 1(Ce) and 15 (Cf) min. The arrow in (Ca) implies the direction of alignment for the underlying stained with antiNF200 (Reprinted with permission from [80]). (D) Effect of fiber diameter on neurite bundles of varying diameter. Immunocytochemical staining for -III tubulin, S-100 and DAPI to label cell nuclei in blue. Neurite outgrowth and Schwann cell behavior on 500 m-diameter (Da,c) and 5 m-diameter (Db,d,e) (Reprinted with permission from [79]). In contrast to the majority of the studies focusing on the effect of the fiber orientation on the neurite outgrowth orientation, the remarkable report of Xie et al. sheds light on the impact of other fiber 20 parameters, i.e. fiber density, surface chemistry 21 surface property of the supporting substrate. By increasing the fiber density (via the increase of electrospinning time) the neurites tended to form bundles and grow perpendicular be also observed (Figure 4Ca- and neurite branching could a relatively high density, the projection of the neurites would switch from perpendicular to parallel outgrowth (Figure 4Cc-d) [80]. In this case, the neurites were observed to adhere well on the fibers and iting increased neurite length. When DRGs were grown on fibers deposited on glass substrate, which had been pre-coated with the protein-repellant Poly(ethylene glycol) (PEG), neurons tended to grow along a di (Figure 4Ce-f) [80]. Due to the cell repellenency property of PEG, neurites tended to form fascicles to avoid the interaction with the substrate. It is the first study that reports a perpendicular neurite outgrowth with respect to the alignment of the fibers, emphasizing the complexity of axonal guidance when multiple cues are presented to the neurons. 21 . f e R 2 2 e s n o p s e r r a l u l l e C / t n e m t a e r T n i e t o r P g n i t a o C / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i / l a i r e t a m o i B e r u t a e F n o i t a c i r b a F e u q i n h c e T e p y T s e i h p a r g o p o T s u o u n i t n o C - s l l e c e v r e n n o s e i h p a r g o p o t o n a n d n a - : a 2 e l b a T ] 8 8 [ t u b s e v o o r g o t l e l l a r a p w e r g s e t i r u e n s u p o n e X : s e i c e p s f o t c e f f E h t w o r g e t i r u e n f o n o i t c e r i d r i e h t d e t a l u g e r s e t i r u e n l a p m a c o p p i h : l a p m a c o p p i H e n i s y l - L - y l o P , w o l l a h s o t r a l u c i d n e p r e p t u b s e v o o r g e d i w , p e e d o t l e l l a r a p w e r g s e t i r u e n l a p m a c o p p i H : ) h t p e d ( e z i s e r u t a e f f o t c e f f E e n o N . s n o i s n e m i d e v o o r g n o g n i d n e p e d : s n o r u e n l a n i p S s u p o n e X c i n o y r b m E s n o r u e n d r o c l a n i p s l a p m a c o p p i h t a r d n a / ) 9 1 & 6 1 E ( s n o r u e n h t w o r g e t i r u e N r a l u c i d n e p r e p f o y c n e u q e r f e h T : e g a l a t n e m p o l e v e d f o t c e f f E 9 1 E . s v 6 1 E r o f r e h g i h s i s e t i r u e n l a p m a c o p p i h f o t n e m n g i l a . s n o r u e n l a p m a c o p p i h . s e n o w o r r a n h t d i w g n i s a e r c n i h t i w ] 3 8 [ d e c u d e r s a w t n e m n g i l a e t i r u e N : ) h t d i w ( e z i s e r u t a e f f o t c e f f E e n o N e h t s e g d i r / s e v o o r g w o r r a n n O : ) h t p e d ( e z i s e r u t a e f f o t c e f f E n o e l i h w , y l s u o e n a t l u m i s s e g d i r l a r e v e s f o p o t n o w e r g s n o x a d n a r e t e m a i d n o x a f o n o i t a l e r e h T : ) h t d i w ( e z i s e r u t a e f f o t c e f f E s e g d i r e l g n i s n o d n u o f e r e w s n o x a e h t s e g d i r / s e v o o r g r e d i w e c n a d i u g n o x a r o f l a i c u r c e b o t s m e e s h t d i w e g d i r / e v o o r g ] 5 8 [ s e v o o r g n i t o n d n a s g n i c a p s r o s e g d i r n o w e r g s n o x a l l A l e g i r t a M r e t f a n o i t a i t n e r e f f i D / s l l e c 2 1 C P F G N s G C S & s G R D / s t n a l p x e e l o h w e c n a d i u g n o x A ] 9 8 [ m o r f t f i h s d e t i b i h x e y g o l o h p r o m l l e c 2 1 C P : t n e m n g i l a f o t c e f f E e v o o r g e h t s a , y g o l o h p r o m r a l o p i t l u m o t r a l o p i b d n a r a l o p o n o m 2 2 s A F e h t f o n o i t a l u d o m r a l u g n a e v i t c e f f E . n o i s n e m i d l a c i t i r c m n . m n 0 5 7 o t p u g n i c a p s e v o o r g h t i w 0 0 0 1 h t i w m n 0 0 0 2 o t 0 0 0 5 m o r f d e t f i h s h t d i w e g d i r r o g n i c a p s r e t f a n o i t a i t n e r e f f i D / s l l e c 2 1 C P F G N m µ 4 r o 2 , 1 m n 0 0 1 . 1 - 4 1 m a e b n o r t c e l E : h t d i W y h p a r g o h t i l : h t p e D / z t r a u q d e s u F s e v o o r G m n 0 5 7 , 0 0 5 m n 0 0 2 : h t d i W - o n a n l a m r e h T / y h p a r g o h t i l t n i r p m i : h t p e D S P e r u t l u c e u s s i T s e v o o r G : h t p e D d e r e v o c A M M P s e v o o r G m : h t d i W 4 . 0 - 1 . 0 : g n i c a p S m 2 - 2 . 0 m n 0 5 3 : h t p e D m n 0 0 5 : h t d i W m n 0 0 0 2 - 0 0 5 : g n i c a p S t n i r p m i o n a N y h p a r g o h t i l / i S r e m y l o p o c l a m r e h / t n i r p m i o n a n y h p a r g o h t i l n i f e l o c i l c y C s e v o o r G . f e R 3 2 e s n o p s e r r a l u l l e C s e i h p a r g o p o T s u o u n i t n o C - s l l e c e v r e n n o s e i h p a r g o p o t o n a n d n a - / t n e m t a e r T g n i t a o C n i e t o r P / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i / l a i r e t a m o i B e r u t a e F n o i t a c i r b a F e u q i n h c e T e p y T : b 2 e l b a T . y l s u o e n a t l u m i s d e t n e s e r p e r e w s e u c h t o b n e h w n e e s s e z i s e r u t a e f e c a f r u s / s l l e c m n 0 0 8 , 0 0 4 d e z i l i b o m m i s l l e c f o e g a t n e c r e p e h t , d e s a e r c n i h t p e d s A : ) h t p e d ( e z i s e r u t a e f f o t c e f f E t s o m e h t d a h s e u c l a c i h p a r g o p o t e h T : F G N f o g n i d n i b c i f i c e p s f o t c e f f E s u o e n a t l u m i s e h t f o s s e l d r a g e r , n o i t a z i r a l o p n o t c e f f e d e c n u o n o r p y b d e s a e r c n i s a w h t g n e l n o x a , t s a r t n o c n I . F G N f o e c n e s b a r o e c n e s e r p s a w t c e f f e g n i c n a h n e n a h g u o h t , y h p a r g o p o t y b t o n d n a F G N d e r e h t e t d e s a e r c n i s l e n n a h c o r c i m e h t o t l e l l a r a p g n i w o r g d n a n o i t a z i r a l o P h t g n e l n o x a : h t d i W / F G N a c i l p e R g n i d l o m ] 4 5 [ n o t n e d n e p e d r e n n a m a n i e c n a d i u g t c a t n o c l e l l a r a p r o r a l u c i d n e p r e P l a p m a c o p p i h t a r 8 1 E : h t p e D h t i w S M D P s e v o o r G ] 4 8 [ r e g r a l n o n a h t l l a m s n o h t w o r g l e l l a r a p e r o M : e z i s e r u t a e f f o t c e f f E I e p y t n e g a l l o C / s l l e c 2 1 C P - 2 : h t p e D / A G L P s e v o o r G . s e v o o r g n i t o n d n a s e c a p s n o h t w o r g l a i t n e r e f e r P . s e z i s e v o o r g n i n i m a l r o r e t f a n o i t a i t n e r e f f i D S G K L L M L F P P ( e d i t p e p ) R T F G N n o i t a l b a r e s a L ] 7 8 [ - 8 d n a , e v o o r g - 2 n o s e t i r u e N : ) h t p e d ( e z i s e r u t a e f f o t c e f f E e n i s y l - L - y l o P k c i h c c i n o y r b m E : h t p e D h p a r g o h t i l o t o h P s e v o o r G ] 0 9 [ r e g n o l s a w h t g n e l l a n o x A . t c e f f e e c n a d i u g g n o r t S : e p y t e r u t a e f f o t c e f f E n i n i m a L l a p m a c o p p i H . s n r e t t a p p e e d - s e t i r u e n , y r a r t n o c e h t n O . d e t c e f f a n u e r e w h t p e d m 1 f o s n r e t t a p d e g d i r n o r u e n l a p m a c o p p i h h t w o r g e t i r u e N / ) 8 E ( : g n i c a p S m 0 2 : h t d i W m µ 8 s a w g n i h c n a r b e t i r d n e D . e t a r t s b u s t a l f e h t n o e n o e h t o t d e r a p m o c . t a l f e h t n o t a h t o t d e r a p m o c d e t i m i l h t w o r g t u o e t i r u e N ) C ( m / s n o r u e n d n a ) L ( m 2 5 : g n i c a p S m 5 1 - 2 m 2 r o 1 : h t d i W y / S M D P a c i l p e R g n i d l o m r a e n i L ) L ( & r a l u c r i C s g n i t a r g ) C ( ] 5 7 [ d e c u d e r s a w e r u t c e t i h c r a l a n o r u e n f o y t i x e l p m o c e h t d n a s l l a w l e n n a h c o t l e l l a r a p n o i t c e r i d a n i d e t o m o r p s a w s e t i r u e n e h t f o n o i t a t n e i r o h t w o r g e h t , d e s a e r c e d h t d i w s A : ) h t d i w ( e z i s e r u t a e f f o t c e f f E n e g a l l o C / s l l e c 2 1 C P r e t f a n o i t a i t n e r e f f i D 0 2 : h t d i W e v i t i s n e s o t o h P - o r c i M / e d i m i y l o p s l e n n a h c F G N m 0 1 : g n i c a p S - o t o h P y h p a r g o h t i l 3 2 . f e R 4 2 e s n o p s e r r a l u l l e C s e i h p a r g o p o T s u o u n ] 2 8 [ d e s a e r c e d h t p e d e v o o r g s a s p e t s e h t s s o r c o t d n u o f e r e w s n o x A : h t p e d f o t c e f f E ) m 1 1 t a e s n o p s e r e t a i d e m r e t n i ( c i n o y r b m e e n i r u M s n o r u e n l a c i t r o c e c n a d i u G n o x A / ) 4 1 E - 1 1 E ( : h t d i W - 0 5 - 5 . 2 g n i d l o m a c i l p e r : h t p e D / S M D P s e v o o r G n o i t c e r i d a n i s e t i r u e n d e d n e t x e G R D f o y t i r o j a m e h T : y p o r t o s i n a f o t c e f f E s e t i r u e n d e d n e t x e G R D f o t e s b u s a d n a , n r e t t a p e v o o r g e h t o t l e l l a r a p s e g d i r b f o r e b m u n e h t , s e s a e r c n i h t d i w s A : s n o i s n e m i d e r u t a e f f o t c e f f E ) 0 3 / 0 0 2 = e v o o r g w / g n i c a p s w r o f s e g d i r b f o s r e b m u n t s e h g i h e h T ( l l e w s a s e s a e r c e d . n r e t t a p e v o o r g e h t g n i g d i r b / o t r a l u c i d n e p r e p ] 3 5 [ n o i t a m r o f g n i g d i r b e t i r u e N d e t a i c o s s i D s G R D l a p m a c o p p i H / s n o r u e n g n i g d i r b e t i r u e N o t n g i l a o t d e n r u t t u b y l l a i d a r d e t u o r p s s e t i r u e n , s r e b i f d e n g i l a n O . t n e m n g i l a o t e v i t a l e r s r e b i f d e n g i l a n o d e s a e r c n i h t g n e l e t i r u e n d n a , t c a t n o c n o p u s r e b i f ] 9 6 [ r e b i f f o n o i t c e r i d e h t n i d e t a g n o l e e r e w a i l g n a G : y p o r t o s i n a f o t c e f f E n e g a l l o C h t w o r g t u o e t i r u e N G R D t a r 5 1 E / s t n a l p x e s r e b i f e t a i d e m r e t n i d n a m o d n a r m n 4 2 5 ~ g n i n n i p s o r t c e l E m : h t d i W 0 0 1 - 0 3 : g n i c a p S - 0 3 : r e t e m a i D / A L L P s r e b i F g n i d l o m a c i l p e R : h t p e D / S M D P s e v o o r G . s r e b e h t h t i w l e l l a r a p n i ] 9 5 [ d e t a t n e i r o y l e g r a l s a w n o i t a g n o l e e t i r u e n f o n o i t c e r i d e h T : y p o r t o s i n a f o t c e f f E o N h t w o r g t u o e t i r u e N / s t n a l p x e n e g a l l o C L C P / / d n e l b g n i n n i p s o r t c e l E G R D k c i h c 0 1 E : r e t e m a i D d n a L C P s r e b i F ] 3 9 [ d n a s r e b i f d e n g i l a f o e c n a d i u g e h t d e w o l l o f s e t i r u e N : y p o r t o s i n a f o t c e f f E n i n i m a L / s t n a l p x e G R D 5 - 4 P : r e t e m a i D / A L L P s r e b i F ] 8 4 [ e h t n o s G R D e h t m o r f h t w o r g t u o e t i r u e n f o y t i r o j a m e h T : y p o r t o s i n a f o t c e f f E . s r e b i f d e n g i l a e h t o t l e l l a r a p , y l l a n o i t c e r i d i n u d e d n e t x e m l i f r e b i f d e n g i l a s r e b i f d e t a o c n u e h t o t d e r a p m o c s e u c l a c i m e h c o i b d e z i l i b o m m i h t i w s r e b i f o N d e h t n o h t w o r g t u o e t i r u e n r e h g i h y l t n a c i f i n g i S : s e u c l a c i m e h c o i b f o t c e f f E e z i l i b o m m i s t n a l p x e G R D t a r 1 P h t w o r g t u o e t i r u e N / m n 0 0 6 - 0 0 4 : r e t e m a i D g n i n n i p s o r t c e l E / - A M N A P ] 0 7 [ n i t c e n o r b i f d e c n e u l f n i s r e b i f - A M N A P d e n g i l A : y p o r t o s i n a f o t c e f f E 4 2 o t d e r a p m o c n o i t a m r o f k r o w t e n n i t c e n o r b i f d e n g i l a d e t o m o r p d n a , n o i t u b i r t s i d / s t n a l p x e G R D 1 P h t w o r g t u o e t i r u e N : r e t e m a i D / - A M N A P m 8 . 0 g n i n n i p s r o t c e l E s r e b i F s r e b i F s r e b i f m o d n a r n o t a h t o t d e r a p m o c g n i h c n a r b o n o t e l t t i l d e t i b i h x e F G F h t w o r g t u o e t i r u e N g n i n n i p s o r t c e l E / t n e m t a e r T n i e t o r P g n i t a o C - D - y l o P e n i s y L i t n o C - s l l e c e v r e n n o s e i h p a r g o p o t o n a n d n a - / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i / l a i r e t a m o i B e r u t a e F n o i t a c i r b a F e u q i n h c e T e p y T : c 2 s e l b a T 5 2 o t s r e b i f e h t g n o l a d e d n e t x e s n o x a n w o r g t u O . s m l i f - A M N A P h t o o m s . f e R e s n o p s e r r a l u l l e C / t n e m t a e r T n i e t o r P g n i t a o C / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i / l a i r e t a m o i B e r u t a e F n o i t a c i r b a F e u q i n h c e T e p y T ] 8 6 [ d e y a l p s i d s e c i r t a m d e n g i l a n o n w o r g s e t i r u e N : y p o r t o s i n a f o t c e f f E s t n a l p x e G R D t a r 6 1 E : r e t e m a i D e n o n a x o i d y l o P s r e b i F s m l i f h t o o m s o t d e r a p m o c t n e t x e r e t a e r g y l t n a c i f i n g i s s e i h p a r g o p o T s u o u n i t n o C - s l l e c e v r e n n o s e i h p a r g o p o t o n a n d n a - : d 2 e l b a T ] 9 7 [ e h t f o s i x a g n o l e h t g n o l a w e r g s e t i r u e N : y p o r t o s i n a f o t c e f f E . n o i t a t n e i r o r e b i f g n i y l r e d n u e h t f o t a h t s c i m m i t a h t y t i l a n o i t c e r i d d n a y l t s u b o r e r o m w e r g s e t y c o r t s a f o e t a r t s b u s a n o d e r u t l u c s G R D x i r t a m e e r f - a i l g a n o n w o r g n e h w n a h t s e s s e c o r p r e g n o l d e d n e t x e r o n i t c e n o r b i F / ) e t a r t s b u s l a i l g s a ( h t w o r g t u o e t i r u e N s e t y c o r t s a 3 P - 2 ~ / s t n a l p x e G R D : r e t e m a i D s e l d n u b t n e m a l e h t f o s i x a g n o l e h t o t l e l l a r a p n o i t c e r i d a n i d e t o m o r p s a w h t w o r g t u o e t i r u e n e h t , d e s a e : r e t e m a i d f o t c e f f E e n o n r e t e m a i d t n e m a l i f e h t f o s s e l d r a g e r , s e l d n u b t n e m a l r o n i n i m a L h t w o r g t u o e t i r u e N , 0 0 2 , 0 0 1 , 0 3 , 5 g n i n n i p s o r t c e l E / l a m r e h T n o i s u r t x e / P P t n e m a l i F s e l d n u b ] 1 7 [ h t w o r g t u o e t i r u e n d e d i u g - y h p a r g o p o T : y p o r t o s i n a f o t c e f f E n i d e s a e r c n i y l t n a c i f i n g i s s a w h t w o r g t u o e t i r u e N : g n i t a o c f o t c e f f E e n i s y L L y l o P r o n i n i m a L & h t w o r g t u o e t i r u e N / n o i s u r t x e t l e M s t n a l p x e e l o h w G R D : r e t e m a i D / A L L P s t n e m a l i F e g d e g n i d a e l l l e c n n a w h c S e h t t s a p , n i n i m a l f o e c n e s e r p e h t n o i t a r g i m l l e c n n a w h c S ] 0 8 [ r a l u c i d n e p r e p d n a l e l l a r a p h t o B : g n i t a o c d n a y t i s n e d r e b i f f o t c e f f E s r e t e m a r a p r e b i f e h t n o g n i d n e p e d d e v r e s b o e r e w e c n a d i u g t c a t n o c g n i t a o c r o n i n i m a L / s t n a l p x e G R D 8 E h t w o r g t u o e t i r u e N 0 0 0 . 3 0 0 1 g n i n n i p s o r t c e l E : y t i s n e d r e b i F / L C P s r e b i F G E P h t i w m m / s r e b : - A M N A P ; 1 y a D l a t a n t s o P : 1 P ; r o t c a f h t w o r g e v r e N : F G N ; t n e m a l i f o r u e N : F N ; 8 1 y a D c i n o y r b m E : 8 1 E ; n o i l g n a G t o o R l a s r o D : G R D : s n o i t a i v e r b b A * ; e d i t c a l - L - y l o p : A L L P ; ) d i c a c i l o c y l g - o c - c i t c a l ( y l o p : A G L P ; e n a x o l i s l y h t e m i d y l o P : S M D P ; e n o t c a l o r p a c y l o P : L C P ; e t a l y r c a l y h t e m - o c - e l i r t i n o l y r c a ( y l o p e v a h s r o h t u a e h t f i n e v e ( e v o o r g e h t o t t c e p s e r h t i w d e t n e s e r p e r a s n r e t t a p - d e g d i r / - d e v o o r g e h t , n o s i r a p m o c f o e s a e d n a y t i r a l c f o s n o s a e r r o F : t n e m m o C 5 2 . ) s e g d i r e h t o t t c e p s e r h t i w m e h t d e t r o p e r n o c i l i S : i S ; n o i l g n a G l a c i v r e C r o i r e p u S : G C S ; e n e r y t s y l o P : S P ; e n e l y p o r p y l o P : P P ; ) e t a l y r c a h t e m l y h t e m ( y l o P : A M M P 26 2.2 The effect of discontinuous topographies on nerve cells Effect on neuron outgrowth and neurite outgrowth orientation In general, the effect of discontinuous isotropic topographies in the form of micro- or submicron-scale pillars, posts, cones and holes have been investigated. Generally, the feature height ranged from 1-10 m. According to these studies, feature spacing was the critical parameter for oriented neurite outgrowth. As the spacing between pillars increased, the fidelity of alignment decreased. Hippocampal neurons were cultured on square-shaped silicon pillars of 0.5 or 2 m diameter and spacing ranging from 0.5 to 5 m. The neurites tended to span the smallest distance between pillars, aligning either at 0° or 90°, with the highest alignment with the larger pillars at the smallest spacing. As the spacing between pillars increased, the fidelity of alignment decreased, and at surface (Figure 5A) [36]. In another study, hippocampal neurons have been cultured on conical posts of 10-100 m diameter and respective height of 1/10 of the diameter and the edge-to-edge spacing ranging from 10- 200 m [55]. In this study, diameter and spacing were found to be two critical parameters in the neuron outgrowth process (Figure 5B). Neurite processes on surfaces with smaller features and smaller intercone spacings were aligned and connected in straight lines between adjacent pillars. However, as microcone diameters started to increase, neurites wrapped around the post they were processes on surfaces with more than 200 m inter-feature spacing exhibited random outgrowth and wrapping. Therefore, as the feature size and spacing was increased, a transition from aligned to random growth occurred. The importance of interpillar spacing (or pitch) for cell polarization and alignment has been strongly emphasized in the study of Bucaro et al. using silicon arrays of pillars with 10 m height and varying pitch (i.e. from 0.8 to 5 m). PC12 cells exhibited polarization and alignment on all of the substrates with an interpillar distance in the range of 1.6 to 2 [94]. Interestingly, the same morphological cell response was to be seen on Si (Young's modulus of 180 GPa) and epoxy (Young's modulus of 1-2 GPa) substrates, suggesting that interpillar spacing may have a more the extent of cell polarization [94]. Square arrays of silicon pillars exhibiting diameter of 461 nm, spacing of 339 nm and height of 1200 nm could support PC12 cell differentiation towards a highly branched neurite phenotype after treatment with NGF [73]. In the case of the interpillar distance at nanoscale, PC12 cell differentiation has been shown to be inhibited or limited. Specifically, gold ects only 26 27 nanopillars of PC12 cells exhibiting fewer and shorter neurites compared to the cells on smooth substrates [74]. Kang et al. used recently a model of vertically grown silicon nanowires (vg-SiNWs) of 72 ± 8 2 density. An accelerated polarization of embryonic hippocampal neurons at early timepoints was shown compared to the growth on the control substrates [95]. The neuronal networks formed were functional as was evaluated by somatic calcium imaging. The same group used also another topographical model of silica bead monolayer and pitch ranging from 700 to 1800 nm [96]. Increased hippocampal neurite length at early timepoints was reported with increasing pitch compared to the control substrates. Figure 5: Effect of discontinuous topographies on neurite outgrowth. (Aa) Schematic representation of the different pillared topographies (with varying inter-pillar distances); (Ab) Fluorescence images demonstrate the effects of pillar width of 2.0 µm at increasing pillar gap sizes on neurite outgrowth; (Ac) SEM images of hippocampal neurons on the different substrates. SEM images show that most axons (white arrowhead) and dendrites (black arrowhead) that grew out from the cell body tended to follow the tops of pillars with widths of 2.0 µm and gaps of 1.5 µm (Ac). Processes grew at the base of pillars with widths of 0.5 µm and gaps of 1.5 µm (Ad), as well as on pillars with widths of 2.0 µm and gaps of 4.5 µm (Ae). Scale bar = 20 µm. Growth cones (black arrows) on pillars with widths of 2.0 µm and gaps of 1.5 µ indicative of rapid growth (Af). In contrast, growth cones (black arrows) on pillars with widths of 2.0 µm and growth (Ag), Scale bar (Af-g) = 10 µm. An increase in process formation and bundling was observed on pillars with widths of 2.0 µm and gaps of 1.5 µm (Ah). Smaller processes arose from a main process at consecutive pillars, Scale bar (Ah) = 2 µm (Reprinted with permission from [36]). (Ba) PDMS conical post array captured using tapping mode Atomic Fluorescence Microscopy (AFM); (Bb-e) Optical micrographs of hippocampal neurons plated on glass substrates patterned with conical posts of PDMS on the pillared arrays of 10 m diameter/10 m spacing (Bb); 20 m diameter/ 40 m spacing (Bc), 50 m diameter/100 m spacing (Bd) and 100 m diameter/200 m spacing (Be); scale bar: 90 m (Reprinted with permission from [55]). (C) Axonal outgrowth of hippocampal neuron cells on 10x10 m2 dots with 10 m spacing (Ca), triangles with 5 m space in Y direction (Cb) and flat surface (Cc) after 2DIV; scale bar: 100 m [90]. (D) SEM images of PC12 cells grown (Da) on high-aspect-ratio Si NPs (r = 100 nm, and (Db) on an OG142 polymer micropillar array (r = 750 nm, h =1 permission from [94]). (Ea-f): SEM images in tilted view (a-c) and top view (d-f) of micropatterned Si substrates of low (Ea,d), medium (Eb,e) and high (Ec,f) roughness; (Eg-i) positive sympathetic neurons grown on low (g), medium (h) and high (i) roughness micropatterned Si substrates 27 28 for 6 days, scale bar: 150 m. (Reprinted with permission from [58]). Li et al. introduced a large library of continuous and discontinuous microscale size features ranging from 2 to 15 m in size and varying pitch (from 0.5 to 20 m) [90]. Regarding discontinuous patterns, both isotropic (i.e. in the form of rectangles and dots) and anisotropic (i.e. in the form of semi- circles and triangles) of 1 m height and with varying feature characteristics (i.e. width, diameter) and interfeature distances were replica molded with PDMS polymer have been developed and further used as cell culture platforms to grow hippocampal neuronal cells [90]. Regarding axonal guidance, the axons were only guided along a few dots or triangles before losing tracks after a short distance (Figure 5C). After 2DIV, the axonal length on the discontinuous patterns was smaller compared to that on the continuous topographical features (183 m - although still longer than the one on flat). ranching was lower in all the other cases of the discontinuous patterns compared to that on the flat surface [90]. We have recently explored the use of a discontinuous anisotropic surface micropattern for controlled directional outgrowth of neuronal cells and respective intracellular networks. Micropatterned surfaces had been fabricated via ultra-short pulsed laser processing on silicon and comprised arrays of microcones (MCs) that were either of elliptical or arbitrary-shaped cross-sections (Figure 5Ea-f) [58]. Although very few sympathetic neurons could micropatterned Si substrates equally supported extended neuronal outgrowth, depicting the importance of surface roughness over nerve cell outgrowth and network formation. Furthermore, neurons on the silicon MCs of elliptical cross-section and parallel orientation (medium and high roughness substrates) exhibited a preferential parallel alignment, while neurons on the MCs of arbitrary cross-section and random orientation (low roughness) formed a highly branched network exhibiting no preferred orientation (Figure 5Eg-i). Apparently, the preferential orientation of neuronal processes matched the direction of the major axis of the elliptical MCs, suggesting a dependence of the axonal outgrowth pattern on the underlying topography [58]. This topography-induced guidance effect was also observed in the more complex cell culture system of whole DRG explant, suggesting that even a discontinuous topographical pattern can promote axonal alignment, provided that it hosts anisotropic geometrical features, even though the sizes of those range at the subcellular lengthscale. The same topographical model has been used to study the PC12 differentiation after treatment with NGF. It was shown that, unlike low and medium roughness surfaces, highly rough ones exhibiting large distances between MCs did not support PC12 cell differentiation, although cells had been stimulated with NGF [37]. 28 29 2.3 The effect of random topographies on nerve cells Beyond the well-defined microscale topographies, there is an increasing interest on more random (or stochastic) topographies that resemble better the nanotopographies of the in vivo extracellular environment. Blumenthal et al. have used assembly of monodispersed silica colloids of increasing size in an attempt to mimic the topography from the level of receptor clusters to ECM features. Among the different substrates (i.e. corresponding surface roughness ranged from 12 to 80 nm), PC12 cells after treatment with NGF exhibited increased differentiation - in terms of polarity and neurite length- and associated functional traits on a specific Ra of 32 nm (Figure 6A). Remarkably, primary hippocampal neurons also responded to roughness in a manner similar to PC12 cells exhibiting prominent, axon-like polarized structures on exactly the same roughness value of 32 nm. Furthermore, it was shown that this stochastic nanoroughness can modulate the function of hippocampal neurons and their relationship with astrocytes. Specifically, although on the rough substrates neurons were predominantly found associated with astrocytes, for a critical roughness value of Ra of 32 nm, neurons were dissociated from astrocytes and continued to survive independently even up to 6 weeks (Figure 6B) [35]. Interestingly, this optimal roughness regime coincides with the roughness value of the astrocyte surface (i.e. Rq = 26 28 nm) measured by AFM; this could explain the dissociation of the neurons from the astrocytes at this specific roughness regime. In two other studies, neuronal adhesion and survival were affected in a unimodal manner when cultured on nanotextured silicon. In the first study, an intermediate value of Ra = 64 nm promoted an optimal response of embryonic primary cortical neurons and both higher and lower roughness reduced this response (Figure 6C) [97]. In the study of Fan et al., substantia nigra neurons survived for over 5 days with normal morphology and expressed neuronal tyrosine hydroxylase when grown on surfaces with a Ra ranging from 20 to 50 nm. However, cell adherence was adversely affected on surfaces with Ra less than 10 nm and above 70 nm (Figure 6D) [98]. Though nanoscale materials have received a great deal of attention in recent years, it is not yet clear the role they could play in directing neuronal growth for tissue engineered applications [11]. Cho et al. used a model of aluminium oxide concave nanostructures with a pitch of 60 (small) and 400 (large) nm, which were fabricated by electropolishing and electrochemical anodization. The large pitch was reported to exert an accelerating effect on hippocampal neuronal polarization or axon formation [99]. Figure 6: Effect of random topographies on neurite outgrowth. (A) PC12 cells on nanorough substrates. (Aa) 29 30 AFM of SNP modified substrates with the corresponding surface roughness values; (Ab-d) Morphology of PC- 12 cells on Rq= 3.5 nm (Ab), Rq= 32 nm (Ac), and R= 80 nm (Ad) visualized by staining for F-actin; (Ae-f) Impact of nanoroughness on PC12 polarization as assessed by determining number of neurites per cell (Ae) and neurite length (Af) (Reprinted with permission from [35]). (B) Neuron astrocyte interaction on smooth glass (Ba) and substrate of R= 32 nm (Bb). Astrocytes and neurons were visualized using antibody against GFAP (blue) and MAP-2 (red), respectively. Quantification of neuron astrocyte association in short-term (5 d) (Bc) and long-term (6 wk) (Bd) cultures; (Be-f) Calcium-sensitive FURA-2 imaging in hippocampal neurons on smooth glass substrates and surfaces with R of 32 nm: (Be) change in intracellular calcium level as assessed by FURA-2 intensity and (Bf) rate of depolarization as determined by the slope of the depolarization portion of the curve (Reprinted with permission from [35]). (C) Dependence of cell viability (6 days after inoculation) on Ra of Si wafers [97]. (D) Dependence of cell viability (5 days after inoculation) on the Ra of Si wafers (Reprinted with permission from [98]). 30 . f e R ] 4 9 [ ] 6 3 [ 1 3 2 t a t n e m n g i l a d n a n o i t a z i r a l o p d e t i b i h x e s l l e C : g n i c a p s f o t c e f f E t n e m t a e r t e r P / s l l e c 2 1 C P , m n 0 0 4 : r e t e m a i D / i S l a t s y r c e l g n i S s r a l l i P e s n o p s e r r a l u l l e C / t n e m t a e r T n i e t o r P g n i t a o C / e p y t l l e C y a s s a l l e C s n o i s n e m d e r u t a e F i s e i h p a r g o p o T s u o u n i t n o c s i D - s l l e c e v r e n n o s e i h p a r g o p o t o n a n d n a - / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T e r u t a e F e p y T : a 3 e l b a T t s e h g i h e h t h t i w , ° 0 9 r o ° 0 t a r e h t i e g n i n g i l a , s r a l l i p n e e w t e b e h t s A . g n i c a p s t s e l l a m s e h t t a s r a l l i p r e g r a l e h t h t i w t n e m n g i l a t n e m n g i l a f o y t i l e d i f e h t , d e s a e r c n i s r a l l i p n e e w t e b g n i c a p s e c n a t s i d t s e l l a m s e h t n a p s o t d e d n e t s e t i r u e n e h T : g n i c a p s f o t c e f f E e n i s y l - L - y l o P l o n a h t e h t i w m o d n a r h t o . g n i p p a r w d n a h t w o r g t u o . d e v r e s b o s a w t s o p e h t : g n i c a p s f o t c e f f E t h g i a r t s n i d e t c e n n o c d n a d e n g i l a s e s s e c o r p , s g n i c a p s r e l l a m s d n a e l g n i s a d e w o l l o f y l t s o m d n a s r a l l i p t n e c a j d a n e e w t e b s e n i l . n o i t c e r i d r a l u c i d n e p r e p e h t n i g n i h c n a r b y l l a n o i s a c c o y b n o i t c e r i d d n u o r a g n i p p a r w e t i r u e n , d e s a e r c n i s r e t e m a i d e r u t a e f s a , r e v e w o H ] 5 5 [ s e r u t a e f r e l l a m s h t i w s e c a f r u s n O : ) r e t e m a i d ( e z i s e r u t a e f f o t c e f f E e n i s y l - D - y l o P . e c a f r u s t a l f a n o d n u o f t a h t o t r a l i m i s s a w l a p m a c o p p i H / s n o r u e n l a n o x A & e t i r d n e D - 5 . 0 : g n i c a p S - 8 0 . : g n i c a p S 1 : 2 1 : o i t a r t c e p s A n o i e v i t c a e r p e e d g n i h c t e / n o c i l i S y h p a r g h o t i l o t o h P & y h p a r g o h t i l y g o l o h p r o m l l e C r e p p e t s V U s s e c o r p n o r u e N l a p m a c o p p i H / s n o r u e n h t w o r g t u o h t w o r g t u o ) r e t e m a i d ( 0 1 / 1 ~ : g n i c a p S 0 0 1 - 0 1 - 0 1 t f o s s s e l r e t s a M y h p a r g o h t i l / s s a l g : r e t e m a i D n o s t s o p S M D P : t h g i e H d o h t e m r e f s n a r t - e r a u q S d e p a h s s r a l l i P l a c i n o C s t s o P . m e h t g n i d n e b d n a n o g n i l l u p s a h c u s , s e r i w o n a n e h t h t i w ] 0 0 1 [ s e c a f r u s l o r t n o c o t e v i t a l e r l a v i v r u s l a n o r u e N : y h p a r g o p o t f o t c e f f E s n o i t c a r e t n i x e l p m o c t n e w r e d n u s n o r u e n e h t d n a , d e s a e r c n i s a w h t w o r g l a n o r u e N / s n o r u e n G R D D V C / e d i h p s o h p m n 0 5 : r e t e m a i D m u i l l a G s r a l l i P ] 0 9 [ , s n r e t t a p t o d d n a r a l u c r i c e h t f o e s a c e h t n I : e p y t e r u t a e f f o t c e f f E . d e t i m i l e r e w g n i h c n a r b d n a h t g n e l n o x a h t o b n i n i m a L d n a e n i s y l y l o P e t i r u e N / s n o r u e n m 5 1 - 2 g n i d l o m a c i l p e R s e r a u q s & l a p m a c o p p i H : r e t e m a i D r o h t d i W / S M D P , s d i r g , s t o D h t w o r g t u o m 0 2 - 5 0 . : g n i c a p S ] 5 9 [ s n o r u e n e h t f o n o i t a z i r a l o p d e t a r e l e c c A : y h p a r g o p o t f o t c e f f E e n i s y l - L - y l o P l a p m a c o p p i h t a r 8 1 E : r e t e m a i D ; m n 8 ± 2 7 / i S d e r e v o c - u A f o h t w o r g ( D V C r o j a m d e t a g n o l e y l e m e r t x e , e l g n i s a d e m r o f s W N S i n o s n o r u e N & h t w o r g t u o l a n o i t c n u f f o n o i t a m r o F . s e t i r u e n r o n i m n a h t r e i l r a e e t i r u e n s k r o w t e n l a n o r u e n s y a s s a y t i l a n o i t c n u F ) g n i g a m i m u i c l a c ( . s t n i o p e m i t y l r a e t a p i l s r e v o c n o h t w o r g e h t o t d e r a p m o c e t i r u e N / s n o r u e n 0 1 o t 7 : h t g n e L & ) s e r i w o n a n 2 : y t i s n e D y h p a r g o h t i l o t o h P ) n r e t t a p e n i l ( s e r i w o n a N . f e R ] 4 7 [ ] 1 0 1 [ ] 8 5 [ ] 7 3 [ f o b e w e t a r o b a l e n a d e m r o f s r a l l i p e h t n o s n o r u e n e h t , e t a r t s b u s t a l f e h t . s t n e m e l e l a i l g f o e c n e s b a e h t n i s e s s e c o r p c i m s a l p o t y c n o d e v i v r u s s n o r u e n w e f a y l n o h g u o h t l A : s s e n h g u o r e c a f r u s f o t c e f f E n i a r b e s u o m c i n o y r b m E / s l l e c h t w o r g t u o l l e C : t h g i e H m 2 - 1 ~ m n 0 4 ~ : g n i c a p s m n 0 0 2 : r e t e m a i d s e r o P e h t o t n o d e t n e i r o l e l l a r a p e r e w s n o r u e N : s s e n h g u o r e c a f r u s f o t c e f f E o t n o n o i t a t n e i r o m o d n a r a d e t i b i h x e y e h t e l i h w , e p a h s l a c i t p i l l e f o s C M n e e s e b o t s a w e s n o p s e r e c n a d i u g e m a s e h T . e p a h s y r a r t i b r a f o s C M e h t . s t n a l p x e G R D e l o h w e h t o t h t i w d e t a e r t n e e b d a h s l l e c 2 1 C P h g u o h t l A : s s e n h g u o r e c a f r u s f o s e t a r t s b u s s s e n h g u o r - h g i h e h t n o e t a i t n e r e f f i d o t d e l i a f y e h t t c e f f E , F G N n e g a l l o C n e g a l l o C s C M n e e w t e b s e c n a t s i d e g r a l g n i t i b i h x e . s g n i t a o c l a c i m e h c d e t a i c o s s i D c i t e h t a p m y s n o i t a m r o f k r o w t e n & G R D & s n o r u e n / s t n a l p x e e l o h w / s l l e c 2 1 C P n o i t a m r o f k r o w t e n o i t a i t n e r e f f i D F G N r e t f a ] 6 9 [ w o l e h t r o p i l s r e v o c e h t n o s l l e C : y h p a r g o p o t / s s e n h g u o r f o t c e f f E a m s a l P t a r 8 1 E h t i w d e s a e r c n i e t i r u e n t s e g n o l e h t f o h t g n e l e h T : r e t e m a i d d a e b f o t c e f f E . ) I V D 1 t a m 0 7 ~ ( m n 0 0 0 1 t a d l o h s e r h t a g n i t i b i h x e r e t e m a i d d a e b e h t ) I V D 1 t a ( s e t i r u e n d e t u o r p s y l e r a b d a h r e y a l o n o m d a e b r e t e m a i d & t n e m t a e r t l a p m a c o p p i H - D - y l o P e n i s y L s n o r u e n h c t i p m 6 . 2 f o ( n o i t c e s m t h g i e h m 3 . 1 d n a - s s o r c l a c i t p i l l e ) i i . 5 6 / 7 . 4 f o ( n o i t c e s m 6 . 8 / 7 / 3 d n a h c t i p - s s o r c y r a r t i r b r a ) i f o s e n o c o r c i M : r e t e m a i d d a e B m n 0 0 8 1 - 0 0 7 ) t h g i e h 2 3 s e i h p a r g o p o T s u o u n i t n o c s i D - s l l e c e v r e n n o s e i h p a r g o p o t o n a n d n a - e s n o p s e r r a l u l l e C r o d e t i b i h n i s a w h t w o r g t u o e t i r u e n l l e c 2 1 C P : e z i s e r u t a e f f o t c e f f E s e t a r t s b u s h t o o m s e h t o t d e r a p m o c s e r o p & s r a l l i p e h t n o d e t i m i l / t n e m t a e r T n i e t o r P g n i t a o C / e p y t l l e C y a s s a l l e C - L - y l o P e n i s y L n o i t a i t n e r e f f i D / s l l e c 2 1 C P s n o i s n e m d e r u t a e F i / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T e r u t a e F e p y T b 3 s e l b a T F G N r e t f a m n 0 7 : g n i c a p s n o i t i s o p e d m n 0 0 2 : r e t e m a i d s r a l l i P / d l o G - o r t c e l E & s r a l l i P s e r o P n o c i l i S : i S ; e n a x o l i s l y h t e m i d y l o P : S M D P ; r o t c a f h t w o r g e v r e N : F G N ; n o i l g n a G t o o R l a s r o D : G R D ; n o i t i s o p e D r o p a V l a c i m e h C : D V C : s n o i t a i v e r b b A * i S l a t s y r c e l g n i S d n o c e s o t m e F / r e s a l d e s l u p g n i s s e c o r p i S l a t s y r c e l g n i S d n o c e s o t m e F / r e s a l d e s l u p g n i s s e c o r p - l a c i n o C d e p a h s s r a l l i p s e n o c o r c i M ) s C M ( a c i l i S r e y a l o n o m d a e B ] 7 9 [ n a e . i , e s n o p s e r l l e c l a d o m i n U : s s e n h g u o r o n a n f o t c e f f E l a m i t p o n a d e t o m o r p m n 4 6 = a R f o e u l a v e t a i d e m r e t n i s i h t d e c u d e r s s e n h g u o r r e w o l d n a r e h g i h h t o b d n a , e s n o p s e r o t d e u n i t n o c d n a s e t y c o r t s a m o r f d e t a i c o s s i d e r e w s n o r u e n s k e e w 6 o t p u n e v e y l t n e d n e p e d n i e v i v r u s e s n o p s e r ] 8 9 [ n a ( e s n o p s e r l l e c l a d o m i n U : s s e n h g u o r o n a n f o t c e f f E n a d e t o m o r p m n 0 5 - 0 2 = a R f o e g n a r e u l a v e t a i d e m r e t n i s s e n h g u o r r e w o l d n a r e h g i h h t o b d n a , e s n o p s e r l a m i t p o ) e s n o p s e r s i h t d e c u d e r m n 2 3 f o e u l a v s s e n h g u o r s i h t r o f , s e t y c o r t s a h t i w d e t a i c o s s a n o h g u o h t l A . m n 2 3 : q R e m a s e h t n o s e r u t c u r t s d e z i r a l o p e k i l d n u o f y l t n a n i m o d e r p e r e w s n o r u e n s e t a r t s b u s h g u o r e h t - n o x a , t n e n i m o r p d e t i b i h x e s n o r u e : s s e n h g u o r o n a n f o t c e f f E . m n 2 3 : q R c i f i c e p s e h t n o s t i a r t l a n o i t c n u f d e t a i c o s s a ] 9 9 [ h c u m e b o t d n u o f s a w t n e m p o l e v e d e t i r u e N : h c t i p f o t c e f f E h t i w g n i t a o C a h t i w s e c a f r u s n o n a h t h c t i p m n 0 0 4 a h t i w s e c a f r u s n o r e t s a f e h t m o r f r e i l r a e t u o d e h c t e r t s s e t i r u e n r o j a M . h c t i p m n 0 6 d n a e v a c n o c - e g r a l e h t n o y l s u o r o g i v e r o m w e r g d n a a m o s s e t a r t s b u s s u o r o p o n a n - 3 - ) l y h t e o n i m a i r t l y p o r p o n i m a e n a l i s y x o h t e m - 2 ( - N e n o N / s n o r u e n l a c i t r o c t a r c i n o y r b m E e c n e r e h d a l l e C e n o N a r g i N a i t n a t s b u S s e t y c o r t s a - n o r u e n s p i h s n o i t a l e r & h t w o r g t u o e t i r u e N l a p m a c o p p i h / s n o r u e n y r a m i r P F G N d n a e c n e r e h d a l l e C / s n o r u e n y t i l i b a i v l a p m a c o p p i h t a r 8 1 E h t w o r g t u o e t i r u e N / s n o r u e n . f e R ] 5 3 [ d e s a e r c n i n a d e t i b i h x e s l l e c 2 1 C P : s s e n h g u o r o n a n f o t c e f f E d n a - h t g n e l e t i r u e n d n a y t i r a l o p f o s m r e t n i - n o i t a i t n e r e f f i d n e g a l l o C r e t f a n o i t a i t n e r e f f i D / s l l e c 2 1 C P a c i l i s d e s r e p s i d o n o M g n i s a e r c n i f o s d i o l l o c / s e l c i t r a p o n a n ) 2 i O S ( a c i l i S e s n o p s e r r a l u l l e C / t n e m t a e r T n i e t o r P g n i t a o C y a s s a l l e C / e p y t l l e C s n o i s n e m d e r u t a e F i / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T 3 3 s e i h p a r g o p o T m o d n a R - s l l e c e v r e n n o s e i h p a r g o p o t o n a n d n a - o r c i m l a i c i t r a f o s t c e f f e e h T : 4 e l b a T m n 4 0 2 & 4 6 , 8 1 : a R m n 0 1 8 - 2 : a R m n 0 8 - 2 1 : q R e z i s s s e c o r p r e b ö t S i S l a t s y r c e l g n i S i S l a t s y r c e l g n i S - o n a N g n i h c t e l a c i m e h C / s s e n h g u o r g n i h c t e l a c i m e h C / s s e n h g u o r e r u t a e F e p y T s s e n h g u o r - o n a N - o n a N f o r e t n e c e h t t a s l e n n a h c s e r u t a e f e v a c n o c e g r a l e t a r t s b u s s u o r o p o n a N e r o p l a c i r d n i l y c h t i w l a c i m e h c o r t c e l e n o i t a z i d o n a e v a c n o c l l a m S e v a c n o c e g r a L m n 0 5 4 : h c t i P m n 0 6 : h c t i P g n i h s i l o p o r t c e l E & m u i n i m u l A & e v a c n o C / e d i x O s u o r o p o n a N n o c i l i S : i S ; r e t e m a r a p s s e n h g u o r e r a u q s n a e m t o o R : q R ; r e t e m a r a p s s e n h g u o r e g a r e v a c i t e m h t i r a : a R ; r o t c a f h t w o r g e v r e N : F G N : s n o i t a i v e r b b A * 34 2.4 The effect of topography on nerve cells - An insight into the mechanisms According to the previous sections, a considerable amount of valuable information on the nerve cell responses has been gathered from a plethora of the different topographical models (i.e. in terms of geometries, materials, cell types and assays). However, an insight into the mechanism of topography sensing of the neuronal cells is still lacking. Cell responses to topographical cues can be mainly classified into two main classes: a) The direct effect on cytoskeleton and b) the indirect effect on signaling. In the case of non-neuronal cells it has generally been accepted that the cellular machinery that recognizes the physical and topographical characteristics of the extracellular milieu are the intrgrin-based adhesion molecules [102]. Integrin activation leads in turn to the assembly of focal adhesions (FAs) which link the integrins to the actin filaments of the cytoskeleton [103,104]. FAs include the scaffolding proteins, like vinculin or talin, which transduce forces through the actin- myosin cytoskeletal network due to cytoskeletal tension; thus they can directly result in changes in cytoskeletal organization, structure and finally cell shape [102,105]. Furthermore, FAs include the signaling proteins, such as paxillin, that initiate biochemical signal cascades [104,105]. Such cascades include the activation of phosphorylation- and G-protein (like Rho)-mediated pathways (e.g. ROCK), which can lead to long-term changes in transcriptional regulation, cell proliferation, differentiation and survival [102]. According to this hypothesis, topography can play a critical role in the way the proteins (e.g. from the serum) deposit onto the surface which then influences the cell binding via the integrin activation [104]. Especially in the case of nanotopographical features, varying features at the scale of individual adhesions may alter the clustering and reorganization of integrins thus influencing the number and distribution of FAs [106]. In the case of neuronal cells, integrins are widely expressed in the nervous system and their presence has been linked with various developmental processes such as neuronal migration, axonal growth, and guidance by attaching to ECM proteins such as laminin, fibronectin, and tenascin [107 109]. Furthermore, many of these focal adhesion proteins, such as vinculin and talin have been detected in growth cones [109,110], suggesting a role for integrin-cytoskeletal coupling in growth cone motility and neuronal processes in general. Accordingly, the majority of the studies investigating the effect of topography on nerve cells have focused on integrin activation, FAs, cytoskseleton and ROCK signaling [80,88,96]. 34 35 Beyond its effect on integrin activation, topography may influence cell functions via other mechanisms, like triggering mechanosensitive ion channels or through secondary effects, such as alterations in the effective stiffness of the substrate. However, these remain still largely unexplored. As the cellular mechanisms of topography sensing have been very explicitly discussed in the review of Hoffman-Kim et al. [11], this section of the present review article is focused on the in vitro studies with nerve cells addressing specific questions of nerve cell contact guidance and neurite outgrowth, i.e. neuronal polarity, perpendicular vs. parallel contact guidance and neuronal function via mechanosensitive channels, in an attempt to decipher the mechanisms controlling these responses. Study of neuronal polarity selection Ferrari et al. have monitored PC12 cells after NGF stimulation by time-lapse, high- resolution microscopy to evaluate neurite outgrowth and polarity during the early stages of neurite outgrowth on nanogratings. Although inititially randomly oriented neurites emerged from an unpolarized cell body, these were in time retracted, as the cell body was acquiring a spindle- like shape and aligned its two poles to the nanograting. These results suggest that the transition from multipolar to bipolar shapes was highly guided by the nanogratings. Neurites emerging from the poles explored a restricted range of alignment angles. This bipolarity and neurite alignment were maintained after reaching this polarity. Focusing on the dynamics of focal adhesion (FA) formation and maturation they showed that FAs were preferentially grown on the ridges and developed preferentially along the direction of neurite extension (i.e. FA length). This topographical constraint imposed to the FAs by the ridge width resulted in a di erence in size between FAs formed at the tip of aligned neurites and those generated at the tip of misaligned ones. Accordingly, the maturation of aligned FAs was enhanced, while the maturation of the misaligned ones was lost. In an attempt to shed light on the mechanism, it was shown that this topography-driven polarity resulting into the maturation of the FAs, requires ROCK and myosin- II -mediated cell contractility. Inhibition of ROCK or of myosin-II activity affected PC12 polarity selection on nanograt hampering FA maturation [89]. the fraction of bipolar cells and Micholt et al. have studied the very early stages of neuronal polarity of hippocampal cells 35 36 on micron-sized pillars of 3 m height and varying width and spacing (ie. ranging from 1 to 5.6 m and from 0.6 to 15 m, respectively) framed by flat regions. Using time lapse imaging it was shown that the first sprout, but also the Golgi-centrosome complex, which is used as an indicator of neuronal polarization, were preferentially located on the pillar surfaces and not on the flat region [111]. This preferential orientation towards the pillars was also observed in N-cadherin distribution, whose asymmetric presence has been implicated in cell polarization [112]. All these events were independent of the inter-pillar spacing. However, growth cone morphology was shown to be dependent on inter-pillar spacing and seems to be correlated with the faster outgrowth response - i.e. longer neurites exhibited smaller growth cones- suggesting that the events of polarization and growth may not be influenced in the same way. In order to elucidate the possible signal transduction mechanism, analysis of the phosphotyrosine patches, which indicate areas of high activity for growth signaling in relation to the pillar-axons contact points, have been analyzed. Colocalization of actin filaments and phosphorylated tyrosine positions on the pillars-axon contacts suggests that tyrosine phosphorylation is involved in the regulation of actin remodeling. This behavior was dependent on the pillar spacing and was more striking on pillars with spacing in the range of 1 2 m, which was the dimension range of the longest neurite. Study of perpendicular contact guidance with respect to the cytoskeleton Rajnicek et al. have attempted to elucidate the mechanism for growth cone contact guidance by physical substratum contours via studying the cytoskeleton elements that are implicated in the organization of the cytoskeleton on topographical cues [56]. Use of cytoskeleton inhibitors targeting the particular growth cone elements, i.e. taxol and nocodazole for the microtubules and cytochalasin B for the actin microfilaments, did not influence the parallel and perpendicular contact guidance. Furthermore, drugs that have been shown to block perpendicular neurite guidance did not affect parallel orientation, suggesting that hippocampal neurites appear to use different mechanisms for perpendicular and parallel contact guidance. In an attempt to shed light on the signaling pathway for perpendicular alignment of hippocampal neurites, inhibitors for calcium channels, for G proteins and for protein tyrosine kinases have been used. Among them, essential components in the intracellular signal transduction pathway for perpendicular contact 36 37 guidance [56]. In order to shed light on the perpendicular contact guidance of DRG axons on parallel aligned fibers, Xie. et al. have investigated the pharmacological effect of myosin II inhibition [80]. By adding blebbistatin (i.e. an inhibitior of myosin II) to the culture medium, the perpendicular growth during the initial period of neurite extension from the ganglion would change into parallel growth. On the contrary, the presence of blebbistatin did not influence the direction of the parallel axonal growth although it had an impact on the quality of the interactions among the neurites. These results showed that the perpendicular and not the parallel neurite growth was dependent on the myosin II, suggesting an important role of myosin II in the perpendicular contact guidance of neurite outgrowth on parallel aligned fibers. Cytoskeletal actin dynamics are involved in pitch-dependent neurite outgrowth In an attempt to investigate the pitch-dependent neurite outgrowth of hippocampal neurons, Kang et al. have used three different biochemical inhibitors of neuronal cytoskeletal dynamics, i.e. cytochalasin D (inhibitor of the formation/function of F-actin), nocodazole (inhibitor of microtubule formation) and blebbistatin. In the case of nocodazole the neurons barely developed neurites in any substrate [96]. In the cases of cytochalasin D and blebbistatin, hippocampal neurons developed neurites; however the pitch-dependent outgrowth was lost, suggesting the importance of actin dynamics in this outgrowth response. However, treatment with the Y-27632, which is a selective inhibitor of ROCK of the Rho/ROCK pathway for cytoskeletal dynamics, did not affect this pitch-dependent neurite outgrowth. The authors suggest that biophysical (i.e. topographical) cues may induce intracellular pathways other than the Rho/ROCK pathway to influence neuritogenesis and neurite outgrowth. Nanoroughness modulates neuron function via mechanosensing channels It has been suggested that topography may influence neuron contact guidance and neurite outgrowth via other mechanisms, like triggering mechanosensitive ion channels. The study of Blumenthal et al. suggests a prominent role for the mechanosensitive ion channel the Piezo-1, which is expressed by CNS neurons and not by sensory neurons such as dorsal root ganglia [113], in sensing nanotopography [35]. For that reason, the expression of FAM38A, which is an integrin-activated transmembrane protein that is part of the mechanosensitive ion channel Piezo- 37 38 1 has been evaluated. Whereas FAM38A expression in PC12 cells on glass was predominantly localized at neurite branch points, which would be a region of high cytoskeletal tension, FAM38A expression was distributed more homogenously in cells grown on the substrate of the nanoroughness value which supported an increased differentiation, in terms of polarity and neurite length (i.e. Ra=32 nm). Interestingly, DRGs, which do not express Piezo-1 channels, did not show any morphological changes on nanoroughness substrates. Furthermore, inhibition of mechanosensing cation channels including Piezo-1 diminished the increased sensitivity to depolarization that was observed in hippocampal neurons on mean roughness of 32 nm and generally influenced hippocampal neuron-astrocyte interactions on all substrates. 3. The in vitro effects of topography on neuroglial cell morphology and functions Directional growth of neuroglial cells in response to topographical cues has been in vitro investigated with different topographical models, in an attempt to develop a stimulating and guiding environment for neurons. These studies are reviewed in this chapter. The current status in the literature regarding cell response including morphology, migration and activation- to various continuous and discontinuous topographies and random roughness is summarized in the Tables 5,6 and 7, respectively, where the various studies are presented with increasing topographical feature size. 3.1 The effect of continuous topographies on neuroglial cells Two types of continuous topographies in the form of i) alternating grooves/ridges and ii) aligned fibers have been mainly used to study the effect of topographical anisotropy on neuroglial cell growth and orientation. Growth and orientation of dissociated neuroglial cells on microgrooved substrates In the case of alternating microgrooves, groove width seems to be the key parameter for alignment of Schwann cells. The width of Schwann cells varies from 5 to 10 m and pattern widths or spacings ranging from 2 to 30 found to be optimal for the alignment of Schwann cells. Lietz et al. used PDL/LN-coated silicon chips with microgrooves of 15 m depth and various widths (2-100 m) and studied the effect of groove width on Schwann cell 38 39 morphology. It was shown that, when cultured on the micropatterned surfaces, Schwann cells displayed polar morphologies in parallel to the microgrooves. Specifically, among the different patterned substrates tested, cells were completely aligned on microgrooves of 2- and this high degree of orientation was preserved over culture time (Figure 7A) [114]. In contrast, cells grown on non-structured areas appeared disorganized without preferential orientation. Hsu et al. used a topographical model comprising silicon grooves with depth of subcellular size (i.e. of 1.5 m) groove width ranging from 10-20 m. When the width/spacing of the grooves increased from 10/10 to 20/20 m, the extent of cell alignment at 24 h was enhanced by 1.5 times (Figure 7B). Coating with laminin or collagen increased the percentage of aligned cells [115]. In another study, laminin-coated lines of varying widths or with constant width and increasing spacing (i.e. ranging from 10 to 50 m) were patterned onto PMMA. It was shown that Schwann cells attached and became elongated along the laminin stripes. Pattern widths below 40 m gave rise to increased degree of orientation [60]. Figure 7: Effect of groove width on Schwann cell growth. (Aa-c) SEM micrographs of the microgrooves of varying width (Aa); On microlanes, a longitudinal Schwann cell alignment was evident (arrowhead in (Ab)) but not on the planar surface of the same MStC (arrowhead in Ac); (Ad) chwann cell orientation (Reprinted with permission from [114]). (B) Optical micrographs of Schwann cells stained with anti S-100 on patterned silicon substrates at 24 h: 10/10/1.5 (Ba), 10 /20/1.5 (Bb), 20/10/1.5 (Bc) and 20/20/1.5 m (Bd) silicon/Al (Reprinted with permission from [115]). Apart from the parallel alignment of the cells along the groove axis, Goldner et al. have reported cellular bridging between neighboring microgrooves. In their study, a number of glial cell types including Schwann cells, have been shown to span grooves of a dimension larger than extensions, [53]. These findings are very important because they suggest that cellular response to topography is more complex than simply cytoskeletal restriction. The impact of groove depth is emphasized in the study of Hsu et al. using a topographical model comprising PLGA grooves with groove width and depth ranging from 10-20 m and 0.5-3 m, respectively. They showed that 1.5 m was the minimal groove depth to exert a proper 39 40 guiding effect on dissociated Schwann cells [115]. Coating with laminin improved short-term cell adhesion and alignment, although this effect was decaying with increasing groove depth. Since the use of aligned neuroglial cells as guidance structures for the growing or regenerating neurons is highly envisaged, some studies investigate the guidance effect of preseeded Schwann cells on neurons. Miller et al. used microgrooved polymer substrates made of biodegradable PDLA of 10 m width. Groove width was found to be a significant factor in promoting Schwann cell alignment, and widths and spacings ranging from 10- to be optimal [116]. The presence of preseeded aligned Schwann cells in the grooves were found to promote neurite alignment. The topography-induced orientation effect studied with Schwann cells has also been observed in astrocytes. In particular, rat type-1 astrocytes have been shown to get aligned along the direction of the groove of 3 and 10 m depth and width, respectively. Cell adhesion and at all seeding densities tested [66]. Astrocytes have been shown to orient themselves even in shallower grooves of 250 nm depth [117]. Oligodendrocytes were also sensitive to shallow grooves. In the remarkable study of Webb et al. the effect of anisotropic topographical cues on oligodendrocyte progenitor and mature oligodendrocytes has been studied and compared with that on hippocampal neurons. For that reason, quartz grooves of varying depth (i.e. 0.1-1.17 0.13-4 (i.e. 0.13-8 m) have been used. Both cell types of oligodendroglial lineage were found to be highly aligned by all topographical contours, although this alignment was less striking as inter- groove spacing increased to several micrometres. Remarkably, cells aligned even in the ultrafine topographical cues down to 100 nm, which was not the case for the neurons, suggesting the high degree of sensitivity of the oligodendrocyte lineage to topographical anisotropy [67]. To our knowledge, there is no additional study on the effect of oligodendrocyes onto micropatterned substrate. Growth and orientation of dissociated neuroglial cells on fibrous substrates Studies with parallel aligned electrospun polymeric fibers emphasize the importance of anisotropy in cell orientation. When Schwann cells were cultured on aligned electrospun fibers of submicron-to-micron diameter (0.3-1 m), both cell cytoskeleton and nuclei were elongated 40 41 and aligned along the fiber axes. On the contrary, cells cultured on random fibers exhibited random orientation [62,71]. Gnavi et al. showed that both primary Schwann cells and cells of a Schwann cell line were elongated along the axis of the aligned gelatin fibers of ~200-250 nm. Interestingly, cell adhesion and proliferation were deteriorated on the aligned vs. the random fibers [63]. When Schwann cells were cultured on aligned electrospun chitosan fibers of wider diameter (i.e. 15 via spiraling along the fibers and exhibited two shapes: spherical and long olivary. Furthermore, it was observed that the long olivary cells inclined to in a 3D fashion after 14 days [81]. Remarkably, this Schwann cell morphology change described above seems to be accompanied by a change in maturation and functionality. Two studies report that, compared to random a -regulated the expression of early myelination markers (i.e. myelin-associated glycoprotein and myelin protein zero, the cell adhesion molecule neural cadherin and the extracellular matrix molecule neurocan). This was accompanied by a down-regulation of non-myelinating immature Schwann cell markers (i.e. neural cell adhesion molecule) [64]. The effect of anisotropy imposed by parallel aligned fibers on cell directionality has been also reported with astrocytes [68]. Cao et al. have grown rat astrocytes on randomly oriented electrospun polymeric fibers of 665 nm diameter and studied their morphology, proliferation and a suppressed proliferation and increased apoptosis, exhibiting elongated and ramified cell shapes [65]. Furthermore, the Chondroitin polymerization factor (ChPF) siRNA uptake efficiency of the cells has been investigated for their potential on gene-silencing applications, since silencing the astrocytes Directional Schwann cell migration from whole explants on fibrous substrates Electrospun polymeric fibers have been also used to study the guidance effect of anisotropy on the Schwann cell migration from whole DRG explants. In these studies, fibers of parallel aligned and/or random orientation with constant fiber diameter of submicron (i.e. up to 500 nm) to micron (i.e. up to 3 m) scale were developed and the effect of anisotropy on Schwann cell migration was investigated. According to these studies, although cells grown on 41 42 random electrospun polymer matrices showed no directional preference, when the explants were grown on aligned matrices, Schwann cell migration displayed directionality that mimics that of the underlying fiber orientation [48,59,69 71]. Additionally, Schwann cells demonstrate the In another study, Wen and Tresco reported the effect of filament diameter on DRG outgrowth using PP fibers of varying diameter ranging from supracellular and beyond (500 to 100 m), cellular (30 m), down to subcellular size (5 m). An increasing degree of alignment of decreasing the Schwann cells parallel to the long axis of the Treatment of the filaments with laminin or fibronectin increased the maximum migration distance of the cells in all types of fibers. This amplifying effect was more obvious while the highly aligned along the 5 [79]. fiber diameter was decreased. 42 ] 5 1 1 [ e r o m e r e w s e c a f r u s d e n r e t t a p n o s e i d o b l l e C : y p o r t o s i n a f o t c e f f E r o n e g a l l o C n n a w h c S t a R e h t n o d e v r e s b o s l l e c r e d n u o r e h t h t i w d e r a p m o c d e t a g n o l e n i n i m a l s l l e c m 3 4 . f e R e s n o p s e r r a l u l l e C s e i h p a r g o p o T s u o u n i t n o C - s l l e c l a i l g o r u e n n o s e i h p a r g o p o t o n a n d n a - / t n e m t a e r T n i e t o r P g n i t a o C / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i ] 7 1 1 [ e h t o t n o d e h c a t t a y l l a i t n e r e f e r p e r e w s l l e C : y h p a r g o p o t f o t c e f f E , n o i g e r t a l f e h t n a h t r e h t a r n r e t t a p n a m u h y r a m i r P s e y c o r t s a : h t p e D m n 0 5 2 : h t d i W m 1 / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T e r u t a e F e p y T e n e z n e b o z A l a d i o s u n i S : a 5 s e l b a T / r e m y l o p o c c i h p a r g o l o h f e i l e r e c a f r u s g n i t a r g s e v o o r g ] 7 6 [ e r e w e g a e n i l l a i l g o r d n e d o g i l o e h t f o s l l e C : y p o r t o s i n a f o t c e f f E . s r u o t n o c l a c i h p a r g o p o t l l a y b d e n g i l a y l h g i h e b o t d n u o f o t d e s a e r c n i g n i c a p s e v o o r g - r e t n i s A : h t d i w e r u t a e f f o t c e f f E e g a e n i l l a i l g o r d n e d o g i l o f o s e p y t l l e c h t o b , s e r t e m o r c i m l a r e v e s . s n r e t t a p e h t y b d e n g i l a s s e l e m a c e b e n i s y L D y l o P e t y c o r d n e d o g i l O , r o t i n e g o r p e r u t a m s e t y c o r d n e d o g i l o s e t y c o r t s a & m : h t d i W 4 - 3 1 . 0 m 8 - 3 1 . 0 : g n i c a p S d o h t e m m 7 1 . 1 - 1 . 0 c i h p a r g o l o h r e s a L : h t p e D / z t r a u Q s e v o o r G : h t p e D 5 . 1 - 5 . 0 : h t d i W m 0 2 - 0 1 : g n i c a p S m 0 2 - 0 1 y h p a r g o h t i l o r c i M y h p a r g o h t i l t f o S / A G L P / n o c i l i S s e v o o G s e v o o r g e h t f o g n i c a p s / h t d i w e h t n e h W : g n i c a p s & h t d i w f o t c e f f E s a w t n e m n g i l a l l e c f o t n e t x e e h t , m 0 2 / 0 2 o t 0 1 / 0 1 m o r f d e s a e r c n i s e t a r t s b u s d e n r e t t a p n u e m i t 5 . 1 y b d e c n a h n e a e d i v o r p o t h t p e d e v o o r g l a m i n i m e h T : h t p e d e v o o r g f o t c e f f E m 5 . 1 s a w t c e f f e g n i d i u g r e p o r p - t r o h s d e v o r p m i n i n i m a l h t i w g n i t a o C : g n i t a o c n i e t o r p f o t c e f f E h t i w d e n i l c e d t c e f f e s i h t h g u o h t l a , t n e m n g i l a d n a n o i s e h d a l l e c m r e t . h t p e d e v o o r g g n i s a e r c n i ] 6 1 1 [ d e t o m o r p s l l e c n n a w h c S d e n g i l a d e d e e s e r P : y p o r t o s i n a f o t c e f f E n i n i m a L n n a w h c S t a R m 4 - 1 : h t p e D / A L D P s e v o o r G . t n e m n g i l a e t i r u e n / s l l e c m 0 1 : h t d i W 3 4 t n e m n g i l a e t i r u e n n o t c e f f E m 0 2 / 0 1 : g n i c a p S n o i s s e r p m o C r o g n i d l o m g n i t s a c t n e v l o s 4 4 . f e R e s n o p s e r r a l u l l e C / t n e m t a e r T n i e t o r P g n i t a o C / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i s e i h p a r g o p o T s u o u n i t n o C - s l l e c l a i l g o r u e n n o s e i h p a r g o p o t o n a n d n a - l a t e l e k s o t y c f o g n i d a e r p s d n a n o i s e h d a l l e C : g n i t a o c f o t c e f f E . s e v o o r g e h t f o n o i t c e r i d e h t n i d e n g i l a ] 6 6 [ e r e w s e t a r t s b u s d e n r e t t a p n u n o s e t y c o r t s a e l i h W : y p o r t o s i n a f o t c e f f E e r e w s e t a r t s b u s d e n r e t t a p e h t n o s l l e c e h t f o t s o m , d e t n e i r o y l m o d n a r n i n i m a L n o i t u l o s 1 - e p y t t a R s e t y c o r t s a e r e w s t n e m a l . s e t a r t s b u s h t i w d e s a e r c n i s a w n o i t a t n e i r o l l e C : h t d i w e v o o r g f o t c e f f E ] 4 1 1 [ l e l l a r a p n i s e i g o l o h p r o m r a l o p d e y a l p s i d s l l e C : y p o r t o s i n a f o t c e f f E e n i s y l - D - y l o P n n a w h c S t a R d e z i n a g r o s i d d e r a e p p a s a e r a d e r u t c u r t s - n o n n o s l l e c s a e r e h w , s e n a l o t n i n i m a L / c i t a i c s m o r f s l l e c . n o i t a t n e i r o l a i t n e r e f e r p t u o h t i w s e v r e n . h t d i w g n i s a e r c e d 4 4 ] 0 6 [ e h t g n o l a d e t a g n o l e d n a d e h c a t t a s l l e c n n a w h c S : y p o r t o s i n a f o t c e f f E . s e p i r t s n i n i m a l e h t d e s a e r c n i ) m 0 4 < ( s h t d i w n r e t t a p r e l l a m S : h t d i w f o t c e f f E . ) s y a d 4 d n a h 4 . e i ( s t n i o p e m i t o w t e h t n e e w t e b d e v r e s b o o N : e m i t e r u t l u c f o t c e f f E . n o i t a t n e i r o f o e e r g e d / t n e m t a e r t n i n i m a L n e g y x O a m s a l p n n a w h c S t a R s l l e c m 0 1 : h t d i W 0 2 : g n i c a p S m m : h t d i W 0 0 1 - 2 2 ± 5 1 : h t p e D m 0 5 - 0 1 : g n i c a p S : h t d i W 0 5 - 0 1 I e p y T e n i l d e t a u d a r g r o / d n a s h t d i w g n i c a p s I I e p y T d n a e n i l t n a t s n o c h t d i w g n i c a p s e e h T : b 5 s e l b a T / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T e r u t a e F e p y T g n i t s a c t n e v l o S / S P s e v o o r G n o i e v i t c a e r p e e D g n i h c t e / n o c i l i S s e v o o r G t c a t n o c o r c i M / A M M P g n i t n i r p s e n i L 5 4 . f e R e s n o p s e r r a l u l l e C / t n e m t a e r T n i e t o r P g n i t a o C / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i ] 8 1 1 [ l e l l a r a p d n a y t i l a n o i t c e r i d l l e c n n a w h c S : y p o r t o s i n a f o t c e f f E a m s a l P t a l f o t d e r a p m o c d e v o o r g o r c i m n o r e h g i h e r e w y t i c o l e v n o i t c e r i d / t n e m t a e r t n o e s o h t n a h t r e t s a f d e l e v a r t s e v o o r g n o s l l e C : e p y t e r u t a e f f o t c e f f E g n i t c a t n o c e m i t e r o m y l t n a c i f i n g i s t n e p s s u a e t a l p n o s l l e C . s u a e t a l p e r u t a e f e h t g n i t c a t n o c t n e p s s e v o o r g n o s l l e c n a h t s e g d e e r u t a e f e h t s e r u t a e f r e w o r r a n e h t n o t n e m n g i l a d e v o r p m I : h t d i w f o t c e f f E . s e t a r t s b u s - L - y l o P e n i s y l s l l e c n n a w h c S t a r t l u d a m o r f / e v r e n c i t a i c s e s p a l e m T i y p o c s o r c i m : h t p e D : h t d i W 0 6 / 0 3 m 0 6 / 0 3 t h g i l s a h t i W ( ) e r u t a v r u c g n i d l o m a c i l p e R / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T e r u t a e F e p y T / S M D P s e v o o r G s e i h p a r g o p o T s u o u n i t n o C - s l l e c l a i l g o r u e n n o s e i h p a r g o p o t o n a n d n a - : c 5 e l b a T ] 3 5 [ ) m 0 6 / 0 6 / 0 5 : s / w / d t a ( g n i g d i r b n o i s n e t x e r a l u l l e c f o n o i t a m r o F . s l l a w ] 3 6 [ n o i t a r e f i l o r p d n a n o i s e h d a l l e C . s t n e m a l i f n i t c a d e n g i l a g n i t i b i h x e , d e t a g n o l e e r e w s r e b i f d e n g i l a e h t n o s l l e C : y p o r t o s i n a f o t c e f f E . s r e b i f m o d n a r e h t o t d e r a p m o c d e n g i l a e h t n o d e s a e r c e d s a w ] 4 6 [ l a c i r e h p s d e t i b i h x e s r e b i f m o d n a r n o s l l e C : y p o r t o s i n a f o t c e f f E y g o l o h p r o m d a e r p s d e w o h s s r e b i f d e n g i l a n o s l l e c s a e r e h w e p a h s f o n o i s s e r p x e e h t d e t a l u g e r - . t n e m n g i l a f o n o i t c e r i d e h t g n o l a s r e b i f d e n g i l a n o s l l e C s r e b i f m o d n a r d n a m l i f o t d e r a p m o c d l o s r e k r a m n o i t a n i l e y m y l r a e ] 8 4 [ m o r f n o i t a r g i m l l e c n n a w h c S f o y t i r o j a m e h T : y p o r t o s i n a f o t c e f f E l e l l a r a p , y l l a n o i t c e r i d i n u d e d n e t x e m l i f r e b i f d e n g i l a e h t n o s G R D e h t l l e c n n a w h c S f o n o i t a t n e i r o e h t , y r a r t n o c e h t n O . s r e b i f d e n g i l a e h t o t 5 4 d e t u b i r t s i d y l m o d n a r s a w s m l i f r e b i f m o d n a r e h t n o n o i t a r g i m - L - y l o P e n i s y L s l l e C n n a w h c S c i t a i c s t a r m o r f s e v r e n o N e n i l l l e c n n a w h c S T 2 P 6 D - 4 T R ) 1 t l u d a y r a m i r p ) 2 s l l e C n n a w h c S s e v r e n c i t a i c s m o r f m n 4 7 ± 9 . 8 3 2 M T P G _ O E P L G / ( g n i n n i p s o r t c e l E / ) S m 0 0 2 - 0 3 : h t d i W g n i d l o m a c i l p e R / S M D P s e v o o r G : g n i c a p S 0 0 0 1 - 0 3 : r e t e m a i D n i t a l e G s r e b i F o N s l l e c n n a w h c S t a R : r e t e m a i D / A G L P s r e b i F ) 5 6 7 2 - L R C ( m n 0 5 ± 0 5 3 g n i n n i p s o r t c e l E o N G R D t a r 1 P / s t n a l p x e l l e c n n a w h c S n o i t a r g i m m n 0 0 6 - 0 0 4 : r e t e m a i D g n i n n i p s o r t c e l E / - A M N A P s r e b i F . f e R ] 9 5 [ e s n o p s e r r a l u l l e C y l e g r a l s a w n o i t a r g i m l l e c f o n o i t c e r i d e h T : y p o r t o s i n a f o t c e f f E ] 9 6 [ s s a l g r a n a l p o t g n i r e h d a s l l e c n n a w h c S : y p o r t o s i n a f o t c e f f E d w o r r a n y l e m e r t x e e m a c e b d n a d e t i b i h x e t n e m t a e r T / n i e t o r P g n i t a o C y l e v i t c e p s e r , d e c n a h n e d n a d e s a e r c e d s a w s e t a r t s b u s s u o r b i f e t y c o r t s a d e c n a h n e y h p a r g o p o t r e b i F . m l e h t o t d e r a p m o c ] 5 6 [ e h t n o s i s o t p o p a d n a n o i t a r e f i l o r p e t y c o r t s A : y h p a r g o p o t f o t c e f f E o N . n o i t a l l e t s d n a n o i t a g n o l e - e n e G n n a w h c S / s t n a l p x e n o i t a r g i m l l e c G R D t a r 5 1 E l l e c n n a w h c S / n o i t a r g i m s t n a l p x e d n a y g o l o h r p o m e k a t p u A N R i s , h t w o r g l l e C / y c n e i c i f f e l a c i t r o c t a R s e t y c o r t s a G R D k c i h c 0 1 E m n 4 2 5 ~ g n i n n i p s o r t c e l E : r e t e m a i D 5 . 0 ~ n e g a l l o C L C P / / d n e l b g n i n n i p s o r t c e l E & L C P s r e b i F : r e t e m a i D / A L L P s r e b i F m n 1 1 ± 5 6 6 : r e t e m a i D g n i n n i p s o r t c e l E / P E E L C P s r e b i F 6 4 s e i h p a r g o p o T s u o u n i t n o C - s l l e c l a i l g o r u e n n o s e i h p a r g o p o t o n a n d n a - / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T e r u t a e F e p y T : d 5 e l b a T n o i t a m r o f k r o w t e n n i t c e n o r b i f d e n g i l a d e t o m o r p d n a , n o i t u b i r t s i d . s m l i f - A M N A P h t o o m s o t d e r a p m o c d e t a i c o s s i d d n a / s l l e c n n a w h c S d n a n o i t a r g i m l l e c n n a w h c S 8 . 0 g n i n n i p s o r t c e l E ] 0 7 [ n i t c e n o r b i f d e c n e u l f n i s r e b i f - A M N A P d e n g i l A : y p o r t o s i n a f o t c e f f E s t n a l p x e G R D 1 P : r e t e m a i D / - A M N A P s r e b i F n o i t a z i n a g r o M C E 6 4 d n a n o t e l e k s o t y c e h t , s r e b i f d e n g i l a n o w e r g s l l e c n e h W h t o b n o s l l e C . s e x a r e b i f e h t g n o l a d e t a g n o l e d n a d e n g i l a . s r e b i f i e l c u n ) m l i f ( e t a r t s b u s l o r t n o c e h t o t d e r a p m o c s r e b i f d e n g i l a d n a m o d n a r n o d e t a l u g e r - n w o d s a w s r o t c a f c i h p o r t o r u e n e h t f o n o i s s e r p x e e n e G . e t a t s g n i t a n i l e y m - o r p a s d r a w o t d e t c e r i d e r e w . m l i f e h t o t d e r a p m o c s r e b i f n u p s o r t c e l e e h t m o r f s l l e c n n a w h c S / s e v r e n c i t a i c s s u t e f d n a y g o l o h p r o m l l e C n o i s s e r p x e e n e g m m . 8 0 0 ± 6 2 . 2 ) d e n g i l a ( . 3 0 0 ± 3 0 . 1 g n i n n i p s o r t c e l E ) d e t n e i r o ( ] 2 6 [ g n i y l r e d n u e h t f o n o i t a t n e i r o e h t d e w o l l o f s l l e C : y p o r t o s i n a f o t c e f f E o N n a m u h y r a m i r P r e t e m a i D / L C P s r e b i F 7 4 . f e R e s n o p s e r r a l u l l e C / t n e m t a e r T n i e t o r P g n i t a o C / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i s e i h p a r g o p o T s u o u n i t n o C - s l l e c l a i l g o r u e n n o s e i h p a r g o p o t o n a n d n a - / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T e r u t a e F e p y T : e 5 e l b a T . y t i l a n o i t c e r i d d e y a l p s i d s e c i r t a m ] 8 6 [ d e w o h s s e c i r t a m m o d n a r n o n w o r g s e t y c o r t s A : y p o r t o s i n a f o t c e f f E d e n g i l a n o n w o r g s e t y c o r t s a s a e r e h w , e c n e r e f e r p l a n o i t c e r i d o n / s e t y c o r t s a t a r 3 P y g o l o h p r o m l l e C : r e t e m a i D 3 - 2 ~ / e n o n a x o i d y l o P g n i n n i p s o r t c e l E s r e b i F ] 1 8 [ . y r a v i l o g n o l d n a l a c i r e h p s : s e p a h s t n e r e f f i d o w t h t i w s l a i r e t a m n a s o t i h c o t n o w e r g s l l e c n n a w h c S : y p o r t o s i n a f o t c e f f E o N s l l e c n n a w h c S t a R ] 9 7 [ l l e c n n a w h c A : r e t e m a i d f o t c e f f E e r o m s a w s e l d n u b t n e m a l i f e h t f o s i x a g n o l e h t o t l e l l a r a p t n e m e n g i l a n o r e t e m a i d r e b i f g n i s a e r c n i h t i w d e s a e r c e d e c n a t s i d n o i t a r g i m l l e C y b y l l a u q e d e c n a h n e s a w n o i t a r g i m l l e C : g n i t a o c n i e t o r p f o t c e f f E . s r e b i f d e t a o c - n i e t o r p d n a d e t a o c n u h t o b . e l b a k r a m e r r o n i n i m a L n i t c e n o r b i f / s t n a l p x e G R D l l e c n n a w h c S n o i t a r g i m N L d n a N F : r e t e m a i D m 5 1 : r e t e m a i D - 5 n o i s u r t x e l a m r e h T n a s o t i h C s r e b i F / P P t n e m a l i F s e l d n u b ] 1 7 [ h t w o r g t u o e t i r u e n d e d i u g - y h p a r g o p o T : y p o r t o s i n a f o t c e f f E n i n i m a L n o w o r g o t d n u o f e r e w s l l e c n n a w h c S : g n i t a o c f o t c e f f E . s t n e m a l i f f o s e p y t l l a r o / d n a L L P e l o h w G R D / s t n a l p x e l l e c n n a w h c S n o i t a r g i m : r e t e m a i D n o i s u r t x e t l e M / A L L P s t n e m a l i F 7 4 e h t f i n e v e ( e v o o r g e h t o t t c e p s e r h t i w d e t n e s e r p e r a s n r e t t a p - d e g d i r / - d e v o o r g e h t , n o s i r a p m o c f o e s a e d n a y t i r a l c f o s n o s a e r r o F : t n e m m o C n o c i l i S : i S ; n o i l g n a G l a c i v r e C r o i r e p u S : G C S ; e n e r y t s y l o P : S P ; e n e l y p o r p y l o P : P P ; ) e t a l y r c a h t e m . ) s e g d i r e h t o t t c e p s e r h t i w m e h t d e t r o p e r e v a h s r o h t u a l y h t e m ( y l o P : A M M P ; e d i t c a l - L - y l o p : A L L P ; ) d i c a c i l o c y l g - o c - c i t c a l ( y l o p : A G L P ; e n a x o l i s l y h t e m i d y l o P : S M D P ; ] ) e t a h p s o h p e n e l y h t e l y h t e ( - o c - e n o t c a l o r p a c [ y l o p : P E E L C P ; e n o t c a l o r p a c y l o P : L C P ; e t a l y r c a l y h t e m - o c - e l i r t i n o l y r c a ( y l o p : - A M N A P ; 1 y a D l a t a n t s o P : 1 P ; r o t c a f h t w o r g e v r e N : F G N ; t n e m a l i f o r u e N : F N ; n i t c e n o r b i F : N F ; n i n i m a L : N L ; 8 1 y a D c i n o y r b m E : 8 1 E ; n o i l g n a G t o o R l a s r o D : G R D : s n o i t a i v e r b b A * 48 3.2 The effect of discontinuous topographies on neuroglial cells The effect on cell growth and orientation Although the effect of anisotropic continuous geometries on Schwann cell alignment has been extensively studied, there are not many studies reporting on the engineering of cell alignment with discontinuous geometries. We have recently suggested that even a discontinuous topographical pattern can promote Schwann cell alignment, provided that it displays anisotropic features in a parallel orientation with interfeature distance at the subcellular level [58]. We have particularly shown that the parallel oriented microcones (MC) of elliptical cross section promoted oriented Schwann cell outgrowth along the major axis of the ellipse. Interestingly, Schwann cells followed and aligned with the orientation of this MC feature, which was more pronounced as the total roughness increased. On the contrary, cells on low roughness substrates, exhibiting a discontinuous but random pattern, showed an isotropic orientation, resembling to the 8A). This guidance effect was also observed in the migrating Schwann cells out of the DRG explants [58]. Lee et al. cultured an astrocyte cell line (C6 glioma cells) onto nanodotted-arrays with dot diameter ranging from 10 to 200 nm. They showed that cell viability, morphology and adhesion varied on the different substrates and exhibited optimal cell growth on 50 nm nanodots. Except for the increased cell surface area, cells on the substrate with 50 nm nanodots exhibited also a highly branched and complex network (syncytium), which is considered as a structural support for neurons with respect to cell-to-cell signaling (Figure 8B) [119]. In an attempt to provide an insight into the impact on the interecellular communication on the various substrates, it was shown that the expression level and cellular transport of the gap junction protein Cx43 in C6 glioma cells varied depending on the diameter of the dotted-arrays used. Figure 8: Effect of discontinuous topographies on neuroglial cell growth. (A) Effect on Schwann cells: (Aa) Confocal microscopy images of S100b positive Schwann cells grown on different silicon substrates for 5 days of culture, scale bar: 150 m; (Ab) Schwann cell orientation expressed in terms of the orientation angles' (frequency) distribution. The orientation angle of the nucleus, that is approximated as elliptical, was measured as the angle between the major axis of the ellipse and the vertical axis of the image plane. The number of cells exhibiting an orientation angle value within a from [58]). (B) Effect on glioma cells: (Ba) Confocal microscopy images of C6 glioma cells seeded on (Reprinted with permission 48 nadot arrays and incubated for 24, 72, and 120 h (green: vinculin and red: GFAP; Scale bar: 25 m); (Bb) Graph illustrating the density of branching (left) and the density of the meshes (right) against the diameter of the nanodots and grouped by incubation time (Reprinted with permission from [119]). 49 3.3 The effect of random topographies on neuroglial cells The effect on cell growth and orientation Topographical models of random nanoroughness have been used for the study of neuron- astrocyte interactions in order to shed light on the mechanisms of astrogliosis. Blumenthal et al. studied the hippocampal neurons-astrocytes interactions onto substrates of increasing nanoroughness made of monodispersed silica colloids. Although neurons on the rough substrates were predominantly found associated with astrocytes, there was a critical roughness value of Rq= 32 nm at which neurons were dissociated from astrocytes and continued to survive independently even up to 6 weeks [35]. In an attempt to investigate this decoupling of the hippocampal neurons from the astrocytes, the morphology of the astrocytes on the different roughness substrates has been investigated. Interestingly, astrocytes onto the critical roughness value of 32 nm exhibited a more migratory phenotype, accompanied by a significant increase in astrocyte surface roughness (Figure 9A). This remarkable finding provides a direct link between the physical properties of the cell surface and the underlying substrate roughness. Chapman et al. reported the novel ability of nanoporous gold (np-Au) to reduce astrocytic coverage through topographical cues. Specifically, they used a model of npAu comprising ligaments of average width of ~30.6 to 88.6 nm and pores of diameter: ~87-149 nm. Using an in vitro -culture model, they showed that np-Au selectively suppressed astrocytic coverage while maintaining high neuronal coverage [34]. This finding was independent of ligament width and pore diameter. By systematically discriminating between chemical and topographical effects, the cytotoxicity of the substrates as a possible culprit for the astrocytic reduction was excluded. In an attempt to give an insight into the possible mechanism, it was shown that the np-Au surface topography inhibits the initial spreading of astrocytes across the material surface compared to the unstructured planar Au (Figure 9B). Figure 9: Effect of random topographies on astrocytes growth. (A) Nanoroughness alters physical attributes of astrocytes: (Aa) Graph showing the form factor of astroyctes on the rough substrates of increasing nanoroughness. Decrease in form factor on Rq= 32 nm surfaces is consistent with a more 49 50 motile phenotype (Inset); (Ab) Morphological changes to astrocyte cell surface on nanorough surfaces: AFM images of astrocytes grown on glass (left) and on Rq = 32 nm surface (right) (Reprinted with permission from [35]). (Ba) SEM images of four di erent np- prepared by dealloying for 15 min to 24 h; (B astrocytes at in vitro day 1, 7, and 12 on np-Au and unstructured pl-Au. Astrocytes are highlighted by a red outline in the in vitro day 1 and 7 images. Astrocytes and neurons were visually di erentiated through colocalization of tubulin- (Reprinted with permission from [34]). 50 1 5 . f e R e s n o p s e r r a l u l l e C / t n e m t a e r T n i e t o r P g n i t a o c / e p y t l l e C y a s s a l l e C s n o i s n e m d e r u t a e F i / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T s e i h p a r g o p o t s u o u n i t n o c s i D - s l l e c l a i l g o r u e n n o s e r u t a e f l a c i h p a r g o p o t o n a n d n a - 6 e l b a T e r u t a e F e p y T ] 9 1 1 [ , n o t e l e k s o t y c , y g o l o h p r o m , y t i l i b a i v l l e C : r e t e m a i d f o t c e f f E . s t o d o n a n m n - 0 5 n o h t w o r g l l e c l a m i t p o d e w o h s n o i s e h d a d n a m n 0 5 e h t n o y t i x e l p m o c m u m i x a m d e t i b i h x e m u i t y c n y s l l e C n o i t c n u j p a g e h t f o t r o p s n a r t r a l u l l e c d n a l e v e l n o i s s e r p x E t n e r e f f i d e h t n o d e i r a v s l l e c a m o i l g 6 C n i n i e t o r p 3 4 x C n i e t o r p . s y a r r a - d e t t o d r e t e m a i d . s t o d o n a n o N t a r a m o t y c o r t s a - a m o i l g 6 C g n i k r o w t e n ) m u i t y c n y s ( / e n i l l l e c , h t w o r G m n 0 0 2 - 0 1 g n i s s e c o r p e d i x O m u i n i m u l A c i d o n A : r e t e m a i d t o D / e d i x o m u l a t n a T s t o d o n a N ] 8 5 [ f o ) C M ( s e n o c o r c i m h t i w s e c a f r u s d e n r e t t a p o r c i m e h t n o e c a l p 1 5 l a c i t i p i l l e f o s C M h t i w s e t a r t s b u s e h t n o C M e h t f o s i x a r o j a m e h t g n o l a d e n g i l a l e l l a r a p e r e w s l l e c , n o i t c e s - s s o r c y r a r t i b r a e h t t a d e v r e s b o s a w e s n o p s e r e c n a d i u g - y h p a r g o p o t e m a s e h T . n o i t c e s - s s o r c . s t n a l p x e G R D e l o h w s e k a t h t w o r g l l e c y r a r t i b r a h g u o h t l A : y p o r t o s i n a f o t c e f f E n e g a l l o C t a r d e t a i c o s s i D s l l e c n n a w h c S c i t a i c s m o r f e s u o m & s e v r e n e l o h w G R D - s s o r c y r a r t i r b r a ) i m 6 . 2 f o ( n o i t c e s m 3 . 1 d n a h c t i p f o s e n o c o r c i M t h g i e h / t n a l p x e - s s o r c l a c i t p i l l e ) i i & h t w o r g t u o l l e C m . 5 6 / 7 . 4 f o ( n o i t c e s n o i t a r g i m l l e C m . 6 8 / 7 / 3 d n a h c t i p ) t h g i e h i S l a t s y r c e l g n i S d n o c e s o t m e F r e s a l d e s l u p g n i s s e c o r p / s e n o c o r c i M . f e R ] 5 3 [ 2 5 e s n o p s e r r a l u l l e C / t n e m t a e r T g n i t a o c n i e t o r P / e p y t l l e C y a s s a l l e C e r u t a e F s n o i s n e m d i s e i h p a r g o p o t m o d n a R - s l l e c l a i l g o r u e n n o s e r u t a e f l a c i h p a r g o p o t o n a n d n a - / l a i r e t a m o i B n o i t a c i r b a F e u q i n h c e T : s n o i t c a r e t n i s e t y c o r t s a - s n o r u e n n o s s e n h g u o r o n a n f o t c e f f E y l t n a n i m o d e r p e r e w s n o r u e n s e t a r t s b u s h g u o r e h t n o h g u o h t l A n e g a l l o C l a p m a c o p p i h g n i s a e r c n i f o s d i o l l o c y r a m i r P a c i l i s d e s r e p s i d o n o M / s e l c i t r a p o n a n ) 2 i O S ( a c i l i S e v i v r u s e r e w s n o r u e n m n 2 3 = q R o t d e u n i t n o c t a , s e t y c o r t s a h t i w d e t a i c o s s a d n u o f d n a s e t y c o r t s a m o r f d e t a i c o s s i d . k e e w 6 o t p u n e v e y l t n e d n e p e d n i d e r e t l a s s e n h g u o r o n a N : s e t y c o r t s a n o s s e n h g u o r o n a n f o t c e f f E . s e t y c o r t s a y h t l a e h f o s s e n h g u o r l a c i p a e h t d n a s n o r u e n / s e t y c o r t s a s e t y c o r t s a - n o r u e N s p i h s n o i t a l e r m n 0 8 - 2 1 : q R e z i s s s e c o r p r e b ö t S 7 e l b a T e r u t a e F e p y T s s e n h g u o r - o n a N ] 4 3 [ : s n o i t c a r e t n i s e t y c o r t s a - s n o r u e n n o s s e n h g u o r o n a n f o t c e f f E e l i h w e g a r e v o c c i t y c o r t s a d e s s e r p p u s y l e v i t c e l e s u A - p N s e l p m a s e m o S d e t a o c e r e w l a c i t r o c t a R t n e m a g i l e g a r e v A / d l o G / s l l e c : ) m n ( h t d i w n o i s o r r o c y o l l A s u o r o p - o n a N s s o r c a s e t y c o r t s a f o g n i d a e r p s l a i t i n i e h t s t i b i h n i y h p a r g o p o t . u A r a n a l p d e r u t c u r t s n u e h t o t d e r a p m o c e c a f r u s l a i r e t a m e h t e c a f r u s u A - p N : s e t y c o r t s a n o s s e n h g u o r o n a n f o t c e f f E . r e t e m a i d e r o p d n a h t d i w t n e m a g i l f o t n e d n e p e d n i , e g a r e v o c l a n o r u e n h g i h g n i n i a t n i a m m l 3 O 2 l A h t i w l e d o m e r u t l u c o c a i l g - n o r u e N . o t 4 2 1 ± 6 . 9 8 4 ± 1 6 . 0 3 . 8 8 s s e c o r p 2 5 ] 6 8 [ s a w a e r a l l e c e g a r e v a e l i h W : s s e n h g u o r o n a n f o t c e f f E a d e y a l p s i d t i , e g a r e v o c d i o v t n e c r e p o t l a n o i t r o p o r p y l e s r e v n i a m s a l p r i A t n e m t a e r t . a e r a d i o v e g a r e v a o t e c n e d n e p e d c i n o t o n o m - n o n s l l e c e t y c o r t s a . % 8 0 ± 5 2 3 . o t 0 s s e c o r p g n i y o l l a e D : ) m n ( r e t e m a i d e r o p e g a r e v A o t 5 5 . 4 ± 1 1 . 7 8 9 . 9 ± 4 2 . 9 4 1 e n i r u M : e g a r e v o c d i o V / d l o G 2 : a e r a d i o v e g a r e v A m n 0 0 0 2 ± 0 0 7 3 2 o t 0 . s u o r o p - o n a N 53 3.4 The effect of topography on neuroglial cells: An insight into the mechanisms In the above sections the morphological and functional responses of neuroglial cells on various topographical models have been presented. However, an insight into the mechanism is still lacking. Although there is an increasing number of studies investigating the mechanism mediating the contact guidance effects of cell alignment on nerve cells (Section 2.4) either by studying the cytoskeleton elements that are implicated in the organization of the cytoskeleton on topographical cues, or by studying the indirect effect on signaling, similar studies with neuroglial cells are still lacking. In an attempt to shed light on the Schwann cell motility on microgrooved and flat substrates, Mitchell and Hoffman-Kim studied the effect of anisotropic topography on rat Schwann cells cultured on laminin- monitored cell trajectories revealing information about the cell motility and preference between the groove and plateaus [118]. The interfeature distance was shown to be important, while the smaller distance 30 was shown to reduce the prevalence of a multipolar morphology, giving rise to a more oriented movement. Through analysis of the velocities, they found that the parallel direction velocity was higher for the the microgrooved vs. the flat substrates while the reverse effect was recorded for the perpendicular direction velocity (Figure 10b). However, the overall cell velocity was relatively similar regardless of the topographical cue. The authors suggest that Schwann cells may respond to topography by orienting without altering their motility machinery. Furthermore, plateaus and groove floors exhibited distinct cues promoting differential motility and variable interaction with the topographical features. More specifically, it was shown that Schwann cells on grooves traveled faster than those on plateaus, where the cells spent significantly more time contacting the feature edges than contacting the feature walls on the grooves (Figure 10c). These findings imply that Schwann cells are initially oriented by the topographical cues, but this alignment can be preserved with occasional or periodic additional contact with the guiding feature, suggesting that the establishment of polarity may be a significant feature of the contact guidance mechanism in this topographical model. The complexity in motility was accompanied by a complexity in the phenotype, according to which an individual Schwann cell could exhibit multiple distinct motile phenotypes, with a rare unipolar morphology being correlated with a significantly increased velocity (Figure 10d) [118]. Figure 10: Effect of groove width on Schwann cell growth. (a) Schematic illustration of top and side 53 54 views of culture platforms, consisting of alternating raised (plateau, P) and indented (groove, G) regions. Arrows indicate directionality in motility studies: movement parallel to the topography occurs along the x axis, and movement perpendicular to the topography occurs along the y axis. Cells on plateaus encounter edge topography, while cells on grooves encounter corner and wall topography; (b) Parallel (x) and perpendicular (y) components of cell velocity on flat and microgrooved substrates; (c) Graph showing Schwan cell velocity as a function of the number of extensions (0 4). For all conditions, Schwann cells exhibiting one extension (unipolar morphology) migrated with a significantly higher velocity than cells with zero, two, three, or four extensions; (d) Phase contrast micrographs show cells with unipolar, bipolar, and multipolar morphologies (Reprinted with permission from [118]). 4. Discussion and Conclusions In the previous sections, the nerve and neuroglial cell responses to the three topographies (i.e. continuous, discontinuous and random topographies) have been presented. In the following, the results from these studies will be compared and correlated. It has to be emphasized that any attempt to compare the results among the various studies using similar topographies must take into consideration the specific cell subtypes used (e.g. PNS or CNS neurons/neuroglial cells; primary or cell lines), but also the age of the animal from which the cells had been isolated. 4.1 Effect on neurons Continuous topographies in the form of (i) alternating grooves/ridges and (ii) parallel aligned fibers have been shown to strongly enhance axonal guidance and orientation along groove and fiber axis, respectively. In the case of the grooved patterns the groove depth and width are shown to be the critical parameters for studying axonal guidance/oriented neurite outgrowth and neuronal polarization/branching, respectively [37,75,87 89]. For grooves of subcellular width size, the respective studies show that the majority of the hippocampal neurons and the PC12 cells can orient nm /500 nm (depth/width), respectively [54,87,88]. While, for grooves of cellular to supracellular width size, where neurons can grow in the channels among the grooves, the growth orientation of the neurites is promoted towards a direction parallel to channel walls with decreased neuronal branching as 54 55 step, along the groove axis or to bridge among grooves in the absence of underlying solid support [82]. This variety of distinct responses emphasizes the dynamic and complex nature of the topography-induced neuronal growth and axonal guidance. Concerning the parallel oriented fibers, strong topography-induced axonal guidance can be obtained using fibers of submicron-sized diameter (i.e. 200-500 nm), which resembles the diameter of the neurites [59,69,77,91 93]. This topography-induced guidance effect is observed on both dissociated primary neurons and in more complex cell culture systems, such DRG explant, and is amplified with the incorporation of biochemical signals factors (e.g. ECM proteins or growth factors) covalently bound on the fiber surface [93]. Although the strong axon alignment has been correlated with the underlying fiber alignment, recent studies demonstrate that the neurite orientation can be influenced by a set of additional parameters, including the surface chemistry of the fibers or the substrate and the fiber density [80]. Furthermore, axonal outgrowth fasciculation seems to be affected by the fiber diameter/thickness. DRG neurons on polymeric fibers of cellular to supracellular dimension showed decreased alignment and increased fas These results emphasize that an ensemble of neuronal processes (i.e. orientation, parallel contact guidance, fasciculation) can be topography-guided and provide new guidelines in designing fiber-based scaffolds for peripheral nerve regeneration. Besides continuous topographies, substrates with isotropic discontinuous topographies in the form of pillars and cones of subcellular scale can control the outgrowth of neuronal processes [36,55,73,90,94]. In these topographical models, neurons grow on top of the features forming extensive 2D networks and the feature spacing or pitch seems to be the critical parameter for oriented neurite outgrowth, with a spacing range of ~0.5- [36,94]. Neurons on discontinuous topographical features of larger feature sizes (i.e. tens to hundreds microns) exhibit wrapping around the feature promoting the formation of networks of 3D architecture [55]. In this case, both feature size and interspacing are the critical parameters for oriented neurite outgrowth. Anisotropic discontinuous topographies in the form of ellipitical microcones at subcellular lengthscale have been strikingly shown to enhance parallel alignment of PNS neurons along the major axis of the features [58]. These results could provide a new research dimension towards topography-controlled nerve cell responses. Random topographies have been additionally developed, as an approach to simulate the random nanoroughness of the ECM macromolecules and to study the relation between nanotopography alterations imposed by macromolecules and neuronal functions at specific pathological states (e.g. the 55 [35,97,99]. Recent studies show that both hippocampal neurons and PC12 cells seem to selectively sense a specific range of nanoroughness [35]. However, the establishment of a link between the nanoroughness of the substrate and the neural processes remains quite premature. In this context, future studies on hierarchical topographies combining random nanoroughness patterns with deterministic topographies like the grooved topographies are highly desirable. Studies with such platforms could provide a new insight into the topography-induced nerve 56 cell responses. Although an insight into the mechanism of topography sensing is still lacking, some first conclusions can be drawn based on recent findings [35,56,80,89,96,111]. The topography-driven neuronal polarity on grooved substrates has been associated with ROCK and myosin II- mediated contractility. Studies with micron-sized pillars suggest that neuronal polarization and growth may depend on different mechanisms [111]. Cytoskeletal actin dynamics but not via the Rho/ROCK pathway seem to be involved in pitch-dependent neurite outgrowth in a discontinuous topographical model [96]. Perpendicular neurite contact guidance, which was reported in hippocampal and sensory neurons on continuous anisotropic topographies, seems to have different mechanism from that of the parallel contact guidance [56,80]. In the case of sensory neurons, myosin II seem to be involved in the mechanism of perpendicular contact guidance [80]. While, random nanoroughness has been shown to modulate hippocampal neuronal function via integrin-activated triggering of mechanosensitive ion channels [35]. Most of the studies presented report also on the sensitivity of the neurons to the topographical cues, which can be evident in any interface between a smooth and a topographical patterned substrate, where neurons change its shape, configuration, branching etc. Although it is still challenging to compare the results among the different topographies, it seems that the continuous anisotropic topographies can support a stronger axonal guidance compared with the discontinuous ones of the same lengthscale [90]. Perhaps an interesting case to be explored is the discontinuous anisotropic topographies which have shown a strong topography-induced neuron orientation but this remains to be verified by further studies. 4.2 Effect on neuroglial cells Continuous anisotropic features, including (i) grooves and (ii) fibers, can guide the Schwann cell alignment, migration and influence their functionality. Schwann cells can orient themselves and align on grooved substrates, although the number of 56 studies remains limited compared with these with neuronal cells [60,114 116]. The minimal groove depth required to provide a guidance Neuroglial cells of CNS, including astrocytes and oligodendrocytes tend to sense much shallower grooved features, compared to Schwann cells and also to their neuronal counterparts, such as the hippocampal neurons. Specifically, oligodendrocytes and astrocytes align along grooves down to 100 and 250 nm depth, respectively [67,117]. At the same time, feature width is also a critical parameter for orientation and alignment of 57 the neuroglial cells. Indeed, pattern widths or inter- to be optimal for the alignment of Schwann cells. It should be noted that, topography-induced alignment can be strongly enhanced by an intermediate protein coating, for example laminin [115]. Schwann cells on parallel aligned fibers of subcellular size change the shape and orientation state they exhibit on the smooth films and become oriented and aligned along the fibers axis [59,62,63,81]. Interestingly, Schwann cells on larger fibers may encircle the fibers forming chains [81]. While, in whole explant models (e.g. DRG), the topographical anisotropy of the parallel fibers influences the direction of cell migration [48,59,69,70]. Such strong topography-induced guidance of both migrating Schwann cells and outgrowing axons has been investigated in the design of polymeric scaffolds for nerve tissue regeneration after peripheral nerve injury. Remarkably, this topography-guided orientation can be accompanied by changes in maturation and functionality of the Schwann cells [64]. Although the study of the topography-induced guidance effect on neuroglial cells with discontinuous topographies remains very limited, neuroglial cells seem to sense discontinuous features of subcellular lengthscale [58,119]. In particular, parallel oriented elliptical microcones with along the major axis of the ellipse, suggesting that even a discontinuous topographical pattern can promote Schwann cell alignment, provided that it displays a periodic distribution of anisotropic features in a parallel orientation [58]. Recent studies report that random topographies of specific roughness value can affect neuroglial cells (i.e. astrocytes) adhesion and spreading inducing the dissociation of the astrocytes from the neurons [34,35]. Such response has been correlated either with the topography-induced alteration of the astrocyte morphology [35] or the topography-induced selection of the neurons against the astrocytes due to their different morphologies [34]. These studies provide an insight into neurons-neuroglial interactions which can be exploited for the development of proper electrode coatings in neural prosthetic applications where a glial scar tissue formation hinders the neuron-electrode coupling. Finally, it should be pointed out that although there is an increasing number of studies 57 investigating the topography-sensing mechanism of nerve cells by either studying the cytoskeleton elements or by studying the indirect effect on signaling, similar studies with neuroglial cells are still 58 lacking. 4.3 Other Issues Except for the topographical parameters that have been presented in the previous sections, additional parameters which were beyond the scope of the present review can influence cell- biomaterial interactions. Material parameters can have a strong impact on the cell responses. For instance, the material type selected determines its physico-mechanical characteristics, for example its stiffness. Indeed, an increasing number of studies reported on the mechanical regulation of cell functions by the substrate elasticity [3,50,120]. Another aspect that has to be taken into consideration is the interdependency between topography and physico-mechanical properties of the substrate. Importantly, the elasticity of the substrate can be controlled geometrically by changing the aspect ratio (geometry-induced elasticity) [94,121], an effect which becomes significant at nanoscale topographical or of high aspect ratio features. Prominent examples are surfaces comprising micro-to-nanoscale post arrays which have been used to study cellular traction forces [122]. Future studies shall systematically study the effect of combination of the topographical parameters with the physico-mechanical ones on cell shape and functions [123]. Topographical models of both non-degradable, (e.g. PDMS, PMMA, etc.) and degradable (i.e. PLGA) polymeric substrates have been presented in this review article. In the case of the degradable substrates, degradation is an important aspect that must be emphasized reflecting the complexity of cell-biomaterials interactions. It is of critical importance to understand both the potential impact of the degradation of the polymers (e.g. type of by-products, rate of degradation, etc.) to the cells and the effect of the cellular environment to the degradation [124,125]. In this context, future studies to evaluate and compare the degradation behaviour of the substrates - in terms of the change of mass, molecular weight and other properties- are highly envisaged [126]. Finally, an important aspect in cell-biomaterials interactions, is protein adsorption. By the time cells reach the surface, the material has already been coated with a monolayer of proteins adsorbed by the supernatant. Hence, the host cells do not actually interface the material, but instead a dynamic layer of proteins [24]. Surface topography at the nanoscale, in the form of either deterministic or random topographies, has a great impact on protein adsorption, in terms of the speciation, conformation and 58 orientation of the surface-bound proteins (reviewed in [127]). The orientation and conformation can have a strong impact on the cell recognition via the integrins. In this context, a future design based on surface nanotopography should focus on the optimization of the protein adsorption [127]. 59 In conclusion New types of cell culture platforms with various topographical patterns at the micro and the nanoscale have been used to study in vitro the effect of surface topography on nerve, neuroglial and neural stem cells. Continuous topographies in the form of (i) alternating grooves/ridges of subcellular and cellular width and (ii) parallel aligned fibers of subcellular diameters have been shown to strongly enhance guidance and the orientation of neuronal and neuroglial cells along groove and fiber axis, respectively. Discontinuous isotropic or anisotropic topographies in the form of pillars, posts and cones, with diameters and inter-feature spacing at subcellular to cellular scale, have been shown to orient neurons and neuroglial cells. Random topographies with nano-sized features have been shown to influence specific functions of neurons and astrocytes, as well as their interactions. Regarding the underlying mechanisms of topography sensing, studies have either focused on the focal adhesion and cytoskeleton elements, or on the indirect effect on cell signaling (for instance, Rho- mediated pathways). A broader insight into such mechanisms is highly envisaged from the future studies. This knowledge will aid in controlling the outgrowth and patterning of nerve and neuroglial cells which is important for a broad spectrum of neuroscience subfields, ranging from basic neurobiological research to tissue engineering. Acknowledgements This work was supported by the European Research Infrastructure NFFA-Europe, funded by EU's H2020 framework programme for research and innovation under grant agreement n. 654360. The help of Zografia Karekou in the design of Figure 1 is acknowledged. 5. References [1] J. Floege, T. Ostendorf, G. Wolf, Growth factors and cytokines., Immunol. Ren. Dis. 1 (2001) 415 463. 59 60 [2] Differentiation, and Repair, Dev. Neurobiol. 71 (2011) 1090 1101. [3] K. Franze, P.A. Janmey, J. Guck, Mechanics in neuronal development and repair, Annu. Rev. Biomed. Eng. 15 (2013) 227 251. doi:10.1146/annurev-bioeng-071811-150045. [4] D.H. Kim, P.P. Provenzano, C.L. Smith, A. Levchenko, Matrix nanotopography as a regulator of cell function, J Cell Biol. 197 (2012) 351 360. doi:10.1083/jcb.201108062. [5] S. Vainio, U. Muller, Inductive tissue interactions, cell signaling, and the control of kidney organogenesis, Cell. 90 (1997) 975 978. http://www.ncbi.nlm.nih.gov/pubmed/9323125. [6] J.C. Adams, F.M. Watt, Regulation of development and differentiation by the extracellular matrix, Development. 117 (1993) 1183 1198. http://www.ncbi.nlm.nih.gov/pubmed/8404525. [7] F. Rosso, A. Giordano, M. Barbarisi, A. Barbarisi, From cell-ECM interactions to tissue engineering, J Cell Physiol. 199 (2004) 174 180. doi:10.1002/jcp.10471. [8] J. Lofberg, K. Ahlfors, C. Fallstrom, Neural crest cell migration in relation to extracellular matrix organization in the embryonic axolotl trunk, Dev Biol. 75 (1980) 148 167. http://www.ncbi.nlm.nih.gov/pubmed/7371990. [9] R. Perris, D. Perissinotto, Role of the extracellular matrix during neural crest cell migration, Mech Dev. 95 (2000) 3 21. http://www.ncbi.nlm.nih.gov/pubmed/10906446. [10] A.C. Bellail, S.B. Hunter, D.J. Brat, C. Tan, E.G. Van Meir, Microregional extracellular matrix heterogeneity in brain modulates glioma cell invasion, Int J Biochem Cell Biol. 36 (2004) 1046 1069. doi:10.1016/j.biocel.2004.01.013. [11] D. Hoffman-Kim, J.A. Mitchel, R. V Bellamkonda, Topography, Cell Response, and Nerve Regeneration, Annu. Rev. Biomed. Eng. 12 (2010) 203 231. doi:DOI 10.1146/annurev-bioeng- 070909-105351. [12] G.A. Dunn, J.P. Heath, A new hypothesis of contact guidance in tissue cells, Exp Cell Res. 101 (1976) 1 14. http://www.ncbi.nlm.nih.gov/pubmed/182511. [13] H. Andersson, A. van den Berg, Microfabrication and microfluidics for tissue engineering: state of the art and future opportunities, Lab Chip. 4 (2004) 98 103. doi:10.1039/b314469k. [14] A. Curtis, C. Wilkinson, Topographical control of cells, Biomaterials. 18 (1997) 1573 1583. http://www.ncbi.nlm.nih.gov/pubmed/9613804. [15] T.H. Park, M.L. Shuler, Integration of cell culture and microfabrication technology, Biotechnol Prog. 19 (2003) 243 253. doi:10.1021/bp020143k. [16] J. Voldman, M.L. Gray, M.A. Schmidt, Microfabrication in biology and medicine, Annu. Rev. 60 61 Biomed. Eng. 1 (1999) 401 425. doi:10.1146/annurev.bioeng.1.1.401. [17] P.J. Hollenbeck, J.R. Bamburg, Comparing the properties of neuronal culture systems: a shopping guide for the cell biologist, Methods Cell Biol. 71 (2003) 1 16. http://www.ncbi.nlm.nih.gov/pubmed/12884683. [18] J. Brown, P.C. Bridgman, Role of myosin II in axon outgrowth, J Histochem Cytochem. 51 (2003) 421 428. http://www.ncbi.nlm.nih.gov/pubmed/12642620. [19] L.A. Lowery, D. Van Vactor, The trip of the tip: understanding the growth cone machinery, Nat Rev Mol Cell Biol. 10 (2009) 332 343. doi:10.1038/nrm2679. [20] E.W. Dent, F. Tang, K. Kalil, Axon guidance by growth cones and branches: common cytoskeletal and signaling mechanisms, Neuroscientist. 9 (2003) 343 353. http://www.ncbi.nlm.nih.gov/pubmed/14580119. [21] K. Kalil, E.W. Dent, Branch management: mechanisms of axon branching in the developing vertebrate CNS, Nat Rev Neurosci. 15 (2014) 7 18. doi:10.1038/nrn3650. [22] M. Tessier-Lavigne, C.S. Goodman, The molecular biology of axon guidance, Science (80-. ). 274 (1996) 1123 1133. http://www.ncbi.nlm.nih.gov/pubmed/8895455. [23] M.E. Hatten, Riding the glial monorail: a common mechanism for glial-guided neuronal migration in different regions of the developing mammalian brain, Trends Neurosci. 13 (1990) 179 184. http://www.ncbi.nlm.nih.gov/pubmed/1693236. [24] B. Ratner, A. Hoffman, F. Schoen, J. Lemons, Biomaterials Science; An Introduction to Materials in Medicine, in: Academic Press, 2004. [25] J.J. Norman, T.A. Desai, Methods for fabrication of nanoscale topography for tissue engineering scaffolds, Ann Biomed Eng. 34 (2006) 89 101. doi:10.1007/s10439-005-9005-4. [26] M.M. Stevens, J.H. George, Exploring and engineering the cell surface interface, Science (80-. ). 310 (2005) 1135 1138. doi:10.1126/science.1106587. [27] H.-J. Choi, Vapor Liquid Solid Growth of Semiconductor Nanowire, in: G.-C. Yi (Ed.), Semicond. Nanostructures Optoelectron. Devices, 2012: pp. 1 36. [28] A. Noori, S. Upadhyaya, P.R. Selvanganapathy, Materials and Microfabrication Processes for Microfluidic Devices , in: Springer (Ed.), Microfluid. Biol. Appl., 2008. [29] V. Saile, U. Wallrabe, O. Tabata, J.G. Korvink, Introduction: LIGA and Its Applications , in: & C. GmbH (Ed.), LIGA Its Appl., 2009. [30] D. Qin, Y. Xia, J.A. Rogers, R.J. Jackman, X.-M. Zhao, G.M. Whitesides, Microfabrication, Microstructures and Microsystems, Top. Curr. Chem. 194 (1998). 61 [31] E. Stratakis, A. Ranella, C. Fotakis, Biomimetic micronanostructured functional surfaces for microfluidic and tissue engineering applications, Biomicrofluidics. 5 (2011) 13411. 62 doi:10.1063/1.3553235. [32] G.M. Whitesides, E. Ostuni, S. Takayama, X. Jiang, D.E. Ingber, Soft lithography in biology and biochemistry, Annu. Rev. Biomed. Eng. 3 (2001) 335 373. doi:10.1146/annurev.bioeng.3.1.335. [33] M.A. Wood, Colloidal lithography and current fabrication techniques producing in-plane nanotopography for biological applications, J R Soc Interface. 4 (2007) 1 17. doi:10.1098/rsif.2006.0149. [34] C.A.R. Chapman, H. Chen, M. Stamou, J. Biener, M.M. Biener, P.J. Lein, E. Seker, Nanoporous Gold as a Neural Interface Coating: Effects of Topography, Surface Chemistry, and Feature Size, ACS Appl. Mater. Interfaces. 7 (2015) 7093 7100. [35] N.R. Blumenthal, O. Hermanson, B. Heimrich, V.P. Shastri, Stochastic nanoroughness modulates neuron-astrocyte interactions and function via mechanosensing cation channels, Proc Natl Acad Sci U S A. 111 (2014) 16124 16129. doi:10.1073/pnas.1412740111. [36] N.M. Dowell-Mes n, M.-A. Abdul-Karim, A.M. Turner, S. Schanz, H.G. Craighead et al, Neural Eng. 1 (2004). [37] C. Simitzi, E. Stratakis, C. Fotakis, I. Athanassakis, A. Ranella, Microconical silicon structures influence NGF-induced PC12 cell morphology, J Tissue Eng Regen Med. (2014). doi:10.1002/term.1853. [38] A.T. Nguyen, S.R. Sathe, E.K.F. Yim, From nano to micro: topographical scale and its impact on cell adhesion, morphology and contact guidance., J. Phys. Condens. Matter. 28 (2016) 183001. doi:10.1088/0953-8984/28/18/183001. [39] M.G. Burnett, E.L. Zager, Pathophysiology of peripheral nerve injury: a brief review, Neurosurg Focus. 16 (2004) E1. http://www.ncbi.nlm.nih.gov/pubmed/15174821. [40] V.T. Ribeiro-Resende, B. Koenig, S. Nichterwitz, S. Oberhoffner, B. Schlosshauer, Strategies for inducing the formation of bands of Bungner in peripheral nerve regeneration, Biomaterials. 30 (2009) 5251 5259. doi:10.1016/j.biomaterials.2009.07.007. [41] Y. Li, G. Huang, X. Zhang, L. Wang, Y. Du, T.J. Lu, F. Xu, Engineering cell alignment in vitro, Biotechnol Adv. 32 (2014) 347 365. doi:10.1016/j.biotechadv.2013.11.007. [42] D.T. Theodosis, D.A. Poulain, S.H. Oliet, Activity-dependent structural and functional plasticity of astrocyte-neuron interactions, Physiol Rev. 88 (2008) 983 1008. 62 63 doi:10.1152/physrev.00036.2007. [43] R.B. Campenot, NGF and the local control of nerve terminal growth, J Neurobiol. 25 (1994) 599 611. doi:10.1002/neu.480250603. [44] J. Goldberg, How does an axon grow?, Genes Dev. 17 (2003) 941 958. [45] J. Park, H. Koito, J. Li, A. Han, Microfluidic compartmentalized co-culture platform for CNS axon myelination research, Biomed. Microdevices. 11 (2009) 1145 1153. doi:10.1007/s10544- 009-9331-7. [46] M.P. Maher, J. Pine, J. Wright, Y.C. Tai, The neurochip: a new multielectrode device for stimulating and recording from cultured neurons, J Neurosci Methods. 87 (1999) 45 56. http://www.ncbi.nlm.nih.gov/pubmed/10065993. [47] M. Merz, P. Fromherz, Silicon chip interfaced with a geometrically defined net of snail neurons, Adv. Funct. Mater. 15 (2005) 739 744. [48] Y.T. Kim, V.K. Haftel, S. Kumar, R. V Bellamkonda, The role of aligned polymer fiber-based constructs in the bridging of long peripheral nerve gaps, Biomaterials. 29 (2008) 3117 3127. doi:10.1016/j.biomaterials.2008.03.042. [49] W.M. Grill, S.E. Norman, R. V Bellamkonda, Implanted neural interfaces: biochallenges and engineered solutions, Annu. Rev. Biomed. Eng. 11 (2009) 1 24. doi:10.1146/annurev-bioeng- 061008-124927. [50] F. Han, C. Zhu, Q. Guo, H. Yang, B. Li, Cellular modulation by the elasticity of biomaterials, J. Mater. Chem. B. 4 (2016) 9 26. [51] L.A. Greene, A.S. Tischler, Establishment of a noradrenergic clonal line of rat adrenal pheochromocytoma cells which respond to nerve growth factor, (1976). [52] L.A. Cyster, K.G. Parker, T.L. Parker, D.M. Grant, The effect of surface chemistry and nanotopography of titanium nitride (TiN) films on primary hippocampal neurones, Biomaterials. 25 (2004) 97 107. doi:10.1016/S0142-9612(03)00480-0. [53] J.S. Goldner, J.M. Bruder, G. Li, D. Gazzola, D. Hoffman-Kim, Neurite bridging across micropatterned grooves, Biomaterials. 27 (2006) 460 472. doi:10.1016/j.biomaterials.2005.06.035. [54] N. Gomez, Y. Lu, S.C. Chen, C.E. Schmidt, Immobilized nerve growth factor and microtopography have distinct effects on polarization versus axon elongation in hippocampal cells in culture, Biomaterials. 28 (2007) 271 284. doi:10.1016/j.biomaterials.2006.07.043. [55] J.N. Hanson, M.J. Motala, M.L. Heien, M. Gillette, J. Sweedler, R.G. Nuzzo, Textural guidance 63 64 cues for controlling process outgrowth of mammalian neurons, Lab Chip. 9 (2009) 122 131. [56] A. Rajnicek, C. McCaig, Guidance of CNS growth cones by substratum grooves and ridges: effects of inhibitors of the cytoskeleton, calcium channels and signal transduction pathways, J Cell Sci. 110 ( Pt 2 (1997) 2915 2924. https://www.ncbi.nlm.nih.gov/pubmed/9359874. [57] N. Arimura, K. Kaibuchi, Key regulators in neuronal polarity, Neuron. 48 (2005) 881 884. doi:10.1016/j.neuron.2005.11.007. [58] C. Simitzi, P. Efstathopoulos, A. Kourgiantaki, A. Ranella, I. Charalampopoulos, C. Fotakis, I. Athanassakis, E. Stratakis, A. Gravanis, Laser fabricated discontinuous anisotropic microconical substrates as a new model scaffold to control the directionality of neuronal network outgrowth, Biomaterials. 67 (2015) 115 128. http://dx.doi.org/10.1016/j.biomaterials.2015.07.008. [59] E. Schnell, K. Klinkhammer, S. Balzer, G. Brook, D. Klee, P. Dalton, J. Mey, Guidance of glial cell migration and axonal growth on electrospun nano bers of poly-e-caprolactone and a collagen/poly-e-caprolactone blend, Biomaterials. 28 (2007) 3012 3025. [60] K.E. Schmalenberg, K.E. Uhrich, Micropatterned polymer substrates control alignment of proliferating Schwann cells to direct neuronal regeneration, Biomaterials. 26 (2005) 1423 1430. doi:10.1016/j.biomaterials.2004.04.046. [61] J.A. Richardson, C.W. Rementer, J.M. Bruder, D. Hoffman-Kim, Guidance of dorsal root ganglion neurites and Schwann cells by isolated Schwann cell topography on poly(dimethyl siloxane) conduits and films, J Neural Eng. 8 (2011) 46015. doi:10.1088/1741-2560/8/4/046015. [62] S.Y. Chew, R. Mi, A. Hoke, K.W. Leong, The effect of the alignment of electrospun fibrous scaffolds on Schwann cell maturation, Biomaterials. 29 (2008) 653 661. [63] S. Gnavi, B.E. Fornasari, C. Tonda-Turo, R. Laurano, M. Zanetti, G. Ciardelli, S. Geuna, The Effect of Electrospun Gelatin Fibers Alignment on Schwann Cell and Axon Behavior and Organization in the Perspective of Artificial Nerve Design, Int J Mol Sci. 16 (2015) 12925 12942. doi:10.3390/ijms160612925. [64] J. Radhakrishnan, A.A. Kuppuswamy, S. Sethuraman, A. Subramanian, Topographic Cue from Electrospun Scaffolds Regulate Myelin-Related Gene Expressions in Schwann Cells, J Biomed Nanotechnol. 11 (2015) 512 521. http://www.ncbi.nlm.nih.gov/pubmed/26307833. [65] 666 674. [66] J.B. Recknor, J.C. Recknor, D.S. Sakaguchi, S.K. Mallapragada, Oriented astroglial cell growth 64 65 on micropatterned polystyrene substrates, Biomaterials. 25 (2004) 2753 2767. doi:10.1016/j.biomaterials.2003.11.045. [67] A. Webb, P. Clark, J. Skepper, A. Compston, A. Wood, Guidance of oligodendrocytes and their progenitors by substratum topography, J. Cell Sci. 108 (1995) 2747 2760. [68] W.N. Chow, D.G. Simpson, J.W. Bigbee, R.J. Colello, Evaluating neuronal and glial growth on electrospun polarized matrices: bridging the gap in percussive spinal cord injuries, Neuron Glia Biol. 3 (2007) 119 126. doi:10.1017/S1740925X07000580. [69] J.M. Corey, D.Y. Lin, K.B. Mycek, Q. Chen, S. Samuel, E.L. Feldman, D.C. Martin, Aligned electrospun nanofibers specify the direction of dorsal root ganglia neurite growth, J. Biomed. Mater. Res. Part A. 83A (2007) 636 645. doi:10.1002/jbm.a.31285. [70] V. Mukhatyar, M. Salmerón-Sánchez, S. Rudra, S. Mukhopadaya, T.H. Barker, A.J. García, R. V Bellamkonda, Role of fibronectin in topographical guidance of neurite extension on electrospun fibers, Biomaterials. 32 (2011) 3958 3968. [71] N. Rangappa, A. Romero, K.D. Nelson, R.C. Eberhart, G.M. Smith, Laminin-coated poly(L- lactide) filaments induce robust neurite growth while providing directional orientation, J Biomed Mater Res. 51 (2000) 625 634. http://www.ncbi.nlm.nih.gov/pubmed/10880110. [72] J.D. Foley, E.W. Grunwald, P.F. Nealey, C.J. Murphy, Cooperative modulation of neuritogenesis by PC12 cells by topography and nerve growth factor, Biomaterials. 26 (2005) 3639 3644. doi:10.1016/j.biomaterials.2004.09.048. [73] W.T. Su, Y.F. Liao, T.W. Wu, B.J. Wang, Y.Y. Shih, Microgrooved patterns enhanced PC12 cell growth, orientation, neurite elongation, and neuritogenesis, J. Biomed. Mater. Res. Part A. 101 (2013) 185 194. doi:10.1002/jbm.a.34318. [74] F. Haq, V. Anandan, C. Keith, G. Zhang, Neurite development in PC12 cells cultured on nanopillars and nanopores with sizes comparable with filopodia, Int. J. Nanomedicine. 2 (2007) 107 115. doi:DOI 10.2147/nano.2007.2.1.107. [75] M.J. Mahoney, R.R. Chen, J. Tan, W. Mark Saltzman, The influence of microchannels on neurite growth and architecture, Biomaterials. 26 (2005) 771 778. doi:10.1016/j.biomaterials.2004.03.015. [76] K.P. Das, T.M. Freudenrich, W.R. Mundy, Assessment of PC12 cell differentiation and neurite growth: a comparison of morphological and neurochemical measures, (2004). [77] H.S. Koh, T. Yong, C.K. Chan, S. Ramakrishna, Enhancement of neurite outgrowth using nano- structured scaffolds coupled with laminin, Biomaterials. 29 (2008) 3574 3582. 65 66 doi:10.1016/j.biomaterials.2008.05.014. [78] A. Hurtado, J.M. Cregg, H.B. Wang, D.F. Wendell, M. Oudega, R.J. Gilbert, J.W. McDonald, Robust CNS regeneration after complete spinal cord transection using aligned poly-L-lactic acid microfibers, Biomaterials. 32 (2011) 6068 6079. doi:10.1016/j.biomaterials.2011.05.006. [79] X. Wen, P.A. Tresco, Effect of filament diameter and extracellular matrix molecule precoating on neurite outgrowth and Schwann cell behavior on multifilament entubulation bridging device in vitro, J Biomed Mater Res A. 76 (2006) 626 637. doi:10.1002/jbm.a.30520. [80] J. Xie, W. Liu, M.R. MacEwan, P.C. Bridgman, Y. Xia, Neurite outgrowth on electrospun nanofibers with uniaxial alignment: the effects of fiber density, surface coating, and supporting substrate, ACS Nano. 8 (2014) 1878 1885. doi:10.1021/nn406363j. [81] Y. Yuan, P. Zhang, Y. Yang, X. Wang, X. Gu, The interaction of Schwann cells with chitosan membranes and fibers in vitro, Biomaterials. 25 (2004) 4273 4278. doi:10.1016/j.biomaterials.2003.11.029. [82] N.Z. Li, A. Folch, Integration of topographical and biochemical cues by axons during growth on microfabricated 3-D substrates, Exp Cell Res. 311 (2005) 307 316. doi:10.1016/j.yexcr.2005.10.007. [83] M. Cecchini, G. Bumma, M. Serresi, F. Beltram, PC12 differentiation on biopolymer nanostructures, Nanotechnology. 18 (2007) 505103. doi:10.1088/0957-4484/18/50/505103. [84] A. Gorman, A. Windebank, A. Pandit, Effect of functionalized micropatterned PLGA on guided neurite growth, Acta Biomater. 5 (2009) 580 588. doi:10.1016/j.actbio.2008.09.002. [85] F. Johansson, P. Carlberg, N. Danielsen, L. Montelius, M. Kanje, Axonal outgrowth on nano- imprinted patterns, Biomaterials. 27 (2006) 1251 1258. [86] O. Kurtulus, E. Seker, Nanotopography effects on astrocyte attachment to nanoporous gold surfaces, Conf Proc IEEE Eng Med Biol Soc. 2012 (2012) 6568 6571. doi:10.1109/EMBC.2012.6347499. [87] P. Clark, P. Connolly, A.S. Curtis, J.A. Dow, C.D. Wilkinson, Topographical control of cell behaviour: II. Multiple grooved substrata, Development. 108 (1990) 635 644. http://www.ncbi.nlm.nih.gov/pubmed/2387239. [88] A.M. Rajnicek, S. Britland, C.D. McCaig, Contact guidance of CNS neurites on grooved quartz: in uence of groove dimensions, neuronal age and cell type, J. Cell Sci. 110 (1997) 2905 2913. [89] A. Ferrari, M. Cecchini, A. Dhawan, S. Micera, I. Tonazzini, Nanotopographic control of 66 67 neuronal polarity, Nano Lett. 11 (2011) 505. [90] W. Li, Q.Y. Tang, A.D. Jadhav, A. Narang, W.X. Qian, P. Shi, S.W. Pang, Large-scale topographical screen for investigation of physical neural-guidance cues, Sci Rep. 5 (2015) 8644. doi:10.1038/srep08644. [91] M.K. Leach, Z.-Q. Feng, C.C. Gertz, S.J. Tuck, T.M. Regan et al, The Culture of Primary Motor and Sensory Neurons in Defined Media on Electrospun Poly-L-lactide Nanofiber Scaffold, J. Vis. Exp. (2011). [92] G.G. Genchi, G. Ciofani, A. Polini, I. Liakos, D. Iandolo, A. Athanassiou, D. Pisignano, V. Mattoli, A. Menciassi, PC12 neuron-like cell response to electrospun poly( 3-hydroxybutyrate) substrates, J Tissue Eng Regen Med. 9 (2015) 151 161. doi:10.1002/term.1623. [93] S. Patel, K. Kurpinski, R. Quigley, H. Gao, B.S. Hsiao, M.M. Poo, S. Li, Bioactive nanofibers: synergistic effects of nanotopography and chemical signaling on cell guidance, Nano Lett. 7 (2007) 2122 2128. doi:10.1021/nl071182z. [94] M.A. Bucaro, Y. Vasquez, B.D. Hatton, J. Aizenberg, Fine-tuning the degree of stem cell polarization and alignment on ordered arrays of high-aspect-ratio nanopillars, ACS Nano. 6 (2012) 6222 6230. doi:10.1021/nn301654e. [95] K. Kang, Y.S. Park, M. Park, M.J. Jang, S.M. Kim, J. Lee, J.Y. Choi, H. Jung da, Y.T. Chang, M.H. Yoon, J.S. Lee, Y. Nam, I.S. Choi, Axon-First Neuritogenesis on Vertical Nanowires, Nano Lett. 16 (2016) 675 680. doi:10.1021/acs.nanolett.5b04458. [96] K. Kang, S.Y. Yoon, S.E. Choi, M.H. Kim, M. Park, Y. Nam, J.S. Lee, I.S. Choi, Cytoskeletal actin dynamics are involved in pitch-dependent neurite outgrowth on bead monolayers, Angew Chem Int Ed Engl. 53 (2014) 6075 6079. doi:10.1002/anie.201400653. [97] S.P. Khan, G.G. Auner, G.M. Newaz, Influence of nanoscale surface roughness on neural cell attachment on silicon, Nanomedicine. 1 (2005) 125 129. doi:10.1016/j.nano.2005.03.007. [98] Y.W. Fan, F.Z. Cui, S.P. Hou, Q.Y. Xu, L.N. Chen, I.S. Lee, Culture of neural cells on silicon wafers with nano-scale surface topograph, J. Neurosci. Methods. 120 (2002) 17 23. doi:Pii S0165-0270(02)00181-4Doi 10.1016/S0165-0270(02)00181-4. [99] W.K. Cho, K. Kang, G. Kang, M.J. Jang, Y. Nam, I.S. Choi, Pitch-dependent acceleration of neurite outgrowth on nanostructured anodized aluminum oxide substrates, Angew Chem Int Ed Engl. 49 (2010) 10114 10118. doi:10.1002/anie.201003307. [100] W. Hallstrom, T. Martensson, C. Prinz, P. Gustavsson, L. Montelius, L. Samuelson, M. Kanje, Gallium phosphide nanowires as a substrate for cultured neurons, Nano Lett. 7 (2007) 2960 5. 67 [101] E.L. Papadopoulou, A. Samara, M. Barberoglou, A. Manousaki, S.N. Pagakis, E. Anastasiadou, C. Fotakis, E. Stratakis, Silicon Scaffolds Promoting Three-Dimensional Neuronal Web of Cytoplasmic Processes, Tissue Eng. Part C-Methods. 16 (2010) 497 502. 68 doi:10.1089/ten.tec.2009.0216. [102] B. Geiger, J.P. Spatz, A.D. Bershadsky, Environmental sensing through focal adhesions, Nat. Rev. Mol. Cell Biol. 10 (2009) 21 33. doi:10.1038/nrm2593. [103] S. Di Cio, J.E. Gautrot, Cell sensing of physical properties at the nanoscale: Mechanisms and control of cell adhesion and phenotype, Acta Biomater. 30 (2016) 26 48. doi:10.1016/j.actbio.2015.11.027. [104] K. Kulangara, K.W. Leong, Substrate topography shapes cell function, Soft Matter. 5 (2009) 4072 4076. doi:Doi 10.1039/B910132m. [105] M.J. Dalby, N. Gadegaard, R.O. Oreffo, Harnessing nanotopography and integrin-matrix interactions to influence stem cell fate, Nat Mater. 13 (2014) 558 569. doi:10.1038/nmat3980. [106] F. Guilak, D.M. Cohen, B.T. Estes, J.M. Gimble, W. Liedtke, C.S. Chen, Control of stem cell fate by physical interactions with the extracellular matrix, Cell Stem Cell. 5 (2009) 17 26. doi:10.1016/j.stem.2009.06.016. [107] R.O. Hynes, A.D. Lander, Contact and adhesive specificities in the associations, migrations, and targeting of cells and axons, Cell. 68 (1992) 303 322. http://www.ncbi.nlm.nih.gov/pubmed/1733501. [108] L.F. Reichardt, K.J. Tomaselli, Extracellular matrix molecules and their receptors: functions in neural development, Annu Rev Neurosci. 14 (1991) 531 570. [109] D.M. Suter, P. Forscher, Substrate cytoskeletal coupling as a mechanism for the regulation of growth cone motility and guidance, J. Neurobiol. 44 (2000) 97 113. [110] A.M. Sydor, A.L. Su, F.S. Wang, A. Xu, D.G. Jay, Talin and vinculin play distinct roles in filopodial motility in the neuronal growth cone, J Cell Biol. 134 (1996) 1197 1207. http://www.ncbi.nlm.nih.gov/pubmed/8794861. [111] L. Micholt, A. Gartner, D. Prodanov, D. Braeken, C.G. Dotti, C. Bartic, Substrate topography determines neuronal polarization and growth in vitro, PLoS One. 8 (2013) e66170. doi:10.1371/journal.pone.0066170. [112] A. Gartner, E.F. Fornasiero, S. Munck, K. Vennekens, E. Seuntjens, W.B. Huttner, F. Valtorta, C.G. Dotti, N-cadherin specifies first asymmetry in developing neurons, EMBO J. 31 (2012) 1893 1903. doi:10.1038/emboj.2012.41. 68 [113] Y. Roudaut, A. Lonigro, B. Coste, J. Hao, P. Delmas, M. Crest, Touch sense: functional organization and molecular determinants of mechanosensitive receptors, Channels (Austin). 6 (2012) 234 245. doi:10.4161/chan.22213. [114] M. Lietz, L. Dreesmann, M. Hoss, S. Oberhoffner, B. Schlosshauer, Neuro tissue engineering of glial nerve guides and the impact of different cell types, Biomaterials. 27 (2006) 1425 1436. 69 doi:10.1016/j.biomaterials.2005.08.007. [115] S.H. Hsu, C.Y. Chen, P.S. Lu, C.S. Lai, C.J. Chen, Oriented Schwann cell growth on microgrooved surfaces, Biotechnol Bioeng. 92 (2005) 579 588. doi:10.1002/bit.20634. [116] C. Miller, S. Jeftinija, S. Mallapragada, Micropatterned Schwann cell-seeded biodegradable polymer substrates significantly enhance neurite alignment and outgrowth, Tissue Eng. 7 (2001) 705 715. doi:10.1089/107632701753337663. [117] H. Baac, J.-H. Lee, J.-M. Seo, T.H. Park, H. Chung, S.-D. Lee, S.J. Kim, Submicron-scale topographical control of cell growth using holographic surface relief grating , Mater. Sci. Eng. C. 24 (2004) 209 212. [118] J.A. Mitchel, D. Hoffman-Kim, Cellular scale anisotropic topography guides Schwann cell motility, PLoS One. 6 (2011) e24316. doi:10.1371/journal.pone.0024316. [119] C.H. Lee, Y.W. Cheng, G.S. Huang, Topographical control of cell-cell interaction in C6 glioma by nanodot arrays, Nanoscale Res Lett. 9 (2014) 250. doi:10.1186/1556-276X-9-250. [120] D. Discher, Tissue Cells Feel and Respond to the Stiffness of Their Substrate, Science (80-. ). 310 (2005) 1139. [121] J. Fu, Y.K. Wang, M.T. Yang, R.A. Desai, X. Yu, Z. Liu, C.S. Chen, Mechanical regulation of cell function with geometrically modulated elastomeric substrates, Nat Methods. 7 (2010) 733 736. doi:10.1038/nmeth.1487. [122] M.T. Yang, N.J. Sniadecki, C.S. Chen, Geometric considerations of micro- to nanoscale elastomeric post arrays to study cellular traction forces, Adv. Mater. 19 (2007) 3119 +. doi:10.1002/adma.200701956. [123] S.P. Garland, C.T. McKee, Y.-R. Chang, V.K. Raghunathan, P. Russell, C.J. Murphy, A Cell Culture Substrate with Biologically Relevant Size-Scale Topography and Compliance of the Basement Membrane, Langmuir. 30 (2014) 2101 2108. [124] A. Asti, L. Gioglio, Natural and synthetic biodegradable polymers: different scaffolds for cell expansion and tissue formation, Int J Artif Organs. 37 (2014) 187 205. doi:10.530/ijao.5000307. [125] A. Subramanian, U.M. Krishnan, S. Sethuraman, Development of biomaterial scaffold for nerve 69 tissue engineering: Biomaterial mediated neural regeneration, J Biomed Sci. 16 (2009) 108. 70 doi:10.1186/1423-0127-16-108. [126] Y. Dong, T. Yong, S. Liao, C.K. Chan, M.M. Stevens, S. Ramakrishna, Distinctive degradation behaviors of electrospun polyglycolide, poly(DL-lactide-co-glycolide), and poly(L-lactide-co- epsilon-caprolactone) nanofibers cultured with/without porcine smooth muscle cells, Tissue Eng Part A. 16 (2010) 283 298. doi:10.1089/ten.tea.2008.0537. [127] M.S. Lord, M. Foss, F. Besenbacher, Influence of nanoscale surface topography on protein adsorption and cellular response, Nanotoday. 5 (2010) 66 78. 70 Controlling the morphology and outgrowth of nerve and neuroglial cells: The effect of surface topography C. Simitzi*, A. Ranella and E. Stratakis* Institute of Electronic Structure and Laser (IESL), Foundation for Research and Technology-Hellas (FORTH), Heraklion, 71003, Greece Statement of significance There is increasing evidence that physical cues, such as topography, can have a significant impact on the neural cell functions. With the aid of micro-and nanofabrication techniques, new types of cell culture platforms are developed and the effect of surface topography on the cells has been studied. The present review article aims at reviewing the existing body of literature reporting on the use of various topographies to study and control the morphology and functions of cells from nervous tissue, i.e. the neuronal and the neuroglial cells. The cell responses from phenomenology to investigation of the underlying mechanisms- on the different topographies, including both deterministic and random ones, are summarized. * [email protected]; [email protected]
1609.05052
2
1609
2016-12-05T11:16:33
Uncertainty of current understanding regarding OBT formation in plants
[ "physics.bio-ph", "q-bio.TO" ]
Radiological impact models are important tools that support nuclear safety. For tritium, a special radionuclide that readily enters the life cycle, the processes involved in its transport into the environment are complex and inadequately understood. For example, tritiated water (HTO) enters plants by leaf and root uptake and is converted to organically bound tritium (OBT) in exchangeable and non-exchangeable forms; however, the observed OBT/HTO ratios in crops exhibit large variability and contradict the current models for routine releases. Non-routine or spike releases of tritium further complicate the prediction of OBT formation. The experimental data for a short and intense atmospheric contamination of wheat are presented together with various models predictions. The experimental data on wheat demonstrate that the OBT formation is a long process, it is dependent on receptor location and stack dynamics, there are differences between night and day releases, and the HTO dynamics in leaf and ear is a very important contributor to OBT formation.
physics.bio-ph
physics
Uncertainty of current understanding regarding OBT formation in plants A. Melintescu*, D. Galeriu "Horia Hulubei" National Institute for Physics and Nuclear Engineering, Department of Life and Environmental Physics, 30 Reactorului St., P.O. Box MG-6, Bucharest – Magurele, Romania, RO-077125 * Corresponding author: Phone: +40 21 4042359; Fax: +40 21 4574440; email: [email protected], [email protected] Abstract Radiological impact models are important tools that support nuclear safety. For tritium, a special radionuclide that readily enters the life cycle, the processes involved in its transport into the environment are complex and inadequately understood. For example, tritiated water (HTO) enters plants by leaf and root uptake and is converted to organically bound tritium (OBT) in exchangeable and non-exchangeable forms; however, the observed OBT/HTO ratios in crops exhibit large variability and contradict the current models for routine releases. Non-routine or spike releases of tritium further complicate the prediction of OBT formation. The experimental data for a short and intense atmospheric contamination of wheat are presented together with various models' predictions. The experimental data on wheat demonstrate that the OBT formation is a long process, it is dependent on receptor location and stack dynamics, there are differences between night and day releases, and the HTO dynamics in leaf and ear is a very important contributor to OBT formation. 1. Introduction Tritium (3H) is present in the environment as a result of both natural and anthropogenic sources. Large quantities of tritium are currently produced in heavy water reactors and fuel reprocessing plants, and it is anticipated that the development of fusion energy will increase environmental releases. Romania develops nuclear energy using Canadian heavy water reactors and two CANDU 6 units are in operation having large tritium loads. The past and present Romanian research studies cover environmental tritium monitoring (Paunescu et al., 2012), environmental tritium modelling (Barry et al., 1999), and it was pointed out that tritium has to be considered as a special radionuclide (Galeriu and Melintescu, 2010). Tritiated water (HTO) and tritiated gas (HT) are the major forms initially released in environment and the organic forms (tritiated methane or tritiated formaldehyde) have minor releases (Amano, 1995; Kakiuchi et al.2002). In the process of photosynthesis, plants produce organic matter using solar light as energy source, from carbon dioxide from the air, nutrients from soil, and water from soil or air. Because tritium is larger than hydrogen, the organic forms of tritium are produced less readily than the organic forms of hydrogen. The organic forms of tritium are generically called organically bound tritium (OBT). Following the definitions of Kim et al. (2013), there are three types of OBT: exchangeable OBT, non-exchangeable OBT, and a special form of buried tritium (i.e. tritium included in the hydration shell of biomolecules). How stable tritium is within such organic compounds depends on the nature of the bond between tritium and the organic molecule and on the organic molecule affinity with the different biological tissues (ASN, 2010). When tritium is bound to oxygen, sulphur or nitrogen, it can be easily exchanged with tritium in the HTO (or H2O) and the exchangeable organically bound tritium (E-OBT) is formed. When tritium is covalently bound to carbon, only enzymatic reactions can destroy the bound and non-exchangeable OBT (NE-OBT) is formed. Buried tritium, which is inaccessible because of the physical structure of the organic molecule, quickly exchanges with hydrogen atoms in the body following digestions and, consequently, it increases the amount of tritium in the body water. The time when tritium remains incorporated therefore depends on the biomolecular turnover: fast, in the case of molecules involved in the energy cycle and slow, in the case of structuring molecules or macromolecules such as DNA or energy reserve molecule. Due to longer residence in the organism, NE-OBT is of first concern for health effects of a radiological dose. The methods to measure the HTO in environmental samples are well established but the OBT measurements are expensive and difficult. OBT measurements in all food chain components (human and non-human) are not possible. The alternative is to use radiological /environmental impact assessment models (RIA/EIA). These models can be defined as research grade and decision making models. The latter are used in design, licensing, normal operation, accident prevention and management. The decision making models must meet the following requirements: • Relatively simple; • Transparent; • Easy to program; • Results should be conservative, yet reasonable; • Deterministic calculations possible (worst case assessments); • Probabilistic calculations possible (95% percentile as worst case). Finally, when the models are applied in operational context, they must quickly provide results (i.e. have a short run time). During the working group meeting dedicated to "Tritium Accidents" (WG 7) of Environmental Modelling for Radiation Safety (EMRAS II) programme coordinated by International Atomic Energy Agency (IAEA) (http://wwwns.iaea.org/projects/emras/emras2/working-groups/working-group- seven.asp?s=8), one of the questions was if it is possible to have such models for tritium, due to high complexity of processes involved in the environmental tritium dynamics. A robust model for accidental tritium releases must minimize the uncertainty that can arise from the model structure, model parameters, and those situations that need special attention (see Chapter 15 in IAEA, 2014a). For routine releases of heavy water reactors, a steady state model developed by the Canadian Standard Association (CSA, 2014a) is generally used. Recently, claims have been made that the CSA model is not conservative with regard to OBT concentration in food (Thompson et al., 2015). Romania uses both Canadian (CSA, 2014a) and European practice (https://ec.europa.eu/energy/en/topics/nuclear-energy/radiation- protection) in its radiation protection programmes and some differences between those practices were noted (Galeriu et al., 2009a). The anti-nuclear groups claimed that the current models for routine release largely under-estimate the public dose because it is based on the yearly average of the air concentration and ignores spikes in the releases (Fairlie, 2010). Indeed, this claim is justified because at the receptor, the air concentration fluctuates and the equilibrium conditions are not reached as CSA (2014a) considered. In the extreme situation of a short and intense emission of tritium, a detailed analysis of the uptake processes is still needed, in order to provide a robust prediction of OBT production in crops at harvest. An analysis of the present state of knowledge regarding OBT modelling in plants during normal operation of nuclear utilities, as well as spike releases of tritium emissions are provided in the present study in order to support the current efforts regarding the tritium model improvements (Melintescu et al., 2015; Galeriu and Melintescu, 2016). 2. Materials and Methods 2.1. Uncertainty of OBT/HTO ratios for routine releases Pressurized heavy water reactors (PHWR) were developed in Canada and have higher tritium emissions than other energetic reactors. For calculation of Derived Release Limits (DRL) for assessment of the public dose during normal operation of PHWRs, the standard guide used in Canada and Romania in the past ignored the OBT contribution to the food chain (CSA, 1987). Subsequently, the importance of OBT was pointed out in relation with fusion research (Murphy, 1993) and the consideration of OBT contribution to the food chain was proposed soon afterwards (Galeriu, 1994). The contribution of OBT was considered in the second revision of the Canadian guide (CSA, 2008). The last revision of the guide (CSA, 2014) is based on the same specific activity (SA) approach as the previous guides, assuming full equilibrium in all environmental compartments (CSA, 1987, 2008). The SA approach is also used by IAEA in its coordinated research studies (IAEA, 2009; 2010) and details regarding the CSA (2014a) and IAEA (2009, 2010) approaches are given in the Appendix A. Based on simplifying assumptions of SA approach, the ratio between OBT concentration (measured by water of combustion), COBT and HTO concentration in plant water (leaves), CTFWT is constant. The difference between those two approaches is that CSA considers the total OBT in plant and the isotopic discrimination factor, IDp has a recommended constant value of 0.8 (range of 0.64-1.3), while IAEA considers NE-OBT in plant and the partition factor for plants, Rp has a recommended constant value of 0.54 as a geometric mean (GM) and a geometric standard deviation (GSD) of 1.16. The data considered by IAEA (2009) were for barley, maize and alfalfa, grown in laboratory controlled experiments and had equilibrium values. The experimental values for the partition factor, Rp cover a range of 0.4-0.68. The partition factor, Rp includes both the isotopic discrimination factor (with an average value close to the IDp given by CSA (2014a)) and the contribution of E-OBT to total OBT. For public dose assessment during routine emissions and based on equilibrium assumption, both approaches consider the crop contamination at harvest and they ignore the losses due to storage and food processing. For the assessment of plant tissue free water tritium (TFWT), CSA (2014a) considers the air HTO concentration (with a reduction factor, RFp due to the lower contribution of soil HTO), while IAEA (2009) separately considers the tritium transfer between air and plant (air pathway) and that between soil and plant (root pathway) (for details see Appendix A). The air HTO concentration is a yearly average or an average over the vegetation period. Disregarding the difference between total OBT and NE-OBT, the OBT/HTO ratio for equilibrium conditions is close to 0.7 with a range of 0.4-1.3. Based on a detailed analysis of a large range of experimental values for OBT and HTO in agricultural products, it was observed that the ratio between OBT (water of combustion) and TFWT (HTO in leaf water) largely differs from the equilibrium value of 0.7 (CNSC, 2013; Korolevych et al., 2014; Thompson et al., 2015) with many values higher than 5 and few values higher than 10. Traditionally, this ratio is expressed as OBT/HTO, but in fact, OBT represents tritium concentration in water of combustion and HTO represents tritium concentration in plant/animal water, which is actually TFWT. Theoretically, that ratio should be expressed as OBT/TFWT, but for the sake of simplicity, the present study keeps the traditional expression of OBT/HTO ratio. The uncertainty and under estimates of OBT concentrations come from the equilibrium assumptions based on a long term average of HTO concentration in air. In real field conditions, there is no equilibrium as it was previously pointed out (IAEA, 2008a). In fact, at the receptor where the measurements were carried out, there is a fluctuating HTO concentration in air, the HTO concentration in leaf has a diurnal variability and the OBT accumulation in different plant parts is a slow process. The sampling is usually carried out during the working hours and the HTO concentration in leaf can be low after the plume passage and high during the plume passage. The HTO concentrations in plants reflect conditions in the few hours before sampling, while monitoring results are yearly or monthly. On the other hand, residence times for OBT in plants are sufficiently long that concentrations in these compartments better reflect average air concentrations on longer intervals. Plant OBT/HTO ratios varied between 0.12 and 0.56, when monthly HTO air average at Pickering Nuclear Power Plant (NPP) (Canada) was used (Davis et al., 2005). Modelling attempts regarding these complex situations were reported (IAEA, 2008b) and emphasised in APPENDIX A "Model Performance as a Function of Air Concentration Averaging Time" of the same study (IAEA, 2008b). Further efforts to better link the OBT concentration to air/leaf HTO concentration were reported (Korolevych and Kim, 2013). For leaf, TFWT air concentration was averaged within 1 hour, 1 day, 3 days, and 9 days before sampling, while for NE-OBT, the averaging interval was 15 days, considering the "pod fill" period and 2 weeks before or after. Pod was a generic name for the edible part of tomatoes, red peppers, green peppers, and potatoes and fill period was 22 days after anthesis. Data for 2008, 2010 and 2011 were used to establish the reducing factors, RF for TFWT and NE-OBT and a large variability was observed between years or averaging intervals (see Tables 2 and 3 in Korolevych and Kim, 2013). Due to this large variability and huge costs of monitoring/measurements, that approach was not considered by CSA (2014a) and the "equilibrium" value of the discrimination factor, IDp was kept for total OBT. A series of measurements with frequent sampling were carried out (Korolevych et al., 2014) assuming that the rinsed OBT is NE-OBT. A large variability of the averages of the OBT/HTO ratios up to a factor of 5 and their ranges between different years was observed (see Table 3 in Korolevych et al., 2014). The ranges of OBT/HTO ratio cover two orders of magnitude. It is useful to note that the total OBT (E-OBT + NE-OBT) is higher than the rinsed OBT (NE-OBT) by a factor less than 2 and the measured rinsed OBT can contain buried tritium. Separate experiments were carried out to assess the importance of buried tritium (Kim et al., 2008) showing that buried tritium represents less than 10 % of the rinsed OBT in vegetables. The experimental averages and ranges of OBT/HTO ratios are large and cannot be explained only by the sampling time or measurement of various OBT concentrations as it was demonstrated by a model reconstruction (see Table 4 in Korolevych et al., 2014). The model recommends a generic value of 2 for OBT/HTO ratio, but higher values were observed in certain areas (see Table 4 in Korolevych et al., 2014). Korolevych et al. (2014) concluded that the most high (and low) values of the OBT/HTO ratio arise simply as a result of normal plant sampling procedures, but SA approach does not provide an explanation for OBT/HTO range. In a recent study (Thomson et al., 2015), concentrations of OBT and HTO were measured over two growing seasons in vegetation and soil samples obtained in the vicinity of four nuclear facilities and two background locations in Canada. For tritium processing facilities, the OBT/HTO ratios are slightly higher, with a factor of up to 2 than for CANDU reactors. The OBT/HTO ratios in vegetation have large variations of the ranges between 0.5 and 20. Ratios of the OBT activity concentration in plants to the OBT activity concentration in soils appear to be a good indicator of the long-term behaviour of tritium in soil and vegetation. The results show that some parameters used in environmental transfer models approved for regulatory assessments should be revisited to better account for the behaviour of HTO and OBT in the environment and to ensure that modelled estimates (e.g., plant OBT) are appropriately conservative. This affects not only the OBT/HTO ratios in crops, but also the HTO activity in plants in relation to the HTO activity in soil. Following the regulatory requirements, the assessment of public dose during normal operation of nuclear facilities must be slightly conservative and based on reliable models results and experimental data. Concentrations of HTO and OBT in food items at consumption are needed. Thompson et al. (2015) pointed out few weak points of the Canadian guide (CSA, 2014a):   There is a direct transfer between air and plants and it is supposed that the transfer to soil is only 0.3 from that in air; Relationships for OBT and HTO are too simple and OBT concentration as predicted is not conservative. A recent study (Mihok et al., 2016) presents the experimental results regarding the OBT/HTO ratios in soils and plants near a gaseous tritium light source manufacturing facility in Canada (SRBT), which emits tritium as tritiated hydrogen (HT) and HTO only during the day time and week days. The plants were irrigated with tap water (containing a tritium concentration of 5 Bq L-1), rain water (with a tritium concentration of 80 Bq L-1) and well water (with a high tritium concentration of 11000 Bq L-1). The plants (natural grass, sod, potatoes, beans and Swiss chard) were cultivated in barrels in soil with low tritium contamination (29 Bq L-1 HTO and 105 Bq L-1 OBT). The plants and soil were weekly sampled and analysed (same day and hour). When plants were irrigated with tap and rain water, the OBT concentration in soil did not vary, but the HTO concentration in soil equilibrated with that of irrigation water and with that coming from the atmospheric source (including the HT oxidation in soil). For that irrigation regime (tap and rain water), the OBT/HTO ratio in roots and tubers was about 2 and in the aerial plant parts the OBT/HTO ratio was 8. When plants were irrigated with highly contaminated water from wells, the soil and plants HTO equilibrated with about 25 % from that of irrigation water and OBT/HTO ratio in roots and tubers is 0.4 and in aerial plant parts is 1.4 (see Table 3 in Mihok et al., 2016). The experimental data show clearly that the OBT/HTO ratio is much higher than the equilibrium value of 0.7, but it is close to equilibrium when the irrigation water is highly contaminated with tritium. For tap and rain irrigation experiments, the OBT/HTO ratio increases during the development stage of plants. It is clear that the processes are complex involving photosynthesis, atmospheric and root pathways of tritium transfer and the processes are still not well understood. 2.2. Influence of spike releases on routine emissions The previous experimental results clearly demonstrate that non-equilibrium processes are involved during normal operation of tritium facilities (e.g. CANDU reactors or others). Spike releases can occur due to reactor operation, for example for period of fuel change in classical reactors (Boiled Water Reactors (BWR), Power Water Reaction (PWR)) or due to technological incidents. In Pressurised Heavy Water Reactor (PHWR) as CANDU is, fuel is continuously changed and only technological incidents can produce significant spikes. Consequently, the releases are not constant and there are short periods of increased releases. An operational short term release is defined as a release which is larger than a normal release (i.e. higher than 2 % of 12-monthly actual or expected discharges) and occurs over a relatively short period of time (less than 1 day). For a normally uniform discharge profile, this equates to about 1 week's discharge being released in 1 day or less (IAEA, 2014b). Realistic assumptions should be used for short-term release assessments, if the annual dose exceeds 0.02 mSv (EA, 2002). The anti-nuclear groups claimed that the public dose is largely under estimated because it is based on the yearly average of the releases, disregarding the short time higher releases (Fairlie, 2010). This claim contradicts the past experience regarding 131I and 137Cs releases into the environment stating that "the integral over infinite time of the concentration of a radioactive substance in an environmental compartment per unit of release is numerically equal to its concentration at a future steady-state when that release is repeated continuously at unit rate" (Peterson et al., 1996). In order to clarify this problem, The National Dose Assessment Working Group (NDAWG) (UK) considered in detail the case of a short-term and intense emission comparable with the same total annual emission (NDAWG, 2011). It was considered the same total emission distributed on a short period (a day or less) or distributed during a year. For spike release, three cases were analysed: a) realistic - release during a normal day over 12 hours in neutral meteorological conditions (it can include rain and moderate wind direction changes, in conformity to local multiyear data); b) cautious - release of 30 minutes in neutral meteorological condition (class D), wind direction through production area, including rain; c) cautious - 30 minutes release in stable meteorological condition (class F), wind over production area, no rain. For continuous release over a year, the average UK weather was considered. For ingestion, conservative assumptions were taken: the two foods that make the greatest contribution to dose are assumed to be consumed at the 95th percentile levels and the remaining foods at the 50th percentile levels of a distribution based on national intake rates (see Table A6 in NDAWG, 2011). The methodology for calculating the dose per unit release coming from continuous releases used the following models: ADMS 4 (version 2008) (CERC, 2008) and PC CREAM (version 1997) (Mayall et al., 1997), while for spike release, the following models were used: FARMLAND (version 1995) (Brown and Simmonds, 1995) and SPADE (version 1999) (Mitchell, 1999). In parallel, for tritium, TRIF (acronym for TRitium.transfer Into Food) model (Higgins, 1997) was used. The considered tritium release is 1 TBq for the whole year (continuous) or for the spike release. For the realistic case (atmospheric stability of class D, 12 hours) the wind direction is not uniformly distributed because this case has a very low probability, but it can vary according to local statistics considering successive hours and the same stability class D. The detailed results for each radionuclide of interest and case considered (infant, child, adult) are given (see Tables A10-A21 in NDAWG, 2011) and for tritium, the results are given in Table 1 (present study). In Table 1, it is observed that a short-term emission determines an ingestion dose much higher than that of the same emission distributed during a year and in case of category F of atmospheric stability, the ingestion dose is with a factor of about 35 higher than that in the case of a continuous release. The general conclusion in the report was: "Where there are only annual limits in place, and it is cautiously assumed that discharges occur at these limits over a short period of time, then doses from the assessment of a single realistic short-term release are a factor of about 20 greater than doses from the continuous release assessment" (NDAWG, 2011). 2.3. Uncertainty of tritium dynamics for short term atmospheric exposures For OBT production in crops following a short atmospheric HTO exposure, a review of experimental data was undertaken (Galeriu et al., 2013) and various data for OBT concentration in different plants were compared based on the translocation index (TLI). TLI is defined as the percentage of OBT concentration in plant part (combustion water) at harvest related to the HTO concentration in leaf at the end of exposure. With few exceptions (cherry tomato and tangerine), TLI at night is lower than that at day. It was also observed that the transfer rate between air and leaf at night is 2-10 times lower than that at day emphasising that stomata are not fully closed at night. In WG 7 (Tritium Accidents) in the frame of IAEA EMRAS II programme, the problem of OBT formation in crops was discussed based on experimental data for winter wheat used in a models testing exercise (see Chapter 9 in IAEA, 2014a). The experiments were carried out in Germany in 1995-1996 at Kernforschungszentrum Karlsruhe (KfK now KIT) where winter wheat plants were grown in a small experimental field and short-term (one hour) HTO exposure experiments were conducted during the grain-filling stage of the wheat, at different hours of the day. The details regarding the experiments were given elsewhere (Strack et al., 1998; 2005; 2011; IAEA, 2014a). It is important to note that the soil was isolated and has very low HTO concentration and that experiment considers only the atmospheric pathway. The experimental data in wheat scenario and the analysis of models predictions were partially published (IAEA, 2014a). A detailed analysis of the data and models is presented in the present study emphasising the processes of OBT formation at day and night (see Results and Discussion). 3. Results and Discussion 3.1. Uncertainty of OBT/HTO ratios for routine releases In field conditions the equilibrium conditions are not reached and consequently, the dynamic models need detailed dynamics of HTO concentration in air. The measurements of HTO concentration in air at an hourly time step at a certain distance from a nuclear facility are very difficult and time consuming involving large financial resources. Consequently, various attempts were carried out to reconstruct the air HTO concentration from the measurements on stack emission, meteorological data and data for other radionuclides such as 85Kr for fuel reprocessing plants (Maro et al., 2016) and 41Ar for Nuclear Research Unit (NRU) (Korolevych and Kim, 2013; Korolevych et al., 2014) emitted together with tritium. The reconstruction of the air HTO dynamics is not a part of the present study. In the present study, a simulation of the dynamics of HTO concentration in air is undertaken for the CANDU reactors at Cernavoda NPP (Romania), in order to emphasise the fluctuations of the HTO air concentration and to point out the processes affecting the OBT/HTO ratio. In Romania operating two CANDU reactors, the atmospheric emission is monitored daily at both units starting from 2009. The simulated HTO air concentration is based on the daily measured emissions of both units in 2011 and on the hourly meteorological data base of the utility. Simulations of the dynamics of HTO concentration in air at various receptors were carried out, based on the atmospheric transport model ISC PRIME (Schulman et al., 1997) including the building effects, details on building dimensions and positions, and elevation of the receptor (Cernavoda is a complex site, not a flat horizontal case). For a receptor outside of the exclusion zone at 1.7 km in NW direction, the simulation of the hourly HTO concentration in air is given in Fig. 1 and large fluctuations can be observed. For this simulated example, the yearly average HTO concentration in air is 0.9 Bq m-3. In fact, a series of spikes with amplitudes of up to 100 times larger than the yearly average occurred. If the diurnal distribution (during 24 hours) is considered, the average HTO concentrations in air during the day time of 0.52 Bq m-3 is significantly lower than those during the night time (Fig. 2). It is interesting to observe that for this particular simulation (Fig. 2) the average of air HTO concentration for evening to midnight (of about 1.8 Bq m-3) is higher than that for midnight to sun rise (of about 0.98 Bq m-3). The simulations were carried out for other receptors positions showing a large variability of day/night average, but systematically, the number of hours without plume is much bigger than those with plume above the receptor. The measurements on soil HTO at the monitoring station outside of the exclusion zone show that HTO concentration in soil water are 10-60 Bq L-1 in the upper 20 cm of soil (Bucur, C, personal communication, 2016) and most of the time, the leaf water HTO concentration depends on soil HTO concentration through plant transpiration. Transpiration rate has a diurnal pattern and can be important at night (Caird et al., 2007). Crops have various root lengths during their development stages and the depth of water extraction in soil can be larger than 1 m. The HTO concentration in root soil slowly varies, but the transpiration flux has a strong diurnal pattern and finally, the HTO concentration in leaves varies significantly during the day and night time. For fluctuating HTO concentrations in air, the transfer rate between leaf and air is less than 1 h-1 during the day time and it is 2-10 times lower during the night time (Galeriu et al., 2013). Consequently, the models must consider the day and night transpiration and characteristics of root uptake in more details, in order to better assess the leaf TFWT and OBT dynamics, because the dynamics of leaf TFWT directly affects the OBT formation rate and accumulation. In fact, the OBT concentration in crops at harvest is needed for the assessment of public dose and OBT concentration is affected by the history of the crop contamination during its whole development period and mostly, in the last 1-2 months. For the same leaf HTO concentration, the OBT formation during the night time represents 1/3- 1/10 from that during the day time (Galeriu et al., 2013) and in Fig. 2, it can be observed that night air HTO concentration is higher than that at day. If a plant (or genotype of plant) has a relative high transfer rate between air and leaf (and/or high night transpiration) and the air HTO concentration at night is 2-5 times higher than that at day, it is possible that OBT production at night to be comparable with that at day. The OBT production at night was previously pointed out as a topic of interest for modelling purposes (IAEA 2008a; IAEA 2014a), but in practice, only UFOTRI (Raskob et al., 1996) and IFIN (Galeriu et al., 2000a) models considered it using calibration with experimental data for spring wheat. That calibration was useful for winter wheat and rice and in some limits, for a genotype of soybean (Raskob, 2007). There are many crops of interest and recent experimental data (Shen and Liu, 2016) shows that for at least soybean, during night time conditions, the OBT formation cannot be ignored. Only recently, a model considering OBT production at night without any calibration with experimental data was proposed (Galeriu and Melintescu, 2016) but further studies are still needed, in order to cover various crop types and finally, tests with well- designed experiments. Figs. 1 and 2 are based on simulation of air HTO concentration at an hourly time step and have illustrative purposes. Many simulations were carried out for different receptors directions and many years and the importance of root uptake (plume off) and night OBT formation remain valid. In the previous paragraphs, the importance of transpiration and night OBT production in field conditions with fluctuating air concentration on OBT/HTO ratio were pointed out. Another issue needs a further discussion: when and what has to be measured. The interest resides in crops at harvest and consequently, the plant sampling at an early development stage is not relevant. If the sampling is done when the plant is relatively mature, significant quantity of OBT is accumulated until sampling. Sampling before plume arrival results in low TFWT and high OBT stored before the plume arrival and consequently, high OBT/HTO ratio is obtained. Sampling during the plume or immediately after the maximum of the plume, results in high TFWT and high new OBT production, added to the OBT stored before the plume arrival. The new OBT, compared to the previously stored OBT, is only a fraction and in fact, OBT/HTO ratio is relatively low. It is also important what has to be measured: total OBT or NE-OBT. Interesting results were recently reported (Kim and Korolevych, 2013). For crops with high carbohydrate content, the experiments were carried out in a polluted area and after a first sampling, the crops were irrigated with water containing high HTO concentration (treated case). Weekly average of air HTO was measured and samples were analyzed for total OBT and NE-OBT. In all treaded crops, E-OBT fraction was higher than for the untreated case and for treated and untreated cases, the differences between crops were significant (reflecting crop biomass composition). On average, the E-OBT fraction was 34 % for untreated and 44 % for treated crops. These values are close to past results regarding the fraction of unbound H in carbohydrate, but are lower for crops with high protein contents (grains of cereals) or lipid content (sunflower) (Diabate and Strack, 1993). The HTO concentration in irrigation water was about 20 times higher than the average air concentration and significantly increased the total OBT at sampling time, but marginally increased the E- OBT fraction. It may be concluded that NE-OBT is a variable fraction from total OBT and depends on the plant type, irrigation conditions and sample treatment for OBT measurements. As the data base for the observed OBT/HTO contains total OBT or NE-OBT, this variability can partly explain the large range of observations, far from the equilibrium value. The main factors of variability remain the fast dynamics of TFWT and the OBT formation and its slow accumulation. The previous analysis of OBT/HTO ratio in environmental products (Korolevych et al., 2014; Thompson et al, 2015) were bulked together for all types of environmental samples taken from various sites. In the present study, the experimental data for HTO and OBT concentrations in various environmental samples taken from Wolsong NPP (Korea) (KEPCO, 2012; 2014) were analysed for the same sample type and site. The OBT/HTO ratios were separately analysed for grains (barley and rice), vegetables (cabbage and persimmon) and milk, in order to see if a dependence of the OBT/HTO ratio on HTO concentration in sample occurs. It is observed that for grains (barley and rice), the average OBT/HTO ratio is about 2, slightly higher for barley than for rice with quite large spread of data (Fig. 3). The OBT/HTO ratio for cabbage is close to 0.45 and for persimmon to 0.58 with little spread of data (Fig. 4). It is observed that the OBT/HTO ratio for cabbage and persimmon are close to the expected equilibrium value of 0.7, but for grains is 2-3 times higher. There are no data for grass, but the expected OBT/HTO ratio for milk is 0.25 – 0.4, depending on the contamination of the animal drinking water and based on the equilibrium assumption of CSA (2014a). The average OBT/HTO ratio for milk is 1.2 with a moderate spread of data (Fig. 5) and it is 3 – 5 times higher than the expected one. The results in Figs. 3 – 5 pointed out that the OBT/HTO ratio depends on the crop type and receptor location. From the previous discussion, it results that OBT/HTO ratio largely depends on receptor location, plant type, and history of the dynamics of HTO in air. Consequently, it is very difficult to model these complex processes and the models need major improvements and cannot be simple. The nuclear regulatory bodies need simple models, slightly conservative for public dose assessment and the CSA methodology (CSA, 2014a) was criticised not to be conservative (Thompson et al., 2015). In the present study, a probabilistic application of CSA model (2014a) is proposed with parameters having a probability density distribution, as it was recently reported but not for tritium (Simon-Cornu et al., 2015). The approach regarding the tritium transfer parameters taken by the CSA (2014a) is to derive the pond or well water based on the ratio between soil water and air moisture of tritium content, RFsw, with a default value of 0.3. The uncertainties in the RFsw values can largely contribute to dose uncertainty (about 28 %). A larger set of experimental values from Russia, France, Canada, India, and Japan was used elsewhere (IAEA, 2009) having a log-normal distribution with a geometric mean (GM) of 0.23 and a geometric standard deviation (GSD) of 1.7 (see Table 1 from Miscellaneous Topics in IAEA (2009)). The data in that table suggest that southern regions (e.g. France) or humid regions (e.g. Japan) may have higher values of the ratio of soil water to that of air moisture concentration. Based on these values, a default value of 0.3 is reasonable and consistent with the older recommendation of IAEA (2003). A value of 0.5 is likely to be conservative, although values as high as 1.0 are possible. Consequently, experimental values based on local measurements should be used whenever is possible or selected from general information on climate and precipitation. Tritium transfer from air to plants in the CSA study (2014a) uses the reduction factor, RFp (see equation (A2) in Annex A) with a default value of 0.68, but it has a large variability in practice. Based on the IAEA study (2009), the range of the reduction factor, RFp can be assessed considering the variability of the relative humidity, RH (see equation (A5) in Annex A) and the ratio between tritium concentration in soil water and air moisture. Combining equations (A2) and (A5) from Annex A, the dependence of the reduction factor, RFp on both the relative humidity, RH and the ratio between tritium concentration in soil water and air moisture, Csw/Cam is obtained (see equation (1)): 𝑅𝐹𝑝 = 𝑅𝐻+(1−𝑅𝐻)∗ 𝐶𝑠𝑤 𝐶𝑎𝑚 𝛾 (1) The ratio between tritium concentration in soil water and air moisture, Csw/Cam, as well as the ratio between TFWT in plants and air moisture, CTFWT/Cam for various values of relative humidity, RH is given in Table 2. The bolded values in Table 2 represent the reduction factor, RFp (in conformity with equation (1)). In Table 2, large differences for reduction factor, RFp are observed. For humid areas, the soil water is affected by rain. The scavenging ratio (i.e. the ratio between tritium concentration in rain water and that in surface air moisture) varies between 0.1 and 10 (Ota and Nagai, 2012) and rain affects the soil water concentration. The scavenging ratio depends on the distance between the emission stack and the receptor and the effective release height. For CANDU 6 reactors with low stack height, the HTO concentration in soil water is affected mostly by the HTO concentration in air moisture. Consequently, the plant water is site specific, not generic. In the absence of relevant local data, a probabilistic approach can be taken and for the reduction factor, RFp an average of 0.8 with a SD of 0.1 can be provided. The values of the other parameters involved in equation (A3) in Annex A (dry matter fraction of plant, DWp, isotopic discrimination factor, IDp, and water equivalent of organic matter, WEp) vary with the plant type. For example, based on experimental data, the dry matter fraction of terrestrial plants, DWp has a log-normal distribution with various GMs and GSDs depending on plant type (Table 3). The water equivalents factor, WEp depends on plant composition and has close GMs and GSDs for various plants (Table 4). The water equivalent factor, WEp can be bulked for all plants with a GM of 0.51 and a GSD of 1.1 (see Table 63 in IAEA (2010)). For equilibrium assumptions, the isotopic discrimination factor, IDp (see equations (A1) and (A3) in Annex A) has a range between 0.64 and 1.3 (CSA, 2014a) and a default value of 0.7 is used. It is implicitly assumed that the OBT concentration at sampling time depends on a long term average of HTO concentration in crops. The equilibrium between HTO concentrations in all environmental compartments (air, soil, leaves) is never reached in real field conditions. When the experimental data are used, the ratio between OBT in combustion water and HTO in plant water varies and a generic average value of 2 is recommended (Korolevych et al., 2014), even certain data sets indicate higher values. Based on the previous discussion regarding the OBT/HTO ratio, the isotopic discrimination factor, IDp is rarely close to the equilibrium value, depending on crop, receptor location, and the detailed dynamics of HTO concentration in air. In the absence of site specific values of OBT concentrations in crops, the present study proposes to consider a log-normal distribution for the discrimination factor, IDp (= Rp) with a minimum value of 0.3, a maximum value of. 10, a GM of 1.732, a GSD of 1.77, and a CV of 0.62. The average value is close to the recommended value of 2 (Korolevich et al., 2014), but the maximum value allows a conservative estimate. In order to emphasise the importance of parameters uncertainty, in the present study the case of Cernavoda NPP (Romania) is considered, based on a conservative diet and significant local food consumption. The yearly average concentration of HTO in air at the location of interest is assumed to be 1 Bq m-3. Several scenarios based on the age groups proposed by ICRP (2006) were considered (infant, child and adult):   OBT/HTO ratio has the equilibrium value of 0.7 as in CSA (2014a) and the tritium concentration in the animal drinking water is the same with that of drinking water for humans (4.5 Bq L-1). The results are given in Table 5. OBT/HTO ratio has the equilibrium value of 0.7 as in CSA (2014a) and the tritium concentration in the animal drinking water is 9 times higher than that of drinking water for humans (i.e. pond water with a tritium concentration of 40 Bq L-1)., close with the conservative recommendation in CSA. The results are given in Table 6.  OBT/HTO ratio has a conservative value of 10 and the tritium concentration in the animal drinking water is the same with that of drinking water for humans (4.5 Bq L-1). The results are given in Table 7. In Tables 5-7, it can be observed that the uncertainty introduced by the OBT/HTO ratio is comparable with that introduced by tritium concentration in animals' drinking water and the conservative dose for public is only 2 µSv year-1 for a yearly average concentration of HTO in air of 1 Bq m-3. In all three scenarios the food production was maximised, as well as food intake. In fact, it is important to have site specific food production (16 sectors and radius of 1-5 km) and the consumption of local food. Based on site specific information, the exaggerate conservatism is avoided. In a PHWR as CANDU is, 97 % of atmospheric emission is HTO, during the day time and night time. A CANDU reactor differs from the SRBT case (Mihok et al., 2016) where the emissions are predominately as HT, restricted to working hours (7.00 - 17.00), and stopped during the weekends. There is no such a facility in Romania, but some experience for detritiation facility and other tritium bearing utility exists. There are some peculiarities at SRBT which potentially can explain the results of Mihok et al. (2016) and each of them will be shortly discussed in the following. The experimental errors can be eliminated because special quality assurance was taken (St-Amant, 2016). HT is very little oxidised in contact with the air or leaves (Ichimasa et al., 1999) but strongly oxidised in soil (Ichimasa et al., 1999; Ota et al., 2007) enhancing the HTO concentration in root soil. The increased HTO concentration in the transpiration flow due to HT oxidation in soil is moderate and cannot explain the experimental data provided by Mihok et al. (2016). For tritium handling facilities, various species of soluble tritiated organics were reported (Belot et al., 1993; 1995) and tritiated formaldehyde can be incorporated in leaves increasing the OBT concentration (Belot et al., 1992). Tritiated methane is also a potential substance contributing to leaf OBT (Amano, 1995; Kakiuchi et al., 2002). It seems that at SRBT, the emissions of other tritiated substances than HT and HTO have a very low probability, but cannot be fully excluded (MacDonald, J, personal communication, 2016) and more accurate measurement can be carried out in the future for clarifying this potential cause. The soil around SRBT has a high concentration of OBT due to high tritium emissions occurred in the past and a rain splash process can contaminate the plants (Dunne et al., 2010). The rain amount and intensity was low during the experiment (Clark et al., 2014) and rain splash is not enough to explain the results. If no other impurities excepting HT and HTO were emitted, the explanation might be the fact that the OBT formation in crops is not yet well understood and the present models have large uncertainties for OBT predictions. 3.2. Influence of spike releases in routine emissions The results regarding tritium contamination of foodstuff in case of spike releases considered in the study of NDAWG (2011) are based only on UK standard models. In parallel, TRIF model was investigated. TRIF model (Higgins, 1997) has many compartments with various transfer rates fixed at generic values. It considers only NE-OBT and assumes that there is a loss of plant OBT to HTO with a half time of 10 days. The foodstuffs considered in TRIF model are green and root vegetables, fruits, cow milk, cow meat and liver, sheep meat and liver. TRIF model does not consider the cereals. The results of NDAWG study (2011) for realistic case (class D, 12 hours) can be compared with TRIF model considering integrated food contamination for a similar deposition of tritium of 1240 Bq m-2 and are given in Table 8. In Table 8, it can be observed that NDAWG results (2011) are about 2 times more conservative than TRIF model results (Higgins, 1997). When TRIF model is compared with SA model for routine release, it gives an activity about 30 % higher in food products. NDAWG (2011) finally selected the most conservative results. The routine releases were considered as a special task in the frame of an IAEA coordinated research programme: Biosphere Modelling and Assessment (BIOMASS) (IAEA, 2003). Based on data from Canada, France and Russia, BIOMASS programme provided inter- comparison results between models and observations, including OBT in plants, following chronic atmospheric releases (IAEA, 2003). The final BIOMASS report (IAEA, 2003) concluded that:  The conceptual model of OBT formation may need to be rethought for sites that are subject to a fluctuating airborne plume. None of the models was able to simulate the observed increase in plant OBT concentration with time;  Unlike steady state releases, a non-planned release detected at the stack may not have an effect on all the media compartments. A model inter-comparison exercise was carried out for a hypothetical tritium accident with one hour emission of 3700 TBq (Guetat and Patryl, 2008). The atmospheric transport was imposed to all participants and the radiological doses coming from food ingestion varied on a large range for class D (with rain) and class F (night). If the release is decreased to 1 TB, the radiological doses have a range between 0.27 and 3.4 µSv year-1 for cautious class D and a range between 1.1 and 17 µSv year-1 for cautious class F. The values in Table 1 are in these ranges but close to the maximum values. In the present study, the same conditions as in NDAWG (2011) were considered, but different models for spike or continuous emissions were used. For spike releases, the UFOTRI model (Raskob, 1993) and an upgraded version of IFIN-HH model (Melintescu and Galeriu, 2005) were used but with a time step of one hour. The IFIN model (Melintescu and Galeriu, 2005) is a partially improved version of the tritium module developed in the frame of Real time On-line DecissiOn Support System (RODOS) project (Galeriu et al., 2000a; 2000b). For a normal day case (class D, 12 hours), the hourly meteorological data from Cernavoda NPP (Romania) were used and a case where wind changes directions up to 1500 was selected. For cautious cases (class D and F), the wind direction was chosen to cover the sector with the maximum food production. Recommendations of CSA for accidental releases (see sub-chapters 5.3.3. and 5.3.4 in CSA, 2014b) were followed. For ingestion, as in NDAWG (2011), the most important food items (milk and vegetables) were considered at 95 % from the local intake distribution data and the local consumption was maximised. For continuous release, CSA model (2014a) was used (in the probabilistic approach proposed with conservative assumptions), as well as NORMTRI model (version 1999) (Raskob, 1994). In Table 9, it can see that the occurrence of spikes in routine emissions increases the public dose, but predictions cover a large range, reflecting the assumptions considered by the models. In USA, for routine emissions of tritium, CAP88-PC model is used and gives a public dose 3 times higher than NORMTRI model (Imboden and Overcamp, 2005). All models assessing the public dose needs the air HTO concentration, in order to derive HTO and OBT in crops at harvest and the atmospheric transport calculations are subject to large uncertainty. Comparing the results of CAP88-PC, NORMTRI, and ISC-PRIME models for atmospheric transfer with the measured yearly average of air HTO concentration, those models give good predictions, but for monthly or biweekly average of air HTO concentration, large miss- predictions are observed (Michelotti et al., 2013; Galeriu and Melintescu, 2012). As it was previously discussed in the present study, plant HTO and OBT concentrations depend on air HTO concentration during the previous day (HTO) or month (OBT). Based on the previous considerations, it is clear that the OBT dynamics in plants is a complex process and equilibrium conditions are never attained in practice. Plants used for human consumption have different harvest times and development stages. A spike release can occur any time during the year and crop contamination depends on its development stage at the day of release. For Romanian diet and crops near Cernavoda NPP (Romania), the day of release was varied at each mid-month and the short and intense emission was supposed to occur near mid- day. The IFIN tritium model (Melintescu and Galeriu, 2005) was used in that case and a significant seasonal pattern of the public dose was observed with a maximum value in September (when maize is harvested) and a minimum value in winter (Fig. 6). Local habits regarding food and feed are very important and can vary in each country, as well as the share of the local production in the diet. The Romanian results must be considered as an illustrative example only. In Romania, the atmospheric emission of both CANDU reactors is about 400 TBq y-1 in 2011 - 2015. The HT and HTO releases at stack are monitored daily and monthly values of HTO concentrations in air, water and food are measured at various receptors. The HT emission represents 1-3 % of the total emission. Public dose is assessed in two ways: using conservative estimate of dispersion factors and Canadian standard (CSA, 2014a) or using only monitoring results. When the CSA (2014a) is used, the public dose is higher than when the monitoring results are used, but never exceeds 10 μSv y-1. For a yearly emission of 1 TBq, the public dose is 0.025 µSv year-1, a value much lower than those in Table 9 (fifth column). For all period from 2009 until present, a single significant spike was detected (6 times higher than the weekly average), but it did not occur in the vegetation period and the dose for public marginally increased. The assertion of NADWG (2011) that the public dose from a short term release is much higher than those from the same amount emitted but for the whole year, must be reanalysed considering the high uncertainty of the atmospheric transport for a short term (hour, week), the non-equilibrium processes in field conditions, the mobility of tritium, the fast tritium transfer from air to leaf, and the slow accumulation of OBT, as well as the tritium recycling. Routine models considering equilibrium conditions and simple atmospheric transport models involve large uncertainty. The tritium reemission from soil and plant, as well as respiration process, produce a secondary tritium plume. The OBT accumulated in soil decomposes slowly and is recycled, involving that the processes must be modelled on large space and time domains (but with fine time steps and space grid) and moreover, that all processes are well understood. The uncertainty due to OBT production in crops is only a part of the problem. Due to the insufficient understanding of processes involved in tritium transfer in environment, it is not possible to assess the public dose with low uncertainty and moderate conservatism. Considering the previous results and analysing the present capability of the models to predict the consequences of routine or spike release, it is obvious that further studies are needed in order to clarify if tritium transfer in environment differs from other radionuclides transfer like iodine and caesium and a spike release of tritium gives higher dose than models predict. 3.3. Uncertainty of tritium dynamics for short term atmospheric exposures In wheat scenario (Chapter 9 in IAEA, 2014a), the experimental data on HTO and OBT concentrations in leaves and ears (grains) were used for a series of one hour exposure in an experimental chamber, at different hours of the day. During the exposure, the air HTO concentration, temperature, light and relative humidity were measured. At the end of exposure, the chamber was open, and samples of leaves and ears were taken immediately, but also on the subsequent hours including the normal harvest time (Strack et al., 1998; 2005; 2011). The HTO and NE-OBT data were organised in a data base for each experiment containing the start of one hour exposure, HTO and OBT in leaves and ears at the end of exposure, and the subsequent samplings, including the last one at normal harvest. The data base contains 14 experiments starting at 7.00 in the morning and ending at 23.00 in the evening (denominated as s7, s8, s10, s11, s20p, s23p). In the present study, based on that data base, the OBT and HTO concentrations in leaves at each sampling were analysed in order to assess the dynamics of OBT/HTO ratio in the first 24 hours after the end of exposure (Fig. 7). In Fig. 7, it can be observed that the OBT/HTO ratio has very low values at the end of each exposure and gradually increases up to a value of 10 after 24 hours, demonstrating that the process of OBT formation is complex and not fast and the OBT/HTO ratio strongly depends on sampling time after exposure. In wheat scenario (IAEA, 2014a), the initial conditions for the one hour exposure were used and the modellers were asked to predict, among others, the OBT concentration in grain (ear) at the end of exposure and its dynamics for the next 24 hours and at the harvest (Bq kg-1). Four countries/models, participated at this scenario: - JAEA (Japan Atomic Energy Agency, Japan) (with SOLVEG model ), a complex research grade model (Ota and Nagai, 2011); - CEA (Commissariat of Atomic Energy, France) (with CERES model), a simple model with a constant exchange velocity for day or night and with OBT production depending on the integrated leaf HTO concentration (Patryl and Armand, 2005; 2007); - IFIN ("Horia Hulubei" National Institute for Physics and Nuclear Engineering, Romania) (with FDMH+ model), a research grade and operational model of moderate complexity (Galeriu et al., 2000a; 2000b; Melintescu et al., 2002; Melintescu and Galeriu, 2005), with an intermediate upgraded version (Galeriu et al., 2009b; Melintescu and Galeriu, 2011); - KIT (Karlsruhe Institute of Technology, Germany) (with UFOTRI model), a research grade and operational model of moderate complexity (Raskob, 1993), with an intermediate upgraded version (PLANT-OBT) (Raskob et al., 1996; Raskob, 2011). The OBT concentration in grain at harvest is the most important endpoint and modellers were asked to submit predictions for many observations and time steps, in order to detect the potential compensatory errors and to help the models improvements. For grain at harvest the predicted to observed ratio (P/O) varies between models (Fig. 8). At harvest, the JAEA model exhibited a large spread in its P/O ratios for grain OBT concentrations with practically no OBT in grain predicted at night. The KIT model over- estimated most observations, while the other models (CEA and IFIN) mostly under-estimated. The models predictions are acceptable for that exercise (see Chapter 9 in IAEA, 2014a). For leaf OBT concentration at the end of exposure, KIT model is close to experimental data, but CEA and IFIN models under estimate the experimental data by a factor of 2-10 (Fig. 9). JAEA model gives good predictions excepting the night case (Fig. 9). For leaf OBT at harvest, the models predictions have large variations and any models results are not able to satisfy the statistical criteria for an acceptable performance (Fig. 10). The IFIN model does not predict practically any OBT concentration in leaf at harvest and consequently, the model must be improved for a better understanding of processes. The experimental data on leaf and ear HTO concentration and grain OBT concentration at harvest were analysed (Strack et al., 2011) and an empirical relationship between OBT concentration in grain at harvest (water of combustion) and integrated HTO concentration in leaves and ears was found distinguishing between OBT produced during the day and night time (Strack et al., 2011): 𝐶𝑂𝐵𝑇−𝑔𝑟𝑎𝑖𝑛 = 0.48 ∗ [𝐼𝐶𝑙𝑒𝑎𝑓−𝑑𝑎𝑦 + 0.2 ∗ 𝐼𝐶𝑙𝑒𝑎𝑓−𝑛𝑖𝑔ℎ𝑡 + 0.5 ∗ (𝐼𝐶𝑒𝑎𝑟−𝑑𝑎𝑦 + 0.2 ∗ 𝐼𝐶𝑒𝑎𝑟−𝑛𝑖𝑔ℎ𝑡)] (2) where: COBT-grain is OBT concentration in grain at harvest (Bq mL-1); ICleaf-day, ICleaf-night, ICear- day, ICear-night are HTO integrated concentrations in leaf and ear during the day and night time, respectively (kBq h mL-1); 0.48 is a regression coefficient. The equation (2) indicates that the ear contributes with a half to the overall OBT production in grain at harvest and the OBT produced during the night time is 0.2 of that produced during the day time. The constant slope in equation (2), 0.48 can be justified for the experimental conditions where the grain filling is linear. The value of the slope itself, 0.48 is a characteristic of the German wheat cultivar. It is well known that for cereals, the ear contributes to the photosynthesis in the period before the yellow ripe (Goudriaan and van Laar, 1994) and the 0.5 value in equation (2) is an average for the period considered in the experiment (days 17 – 28 after anthesis and the harvest was 47 days after anthesis). The equation (2) clearly demonstrates the importance of the HTO concentration in leaves (and ears for cereals) for the OBT production. All the models participated to wheat scenario consider the dependence of OBT production on leaf HTO concentration but based on different approaches. In the complex model SOLVEG (Ota and Nagai, 2011) the OBT is produced in leaves in a carbohydrate sub- model depending on leaf HTO concentration and photosynthesis rate. The OBT production rate depends on plant type and growing stage, also. The SOLVEG underestimation of OBT production at night could be due either to very low stomatal conductance at night or to carbohydrate sub-model. The KIT model (PLANT-OBT) (Raskob et al., 1996; Raskob, 2011) considers the dependence of OBT production on HTO concentration in leaf and ear but the model needs too many parameters calibrated for crop type (wheat). The standard KIT model (UFOTRI) (Raskob, 1993) considers the dependence of OBT production on only HTO concentration in leaf, plant type and development stage and includes separately the OBT production at night calibrated with experimental data. UFOTRI was successfully tested for other crops as rice and soybean (Raskob, 2007). The IFIN model (Galeriu et al., 2009b; Melintescu and Galeriu, 2011) considers the dependence of OBT production on HTO concentration in leaf but indirectly considers ear in photosynthesis increasing the leaf area index (LAI). IFIN model as UFOTRI uses calibration for OBT night production. IFIN model considers fast and slow maintenance respiration, dry matter and OBT partition for different plant parts, but largely under-estimates the leaf OBT concentration at harvest. The simple CEA model (CERES) (Patryl and Armand, 2005; 2007) considers a linear dependence of OBT concentration in grains on the integrated HTO concentration in leaf. The slope of that dependence can be adjusted by the duration of crop growth. The CEA model does not consider the crop type and the development stage. It considers a prescribed value of leaf resistance for day (300 s m-1) and night (3000 s m-1). For wheat scenario, CEA model considers the duration of grain growth and not the duration of plant growth. The model is not documented enough in order to explain the OBT dynamics in leaves for cereals. It is difficult to clarify which model depending on its level of complexity better predicts an accidental situation with tritium emission, based only on the experimental data sets for a single crop grown in specific environmental conditions. Consequently, many tests are necessary for model selection covering many plant types and various environmental conditions. The present study analyses the experimental results regarding NE-OBT but in many cases, the measurements consider total OBT. Considering the field conditions at Perch Lake plot garden of Atomic Energy of Canada Limited (AECL) (Kim and Korolevych, 2013), the ratio between NE-OBT and total OBT of various vegetables (cabbage, lettuce, tomato, radish, and beet) varies between 0.46 and 0.74. There are no data immediately after an acute exposure of wheat for total OBT and the results in Fig. 7 emphasise that the formation of NE- OBT is a long process starting with low values of OBT/HTO ratio. A lower value of NE- OBT/Total-OBT ratio can be suspected immediately after the plume passage. Further dedicated controlled experiments would be necessary. 4. Conclusions The variability of OBT/HTO ratio on a large range is a result of no equilibrium conditions in real field situations and of the difference between the HTO and OBT dynamics in crops. Plants used for human consumption have different harvest times and development stages. The OBT/HTO ratio at sampling time depends on the tritium dynamics at stack emission, atmospheric transport at receptor location and crop type. A general dependence of OBT/HTO ratio on HTO concentration in crops cannot be observed for experimental data at Wolsong NPP (Korea), but trends for each crop type and receptor position (KEPCO, 2012; 2014) are foreseen. The Canadian Standard (CSA, 2014a) assuming equilibrium conditions can be used but in a probabilistic approach or in a deterministic case with an increased value of isotopic discrimination factor, IDp. For routine releases, the public dose is still very low (with realistic local habits and food production) and the uncertainties of the model predictions have no consequences on any health effects. The SRBT case (Mihok et al., 2016) is not specific to conditions of a CANDU reactor and can be affected by the emission of organic tritiated species (to be tested), and the uncertainty of the current models (still not crop specific). The spike release influence on the yearly public dose cannot be ignored and is due to non-equilibrium processes in real field conditions. If the spike coincides with the period when most of the crops growing around the nuclear site are about one month before harvest, the food contamination for human consumption can be higher than in the case of a uniform emission (the same emission as in the spike release) during a year. In temperate climate, there is no plant growth in winter. The uncertainty of the present dynamic models does not allow a quantitative assessment. The analysis of wheat experiments (Strack et al., 1998; 2005; 2011) for a short term but intense atmospheric emission emphasises that the OBT formation is a long process (about 24 hours), there are some differences between night and day releases and the leaf (and ear) HTO dynamics is more important than it was previously assumed. That experiment considers the cereals in their linear grain filling period and the models results cannot be generalised for other crops, but point out the models limitations. The requirements for a robust operational model (see Introduction) are not accomplished by any model participated to wheat scenario. Due to the complexity of processes involved in OBT formation and HTO dynamics, further studies are still needed. The efforts must be oriented to a better understanding of the root and air pathway balance on HTO dynamics subject to various environmental conditions and to the potential role of OBT production at night. Further studies highly depend on each country nuclear energy policy, nuclear regulatory body requirements on model uncertainty and mostly, on the available resource. A large international collaboration is needed, directly between various teams or coordinated by IAEA as practiced from many years. Acknowledgements This study is part of the project EXPLORATORY IDEAS 65/2011 financed by the Romanian Authority for Scientific Research. The excellent collaboration with the Radioprotection team of CANDU NPP (Romania) is highly appreciated. The past and present exchange of ideas with WG 7 members of EMRAS II programme is acknowledged. We thank to Dr. Siegfried Strack (Karlsruhe Institute of Technology, Germany, retired) for providing us full access to his data base on wheat experiments. Last but not least, we want to thank to Dr. Sylvie-Ring Peterson (Lawrence Livermore National Laboratory, USA, retired) and Dr. Gretchen Gallegos (Lawrence Livermore National Laboratory, USA) for their unconditional help and useful suggestions to improve the present manuscript. Appendix A CSA and IAEA guidance for tritium in crops The assessment of public dose following an atmospheric routine release of 14C, 36Cl and 3H was carried out by IAEA in its coordinated research studies (IAEA 2009; 2010) based on specific activity (SA) approach used for equilibrium conditions. SA is defined as the radionuclide activity per mass of the stable element. The SA approach for equilibrium conditions is also used by CSA (2014a). CSA (2014a) proposed the following equations for OBT and tissue free water tritium (TFWT) concentrations in plants on a fresh weight basis: but 𝐶𝑂𝐵𝑇 = 𝑅𝐹𝑝∗𝐷𝑊𝑝∗𝐼𝐷𝑝∗𝑊𝐸𝑝∗𝐶𝑎𝑖𝑟 𝐻𝑎 𝐶𝑇𝐹𝑊𝑇 = 𝐶𝑎𝑖𝑟 ∗ 𝑅𝐹𝑝 𝐻𝑎 = 𝐶𝑎𝑚 ∗ 𝑅𝐹𝑝 (A1) (A2) Consequently, the following equation for OBT concentration in plants is obtained: 𝐶𝑂𝐵𝑇 = 𝐷𝑊𝑝 ∗ 𝐼𝐷𝑝 ∗ 𝑊𝐸𝑝 ∗ 𝐶𝑇𝐹𝑊𝑇 (A3) where: RFp is the reduction factor used because tritium concentration in soil is much smaller than that in air (unit less); DWp is the dry matter (dm) fraction of plant (kg dry weight (dw) plant kg–1 fresh weight (fw) plant); IDp is the isotopic discrimination factor due to plant physiology (unit less); WEp is the water equivalent of organic matter (at combustion in OBT measurements); Ha is the absolute atmospheric humidity (L m–3); Cair is the HTO concentration in air (Bq m-3); Cam is the HTO concentration in air moisture (Bq m-3); CTFWT is the HTO concentration in the leaf free water (Bq L-1) The IAEA approach (IAEA, 2009) uses an equation similar to equation (A3) for OBT concentration in plants, but with somewhat different notation: 𝐶𝑂𝐵𝑇 = (1 − 𝑊𝐶𝑝) ∗ 𝑊𝐸𝑄𝑝 ∗ 𝑅𝑝 ∗ 𝐶𝑇𝐹𝑊𝑇 (A4) where: WEQp is the water equivalent factor (kg of water produced per kg dm combusted) - WEQp in equation (A4) is the same as WEp in equation (A3); WCp is the fractional water content of the plant (L kg-1) – 1- WCp in equation (A4) is the same as DWp in equation (A3); Rp is the partition factor for plants – Rp in equation (A4) is the same as IDp in equation (A3). IAEA (2009) recommends the following equation for tritium concentration in plant water, with an explicit distinction between air and root soil pathway: 𝐶𝑇𝐹𝑊𝑇 = 𝑅𝐻∗𝐶𝑎𝑚+(1−𝑅𝐻)∗𝐶𝑠𝑤 𝛾 (A5) where: CTFWT is the HTO concentration in the leaf free water (Bq L-1); RH is the relative humidity of air; Cam is the HTO concentration in air moisture (Bq L-1); Csw is the HTO concentration in the soil water (Bq L-1); γ (= 0.909) is the ratio of the HTO vapour pressure to that of H2O. The ratio between soil water, Csw and air moisture, Cam can vary between various sites and a default value of 0.3 is recommended by IAEA (2009, 2010). For equilibrium conditions and considering OBT as water of combustion, the following equations are obtained: CSA (2014): 𝑂𝐵𝑇 𝑇𝐹𝑊𝑇 = 𝐼𝐷𝑝 (total OBT) (A6) IAEA (2009): 𝑂𝐵𝑇 𝑇𝐹𝑊𝑇 = 𝑅𝑝 (NE-OBT) (A7) The values of the isotopic discrimination factor, IDp given in various studies (McFarlane, 1976; Garland and Ameen, 1979; Kim and Baumgartner, 1988; Hisamatsu et al., 1989; Inoue and Iwakura, 1990; Kotzer and Workman, 1999) range from 0.64 to 1.3. Based on that narrow range, the arithmetic mean of 0.8 should be used as the default value (CSA, 2014a).The partition factor, Rp accounts for the reduction in dry weight (dw) concentration due to the presence of exchangeable hydrogen in combustion water, as well as for isotopic discrimination factor. The values of the partition factor, Rp must be determined empirically for steady-state conditions. Although the number of data points is small, all values are less than one, with a GM of 0.54 for the crops considered (maize, barley and alfalfa). In the absence of other information, this value is assumed to apply to all plant types. Regardless of the plant in question, the TFWT concentration used in Equation (A4) should be the concentration in the plant leaves, the primary location of dry matter production (IAEA, 2009). References Amano, H., 1995. Preliminary measurement on uptake of tritiated methane by plants. Fusion Technol. 28, 797-802. ASN, 2010. Livre blanc du tritium. Autorite de Surete Nucleaire (in French). http://www.asn.fr/sites/tritium/fichiers/Tritium_livre_blanc_integral_web.pdf. Barry, P.J., Watkins, B.M., Belot, Y., Etlund, O., Galeriu, D., Raskob, W., Russel, S., Togawa, O., 1999. Intercomparison of model predictions of tritium concentration in soil and foods following acute airborne HTO exposure. J. Environ. Radioact. 42, 191-207. Belot, Y., Camus, H., Marini, T., 1992. Determination of tritiated formaldehyde in effluents from tritium facilities. Fusion Technol. 21, 556-559. Belot, Y., Camus, H., Marini, T., Raviart, S., 1993. Volatile tritiated organic acids in stack effluents and in air surrounding contaminated materials. J. Fusion. Energ. 12, 71-75. Belot, Y., Camus, H., Raviart, S., Antoniazzi, A.B., Shmayda, W.T., 1995. Production of tritiated organic acids at tritium-bearing stainless steel surfaces exposed to air. Fusion Technol. 28, 1138-1143. Brown, J., Simmonds, J.R., 1995. FARMLAND – A dynamic model for the transfer of radionuclides through terrestrial foodchains. NRPB-R-273. National Radiological Protection Board, Chilton, UK. Caird, M.A., Richards, J.H., Donovan, L.A., 2007. Nighttime stomatal conductance and transpiration in C3 and C3 plants. Plant Physiol. 143, 4-10. CERC, 2008. ADMS Version 4. Cambridge Environmental Research Consultants. http://www.cerc.co.uk/ Clark, I., Mihok, S., Wilk, M., Lapp, A., Dehay-Turner, B., St-Amant, N., Kwamena, N., 2014. HTO – OBT dynamics in plants grown in a high HT environment. 3rd Organically Bound Tritium (OBT) Workshop, September 15 -18, 2014, Ottawa, Canada. CNSC, 2013. Environmental fate of tritium in soil and vegetation. Part of the Tritium Studies Project. Canadian Nuclear Safety Commission. https://www.cnsc- ccsn.gc.ca/eng/pdfs/Reading-Room/healthstudies/Environmental-Fate-of-Tritium-in-Soil-and- Vegetation-eng.pdf. CSA, 1987. Guidelines for calculating derived release limits for radioactive material in airborne and liquid effluents for normal operation of nuclear facilities. Canadian Standards Association Group. CAN/CSA N288.1-M87, Mississauga, Canada. CSA, 2008. Guidelines for calculating derived release limits for radioactive material in airborne and liquid effluents for normal operation of nuclear facilities. Canadian Standards Association Group. CSA N288.1-08, Mississauga, Canada. CSA, 2014a. Guidelines for calculating derived release limits for radioactive material in airborne and liquid effluents for normal operation of nuclear facilities. Canadian Standards Association Group. CSA N288.1-14, Mississauga, Canada. CSA, 2014b. Guidelines for calculating the radiological consequences to the public of a release of airborne radioactive material for nuclear reactor accidents. Canadian Standards Association Group. CSA N288.2-14, Mississauga, Canada. Davis, P.A., Kim, S.B, Chouhan, S.L., Workman, W., J., G., 2005. Observed and modeled tritium concentrations in the terrestrial food chain near a continuous atmospheric source.Fusion Sci. Technol. 48, 504-507. Diabate, S., Strack, S., 1993. Organically bound tritium. Health Phys. 65, 698-712. Dunne, T., Malmon, D.V., Mudd, S.M., 2010. A rain splash transport equation assimilating field and laboratory measurements. J. Geophys. Res. 115, F01001, doi:10.1029/2009JF001302 EA, 2002. Principles for the assessment of prospective public doses arising from authorised discharges of radioactive waste to the environment. Radioactive Substances Regulation under the Radioactive Substances Act (RSA-93) or under the Environmental Permitting Regulations (EPR-10). Environment Agency, UK. https://www.gov.uk/government/uploads/system/uploads/attachment_data/file/296390/geho12 02bklh-e-e.pdf. http://publications.environment-agency.gov.uk/pdf/PMHO1202BKLH-e- e.pdf. Fairlie, I., 2010. Hypothesis to explain childhood cancer near nuclear power plants. Int. J. Occup. Environ. Health 16, 341–350. Galeriu, D., 1994. Transfer parameters for routine release of HTO – Consideration of OBT. AECL-11052. Atomic Energy of Canada Limited, Chalk River, Ontario, Canada. Galeriu, D., Raskob, W., Melintescu, A., Turcanu, C., 2000a. Model description of the Tritium Food Chain and Dose Module FDMH in RODOS PV 4. RODOS (WG3)-TN(99)-54. Galeriu, D., Raskob, W., Melintescu, A., Turcanu, C., 2000b. Documentation of tritium food chain and dose module FDMH in RODOS PV4. RODOS(WG3)-TN(99)-56. Galeriu, D., Melintescu, A., Slavnicu, D., 2009. IFIN - Cernavoda NPP: Consultancy on Canadian Standard CSA-N-288.1-2008 in conformity with norms in European Community and the IAEA guidance for assessing Derived Release Limits for 3H and 14C (in Romanian) Galeriu, D., Melintescu, A., Slavnicu, D., Gheorghiu, D., Simionov, V., 2009. Accidental release of tritiated water - toward a better radiological assessment. Radioprotection 44, 177 – 183. Galeriu, D., Melintescu, A., 2010. Tritium. In: Atwood, D.A. (Ed.), Radionuclides in the environment. John Wiley & Sons Ltd., West Sussex, England, pp. 47-65. Galeriu, D., Melintescu, A., 2012. Research and development of environmental tritium modelling, an update. The 57th Annual Meeting of the Health Physics Society, 22-26 July 2012, Sacramento, CA. http://hpschapters.org/sections/envrad/2012AMpresentations/HP2012Galeriu.pdf. Galeriu, D., Melintescu, A., Strack, S., Atarashi-Andoh, M., Kim, S.B., 2013. An overview of organically bound tritium experiments in plants following a short atmospheric HTO exposure. J. Environ. Radioact. 118, 40-56. Galeriu, D., Melintescu, A., 2016. Relevance of night production of OBT in crops. Oral presentation at 11th International Conference on Tritium Science and Technology (TRITIUM 2016) (http://www.ans.org/meetings/m_213, http://tritium2016.org/), accepted to Fusion Sci. Technol. FST16-174R1. Garland, J.A., Ameen, M., 1979. Incorporation of tritium in grain plants. Health Phys. 36, 35- 38. Goudriaan, J., van Laar, H.H., 1994. Modelling potential crop growth processes. Kluwer Academic Publishers, Dordrecht, The Netherlands. Guetat, P., Patryl, L., 2008. Environmental and radiological impact of accidental tritium release. Fusion Sci. Technol. 54, 273-276. Higgins, N.A., 1997. TRIF - An intermediate approach to environmental tritium modelling. J. Environ. Radioact. 36, 253-267. Hisamatsu, S., Takizawa, Y., Itoh, M., Ueno, K., Katsumata, T., Sakanoue, M., 1989. Fallout 3H in human tissue at Akita, Japan. Health Phys. 57, 559-563. IAEA, 2003. Modelling the environmental transport of tritium in the vicinity of long term atmospheric and sub-surface sources. IAEA-BIOMASS-3. International Atomic Energy Agency, Vienna. IAEA, 2008a. Modelling the Environmental Transfer of Tritium and Carbon-14 to Biota and Man. Report of the Tritium and Carbon-14 Working Group of EMRAS Theme 1. Environmental Modelling for RAdiation Safety (EMRAS) Programme. International Atomic Energy Agency Vienna. http://www-ns.iaea.org/downloads/rw/projects/emras/draft-final- reports/emras-tritium-wg.pdf. IAEA, 2008b. The Pickering Scenario. Final report. Working Group on Modelling of Tritium and Carbon-14 Transfer to Biota and Man. Environmental Modelling for RAdiation Safety (EMRAS) Programme. International Atomic Energy Agency Vienna. http://www- ns.iaea.org/downloads/rw/projects/emras/tritium/pickering-final.pdf. IAEA, 2009. Quantification of radionuclide transfer in terrestrial and freshwater environments for radiological assessments. IAEA-TECDOC-1616. International Atomic Energy Agency, Vienna. IAEA, 2010. Handbook of parameter values for the prediction of radionuclide transfer in terrestrial and freshwater environments. Technical reports series no. 472. IAEA-TRS-472. International Atomic Energy Agency, Vienna. IAEA, 2014a. Transfer of tritium in the environment after accidental releases from nuclear facilities. Report of Working Group 7 of the IAEA's Environmental Modelling for Radiation Safety (EMRAS II) Programme. IAEA-TECDOC-1738. International Atomic Energy Agency Vienna IAEA, 2014b. Radiation protection and safety of radiation sources: International basic safety standards. General safety requirements Part 3 No. GSR Part 3. International Atomic Energy Agency Vienna. ICRP, 2006. Assessing dose of the representative person for the purpose of the radiation protection of the public. ICRP Publication 101a, Annals of the ICRP 36 (3). International Commission on Radiological Protection, Oxford, Elsevier Science. Ichimasa, M., Suzuki, M., Obayashi, H., Sakuma, Y., Ichimasa, Y., 1999. In vitro determination of oxisation of atmospheric tritium gas in vegetation and soil in Ibaraki and Gifu, JAPAN. J. Radiat. Res. 40, 243-251. Imboden, S.F., Overcamp, T.J., 2006. Chronic dose due to a continuous tritium release calculated by CAP88-PC and NORMTRI. Nucl. Technol. 155, 114-118. Inoue, Y., Iwakura, T., 1990. Tritium concentration in Japanese rice. J. Radiat. Res. 31, 311- 323. Kakiuchi, H., Andoh, M.A., Amano, H., 2002. The evaluation of uptake of tritiated methane to plants. International symposium: Transfer of radionuclides in biosphere. Prediction and assessment, Mito, Ibaraki, Japan, 18 – 19 December 2002. http://www.iaea.org/inis/collection/NCLCollectionStore/_Public/36/116/36116269.pdf. KEPCO, 2012. Environmental monitoring and evaluation report around nuclear power plants. Annual report. Korea Electric Power Corporation (in Korean). KEPCO, 2014. Environmental monitoring and evaluation report around nuclear power plants. Annual report. Korea Electric Power Corporation (in Korean). Kim, M.-A., Baumgartner, F., 1988. Validation of tritium measurements in biological materials. Fusion Technol. 14, 1153-1156. Kim, S.B., Baglan, N., Davis, P.A., 2013. Current understanding of organically bound tritium (OBT) in the environment. J. Environ. Radioact. 126, 83-91. Kim, S.B., Workman, W.J.G., Davis, P.A., 2008. Experimental investigation of buried tritium in plant and animal tissues. Fusion Sci. Technol. 54, 257-260. Kim, S.B., Korolevych, V., 2013. Quantification of exchangeable and non-exchangeable organically bound tritium (OBT) in vegetation. J. Environ. Radioact. 118, 9-14. Korolevych, V.Y, Kim, S.B., 2013. Relation between the tritium in continuous atmospheric release and the tritium contents of fruits and tubers. J. Environ. Radioact. 118, 113-120. Korolevych, V.Y., Kim, S.B., Davis, P.A., 2014. OBT/HTO ratio in agricultural produce subjected to routine atmospheric releases of tritium. J. Environ. Radioact. 129, 157-168. Kotzer, T.C., Workman, W.J.G., 1999. Measurements of tritium (HTO, TFWT, OBT) in environmental samples at varying distances from a Nuclear Generation Station. Report No. AECL-12029. Chalk River, Atomic Energy of Canada Limited Maro, D., Vermorel, F., Rozet, M., Aulagnier, C., Hebert, D., Le Dizes, S., Voiseux, C., Solier, L., Cossonet, C., Godinot, C., Fievet, B., Laguionie, P., Connan, O., Cazimajou, O., Morillo, M, Lamothe, M., 2016. The VATO project: an original methodology to study the transfer of tritium as HT and HTO in grassland ecosystem. Submitted to J. Environ. Radioact. Mayall, A., Cabianca, T., Attwood, C.A., Fayers, C.A., Smith, J.G., Penfold, J., Steadman, D., Martin, G., Morris, T.P., Simmonds, J.R., 1997. PC-CREAM installing and using the PC system for assessing the radiological impact of routine releases. EUR 17791 EN, NRPB- SR296, Chilton, UK. McFarlane, J.C., 1976. Tritium fractionation in plants. Environ. Exp. Bot. 16, 9-14. Melintescu, A., Galeriu, D., Marica, E., 2002. Using WOFOST crop model for data base derivation of tritium and terrestrial food chain modules in RODOS. Radioprotection 37, 1242- 1246. Melintescu, A., Galeriu, D., 2005. A versatile model for tritium transfer from atmosphere to plant and soil. Radioprotection 40, S437-S442. Melintescu, A., Galeriu, D., 2011. Exchange velocity approach and the role of photosynthesis for tritium transfer from atmosphere to plants. Fusion Sci. Technol. 60, 1179 – 1182. Melintescu, A., Galeriu, D., Diabaté, S., Strack, S., 2015. Preparatory steps for a robust dynamic model for OBT dynamics in agricultural crops. Fusion Sci. Technol. 67, 479-482. Michelotti, E., Green, A., Whicker, J., Eisele, W., Fuehne, D., McNaughton, M., 2013. Validation test for CAP88 predictions of tritium dispersion at Los Alamos National Laboratory. Health Phys., 105, S176-S181. Mihok, S., Wilk, M., Lapp, A., St-Amant, N., Kwamena, N.-O.A., Clark, I.D., 2016. Tritium dynamics in soil and plants grown under three irrigation regimes at a tritium processing facility in Canada. J. Environ. Radioact. 153, 176-187. Mitchell, N.G., 1999. A handbook for the SPADE suite of models for radionuclides transfer through terrestrial foodchains. In: Mouchel Report 48112 produced for Ministry of Agriculture, Fisheries and Food/Food Standard Agency, UK. Murphy Jr., C.E., 1993. Tritium transport and cycling in the environment. Health Phys. 65, 683-697. NDAWG, 2011. Short-term releases to the atmosphere. NDAWG/2/2011. National Dose Assessment Working Group, UK. Ota, M., Yamazawa, H., Moriizumi, J., Iida, T., 2007. Measurement and modeling of oxidation rate of hydrogen isotopic gases by soil. J. Environ. Radioact. 97, 103-115. Ota, M., Nagai, H., 2011. Development and validation of a dynamical atmosphere – vegetation – soil HTO transport and OBT formation model. J. Environ. Radioact. 102, 813- 823. Ota, M., Nagai, H., 2012. HTO transport and OBT formation after nighttime wet deposition of atmospheric HTO onto land surface. Radioprotection, 46, S417-S422. Patryl, L., Armand, P. 2005. Modelisation des tranferts du tritium a l'environnement. Technical report, E.O.T.P. A-2S000-01-21-31-01. Commissariat a l'Energie Atomique, Bruyeres-le-Chatel, France (in French). Patryl, L., Armand, P., 2007. Key assumptions and modelling approaches of tritium models implemented in GAZAXI 2002 and CERES. CEA report EOTP A-24100-01-01-AW-43. Paunescu, N., Galeriu, D., Mocanu, N., 2002. Environmental tritium around a new CANDU nuclear power plant. Radioprotection, 37, C1-1253-C1-1258. Peterson, S.R., Hoffman, F.O., Köhler, H., 1996. Summary of the BIOMOVS A4 scenario: Testing models of the air-pasture-cow milk pathway using Chernobyl fallout data. Health Phys. 71, 149-159. Raskob, W., 1993. Description of the new version 4.0 of the tritium model UFOTRI including user guide. Report KfK 5194. Kernforschungszentrum Karlsruhe, Germany. Raskob, W., 1994. NORMTRI, A computer program for assessing the off-Site consequences from air-borne releases of tritium during normal operation of nuclear facilities. Report KfK- 5364. Kernforschungszentrum Karlsruhe, Germany. Raskob, W., Diabate, S., Strack, S., 1996. A new approach for modelling the formation and translocation of organically bound tritium in accident consequence assessment codes. Int. Symposium on Ionization Radiation: Protection of the Natural Environment, Stockholm, Sweden, May 20–24, 1996. Raskob, W., 2011. Plant OBT model. The Fifth Meeting (TM) on the Environmental Modelling for Radiation Safety (EMRAS II) Intercomparison and Harmonization Project, January 24 – 28, 2011, Vienna, Austria. http://www- ns.iaea.org/downloads/rw/projects/emras/emras-two/third-technical-meeting/wgroup- seven/presentation-5th-wg7-plant-obt-model.pdf. Raskob, W., 2007. Test and validation studies performed with UFOTRI and NORMTRI. Report FZK 7281, TW5-TSS/SEP2 – deliverable 4. Forschungszentrum Karlsruhe, Germany. Schulman, L.L., Strimaitis, D.G., Scire, J.S., 1997. The PRIME plume rise and building downwash model. Addendum to ICS3 user's guide. Earth Tech., Inc., Concord, MA. https://www3.epa.gov/ttn/scram/7thconf/iscprime/useguide.pdf. Shen, H.-F., Liu, W., 2016. Tritium concentration in soybean plants exposed to atmospheric HTO during nighttime. Nucl. Sci. Tech. 27, 39, doi:10.1007/s41365-016-0031-8. Simon-Cornu, M., Beaugelin-Seiller, K., Boyer, P., Calmon, P., Garcia-Sanchez, L., Mourlon, C., Nicoulaud, V., Sy, M., Gonze, M.A., 2015. Evaluating variability and uncertainty in radiological impact assessment using SYMBIOSE. J. Environ. Radioact. 139, 91-102. St-Aman, N., 2016. An overview of tritium studies at the Canadian Nuclear Safety Commission. 5th Workshop on OBT (Organically Bound Tritium) and its analysis. Le Mans, France, 4 – 7 October 2016 Strack, S., Diabate, S., Raskob, W., 1998. Modellrechnungen zur Biokinetik von Tritium in Pflanzen, in: Winter, M. (Hrsg.), Radioaktivitat in Mensch und Umwelt, 30. FS Jahrestagung, 28.9. – 2.10.98, Lindau, Bd. II, S. 855-860, Koln: TueV Rheinland (in German). Strack, S., Diabate, S., Raskob, W., 2005. Organically bound tritium in plants: insights gained by long-term experience in experimental and modelling research. Fusion Sci. Technol. 48, 767-770. Strack, S., Diabate, S., Raskob, W., 2011. Biokinetics of tritium in wheat plants: measurements and model calculations. Sixth Meeting of the EMRAS II Working Group 7, "Tritium" Accidents, Bucharest, Romania, 12 – 15 September 2011. http://www- ns.iaea.org/downloads/rw/projects/emras/emras-two/first-technical-meeting/sixth-working- group-meeting/working-group-presentations/workgroup-7-presentations/presentation-6th- wg7-experimental-work.pdf. Thompson, P.A., Kwamena, N.-O.A., Ilin, M., Wilk, M., Clark, I.D., 2015. Levels of tritium in soils and vegetation near Canadian nuclear facilities releasing tritium to the atmosphere: implications for environmental models. J. Environ. Radioact. 140, 105-113. FIGURE CAPTION Fig. 1 Simulation of HTO concentration in air during a year Fig. 2 Simulation of HTO concentration in air during 24 hours period Fig.3 OBT/HTO ratio in barley and rice (based on experimental data from KEPKO, 2012, 2014) Fig. 4 OBT/HTO ratio in cabbage and persimmon (based on experimental data from KEPKO, 2012, 2014) Fig. 5 OBT/HTO ratio in milk (based on experimental data from KEPKO, 2012, 2014) Fig. 6 Ingestion dose of adult and infant for a spike release of one hour at mid-day during a year Fig. 7 Experimental OBT/HTO ratio for wheat leaves at various exposures carried out at different hours of the day and night (s7, s8, s10, s11, s20p, s23p) Fig. 8 Predicted to observed ratios (P/O) for concentrations of OBT in grain at harvest Fig. 9 Predicted to observed ratios (P/O) for concentrations of OBT in leaves at the end of exposure Fig. 10 Predicted to observed ratios (P/O) for concentrations of OBT in leaf at harvest ) 3 m / q B ( r i a n i n o i t a r t n e c n o c O T H 180 160 140 120 100 80 60 40 20 0 0 100 200 300 400 Day of year Fig. 1 ) 3 m / q B ( r i a n i n o i t a r t n e c n o c O T H 180 160 140 120 100 80 60 40 20 0 0 5 10 15 20 25 Hour of the day Fig. 2 barley rice O T H / T B O 12.00 10.00 8.00 6.00 4.00 2.00 0.00 0.00 20.00 40.00 60.00 80.00 100.00 HTO conc. in grains (Bq/L) Fig. 3 cabbage persimmon O T H / T B O 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0 50 100 150 200 HTO conc. in edible part (Bq/L) Fig. 4 milk O T H / T B O 1.9 1.7 1.5 1.3 1.1 0.9 0.7 0.5 2 3 4 5 6 7 8 HTO conc. in milk (Bq/L) Fig. 5 infant adult 0.12 0.1 0.08 0.06 0.04 0.02 ) . u . a ( e s o d n o i t s e g n I 0 0 100 200 Julian day 300 400 Fig. 6 O T H / T B O 100 10 1 0.1 0.01 0.001 s7 s8 s10 s11 s20p s23p 0 5 10 15 20 25 30 35 time after beggining of exposure (h) Fig. 7 2.5 2 1.5 O P / 1 0.5 0 5 Fig. 8 KIT (Germany) IFIN (Romania) CEA (France) JAEA (Japan) 10 15 20 25 Start hour of each experiment 2.5 2 1.5 1 0.5 O / P 0 5 Fig. 9 KIT (Germany) IFIN (Romania) CEA (France) JAEA (Japan) 10 15 20 25 Start hour of each experiment 100 10 1 0.1 0.01 O / P 0.001 5 P/O JAEA P/O CEA P/O IFIN 10 15 20 25 Start hour of each experiment Fig. 10 Table 1 The ingestion doses (µSv year-1) for infant, child and adult for an atmospheric tritium emission of 1 TBq for various atmospheric stability conditions and durations Case Normal Category Category Continuous Category F/continuous day 1.33 1.68 1.80 Infant Child Adult D 2.22 2.77 3.96 F 8.24 10.20 14.30 release, 1 year release ratio 0.23 0.24 0.26 35.21 41.80 55.21 Table 2 The ratio between tritium concentration in soil water and air moisture (Csw/Cam) and the ratio between tissue free water tritium concentration in plants (TFWT) and air moisture (CTFWT/Cam) for various values of relative humidity (RH) C 0.6 0.70 0.75 0.79 0.84 0.88 0.92 am sw RH /C 0.1 0.2 0.3 0.4 0.5 0.6 C am TFWT 0.7 /C 0.80 0.84 0.87 0.90 0.94 0.97 0.8 0.90 0.92 0.95 0.97 0.99 1.01 Table 3 Plants All leafy vegetables Root vegetables Dry matter fraction (DWp) of terrestrial plants GSDb 1.312 1.281 1.363 1.214 1.011 1.160 GMa 0.069 0.107 0.104 0.182 0.869 0.285 Fruits Pasture Cereals Silage CVc 0.277 0.252 0.318 0.196 0.011 0.150 minimum maximum 0.030 0.050 0.040 0.100 0.840 0.180 0.160 0.230 0.270 0.330 0.900 0.450 a geometric mean, b geometric standard deviation, c coefficient of variance Table 4 Water equivalent factor (WEp) of terrestrial plants Plants All leafy vegetables All non-leafy vegetables Root vegetables Other vegetables GMa 0.508 0.524 0.497 0.548 GSDb 1.035 1.021 1.045 1.040 CVc 0.034 0.021 0.044 0.040 minimum maximum 0.470 0.500 0.450 0.500 0.550 0.550 0.550 0.600 a geometric mean, b geometric standard deviation, c coefficient of variance Table 5 Public dose (Sv year-1) coming from different tritium contamination pathways for children of 1 year and 10 years old and for an adult when OBT/HTO ratio is 0.7 and tritium concentration in animal drinking water is 4.5 Bq L-1 1 y Contamination pathway 1.97x10-7 8.67x10-8 6.17x10-7 9.08x10-8 9.92 x10-7 10 y 2.71x10-7 5.84x10-8 4.29x10-7 9.79x10-8 8.57x10-7 Adult 2.27x10-7 7.68x10-8 3.32x10-7 8.71x10-8 7.23x10-7 Inhalation Drinking water Food HTO Food OBT Total Table 6 Public dose (Sv year-1) coming from different tritium contamination pathways for children of 1 year and 10 years old and for an adult when OBT/HTO ratio is 0.7 and tritium concentration in animal drinking water is 40 Bq L-1 Contamination pathway 1 y 1.97x10-7 8.67x10-8 1.08x10-6 1.39x10-7 1.50x10-6 10 y 2.71x10-7 5.84x10-8 5.76x10-7 1.15x10-7 1.02x10-6 Adult 2.27x10-7 7.68x10-8 4.48x10-7 9.87x10-8 8.50x10-7 Inhalation Drinking water Food HTO Food OBT Total Table 7 Public dose (Sv year-1) coming from different tritium contamination pathways for children of 1 year and 10 years old and for an adult when OBT/HTO ratio is 10 and tritium concentration in animal drinking water is 4.5 Bq L-1 Contamination pathway 1 y 1.97x10-7 8.67x10-8 6.17x10-7 1.05x10-6 1.95x10-6 10 y 2.71x10-7 5.84x10-8 4.29x10-7 1.19x10-6 1.95x10-6 Adult 2.27x10-7 7.68x10-8 3.32x10-7 1.07x10-6 1.70x10-6 Inhalation Drinking water Food HTO Food OBT Total Table 8 Comparison between NDAWG results (2011) and TRIF model results (Higgins, 1997) for integrated activity (Bq d kg -1) of various food items in case of a normal day (class D, 12 hours) Crop Green vegetables Fruit Root vegetables Cow milk Cow meat Cow liver Sheep meat Sheep liver a not available NDAWG 7.29 x 103 7.29 x 103 7.29 x 103 4.08 x 102 3.52 x 102 3.52 x 102 5.36 x 102 5.36 x 102 TRIF 4.52 x 102 NAa 4.52 x 102 2.07 x 102 1.79 x 102 NAa 2.73 x 102 NAa Table 9 Comparison of model predictions for ingestion dose (µSv year-1) of adults for an atmospheric tritium emission of 1 TBq for various atmospheric stability conditions and durations Model Normal Category Category Continuous Category F/continuous day 1.10 0.40 1.80 D 1.30 0.80 3.96 F 2.7 1.00 14.30 release, 1 year release ratio 0.03 0.11 0.26 90.20 9.10 55.21 UFOTRIa IFIN-HHb NDAWG a NORMTRI used for continuous release; b CSA used for continuous release
1409.0593
4
1409
2015-05-16T09:16:02
A density-independent glass transition in biological tissues
[ "physics.bio-ph", "cond-mat.dis-nn", "cond-mat.soft", "cond-mat.stat-mech" ]
Cell migration is important in many biological processes, including embryonic development, cancer metastasis, and wound healing. In these tissues, a cell's motion is often strongly constrained by its neighbors, leading to glassy dynamics. While self-propelled particle models exhibit a density-driven glass transition, this does not explain liquid-to-solid transitions in confluent tissues, where there are no gaps between cells and therefore the density is constant. Here we demonstrate the existence of a new type of rigidity transition that occurs in the well-studied vertex model for confluent tissue monolayers at constant density. We find the onset of rigidity is governed by a model parameter that encodes single-cell properties such as cell-cell adhesion and cortical tension, providing an explanation for a liquid-to-solid transitions in confluent tissues and making testable predictions about how these transitions differ from those in particulate matter.
physics.bio-ph
physics
A density-independent rigidity transition in biological tissues Dapeng Bi1, J. H. Lopez1, J. M. Schwarz1,2, and M. Lisa Manning1,2 1Department of Physics, Syracuse University, Syracuse, NY 13244, USA 2Syracuse Biomaterials Institute, Syracuse, NY 13244, USA 5 1 0 2 y a M 6 1 ] h p - o i b . s c i s y h p [ 4 v 3 9 5 0 . 9 0 4 1 : v i X r a Cell migration is important in many biological processes, including embryonic development, cancer metasta- sis, and wound healing. In these tissues, a cell’s motion is often strongly constrained by its neighbors, leading to glassy dynamics. While self-propelled particle models exhibit a density-driven glass transition, this does not explain liquid-to-solid transitions in confluent tissues, where there are no gaps between cells and therefore the density is constant. Here we demonstrate the existence of a new type of rigidity transition that occurs in the well-studied vertex model for confluent tissue monolayers at constant density. We find the onset of rigidity is governed by a model parameter that encodes single-cell properties such as cell-cell adhesion and cortical tension, providing an explanation for a liquid-to-solid transitions in confluent tissues and making testable pre- dictions about how these transitions differ from those in particulate matter. Important biological processes such as embryogensis, tu- morigenesis, and wound healing require cells to move col- lectively within a tissue. Recent experiments suggest that when cells are packed ever more densely, they start to exhibit collective motion [1–3] traditionally seen in non-living disor- dered systems such as colloids, granular matter or foams [4– 6]. These collective behaviors exhibit growing timescales and lengthscales associated with rigidity transitions. Many of these effects are also seen in Self-Propelled Par- ticle (SPP) models [7]. In SPP models, overdamped particles experience an active force that causes them to move at a con- stant speed, and particles change direction due to interactions with their neighbors or an external bath. To model cells with a cortical network of actomyosin and adhesive molecules on their surfaces, particles interact as repulsive disks or spheres, sometimes with an additional short-range attraction [8, 9]. These models generically exhibit a glass transition at a critical packing density of particles, φc, where φc < 1 [1, 8, 10, 11], and near the transition point they exhibit collective motion [8] that is very similar to that seen in experiments [12]. An important open question is whether the density-driven glass transition in SPP models explains the glassy behav- ior observed in non-proliferating confluent biological tissues, where there are no gaps between cells and the packing fraction φ is fixed at precisely unity. For example, zebrafish embry- onic explants are confluent three-dimensional tissues where the cells divide slowly and therefore the number of cells per unit volume remains nearly constant. Nevertheless, these tis- sues exhibit hallmarks of glassy dynamics such as caging behavior and viscoelasticity. Furthermore, ectoderm tissues have longer relaxation timescales than mesendoderm tissues, suggesting ectoderm tissues are closer to a glass transition, de- spite the fact that both tissue types have the same density [1]. This indicates that there should be an additional parameter controlling glass transitions in confluent tissues. In this work, we study confluent monolayers using the ver- tex model [13–21], to determine how tissue mechanical re- sponse varies with single-cell properties such as adhesion and cortical tension. We find a new type of rigidity transition that is not controlled by the density, but instead by a dimension- less target shape index that is specified by single-cell proper- ties. This rigidity transition possesses several hallmarks of a second-order phase transition. These findings provide a novel explanation for liquid-to-solid transitions in tissues that re- main at constant density. The vertex model, which agrees remarkably well with ex- perimental data for confluent monolayers [13–21], approx- imates the monolayer as a collection of adjacent columnar cells. The mechanical energy of a single cell labeled ‘i’ is given by [14, 16]: Ei = βi(Ai − Ai0)2 + ξiP2 i + γiPi. (1) The first term results from a combination of 3D cell incom- pressibility and the monolayer’s resistance to height fluctua- tions or cell bulk elasticity [15, 22]. Then βi is a height elas- ticity, and Ai and Ai0 are the actual and preferred cell cross- sectional areas. FIG. 1. Energy barriers for local cellular rearrangements. (a) Illustration of a T1 transition in a confluent tissue and the normalized distribution ρ of normalized energy barrier heights ∆ε/∆ε for a large range of parameters (r = 0.5,1,2 and p0 = 3.2 − 3.7). They have a universal shape well-fit by a k − gamma distribution (solid line), indicating that ∆ε completely specifies the distribution and describe the mechanical . (b) ∆ε as function of the target shape index p0 for various values of the inverse perimeter modulus r. The second term in equation (1) is quadratic in the cell cross-sectional perimeter Pi and models the active contractil- "p0abx="/""⇢("/")r=10r=2r=1r=0.5r=0.1(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:1)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:2)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:3)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:4)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:5)(cid:1)(cid:2)(cid:3)(cid:1)(cid:2)(cid:4)(cid:1)(cid:2)(cid:5)(cid:1)(cid:2)(cid:6)(cid:1)(cid:2)(cid:7)(cid:5)(cid:2)(cid:3)(cid:3)(cid:2)(cid:3)(cid:3)(cid:2)(cid:8)(cid:9)(cid:2)(cid:3)(cid:9)(cid:2)(cid:8)012345610−310−210−1100 2 FIG. 2. A rigidity transition in confluent tissues (a) Critical scaling collapse of the average energy barrier height ∆ε, normalized by multi- plying r/p0 − p∗0β, as a function of z = r/p0 − p∗0∆ for the data shown in Fig. 1(b), confirming the scaling ansatz of equation (3). (b) The rigidity transition is demonstrated in a simple phase diagram as function of p0, snapshots are taken from a typical rigid tissue (p0 = 3.7) and soft tissue (p0 = 3.96). A rigidity transition occurs at p0 = p∗0 ≈ 3.813 for disordered metastable tissue configurations. The line corresponding to the order-to-disorder transition reported by Staple et al [16] is shown for comparison. Below phex , the ground state is a hexagonal lattice and above phex the ground state is disordered. 0 0 ity of the actin-myosin subcellular cortex, with elastic con- stant ξi [14], and the last term represents an interfacial tension γi set by a competition between the cortical tension and the energy of cell-cell adhesion [18, 23] between two contacting cells. γi can be positive if the cortical tension is greater than the adhesive energy, or negative if the adhesion dominates. It is also possible to incorporate strong feedback between adhe- sion and cortical tension in this term [18, 24]. Since only the effective forces – the derivatives of the energy with respect to the degrees of freedom – are physically relevant, equa- 2, tion (1) can be rewritten: Ei = βi(Ai − Ai0)2 + ξi(Pi − Pi0) where Pi0 = −γi/(2ξi) is an effective target shape index. As discussed in [16], when all single-cell properties are equal (βi = β, ξi = ξ, Ai0 = A0, Pi0 = P0), the total me- chanical energy of a tissue containing N cells can be non- (cid:21) dimensionalized: N (cid:20) ( ai − 1)2 + ε = 1 βA2 0 i ∑ Ei = ∑ i ( pi − p0)2 r , (2) where ai = Ai/A0 and pi = Pi/√A0 are the rescaled shape functions for area and perimeter. r = βA2 0/ξ is the inverse perimeter modulus and p0 = P0/√A0 is the target shape in- dex [25] or a preferred perimeter-to-area ratio; geometrically, 0 = 2√2 4√3 ≈ 3.72 and a a regular hexagon corresponds to phex regular pentagon to ppent ≈ 3.81. While we focus on the simple case where cells are identical, the rigidity transition is robust to small variations in cell prop- erties(see Supplementary Materials). 0 = 2√5(5− 2√5) 1/4 In non-biological materials, bulk quantities such as shear/bulk modulus, shear viscosity and yield stress are of- ten used to describe the mechanical response to external per- turbations. However, cells are self-propelled and even in the absence of external forces, cells in confluent tissues regularly intercalate, or exchange neighbors [26, 27]. In an isotropic confluent tissue monolayer where mitosis (cell division) or apoptosis (cell death) are rare, cell neighbor exchange must happen through intercalation processes known as T1 transi- tions [28, 29], where an edge between two cells shrinks to a point and a new edge arises between two neighboring cells as illustrated in Fig. 1 (a). The mechanical response of the tissue is governed by the rate of cell rearrangements, and within the vertex model, the rate of T1 rearrangements is related to the amount of mechanical energy required to execute a T1 transi- tion [29]. Therefore, we first study how these energy barriers change with single-cell properties encoded in the model pa- rameters r and p0. To explore the statistics of energy barriers, we test all pos- sible T1 transition paths (see Methods section) in 10 ran- domly generated disordered samples each consisting of N = 64 cells. For each value of p0 and r tested, we obtained the distribution of energy barrier heights ρ(∆ε). The func- tional form of the distribution becomes universal (Fig. 1 (a)) when scaled by the mean energy barrier height ∆ε(r, p0). The rescaled distribution is well-fit by a k − gamma distribution (kk/(k−1)! xk−1 exp(−kx)) with x = ∆ε/∆ε and k = 2.2±0.2. The k− gamma distribution has been observed in many non- biological systems disordered systems [30–32], and gener- ically results from maximizing the entropy subject to con- straints [31, 32]. This confirms that the distribution of energy barriers depends on the single-cell properties p0 and r only through its average ∆ε. Fig. 1(b) shows the dependence of ∆ε on p0 for various values of r. At p0 (cid:46) 3.8, the energy barriers are always finite, i.e. cells must put in some amount of work in order to deform and rearrange. Here the tissue behaves like a solid; it is a rigid material with a finite yield stress. As p0 is increased, the abr"/p0p⇤0p0<p⇤0p0>p⇤0z=r/p0p⇤0p⇤0=3.813±0.005=4.0±0.4=1.0±0.2●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇��-�������������������-���-���-����������0.5/Cell-cell adhesionp0phex0p⇤0RigidSoftCortical tension3.813.72 (cid:19) (cid:18) energy barriers decrease and become vanishingly small in the vicinity of p0 ≈ 3.8, so that cell shape deformations require no energy. This change in the mechanical behavior as function of p0 is suggestive of a rigidity transition. To better understand the nature of this rigidity transition, we search for a scaling collapse. We use the scaled energy barrier value r∆ε(r, p0) as an order parameter, since r con- trols the overall scale of ∆ε away from the transition. From Fig. 1(b), we also see that r controls the sharpness of the tran- sition, playing a role similar to the magnetic field in the Ising model. This is reasonable because as seen in Eq. 9, r controls the strength of fluctuations in the perimeter. Assuming that the mechanical rigidity of the tissue is controlled by a critical point at some p0 = p∗0, then near the critical point the order parameter r∆ε should be related to the variable that controls the fluctuations r by a universal ansatz[33]: r ∆ε = p0 − p∗0β f± r p0 − p∗0∆ . (3) β Here z = r/p0 − p∗0∆ is the crossover scaling variable, ∆ is the crossover scaling critical exponent, and f−, f+ are the two branches of the crossover scaling functions for p0 < p∗0 and p0 > p∗0, respectively. After re-plotting the data in Fig. 1(b) using equation (3), we find an excellent scaling collapse onto two branches with ∆ = 4.0 ± 0.4, β = 1.0 ± 0.2 and a precise location of the critical point p∗0 = 3.813± 0.005 as shown in Fig. 2. In the mechanically rigid branch (p0 < p∗0), as z → 0, f− is finite, meaning that the energy barrier is finite and scales as ∆ε ∝ (p∗0 − p0) /r. At the critical point (p0 = p∗0), the two branches of the scaling function merge and f+ = f− = zβ/∆ resulting in the scaling ∆ε ∝ rβ/∆−1. The mechanically soft or fluid branch (p0 > p∗0) decays as z0.5 as z → 0 or ∆ε ∝ r−0.5/ (p0 − p∗0). This scaling collapse is similar to those seen in jamming in particulate matter [4, 34] and rigidity percolation on random networks [35–37], suggesting that p∗0 is a critical point anal- ogous to Point J in the jamming transition or the critical oc- cupation probability p∗ in random network models. However, unlike the jamming transition which is density driven, den- sity can not control the rigidity transition in the vertex model because everything takes place at a packing fraction of unity. Instead, this model displays a novel rigidity transition con- trolled by the target shape index, p0. Fig. 2(b) summarizes these results by a simple phase diagram and depicts two snap- shots from rigid and soft simulations. Although we calculate T1 transitions by shortening or lengthening a single cell-cell contact, our analysis of these lo- cal perturbations suggest a critical mechanical response with a growing lengthscale. To confirm and quantify these changes in the macroscopic mechanical response, we study the vibra- tional spectrum of the dynamical matrix [38, 39]. (cid:112) We diagonalize the dynamical matrix to obtain normal modes and their corresponding eigenvalues {λi} (Methods) λi. Besides the trivial and eigenfrequencies ωi = sign(λi) global translation modes which have ω = 0, any other non- positive ω correspond to a soft mode. 3 The cumulative density of states is defined as the cumula- tive distribution function of ω, N(ω) = D(ω(cid:48))dω(cid:48) (4) (cid:90) ω −∞ where D(ω) is the density of states. If N(ω = 0) > 0, there are floppy modes – collective dis- placements of the vertices that cost zero energy – and the sys- tem is a fluid, while if N(ω) → 0 as ω → 0, any linear combi- nation of displacements costs finite energy and the material is a solid. Fig. 3(a) demonstrates that the collective linear response ex- hibits a rigidity transition at p0 = p∗0 = 3.813, which is iden- tical to the transition identified by our local, nonlinear energy barrier analysis. For p0 < p∗0, N(ω) exhibits Debye scaling and approaches zero at zero frequency (N(ω) ∼ ωd = ω2), while for p0 > p∗0, N exhibits a finite plateau at the lowest fre- quencies. In addition, as the system approaches the rigidity transition from the solid phase, the density of states D(ω) ex- hibits a peak that shifts to lower frequencies (Fig. 3(a)), just as the so-called Boson Peak [39–41] in jammed particle pack- ings and glasses. Interestingly, the shape and scaling of the peak is different from those in particulate matter, and this is an interesting avenue for future research. Another standard measure of linear mechanical response is the shear modulus. We probe the tissue near the rigidity tran- sition in response to a quasistatic simple shear strain γxy and calculate the shear modulus Gxy(Methods). The shear modu- lus behaves similarly to ∆ε as function p0 and r. As the p0 is increased towards p∗0, Gxy drops rapidly (Fig. 3(b)) at the rigidity transition, with r controlling the overall magnitude and the sharpness of the transition. Therefore, we propose a similar scaling ansatz for Gxy: rGxy = p0 − p∗0βg± r p0 − p∗0φ , (5) β β−φ similar to equation (3). Again,the data Fig. 3(c) collapses onto two branches, with scaling exponents β = 1.0 ± 0.2 & φ = 5.0 + 0.5. The upper branch corresponds to rigid tissues with a finite shear modulus, i.e. z << 1, g+(z) → const or /r. The lower branch corresponds to soft Gxy ∼ (p∗0 − p0) tissues with vanishing shear modulus, i.e., g−(z) → √z or /√r as z → 0. At the transition, Gxy be- Gxy ∼ (p0 − p∗0) comes independent of p0 − p∗0 and scales as Gxy ∼ rβ/φ. An obvious remaining question is what sets the critical point p∗0 ∼ 3.81. To answer this question, we first study a simple mean-field model for a T1 topological swap. In an infinite confluent tissue, the topological Gauss-Bonnet theo- rem requires each cell to have six neighbors on average [28]. Therefore our mean-field model consists of four adjacent six- sided cells. To mimic the effect of additional neighboring cells, we fix each cell area equal to unity. Equation (9) then becomes: (cid:18) (cid:19) ε4 = ∑ 4 cells ( pi − p0)2; ai = 1. (6) 4 with n > 5) cost no energy and the cells are able to remain in the ground state throughout the transition, requiring zero energy. The estimate p∗0 = ppent , indicated by a red dashed line in Fig. 4(c), does identify the critical target shape index in our mean-field model, and is consistent with the critical point p∗0 = 3.813± 0.005 identified by the scaling collapse of energy barriers in the full vertex model. 0 FIG. 4. A simple four-cell model (a) A four-cell aggregate under- going a T1 topological swap. The thick (green) edge represents the cell-cell interface that is contracted to a point and then resolved in the perpendicular direction. (b) Energy of a four cell aggregate during a T1 transition, which attains a maximum at the transition point. p0 varies from 1.5 to 3.8 in equal increments. (c) Energy barrier height as function of p0 for a four-cell aggregate and a mean-field estimate (dashed) for the value of p0 = ppent at which ∆ε vanishes. 0 0 Is there an even simpler explanation for p∗0 ∼ ppent ? As in other rigidity transitions [6, 35, 36, 40], we expect that the critical shape index should also be related to isostatic- ity. In the vertex model with periodic boundary conditions, cells tile the flat 2D plane, and therefore the total number of vertices V , cells N, and edges E are related through Euler’s formula: 0 = V − E + N. Since each edge is shared by two cells, E is also related to the average coordination number z of cells or E = Nz/2, which yields V = N(z/2 − 1). The degrees of freedom are simply the motions of each vertex in 2D: Mdo f = 2V . Assuming force balance (in both directions) and torque balance on each cell generates three constraints per cell: Mc = 3N. At isostaticity, the number of degrees of free- dom equal the number of constraints: Mdo f = Mc, resulting in ziso = 5 and suggesting a mean-field transition at a shape index of p∗0 (cid:39) 3.812. Although it gives a correct prediction, this iso- static argument makes a strong assumption: that constraints are applied to each cell instead of to each vertex. Therefore, an interesting direction for future research is to understand under what circumstances the energy functional (equation 9) effectively groups vertices into functional units that are cells. Discussion Although the vertex model has been used ex- tensively to model tissues over the past 15 years, there has never been a clear way to connect the model parameters to tissue mechanical properties. Here we show that the vertex model has a new and previously unreported critical rigidity transition that occurs at a critical value of the target shape in- dex p∗0 ∼ 3.81. This criticality is evident in (a) energy barri- FIG. 3. Analysis of mechanical properties. (a) The cumulative vibrational density of states N(ω) exhibits a rigidity transition at r = 0.1 and p0∗ = 3.813 (thick line). Thin lines correspond to r = 0.1 and values of p0 ranging from 3.78 to 3.83 in increments of 10−3. Inset: Vibrational density of states D(ω) for selected p0 values, from left to right: 3.78, 3.79, 3.80, 3.81. At low ω, D(ω) ∼ ω follows Debye scaling before arriving at a Boson peak. As p0 is decreased toward the rigidity transition, the Boson peak also shifts to lower frequencies. (b) The scaled shear modulus, rGxy, as a function of r (top to bottom: r = 0.05,0.1,0.5,1,2,10,20) and p0. Inset: linear scale of the same plot. (c) Scaling of Gxy near the rigidity transition obeys rGxyp0− p∗0−β = g±(rp0− p∗0−φ), where β = 1.0±0.2 and φ = 5± +0.5. Equation (6) is calculated numerically during a T1 rear- rangement (Methods) as shown in Fig. 4(a). The total energy during this process is shown in Fig. 4(b) as the edge length (cid:96) is contracted (negative values) and a new edge is extended (positive values); the energy barrier ∆ε is the difference in energy between the initial and maximum energy state. As p0 increases, ∆ε decreases as shown in Fig. 4(c). The pre- cise value p∗0 at which energy barriers vanish can be estimated by calculating the energy cost of shrinking an edge of length (cid:96) = (cid:96)0 inside a hexagonal lattice, while all other edges remain unchanged. Precisely at the T1 transition, two of the cells are pentagons, while the other two remain hexagonal. There- fore if p0 < ppent √2×33/4 ≈ 3.812, pentagons cost finite energy and therefore the transition necessarily requires finite pentagons (and n-gons energy . 0 = 7+2√7 In contrast, for p0 ≥ ppent 0 ��-���-������-���-����10-210-110010110-310-210-1100!N(!)D(!)p0<p⇤0p0=p⇤0p0>p⇤0!abcr/p0p⇤0rGxy/p0p⇤0/0.5●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□��-�����������������-����������●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□�����������������������������������p0rGxy●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○○□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□□������������������-���-���-�−0.6−0.300.30.600.51.01.5Increasingp0`−0.500.500.511.53.53.63.73.83.910−510−410−310−210−1100p0"`"Increasingp0ab"p0a33.5410−610−410−2100""(`=`0)bc ers to local T1 rearrangements, (b) the vibrational spectrum of the linear response, and (c) the shear modulus of the tis- sue. Unlike SPP models where the liquid-to-solid transition is governed by density, our model has a constant-density glass transition governed by single-cell mechanical properties such as cell-cell adhesion and cortical tension encoded in the target shape index p0. Analyzing only the ground states of the vertex model, the seminal work of Staple et al [16] found an ordered-to- disordered transition at p0 = phex 0 ∼ 3.722. However, be- cause almost all biological tissues are strongly disordered, it remained unclear whether this transition was relevant for the observed glass or jamming transitions in multicellular tissues. Therefore, we explicitly study disordered metastable states and transitions between them. In addition, [16] uses a lin- ear stability analysis of a single cell to suggest that a rigidity transition also occurs at phex 0 . In contrast, our analysis explic- itly includes multicellular interactions (i.e. collective normal modes) and nonlinear effects (i.e. energy barriers). With this more sophisticated analysis, we demonstrate that vertex mod- els exhibit a rigidity transition at a value p0 = ppent 0 ∼ 3.81 that is measurably different from the prediction p0 = phex 0 based on single-cell linear stability. Importantly, predictions based on this critical rigidity tran- sition have recently been verified in experiments [42]. Specif- ically, in both simulations and experiments we can measure the shape index p = P/√A for each cell in a monolayer, where P is the projected cell perimeter and A is the cross-sectional area. In simulations of the vertex model, we find that the me- dian value of the observed shape index p is an order parameter that also exhibits critical scaling: p = p∗0 ∼ 3.81 for rigid or jammed tissues and p becomes increasingly larger than p∗0 as a tissue becomes increasingly unjammed Fig. 2. This prediction is precisely realized in cultures from primary cells in human patients, with implications for asthma pathobiology [42]. We expect that this rigidity framework will help experi- mentalists develop other testable hypotheses about how the mechanical response of tissues depends on single-cell proper- ties. For example, Sadati et al. [43] have proposed a jamming phase diagram where tissues become more solid-like as ad- hesion increases, based on observations of jamming in adhe- sive particulate matter at densities far below confluency. Us- ing standard interpretations of the vertex model (equations (1) and (9)), p0 increases with increasing adhesion, and therefore our model predicts that confluent tissues become more liquid- like as adhesion increases. This highlights the fact that ad- hesion acts differently in particulate and confluent materials; in particulate matter higher adhesion leads to gelation and so- lidification, while in the vertex model larger adhesion leads to larger perimeters, more degrees of freedom, and liquid-like behavior. These ideas suggest that the role of adhesion in tis- sue rheology may be much richer and more interesting than previously thought. In addition, although all published vertex models as- sume three-fold coordinated vertices, there is no proof that such structures are stable for p0 > phex [16]. Addition- 0 5 ally, higher order vertices are apparently stabilized in some anisotropic biological tissues, including Rosette formation in Drosophila [44]. It will be interesting to study what conditions stabilize higher-fold vertices. This work may also be relevant to modeling the Epithelial- to-Mesenchymal Transition (EMT) that occurs during can- cer tumorigenesis. During EMT, epithelial cells with well- defined, compact shapes and small perimeters relative to their areas transition to mesenchymal cells with irregular shapes and large perimeters relative to their areas [45]. Since equa- tion (9) specifies a fixed shape index, one could interpret EMT as an increase in p0 leading to a solid-to-liquid transition, providing a simple mechanical explanation for the role EMT plays in metastasis. In order to explore this idea further, it will be necessary to determine if a similar rigidity transition exists in three dimensions. A simple extension of this model would replace perimeters and areas in Eq. 9 with surface areas and volumes, respectively; this is a promising avenue for future work. We expect that this model may be of interest to scientists independent of its biological relevance. We have shown it ex- hibits a simple rigidity transition with a novel control param- eter, and therefore it might provide a useful bridge between jamming transitions in particulate matter [6, 40] and rigidity transitions in random elastic networks [35–37]. In particu- lar, the potential grouping of vertices into functional cell units could draw an explicit connection between spring networks and particle/cell packings. An open question is whether our model belongs to an existing universality class, and whether the transition is mean-field. Finally, the fact that the vertex model exhibits disordered ground states for p0 > p∗0 suggests that it may be a useful toy model for thermodynamic (as opposed to kinetic) expla- nations of the glass transition in particulate matter. Further- more, these states are predicted to be hyperuniform [46] with a photonic band gap, indicating that they may be useful for designing metamaterials with interesting optical properties. METHODS Simulating a confluent tissue monolayer To simulate confluent monolayers, a Random Sequential Addition point pattern [47] of N points was generated under periodic boundary conditions, with box size L chosen such that the average area per cell is unity. Two methods of gen- erating this initial point pattern were used: a Random Se- quential Addition point pattern [47], and a Poisson point pat- tern. The results presented in this work are independent of the method of initial point pattern generation. A voronoi tes- sellation of this point pattern results in a disordered cellular structure, which was then used as an input to the program Sur- face Evolver [48]. Surface Evolver numerically minimizes the total energy of the system (equation (9)) at fixed topology us- ing gradient descent with respect to the vertices of the cells. If an edge shrinks below a threshold value l∗, a passive T1 transition is allowed if it lowers the energy. All structures are minimized such that the average energy of a cell changes by less than one part in 1010 between consecutive minimization steps, and as in other simulations of the vertex model [15, 16]. Once an initial energy-minimized state is reached, T1 tran- sitions are actively induced at every edge to measure energy barriers [29]. An example of a T1 in the simple four-cell case is shown in Fig. 1(a): the central thick edge is quasi-statically contracted to zero length ((cid:96) = 0) at which point a T1 topo- logical swap is executed. After the T1, the length of the cen- tral edge is then expanded until it reaches the initial length ((cid:96) = (cid:96)0). The total energy of four cells during this process is shown in Fig. 4(b); the edge length is represented by a nega- tive value during contraction and flips sign after the T1. For each active T1 transition in an N-cell system, the energy barrier is defined as the total energy difference between the initial state (cid:96) = (cid:96)0 and the onset of T1 topological swap ((cid:96) = 0). Calculations of energy barriers were repeated for various values of r at decade increments from 0.005 to 200 and p0 ranging from 3 to 4. To calculate the shear modulus, we apply quasistatic simple shear to a tissue using Lee-Edwards periodic boundary condi- tions. The shear modulus is calculated by taking the linear response of the tissue, Gxy = 1 L2 lim γxy→0 ∂2ε ∂γ2 xy , (7) where L is the linear dimension of the tissue. The results for Gxy at each value of r and p0 were obtained by averaging 20 runs. Calculation of the vibration density of states We obtain the vibrational density of states by diagonalizing the Hessian matrix of the system Hiµ jν = ∂2ε ∂riµr jν , (8) where i, j are indices for vertices and µ,ν cartesian coordi- nates, and ε is defined in equation (9). The eigenvalues of equation (8) are {λi}. ACKNOWLEDGEMENTS We would like to thank M. C. Marchetti for useful com- ments on this manuscript. M.L.M. acknowledges support from the Alfred P. Sloan Foundation, and M.L.M and D.B. acknowledge support from NSF-BMMB-1334611 and NSF- DMR-1352184. M.L.M and D.B. also would like to thank the KITP at the University of California Santa Barbara for hospitality and was supported in part by the National Science Foundation under Grant No. NSF PHY11-25915. 6 [1] E.-M. Schoetz, M. Lanio, J. A. Talbot, and M. L. Manning, J. Roy. Soc. Interface 10, 20130726 (2013). [2] T. E. Angelini, E. Hannezo, X. Trepat, M. Marquez, J. J. Fred- berg, and D. A. Weitz, Proceedings of the National Academy of Sciences 108, 4714 (2011). [3] K. D. Nnetu, M. Knorr, J. Kas, and M. Zink, New Journal of [4] T. K. Haxton, M. Schmiedeberg, and A. J. Liu, Phys. Rev. E Physics 14, 115012 (2012). 83, 031503 (2011). [5] A. R. Abate and D. J. Durian, Phys. Rev. E 76, 021306 (2007). [6] A. J. Liu and S. R. Nagel, Annual Review of Condensed Matter Physics 1, 347 (2010). [7] T. Vicsek, A. Czir´ok, E. Ben-Jacob, I. Cohen, and O. Shochet, Phys. Rev. Lett. 75, 1226 (1995). [8] S. Henkes, Y. Fily, and M. C. Marchetti, Phys. Rev. E 84, 040301 (2011). [9] H. Chate, F. Ginelli, G. Gregoire, F. Peruani, and F. Raynaud, Eur. Phys. J. B 64, 451 (2008). [10] L. Berthier, Physical Review Letters 112, 220602 (2014). [11] L. Berthier and J. Kurchan, Nat Phys 9, 310 (2013). [12] L. Petitjean, M. Reffay, E. Grasland-Mongrain, M. Poujade, B. Ladoux, A. Buguin, and P. Silberzan, Biophysical journal 98, 1790 (2010). [13] T. Nagai and H. Honda, Philosophical Magazine Part B 81, 699 (2001). [14] R. Farhadifar, J.-C. R??per, B. Aigouy, S. Eaton, and F. J?licher, Current Biology 17, 2095 (2007). [15] L. Hufnagel, A. A. Teleman, H. Rouault, S. M. Cohen, and B. I. Shraiman, Proceedings of the National Academy of Sciences 104, 3835 (2007). [16] D. B. Staple, R. Farhadifar, J. C. Roeper, B. Aigouy, S. Eaton, and F. Julicher, Eur. Phys. J. E 33, 117 (2010). [17] S. Hilgenfeldt, S. Erisken, and R. W. Carthew, Proceedings of the National Academy of Sciences 105, 907 (2008). [18] M. L. Manning, R. A. Foty, M. S. Steinberg, and E.-M. Schoetz, Proceedings of the National Academy of Sciences 107, 12517 (2010). [19] G. Wang, M. L. Manning, and J. D. Amack, Developmental [20] K. K. Chiou, L. Hufnagel, and B. I. Shraiman, PLoS Comput Biology 370, 52 (2012). Biol 8, e1002512 (2012). [21] A. G. Fletcher, M. Osterfield, R. E. Baker, and S. Y. Shvarts- man, Biophysical Journal, Biophysical Journal 106, 2291 (2014). [22] S. Zehnder, M. Suaris, M. Bellaire, and T. Angelini, Biophysi- cal Journal 108, 247 (2015). [23] F. Graner and J. A. Glazier, Physical Review Letters 69, 2013 (1992). [24] J. D. Amack and M. L. Manning, Science 338, 212 (2012). [25] In this Letter we focus on the regime where p0 > 0, but p0 < 0 is a valid parameter regime where our results hold as well. [26] T. Lecuit, HFSP Journal 2, 72 (2008). [27] C. Guillot and T. Lecuit, Science 340, 1185 (2013). [28] D. L. Weaire and S. Hutzler, The physics of foams (Oxford Uni- [29] D. Bi, J. H. Lopez, J. M. Schwarz, and M. L. Manning, Soft versity Press, 1999). Matter 10, 1885 (2014). [30] K. A. Newhall, I. Jorjadze, E. Vanden-Eijnden, and J. Brujic, Soft Matter 7, 11518 (2011). [31] T. Aste and T. Di Matteo, Phys. Rev. E 77, 021309 (2008). [32] D. Bi, J. Zhang, R. P. Behringer, and B. Chakraborty, EPL (Europhysics Letters) 102, 34002 (2013). [33] L. P. KADANOFF, W. G OTZE, D. HAMBLEN, R. HECHT, E. A. S. LEWIS, V. V. PALCIAUSKAS, M. RAYL, J. SWIFT, D. ASPNES, and J. KANE, Rev. Mod. Phys. 39, 395 (1967). [34] P. Olsson and S. Teitel, Phys. Rev. Lett. 99, 178001 (2007). [35] M. Das, D. A. Quint, and J. M. Schwarz, PLoS ONE 7, e35939 (2012). (2006). [36] C. P. Broedersz, X. Mao, T. C. Lubensky, and F. C. MacKin- tosh, Nature Physics 7, 983 (2011). [37] C. Heussinger and E. Frey, Physical review letters 97, 105501 [38] N. Ashcroft and N. Mermin, Solid State Physics, HRW interna- tional editions (Holt, Rinehart and Winston, 1976). [39] L. E. Silbert, A. J. Liu, and S. R. Nagel, Phys. Rev. Lett. 95, [40] M. Wyart, L. E. Silbert, S. R. Nagel, and T. A. Witten, Phys. [41] M. L. Manning and A. J. Liu, Euro. Phys. Lett. (2013), 098301 (2005). Rev. E 72, 051306 (2005). 1307.5904. [42] J.-A. Park, J. H. Kim, D. Bi, J. A. Mitchel, N. T. Qazvini, K. Tantisira, C. Y. Park, M. McGill, S.-H. Kim, B. Gweon, J. Robert Steward, S. Burger, S. H. Randell, A. Kho, D. Tambe, C. Hardin, S. A. Shore, E. Israel, D. A. Weitz, D. J. Tschumper- lin, E. P. Henske, S. T. Weiss, M. L. Manning, J. P. Butler, J. M. Drazen, and J. J. Fredberg, “Unjamming transition to cellular hypermobility in the asthmatic airway epithelium,” (2015). [43] M. Sadati, N. T. Qazvini, R. Krishnan, C. Y. Park, and J. J. Fredberg, Differentiation 86, 121 (2013), mechanotransduc- tion. [44] K. E. Kasza, D. L. Farrell, and J. A. Zallen, Proc Natl Acad Sci U S A 111, 11732 (2014). tigation 119, 1420 (2009). [45] R. Kalluri, R. A. Weinberg, et al., The Journal of clinical inves- [46] A. Gabrielli and S. Torquato, Phys. Rev. E 70, 041105 (2004). [47] S. Torquato, Author and H. Haslach, Jr, Applied Mechanics Re- views (2002). [48] K. A. Brakke, Experimental mathematics 1, 141 (1992). SUPPLEMENTARY INFORMATION: EFFECT OF VARIATIONS IN SINGLE CELL PROPERTIES To study the effect of non-homogeneous cellular properties on the rigidity transition, we first probe the mechanical prop- erty of a tissue with a randomly distributed ‘preferred’ cell area (cid:20) ( ai − a0)2 + ε = ∑ i (cid:21) ( pi − p0)2 r , (9) where a0 is drawn from a Gaussian distribution with mean 1 and variance σ2 a. We have chosen to use r = 0 for sim- plicity, which does not affect the result. The shear modulus is plotted as function of p0 in Fig. 5(a) for various values of σa. The data at σa = 0 correspond to that shown in the main manuscript. σa introduces more disorder in cell areas 7 and results in more fluctuations near the rigidity transition. As σa is increased, the rigidity transition is ‘softened’ simi- lar to the effect of increasing r. A small value σa of should provide a white noise or thermal-like fluctuation to the tissue and we hypothesize the scaling form for the shear modulus. Gxy/p0 − p∗0φa ∝ σ2 p0/p0 − p∗0∆a. Fig. 5(b) shows that the data from Fig. 5(a) can be scaled to collapse using ∆a = 2.6, φa = 0.77 and at the same value of p∗0 = 3.813. This suggests that adding disorder in preferred cell areas does not change the location of the rigidity transition. A similar numerical calculation was performed for tissues with varying value of p0. With mean of p0 and variance σ2 p0. The same analysis can be carried out to show that the location of the rigidity transition does not shift with σ2 p0 (Fig. 6). FIG. 5. The shear modulus of a tissue with varying preferred cell area. (a) The shear modulus as function of p0 for different values of σa = 0, 0.05, 0.1, 0.15, 0.2 & 0.25 (from bottom to top). (b) The shear modulus obeys a universal scaling function. The location of the rigidity transition is found to be p∗0 = 3.813. FIG. 6. The shear modulus of a tissue with varying p0. (a) The shear modulus as function of p0 for different values of σa = 0, 0.05, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4(from bottom to top). (b) The shear modulus obeys a universal scaling function. The location of the rigidity transition is found to be p∗0 = 3.813. ●●●●●●●●●●●●●●●●●●●●●■■■■■■■■■■■■■■■■■■■■■◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼○○○○○○○○○○○○○○○○○○○○○��������������������������������������������■■■■■■■■■■■■■■■■■■■■■◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼○○○○○○○○○○○○○○○○○○○○○��������������-�������Gxy/p0p⇤00.772a/p0p⇤02.61p0Gxyincreasingaab●●●●●●●●●●●●●●●●●●●●●■■■■■■■■■■■■■■■■■■■■■◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼○○○○○○○○○○○○○○○○○○○○○□□□□□□□□□□□□□□□□□□□□□◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇◇��������������������������������������������■■■■■■■■■■■■■■■■■■■■■◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▲▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼▼○○○○○○○○○○○○○○○○○○○○○□□□□□□□□□□□□□□□□□□□□□��������������-�������Gxy/p0p⇤00.77p0Gxyabincreasingp02p0/p0p⇤02.61
1505.04231
1
1505
2015-05-16T02:13:17
Quantitative Morphological and Biochemical Studies on Human Downy Hairs using 3-D Quantitative Phase Imaging
[ "physics.bio-ph", "physics.optics" ]
This study presents the morphological and biochemical findings on human downy arm hairs using 3-D quantitative phase imaging techniques. 3-D refractive index tomograms and high-resolution 2-D synthetic aperture images of individual downy arm hairs were measured using a Mach-Zehnder laser interferometric microscopy equipped with a two-axis galvanometer mirror. From the measured quantitative images, the biochemical and morphological parameters of downy hairs were non-invasively quantified including the mean refractive index, volume, cylinder, and effective radius of individual hairs. In addition, the effects of hydrogen peroxide on individual downy hairs were investigated.
physics.bio-ph
physics
Quantitative Morphological and Biochemical Studies on Human Downy Hairs using 3-D Quantitative Phase Imaging SangYun Leea, Kyoohyun Kima, Yuhyun Leeb, Sungjin Parkb, Heejae Shinb, Jongwon Yangb, Kwanhong Kob, HyunJoo Parka, and YongKeun Parka aKorea Advanced Institute of Science and Technology, Department of Physics, Daejeon 305-701, Republic of Korea. bDaejeon Dongsin Science High School, Daejeon 300-310, Republic of Korea. Abstract. This study presents the morphological and biochemical findings on human downy arm hairs using 3-D quantitative phase imaging techniques. 3-D refractive index tomograms and high-resolution 2-D synthetic aperture images of individual downy arm hairs were measured using a Mach-Zehnder laser interferometric microscopy equipped with a two-axis galvanometer mirror. From the measured quantitative images, the biochemical and morphological parameters of downy hairs were non-invasively quantified including the mean refractive index, volume, cylinder, and effective radius of individual hairs. In addition, the effects of hydrogen peroxide on individual downy hairs were investigated. Keywords: body hair, quantitative phase imaging, synthetic aperture imaging, optical diffraction tomography. Address all correspondence to: YongKeun Park, Korea Advanced Institute of Science and Technology, Department of Physics, 291 Daehak-Ro Yusung-Gu, Daejeon 305-701, Republic of Korea. Tel: (82) 42-350-2514; Fax: (82) 42- 350-7160; E-mail: [email protected] 1. Introduction Human body hairs exhibit distinct morphologies at their relative sites in the body, from soft downy hairs on the arms to long stiff hairs on the head. The characteristics of different hairs have been regarded as one of the evolutionary consequences incurred by human bipedalism (1, 2). While human body hairs have lost their thermoregulatory roles in maintaining warmth through an evolutionary process, decorative aspects of hairs have gained attentions and led many researchers to investigate the microstructures of downy hairs using diverse microscopic techniques, including atomic force microscopy (3-5), scanning electron microscopy (6-8), transmission electron microscopy (9, 10), confocal microscopy (11, 12) and infrared spectroscopy (13-15) under various experimental conditions. Recently, chemical analyses of hairs have been extensively carried out for drug detection (e.g. alcohol and cocaine), particularly in the fields of forensic science (16-20). However, previous approaches cannot provide non-invasiveness and quantitative imaging capabilities. Furthermore, most previous studies have used expensive instrumentation, which prevents these imaging techniques from being utilized in general laboratories in which body hairs are studied. Here, we present quantitative phase imaging (QPI) as an effective imaging tool to study the morphological and biochemical properties of downy hairs in a non-invasive and quantitative manner. QPI techniques provide quantitative measurements of optical phase delay introduced by intrinsic refractive index (RI) distributions of transparent cells using interferometry (21, 22). QPI techniques have been applied previously for biological studies of cells and tissues including red blood cells (23-27), cell growth monitoring (28), neurons (29), and optical imaging of tissue slices (30). The feasibility of non-invasive and quantitative imaging of QPI were demonstrated with individual downy hairs using the QPI techniques. The high-resolution 2-D holographic synthetic aperture images and 3-D RI distribution maps of individual hairs were measured, from which the morphological and biochemical properties were retrieved, including the mean RI, volume, cylinder, and effective radius of individual hairs. Furthermore, we investigated the effects of hydrogen peroxide (H2O2), which is one of the widely-used hair bleaching agents, on individual downy 1 hairs using the present method. 2. Materials and Methods 2.1 Sample Preparations A total 13 human downy hairs were collected from one arm of a healthy donor (Asian male) [Fig. 1(a1)]. Each hair was gently posed on the top of a coverslip (24 × 50 mm, C024501, Matsunami, LTD, Japan) with oil immersion (n = 1.518), which decreases the RI contrast between the hairs and the medium. The sample was then covered with another coverslip [Fig. 1(a2)]. To study the effects of hydrogen peroxide on the hair structures, four hair samples were treated with 3% H2O2 solution (Sigma-Aldrich, St. Louis, Missouri, US) for 24 hours. The shape of the edges of the hairs cells was imaged before and after the H2O2 treatments. (b) (c1) phase (c2) (rad) 6 amplitude 10 μm (a.u.) 1.5 1 0.5 (c5) x-zplane 10 μm 3-D RI maps n 1.6 1.56 1.52 x-y plane y-zplane 4 2 0 (rad) 6 4 2 0 BS 10 μm (c3) synthetic aperture phase image 10 μm (c4) emulated DIC image (c6) Z 5 μm Y X (n> 1.53) top view 10 μm (a1) (a2) 4 mm Fig. 1 (a) Human downy hair preparation. (a1) Left arm of a healthy donor from which downy hairs were side view collected. (a2) A downy hair with immersion oil loaded between two coverslips. (b) Mach-Zehnder interferometric microscopy equipped with a two-axis galvanometer mirror. BS: beam splitter; GM: galvanometer mirror; CL: condenser lens; OL: objective lens, M: mirror. (c) Holographic image reconstruction process. A set of retrieved (c1) phase and (c2) amplitude maps of complex optical fields. (c3) Synthetic aperture phase image. (c4) Emulated DIC image. (c5) Reconstructed 3-D RI distributions of a hair sample. (c6) 3-D RI isosurface of a hair tomogram. 2.2 Optical Setup for QPI For quantitative measurements of human downy hairs, we employed a Mach-Zehnder interferometric microscope equipped with a two-axis galvanometer mirror (31-33). The schematic of the setup is shown in Fig. 1(b). A diode-pumped solid state laser (λ = 532 nm, 50 mW, Cobolt Co., Solna, Sweden) was used as a coherent light source. A beam splitter (BS, BS016, Thorlabs, U.S.A.) divides a laser beam into two arms: a reference beam and a sample beam. Sample beam passes through a downy hair loaded on the sample stage of an inverted microscope. A two-axis galvanometer mirror (GM, GVS012/M, Thorlabs, U.S.A.) varies the angle of the illumination beam impinging onto the hair sample, from which 2-D optical field images of the sample are obtained with various illumination angles. For the reconstruction of off-axis interference patterns, the reference and the sample beams are recombined with a tilted angle by another beam splitter, and the resultant interferogams of samples are recorded by a high-speed CMOS camera (Neo sCMOS, Andor Inc., 2 Northern Ireland, U.K.). An objective lens [UPLFLN, 60×, Numerical aperture (NA) = 0.9, Olympus Inc., San Diego, C.A., U.S.A.] was used as a condenser lens with the tube lens of a focal length of 200 mm. For the imaging purpose, a high-N.A. objective lens (PLAPON, 60×, Oil immersion, NA = 1.42, Olympus Inc., San Diego, C.A., U.S.A.) was used with an additional telescopic 4-f system, and the total magnification of the imaging system is 250×. The camera has 1,776 × 1,760 pixels with a pixel size of 6.5 μm. The field of view at the sample plane was 46.18 × 45.76 μm2. Each hair sample was illuminated with plane waves at 300 different incidence angles, which were systematically controlled by the two-axis GM at a frame rate of 100 Hz. 2.3 Image Reconstruction Procedures Complex optical fields of the sample were retrieved from the interferogram recorded by the QPI technique, via a retrieval algorithm based on Fourier transform (34, 35). The phase and amplitude maps of a representative hair are shown in Figs. 1(c1) & 1(c2). From a set of retrieved phase images [Fig. 1(c1)] with various illumination angles, the 2-D high- resolution synthetic aperture phase image [Fig. 1(c3)] was constructed with a synthetic aperture imaging algorithm (36, 37). By numerically extending the aperture size, the synthetic aperture algorithm fully used the high spatial frequency information of a sample which cannot be accessed with just a single laser illumination angle. The numerical extension of an aperture was conducted in the 2-D Fourier space, and the resultant 2-D synthetic aperture phase image [Fig. 1(c3)] exhibited a higher spatial resolution and signal- to-noise ratio (SNR), compared to the phase image from a single hair interferogram [Fig. 1(c1)] (38, 39). In principle, the resolution of an imaging system is determined by the NAs of an objective lens. In general, for 2-D QPI with a single illumination, the maximum accessible spatial frequency kmax is determined as 2πNAimag/λ, where NAimag represents the NA of the imaging system. In synthetic aperture imaging, the maximum spatial frequency can be further extended by combining multiple phase images obtained with various incident angles; the maximum spatial frequency is extended to 2π(NAimag+NAillum)/λ, where NAillum is the N.A. of the condenser lens. The SNR of the synthetic aperture phase image is also more significantly enhanced than that of a single phase image, mainly due to speckle noise reduction. A differential interference contrast (DIC) image can also be numerically reconstructed from the synthetic aperture phase image [Fig. 1(c4)]. This emulated DIC image is readily obtained from a synthetic aperture phase image by numerically interfering an original phase image and a slightly translated synthetic aperture phase image with an additional phase shift (40). Because the optical imaging contrast in DIC images is a result of the phase gradient, the imaging contrast is enhanced for objects with high RI changes or gradients, such as subcellular organelles in unstained biological cells or defects in non-biological samples. To reconstruct the 3-D RI distribution of a hair sample from a set of retrieved complex optical fields, we used the optical diffraction tomography (ODT) algorithm (32, 33, 41). Compared to the projection algorithm (42, 43), the ODT algorithm considers light diffraction at samples, and thus provides high- resolution tomographic reconstruction with better image qualities especially for large samples with high RI contrast (32). 2.4 Statistical Analysis P-values are calculated by two-tailed paired t-test comparing quantitative parameters of human arm hairs. All the numbers follow the ± sign in the text are standard deviations. 3. Result and Discussion 3.1 High-resolution 2-D Phase Images and 3-D RI Maps of Human Downy Hairs 3 In order to measure the morphologies of individual downy hairs, we measured and took 3-D RI maps and 2-D synthetic aperture phase images of human downy hairs collected from arms, using the Mach-Zehnder laser interferometry equipped with a two-axis galvanometer mirror (See Materials and Methods). Figure 2 shows the measured images of representative human downy hairs. Measured hair samples were categorized into two groups based on their morphologies; hairs in the soft type exhibit smooth surfaces and internal RI distributions [Figs. 2(a1−a3)], and hairs in the rough type present complex internal structural defects and spiky surfaces [Figs. 2(b1−b3)]. High-resolution 2-D synthetic aperture phase images [Figs. 2(a1) and (b1), top], corresponding emulated DIC images [Figs. 2(a1) and (b1), bottom] and three cross-sectional slices of the reconstructed 3- D RI tomograms [Figs. 2(a2) and (b2)] of the representative downy hairs showed significant morphological differences (See Materials and Methods). The renderings of 3-D RI isosurfaces (n > 1.545) of the hair samples are also shown in Figs. 2(a3) and (b3). (a1) (rad) 4 (a2) 10 μm 3 2 1 0 x-z plane 10 μm (a3) n 1.6 1.58 1.56 1.54 1.52 Z 5 μm x-y plane z-y plane X Y (b1) (rad) 4 (b2) 10 μm 3 2 1 0 x-z plane 10 μm (b3) n 1.6 1.58 1.56 1.54 1.52 Z 5 μm x-y plane z-y plane X Y Fig. 2 Quantitative phase images of representative human downy hairs for soft and rough type ((a) and (b), respectively): (a1) and (b1), the 2-D synthetic aperture phase image (top) and the corresponding emulated DIC image (bottom) of the representative hair samples. (a2) and (b2), the cross-sectional slices of the reconstructed 3-D RI tomogram in x−y, z−y, and x−z plane. (a3) and (b3), perspective view of the representative 3-D hair RI isosurface. RI threshold set as a value of 1.545. Each colored arrow in the figures represents the same fine structures of the measured downy hair. In total, 13 human downy hairs were measured; 6 hair cells were categorized as the soft type, and 7 hair cells were in the rough type. The representative holographic images of the soft type hair [Figs. 2(a1−a3)] had soft and continuously connected cell boundaries without any apparent structural defects. The relatively homogenous RI distribution of the soft type hair [Fig. 2(a2)] reflects the homogeneity of its internal structures. Figures 2(b1−b3) show quantitative phase images of the representative rough type downy hair. Significantly low speckle noise level in the backgrounds of the DIC image indicates that the observed fine structures in the hair sample are from real structural defects. The characteristic fine structures, indicated with the colored arrows in Fig. (b1), are also shown in the reconstructed 3-D RI tomogram [Fig. 2(b2)]. Particularly, the structural defect [the black arrow in Fig. 2(b1)] can be clearly seen in the z − y sectional RI slice. Finally, the RI isosurface of the rough type hair [Fig. 2(b3)] shows more complicated surface 4 structures when compared to the RI isosurface of the soft type hair [Fig. 2(a3)]. 3.2 Retrieval of Quantitative Downy Hair Parameters To demonstrate the quantitative imaging capability of the present method for the study of downy hairs, we retrieved the morphological and biochemical parameters from measured 3-D RI maps of individual hairs. The retrieved parameters included mean RI values, the volumes for 20-μm-length hairs [Fig. 3(a)], cylindrical radii of the hairs as a function of length [Fig. 3(b)], and the effective radii of hair edges [Fig 3(c)]. (b) (pL) 1.2 0.9 0.6 0.3 0 mean refractive index hair volume (20 μm length) (a) 1.6 1.58 1.56 1.54 1.52 1.5 (c) 5 ) i m μ ( s u d a r e v i t c e f f e 4 3 2 1 0 downy hair ) i m μ ( s u d a r r e d n i l y c (d) ) i m μ ( s u d a r e v i t c e f f e 4 3 2 1 0 0 3.5 3.0 2.5 2.8 cylindrical radius effective radius 5 10 15 downy hair axis (μm) hair axis 20 0.79±0.08 3.2 3.6 4 saturated cylindrical radius (μm) 4.4 Fig. 3 Biochemical and morphological parameters of human downy hairs. (a) Retrieved mean RIs and volumes of 20-μm-length hairs. (b) Cylindrical radii of hairs as a function of length. The shaded area represents standard deviations. Inset: a schematic describing the radius and effective radius of a hair. (c) Retrieved effective radius. (d) Correlation map between the effective radii and the cylindrical radii. Each colored dot denotes individual hairs. Boxes, median with upper and lower quartiles; Whiskers, parameter range. To obtain the mean RI value of a hair, we averaged the RI values over the sample area in the focal plane [i.e. the x − y plane in Figs. 2(a2) and (b2)]. Volumes of the 20-μm-length hairs were calculated by integrating voxels higher than the RI threshold (n = 1.53). The mean values of the retrieved mean RI and volume for the 20 μm length hairs were 1.557 ± 0.007 (mean ± std) and 692 ± 173 fL, respectively, which are in good agreement with a previous RI measurement (44). The cylindrical radii of hairs as a function of hair lengths are retrieved assuming a cylindrical symmetry [inset in Fig. 3(b)]. Then, the cylinder radius represents the hair thickness as a function of hair length. The graph in Fig. 3(b) depicts the measured cylindrical radii for 13 downy hairs along their hair axis. The red line denotes the mean cylinder radii, and the shaded region corresponds to standard deviations. The mean cylindrical radii seems to saturate beyond a length of 5 μm. The roundness of the hair tips can be effectively described by measuring the effective radii of the tips [Fig. 3(c)]. In order to retrieve the effective radius of a hair R, the following relation was used: R = (r2 + h2)/2h, where r is the radius of a circular slice orthogonal to the hair axis, h is the distance between the circular slice and the hair tip. For the case of a perfect sphere, any circular slice of the sphere with a different h gives the same R. This is not the case for hairs, and the height h should be determined appropriately in 5 order to have a sphere with an effective radius Reff describing the roundness of the hair edge. We determined h from the plane on the hair axis where the cylindrical radius equals 85% of the saturated cylindrical radius. Results are shown in Fig. 3(c). The mean effective radius of the measured hairs was 3.00 ± 0.37 μm. To address the relationship between the effective radii Reff and the saturated cylindrical radii of the hairs rsat, the correlation of these two parameters were investigated [Fig. 3(d)]. The correlation clearly shows a linear relationship; they are well described by a linear regression model Reff = 0.79 rsat + 0.26 μm with R2 = 0.96. This linear relation of these radii implies that each human arm hair shares a morphological similarity with each other. In addition, the coefficient of proportionality less than one means that hair edges are sharper than ideal hemispheres, in accordance with the measured 3-D RI tomograms. 3.3 Effects of Hydrogen Peroxide on Individual Human Downy Hairs To further show the capability of QPI, we performed experiments to study the effects of hydrogen peroxide on individual hairs. Using the present method, four hairs were systematically measured before and after a 24-hour treatment with a 3% hydrogen peroxide solution (See Methods). Hydrogen peroxide is one of the widely used bleaching agents which induces irreversible alterations in the physicochemical properties of the melanin granules in hair proteins (45). It is also known that physical damage occurs to keratin proteins from oxidization. To investigate possible alterations induced by hydrogen peroxide, we measured the morphological and biochemical properties of hairs. The x−y cross-sectional slices of the reconstructed 3-D RI tomograms for four intact arm hairs [Figs. 4(a1−a4)] and the same hairs treated with 3% H2O2 [Figs. 4(a5−a8)] are shown in Fig. 4(a). The arrows and the dashed region denote the same fine structures, which were used to track the same hairs before and after the H2O2 treatment. Each graph in Fig. 4(b) and Fig. 4(c), respectively, describes the mean RIs and volumes of the 20-µm-length hairs before and after H2O2 treatment. To retrieve these parameters, the same procedures described above were used. The mean values of RI for intact and H2O2 treated hairs were 1.557 ± 0.003 and 1.556 ± 0.005, respectively. In addition, the mean volume for the intact and H2O2 treated hair groups was 724 ± 141 and 731 ± 166 fL, respectively. The paired t-test yielded p-values of 0.7330 for the mean RI and 0.7217 for the volume between the two hair groups, and thus does not show statistical difference before and after the treatment. (a1) (a2) (a3) (a4) r i a h t c a t n i 10 μm t (a5) n e m t a e r t 2 O 2 H (a6) (a7) (a8) n 1.6 1.58 1.56 1.54 1.52 (b) 1.6 x e d n i e v i t c a r f e r n a e m 1.58 1.56 1.54 1.52 1.5 intact hair H2O2 treatment (c) 1.5 ) h t g n e l ) L p ( e m u o v r i l a h m μ 0 2 ( 1.25 1 0.75 0.5 0.25 0 intact hair H2O2 treatment Fig. 4 Effects of hydrogen peroxide on human downy hairs. (a) x−y cross-sectional slices of RI tomograms for (a1−a4) intact hairs and (a5−a8) the same hairs after a 24 hr H2O2 treatment, respectively. (b) Mean RIs and (c) volumes of four hairs before and after the treatment. Same color denotes the same individual downy hair. The statistical indifference between the intact hairs and the H2O2 treated ones, however, does not imply that the edge structures of human downy hairs are not affected by hydrogen peroxide. It seems that alterations in the edge structures of body hairs by hydrogen peroxide are small as predicted by minute morphological changes in Figs. 4(a1−a8), and so comparable with the measurement errors. Noticeably, we could not 6 observe black stains maybe responsible for melanin pigments at hair edges, contrary to observations at thicker parts of the hair. The absence of melanin pigments at the hair edges possibly give rise to minute changes in the measured parameters by hydrogen peroxide. 4. Conclusion Herein, we performed quantitative and non-invasive optical measurements on human downy arm hairs using 3-D quantitative phase imaging. With a Mach-Zehnder interferometer equipped with a dual axis galvanometer mirror, 3-D RI tomograms and 2-D synthetic aperture images of individual hairs were measured. To fully exploit the quantitative imaging capability of the present method, biochemical and morphological parameters including mean RI, volume, cylinder radius, and effective radius for individual downy hairs were retrieved from the measured 3-D RI maps. Finally, the compositional and structural alterations in the downy hairs by hydrogen peroxide were also investigated with the present method. We expect that the unique advantages of the QPI techniques with its high-resolution 3-D RI imaging capabilities can be beneficial in investigating alterations in human hairs from various chemical agents such as bleaching agents (9, 13), shampoos, and rinses. Furthermore, structural variations in hair between different human species (6, 46) or gathered body sites (47, 48) can also be investigated from both compositional and morphological aspects. Furthermore, recently advanced QPI techniques including a readily implementable QPI unit (49, 50), real-time 3-D measurements (51), spectroscopic RI imaging (52- 55), polarization-sensitive imaging (56, 57), and a super-resolution technique (58) will also extend the applicability toward quantitative studies on the dynamic phenomena relevant to hair growths (59, 60) or formations of structural defects under diverse experimental conditions. Acknowledgements This work was supported by KAIST-Khalifa University Project, APCTP, the Korean Ministry of Education, Science and Technology, and the National Research Foundation (2012R1A1A1009082, 2014K1A3A1A09063027, 2013M3C1A3063046, 2012-M3C1A1-048860, 2014M3C1A3052537). P. E. Wheeler, "The evolution of bipedality and loss of functional body hair in hominids," Journal of Human P. E. Wheeler, "The loss of functional body hair in man: the influence of thermal environment, body form Reference 1. Evolution 13(1), 91-98 (1984) 2. and bipedality," Journal of Human Evolution 14(1), 23-28 (1985) 3. S. D O'Connor, K. L. Komisarek and J. D. Baldeschwieler, "Atomic force microscopy of human hair cuticles: a microscopic study of environmental effects on hair morphology," Journal of investigative dermatology 105(1), 96- 99 (1995) C. LaTorre and B. Bhushan, "Investigation of scale effects and directionality dependence on friction and 4. adhesion of human hair using AFM and macroscale friction test apparatus," Ultramicroscopy 106(8), 720-734 (2006) 5. S. Gurden, V. Monteiro, E. Longo and M. Ferreira, "Quantitative analysis and classification of AFM images of human hair," Journal of microscopy 215(1), 13-23 (2004) W. Hess, R. Seegmiller, J. Gardner, J. Allen and S. Barendregt, "Human hair morphology: a scanning 6. electron microscopy study on a male Caucasoid and a computerized classification of regional differences," Scanning microscopy 4(2), 375-386 (1990) 7. B. Lindelöf, B. Forslind and M.-A. Hedblad, "Human hair form: morphology revealed by light and scanning electron microscopy and computer aided three-dimensional reconstruction," Archives of dermatology 124(9), 1359- 1363 (1988) 8. of dermatology 101(3), 316-322 (1970) 9. damage," Okajimas folia anatomica Japonica 88(1), 1-9 (2011) T. IMAI, "The influence of hair bleach on the ultrastructure of human hair with special reference to hair R. Dawber and S. Comaish, "Scanning electron microscopy of normal and abnormal hair shafts," Archives 7 L. Pötsch, G. Skopp and J. Becker, "Ultrastructural alterations and environmental exposure influence the C. Hadjur, G. Daty, G. Madry and P. Corcuff, "Cosmetic assessment of the human hair by confocal J. Strassburger and M. M. Breuer, "Quantitative Fourier transform infrared spectroscopy of oxidized hair," Y. Miyamae, Y. Yamakawa and Y. Ozaki, "Evaluation of physical properties of human hair by diffuse F. Pragst and M. A. Balikova, "State of the art in hair analysis for detection of drug and alcohol abuse," M. JOURLIN and Y. DUVAULT, "3D reconstruction of human hair by confocal microscopy," J. Soc. Cosmet. W. Baumgartner, V. Hill and W. Blahd, "Hair analysis for drugs of abuse," J Forensic Sci 34(6), 1433-1453 10. opiate concentrations in hair of drug addicts," International journal of legal medicine 107(6), 301-305 (1995) 11. Chem 44(1-12 (1993) 12. microscopy," Scanning 24(2), 59-64 (2002) 13. J. Soc. Cosmet. Chem 36(1), 61-74 (1985) 14. D. Ammann, R. Becker, A. Kohl, J. Hänisch and I. Nehls, "Degradation of the ethyl glucuronide content in hair by hydrogen peroxide and a non-destructive assay for oxidative hair treatment using infra-red spectroscopy," Forensic science international 244(30-35 (2014) 15. reflectance near-infrared spectroscopy," Applied spectroscopy 61(2), 212-217 (2007) 16. Clinica Chimica Acta 370(1), 17-49 (2006) P. Kintz, Analytical and practical aspects of drug testing in hair, CRC Press (2006). 17. 18. S. Hartwig, V. Auwärter and F. Pragst, "Effect of hair care and hair cosmetics on the concentrations of fatty acid ethyl esters in hair as markers of chronically elevated alcohol consumption," Forensic science international 131(2), 90-97 (2003) 19. V. Auwärter, F. Sporkert, S. Hartwig, F. Pragst, H. Vater and A. Diefenbacher, "Fatty acid ethyl esters in hair as markers of alcohol consumption. Segmental hair analysis of alcoholics, social drinkers, and teetotalers," Clinical chemistry 47(12), 2114-2123 (2001) 20. (1989) G. Popescu, Quantitative Phase Imaging of Cells and Tissues, McGraw-Hill Professional (2011). 21. 22. K. Lee, K. Kim, J. Jung, J. H. Heo, S. Cho, S. Lee, G. Chang, Y. J. Jo, H. Park and Y. K. Park, "Quantitative phase imaging techniques for the study of cell pathophysiology: from principles to applications," Sensors 13(4), 4170- 4191 (2013) 23. Y. Park, C. A. Best, T. Auth, N. S. Gov, S. A. Safran, G. Popescu, S. Suresh and M. S. Feld, "Metabolic remodeling of the human red blood cell membrane," Proceedings of the National Academy of Sciences 107(4), 1289 (2010) 24. Y. Park, C. A. Best, K. Badizadegan, R. R. Dasari, M. S. Feld, T. Kuriabova, M. L. Henle, A. J. Levine and G. Popescu, "Measurement of red blood cell mechanics during morphological changes," Proceedings of the National Academy of Sciences 107(15), 6731 (2010) 25. N. T. Shaked, L. L. Satterwhite, M. J. Telen, G. A. Truskey and A. Wax, "Quantitative microscopy and nanoscopy of sickle red blood cells performed by wide field digital interferometry," J Biomed Opt 16(3), 030506 (2011) 26. H. Byun, T. R. Hillman, J. M. Higgins, M. Diez-Silva, Z. Peng, M. Dao, R. R. Dasari, S. Suresh and Y. Park, "Optical measurement of biomechanical properties of individual erythrocytes from a sickle cell patient," Acta biomaterialia (2012) 27. cells by digital holography," Cytometry Part A 85(12), 1030-1036 (2014) 28. G. Popescu, Y. Park, N. Lue, C. Best-Popescu, L. Deflores, R. R. Dasari, M. S. Feld and K. Badizadegan, "Optical imaging of cell mass and growth dynamics," American Journal of Physiology - Cell physiology 295(2), C538- 544 (2008) 29. P. Jourdain, N. Pavillon, C. Moratal, D. Boss, B. Rappaz, C. Depeursinge, P. Marquet and P. J. Magistretti, "Determination of transmembrane water fluxes in neurons elicited by glutamate ionotropic receptors and by the cotransporters KCC2 and NKCC1: a digital holographic microscopy study," The Journal of Neuroscience 31(33), 11846-11854 (2011) 30. Opt 16(11), 116017-1160177 (2011) 31. S. Lee, K. Kim, A. Mubarok, A. Panduwirawan, K. Lee, S. Lee, H. Park and Y. Park, "High-Resolution 3- D Refractive Index Tomography and 2-D Synthetic Aperture Imaging of Live Phytoplankton," Journal of the Optical Society of Korea 18(6), 691-697 (2014) Z. Wang, K. Tangella, A. Balla and G. Popescu, "Tissue refractive index as marker of disease," J Biomed P. Memmolo, L. Miccio, F. Merola, O. Gennari, P. A. Netti and P. Ferraro, "3D morphometry of red blood 8 V. Lauer, "New approach to optical diffraction tomography yielding a vector equation of diffraction M. Takeda, H. Ina and S. Kobayashi, "Fourier-transform method of fringe-pattern analysis for computer- K. Lee, H.-D. Kim, K. Kim, Y. Kim, T. R. Hillman, B. Min and Y. Park, "Synthetic Fourier transform light V. Mico, Z. Zalevsky, P. García-Martínez and J. García, "Synthetic aperture superresolution with multiple Y. Sung, W. Choi, C. Fang-Yen, K. Badizadegan, R. R. Dasari and M. S. Feld, "Optical diffraction W. Choi, C. Fang-Yen, K. Badizadegan, S. Oh, N. Lue, R. R. Dasari and M. S. Feld, "Tomographic phase S. K. Debnath and Y. Park, "Real-time quantitative phase imaging with a spatial phase-shifting algorithm," S. A. Alexandrov, T. R. Hillman, T. Gutzler and D. D. Sampson, "Synthetic aperture Fourier holographic K. Kim, H. Yoon, M. Diez-Silva, M. Dao, R. R. Dasari and Y. Park, "High-resolution three-dimensional 32. imaging of red blood cells parasitized by Plasmodium falciparum and in situ hemozoin crystals using optical diffraction tomography," J Biomed Opt 19(1), 011005-011005 (2014) 33. tomography and a novel tomographic microscope," Journal of Microscopy 205(2), 165-176 (2002) 34. based topography and interferometry," JosA 72(1), 156-160 (1982) 35. Optics letters 36(23), 4677-4679 (2011) 36. optical microscopy," Phys Rev Lett 97(16), 168102 (2006) 37. scattering," Optics express 21(19), 22453-22463 (2013) 38. off-axis holograms," JOSA A 23(12), 3162-3170 (2006) 39. T. R. Hillman, T. Gutzler, S. A. Alexandrov and D. D. Sampson, "High-resolution, wide-field object reconstruction with synthetic aperture Fourier holographic optical microscopy," Optics Express 17(10), 7873-7892 (2009) 40. imaging of live cells using fast Fourier phase microscopy," Applied Optics 46(10), 1836-1842 (2007) 41. tomography for high resolution live cell imaging," Optics express 17(1), 266-277 (2009) 42. microscopy," Nature methods 4(9), 717-719 (2007) 43. Y. Park, M. Diez-Silva, G. Popescu, G. Lykotrafitis, W. Choi, M. S. Feld and S. Suresh, "Refractive index maps and membrane dynamics of human red blood cells parasitized by Plasmodium falciparum," Proceedings of the National Academy of Sciences of the United States of America 105(37), 13730-13735 (2008) 44. of Criminal Law and Criminology (1931-1951) 746-752 (1941) 45. (1970) 46. Journal of the American Academy of Dermatology 48(6), S115-S119 (2003) 47. size and distribution in different body sites," Journal of Investigative Dermatology 122(1), 14-19 (2004) 48. anthropology 64(2), 179-184 (1984) 49. 50. tomography using a quantitative phase imaging unit," Optics letters 39(24), 6935-6938 (2014) 51. objects using optical diffraction tomography," Optics Express 21(26), 32269-32278 (2013) 52. quantitative phase imaging," Analytical chemistry 85(21), 10519-10525 (2013) 53. holographic microspectroscopy [Invited]," Applied optics 53(27), G111-G122 (2014) 54. hemoglobin concentrations in intact red blood cells," Optics letters 34(23), 3668-3670 (2009) 55. concentration and dynamic membrane fluctuation in red blood cells," Optics Express 20(9), 9673-9681 (2012) 56. spatio-temporally resolved Jones matrix," Optics Express 20(9), 9948-9955 (2012) 57. K. Lee and Y. Park, "Quantitative phase imaging unit," Optics Letters 39(12), 3630-3633 (2014) K. Kim, Z. Yaqoob, K. Lee, J. W. Kang, Y. Choi, P. Hosseini, P. T. So and Y. Park, "Diffraction optical Z. Wang, L. J. Millet, M. U. Gillette and G. Popescu, "Jones phase microscopy of transparent and anisotropic J. Jung, K. Kim, H. Yu, K. Lee, S. Lee, S. Nahm, H. Park and Y. Park, "Biomedical applications of M. Greenwell, A. Willner and P. L. Kirk, "Human hair studies. III. Refractive index of crown hair," Journal L. J. Wolfram, K. Hall and I. Hui, "The mechanism of hair bleaching," J. Soc. Cosmet. Chem 21(875-900 A. Franbourg, P. Hallegot, F. Baltenneck, C. Toutaina and F. Leroy, "Current research on ethnic hair," Y. Jang, J. Jang and Y. Park, "Dynamic spectroscopic phase microscopy for quantifying hemoglobin Y. Kim, J. Jeong, J. Jang, M. W. Kim and Y. Park, "Polarization holographic microscopy for extracting K. Kim, K. S. Kim, H. Park, J. C. Ye and Y. Park, "Real-time visualization of 3-D dynamic microscopic J.-H. Jung, J. Jang and Y. Park, "Spectro-refractometry of individual microscopic objects using swept-source N. Otberg, H. Richter, H. Schaefer, U. Blume-Peytavi, W. Sterry and J. Lademann, "Variations of hair follicle O. Lunde, "A study of body hair density and distribution in normal women," American journal of physical N. Lue, W. Choi, G. Popescu, T. Ikeda, R. R. Dasari, K. Badizadegan and M. S. Feld, "Quantitative phase Y. Park, T. Yamauchi, W. Choi, R. Dasari and M. S. Feld, "Spectroscopic phase microscopy for quantifying 9 Y. Cotte, F. Toy, P. Jourdain, N. Pavillon, D. Boss, P. Magistretti, P. Marquet and C. Depeursinge, "Marker- samples," Optics letters 33(11), 1270-1272 (2008) 58. free phase nanoscopy," Nature photonics 7(2), 113-117 (2013) 59. R. Hoffmann, W. Eicheler, A. Huth, E. Wenzel and R. Happle, "Cytokines and growth factors influence hair growth in vitro. Possible implications for the pathogenesis and treatment of alopecia areata," Archives of dermatological research 288(3), 153-156 (1996) 60. 471 (1990) M. P. Philpott, M. R. Green and T. Kealey, "Human hair growth in vitro," Journal of cell science 97(3), 463- 10
1505.07413
1
1505
2015-05-27T17:44:04
The free energy cost of reducing noise while maintaining a high sensitivity
[ "physics.bio-ph", "cond-mat.stat-mech", "q-bio.SC" ]
Living systems need to be highly responsive, and also to keep fluctuations low. These goals are incompatible in equilibrium systems due to the Fluctuation Dissipation Theorem (FDT). Here, we show that biological sensory systems, driven far from equilibrium by free energy consumption, can reduce their intrinsic fluctuations while maintaining high responsiveness. By developing a continuum theory of the E. coli chemotaxis pathway, we demonstrate that adaptation can be understood as a non-equilibrium phase transition controlled by free energy dissipation, and it is characterized by a breaking of the FDT. We show that the maximum response at short time is enhanced by free energy dissipation. At the same time, the low frequency fluctuations and the adaptation error decrease with the free energy dissipation algebraically and exponentially, respectively.
physics.bio-ph
physics
The free energy cost of reducing noise while maintaining a high sensitivity 1Max Planck Institute for the Physics of Complex Systems. Noethnitzer Strasse 38 , Pablo Sartori1 and Yuhai Tu2. 2IBM T.J. Watson Research Center, 01187, Dresden, Germany. 1101 Kitchawan Road, Yorktown Heights, New York 10598, USA,. Living systems need to be highly responsive, and also to keep fluctuations low. These goals are incompatible in equilibrium systems due to the Fluctuation Dissipation Theorem (FDT). Here, we show that biological sensory systems, driven far from equilibrium by free energy consumption, can reduce their intrinsic fluctuations while maintaining high responsiveness. By developing a continuum theory of the E. coli chemotaxis pathway, we demonstrate that adaptation can be understood as a non-equilibrium phase transition controlled by free energy dissipation, and it is characterized by a breaking of the FDT. We show that the maximum response at short time is enhanced by free energy dissipation. At the same time, the low frequency fluctuations and the adaptation error decrease with the free energy dissipation algebraically and exponentially, respectively. PACS numbers: 87.10.Vg, 87.18.Tt, 05.70.Ln Living organisms need to respond to external signals with high sensitivity, and at the same time, they also need to control their internal fluctuations in the absence of signal. In equilibrium systems, the fluctuation dissi- pation theorem (FDT) dictates that these two desirable properties, high sensitivity and low fluctuation, can not be satisfied simultaneously. Most sensory and regula- tory functions in biology are carried out by biochemical networks that operate out of equilibrium – metabolic en- ergy is spent to drive the dynamics of the network [1–4]. Thus, in principle they are not constrained by the FDT [5]. However, how fluctuations, energy dissipation, and sensitivity are related for such systems remains not well understood. Here, we address this question by studying a negative feedback network responsible for adaptation in the bacterial chemosensory system [6–9]. A typical adaptive behavior in a small system such as a single cell is shown in Fig. 1A [10]. In response to a change of the signal S, the output y of the sensory system first changes quickly with a fast time scale τy. After the fast response, the output slowly adapts back towards its pre-stimulus level aad with an adaptation time τad (cid:29) τy. The new steady state (adapted) output may differ from the pre-stimulus value, and the difference is quantified by the adaptation error . In our previous work [11], we showed that the negative feedback network responsible for adaptation operates out of equilibrium with a finite free energy dissipation rate W . The average adaptation error (cid:104)(cid:105) was found to decrease exponentially with W τad. However, how the variance σ2  of the error behaves in an adaptive system still remains unknown. This is an im- portant question as adaptive feedback systems are intrin- sically noisy due to the slow adaptation dynamics [12]. tion. For equilibrium systems, under the general assump- tion that response and signal are conjugate variables, R(0) +(cid:82) t In the linear response regime, the output response of a system to an input signal S(t) is given by R(t) = 0 χ(t − t(cid:48))S(t(cid:48))dt(cid:48), where χ is the response func- FIG. 1: Noisy response of feedback adaptation. A) Adaptive output response to a step input signal increase at time 0. After a sharp response in a time τy, the output y recovers back in a time τad to its adapted value aad. The adaptation error is characterized by its average (cid:104)(cid:105), as well as its variance σ. B) Schematic of the feedback adaptation model. Transitions between the active and inactive memory energy landscapes, f1 and f0, are mediated via equilibrium activity transitions with rates, ω0 and ω1. An external energy input µ is used to drive the memory variable uphill in both the active and inactive states. The result is a dissipative loop of probability flow around the adapted memory state mad, which ensures the output to be near aad. the FDT establishes that χ(t) = −β∂tCR(t)Θ(t), where CR(t) ≡ (cid:104)R(t)R(0)(cid:105) − (cid:104)R(cid:105)2 is the auto-correlation func- tion, Θ(t) is the Heaviside function, and β = (kBT )−1 is the inverse thermal energy set to unity hereafter. For a small step stimulus S(t) = S0Θ(t), integration of the FDT leads to a relation between the response and its correlation: R(t) = R(0) − S0(CR(t) − CR(0)). Since for equilibrium systems CR(t) is a monotonically decreasing function of time [13], the response R(t) is also monotonic in time, and thus no adaptation dynam- ics is possible. Furthermore, the long time response ∆R ≡ R(t = ∞) − R(0) is linearly proportional to the variance σ2 In this paper, we show that in a non-equilibrium adap- tive system both the average adaptation error (cid:104)(cid:105) (anal- ogous to ∆R) and its variance σ2 R) are suppressed by the free energy dissipation of the system but in different ways, which results to a nonlinear (log- R = CR(0), i.e., ∆R = S0σ2 R.  (analogous to σ2 ABmemory, mfree energytimerelative outputsignal increase01 arithmic) relationship between them. More importantly, violation of the FDT allows suppression of noise without compromising the strength of the fast response. The continuous model of feed-back adaptation. We start by introducing a discrete adaptation model moti- vated by the E. coli chemotaxis pathway. The system is characterized by its binary receptor activity A = 0, 1, which determines the output y; and an internal con- trol variable M = 0, 1, . . . , N , that corresponds to the chemoreceptor's methylation level in E. coli chemotaxis [9]. For a given external input signal S, the free energy of the system can be written as: FA(M, S) = −(A − 1/2)[(M − Mr)E − (S − Sr)], (1) where Sr is a reference signal at a methylation level Mr, and E sets the methylation energy scale. For E. coli chemotaxis, the signal S depends on the ligand attractant concentration logarithmically [14]. The dynamics of the system is characterized by the transitions between the 2 × (N + 1) states in the A × M phase space. The receptor activity switches at a time scale τa, which is much shorter than the adaptation time scale τad at which the internal variable M is con- trolled. The receptor activity A determines the out- put y of the signaling pathway. In the case of E. coli chemotaxis, this is carried out by the phosphorylation and dephosphorylation reactions of the response regu- lator CheY with an intermediate time scale τy: τad (cid:29) (cid:82) t τy (cid:29) τa. To account for this, we relate A and y by y(t) = τ−1 −∞ e(t(cid:48)−t)/τy A(t(cid:48))dt(cid:48), which averages the fast binary activity over the time scale τy. y According to Eq. (1), a larger signal S favors the inac- tive state A = 0. Thus, an increase in S quickly reduces the system's average activity, at time scale ∼ τa, and output, at time scale ∼ τy, as represented in Fig. 1A. Af- ter this sudden initial response, the system slowly adapts by adjusting its internal variable M to balance the effect of the increased signal. Due to its slow time scale, M effectively serves as a memory of the system. This adap- tation process restores activity and output to a level near its pre-stimulus value (cid:104)A(cid:105) = (cid:104)y(cid:105) ≈ aad. Although highly precise, this adaptation process is imperfect, and its in- accuracies are quantified by the adaptation error , which we define as . (2) y − aad aad  = For E. coli chemotaxis, the adaptive machinery consists of chemical reactions that increase M in the inactive state and decrease it in the active state. Note from Eq. (1) that such regulatory reactions are energetically unfavorable, and thus require a chemical driving force µ, see Fig. 1B. To gain analytical insights about dynamics and ener- getics of adaptation, we consider the limit where N → ∞ and m = M/N ∈ [0, 1] becomes a continuous variable 2 [15]. Note that free energy and bare rates need to be rescaled for the continuum limit to converge (see Supple- mentary Information, SI, for details). Proceeding in this way we obtain two coupled Fokker-Planck equations that describe the chemotaxis pathway dynamics: ∂tp1 = p0ω0 − p1ω1 − ∂mJ1 ∂tp0 = p1ω1 − p0ω0 − ∂mJ0, (3) where p1(m, t) and p0(m, t) are the joint probabilities for the active and inactive states with a given m respectively. The probability currents are given by JA = DA ((−∂mfA − µ)pA − ∂mpA) , A = 0, 1, (4) where fA(m) = −(A − 1/2)[(m − mr)e − (S − Sr)] is the continuum limit of Eq. (1) characterized by the rescaled energy parameter e = N E. The fast transition rates between the active and inactive states, ω0 and ω1, satisfy detailed balance ω0/ω1 = exp(f0−f1). The diffusion-like constants D1 and D0 set the time-scale of m changes for active and inactive states, and thus the adaptation time goes as τad ∼ D−1 A , see SI for details. Our model is analogous to that of an isothermal ratchet [16], where a chemical driving fuels directed motion. Whereas in ratchets µ drives directed motion, here it fuels currents up the energy landscapes f0 and f1 to achieve adaptation. In the absence of external driving, i.e. µ = 0, the system relaxes to a state of thermal equilibrium with no phase-space fluxes J0 = J1 = 0. In this regime adapta- tion is impossible. The chemical driving µ > 0 breaks detailed balance and creates currents that increase m in the active state and decrease it in the inactive state. For large enough µ, the memory variable m can be stabilized (trapped) in a cycle around its adapted state mad, which ensures (cid:104)y(cid:105) ≈ aad as illustrated in Fig. 1B. The free en- ergy dissipation rate W can be computed W ≈ Cµ/τad with C a system specific constant set to unity by our pa- rameter choice, see SI. In the following, we will use the chemical driving µ ≈ τad W to characterize the system's energy dissipation. The dynamics of A, y, and m are illustrated in Fig. 2A. In the power spectra of Fig. 2B, the high fre- quency fluctuation of y is suppressed from that of A by time-averaging. However, the low frequency fluctuations, which are caused by the slow fluctuations of m, are not affected. These low frequency noise can be suppressed by free energy dissipation, as we show later in this paper. Adaptation as a non-equilibrium phase transition. Given the separation of time scales τa (cid:28) τad, we can solve Eqs. (3) by using the adiabatic approximation [13, 17]: p1(m) = a(m)p(m), and p0(m) = (1 − a(m))p(m), with a(m) = (1 + ef1(m)−f0(m))−1 the average equilibrated ac- tivity for a fixed value of m. The distribution of m can be written as p(m) = e−h(m,S)/Z with h the effective potential and Z a normalization constant. We have de- termined the effective potential h analytically (see SI for 3 tem develops a stable fixed point at mad away from the boundaries, indicating the onset of adaptive behaviors [18]. As µ increases this fixed point becomes increasingly stable, and adaptation accuracy increases. The transition of a feedback system to adaptation can thus be loosely understood as a continuous phase transition (see SI for details). Since the control parameter is the free energy dissipation, the transition to adaptation occurs far from equilibrium and a breaking of FDT is to be expected. Breakdown of Fluctuation Dissipation Theorem. In our feedback model, the observable conjugate to the signal is eA = −∂SfA, but see [19–21] for cases where this is not true. The FDT would lead to χ(t) = e∂tCA(t), where χ is the activity response function and CA is the monotonic correlation function. In an adaptive system the integral of χ, which is just the response to a step stimulus, is non-monotonic, therefore FDT is broken. To quantify the departure from equilibrium, we define an effective temperature Teff using the formulation of the FDT in frequency space [5, 22], see inset in Fig. 2D. The frequency-dependence of Teff for µ > 0 implies a breakdown of FDT. As shown in Fig. 2D, while for any value µ (cid:54)= 0 we have Teff (cid:54)= 1, after the transition to the adaptive regime µ ≥ µc a divergence occurs. This cor- responds to the appearance of a frequency region where Im[χ(ω)] < 0. A negative effective viscosity indicates the dominance of the active effects that drive a net current to flow against the gradients of the equilibrium energy landscape fA, something also observed in other biologi- cal systems such as collections of motors [23] or the inner ear hair bundle [5]. The breakdown of FDT means that there is no a priori connection among fluctuations σ2  , chemical driving µ, and long-time response (cid:104)(cid:105). In the following we derive relations linking these three quanti- ties in the adaptive feedback system studied here. The free energy cost of suppressing fluctuations. As evident from the effective potential, increasing the chem- ical driving µ stabilizes the adapted state. In the limit µ → ∞, the system thus goes to its perfectly adapted state with average output aad = D0/(D0 + D1). For finite µ, the output differs from aad, which can be char- acterized by the average error (cid:104)(cid:105) and its variance σ2  . The average adaptation error is (cid:104)(cid:105) = ((cid:104)y(cid:105) − aad)/aad. Summing and integrating Eqs. (3), we have (cid:104)(cid:105) = D1p1(1) + D0p0(1) D0(e/2 − µ) − D1p1(0) + D0p0(0) D0(e/2 − µ) . (6) Thus to obtain the adaptation error we only need to eval- uate the probability at the boundaries. In the limit of µ (cid:29) µc, we have: (cid:104)(cid:105) ≈ ce−kµµc , (7) where k and c are constants with only weak dependence on µ (see SI for derivation). This shows explicitly that the adaptation error goes down exponentially with energy y < ω < τ−1 y −1 ad < ω < τ−1 FIG. 2: Adaptation as a non-equilibrium transition A) Schematic time traces of the binary activity A (blue), the out- put y (black), and the memory M (red) in steady state. The slow M variations induce large fluctuations in the output y, while the fast A switching for a fixed M only produces small fluctuations in y. B) Power spectra of the activity SA and output Sy. The output noise is filtered (reduced) in the high frequency range τ−1 a ; but it remains unfiltered in the range τ . C) Effective memory poten- tial in Eq. 5 for three values of the chemical driving µ (due to the choice D1 = D0 taken here, mad = m∗). At equi- librium, µ = 0, the adapted memory state mad is unstable. At the value µ = µc the system becomes critical. In the re- gion µ > µc the adapted state mad is stable. Inset: Activity response to step signal increase for corresponding values of µ. D) Effective temperature Teff for three different values of the chemical driving µ. After the onset of adaptation a re- gion with "negative friction" develops, at the end of which the effective temperature diverges. Values of µ from lighter to darker blue are µ = 0, µ = 0.65µc, and µ = 20µc (the same as in panel C). The other parameters are from [17], see SI. detailed derivation): h(m, S) = µ µc ln[D0e−(m−m∗)e/2 + D1e(m−m∗)e/2] − ln[e−(m−m∗)e/2 + e(m−m∗)e/2] , (5) where we have defined the critical chemical driving as µc = e/2, and m∗ = mr + (S − Sr)/e. The analytical form of the effective potential is one of the main results of this paper. The effect of energy dissipation and the onset of adaptation can be under- stood intuitively with h(m, S), which contains two terms with similar shapes, see Fig. 2C. The first term (propor- tional to µ/µc) in the right hand side of Eq. (5) comes from chemical driving (non-equilibrium effect) and has a stable free energy minimum. The second term is the equilibrium potential in the absence of driving, and has a maximum at m∗. At equilibrium the only critical point m∗ is unstable, so the system tends to go to the bound- aries without adapting. As µ increases the first part of the potential starts to dominate. For µ > µc, the sys- CDAtimetime10-110-210-310-410-510-610-1100101102103104frequency (s-1)power spectrumOutput noiseActivity noiseBfrequency (s-1)1210-11001011021033step responseeffective potential10memory, m0-1activity, Aoutput, y0N01memory, M dissipation, as found numerically in our previous work for the discrete model [11]. Here, we show this relationship analytically in the continuum limit. Fig. 3A shows results from both the continuum and discrete model. 4 which vanishes when µ → ∞. This is a main result of the paper, which shows that energy dissipation is used to reduce error noise by suppressing slow activity fluc- tuations. Figure 3B compares this expression to several discrete models with increasing N . FIG. 3: Free energy cost of reducing error and noise. A). Dependence of average error with chemical driving for several system sizes. The decay is exponential, in agreement with the infinite size limit (dashed red). Saturation of the decay for finite N is due to finite size effects. B) Adaptation noise as a function of chemical driving for several system sizes, together with the analytical estimate in dashed red. At very large driving the noise saturates to its minimum σym dictated by the intrinsic activity fluctuations. Note that at the critical driving µc the analytical estimate diverges. This divergence is smoothed for finite N . Besides stabilizing the adapted state, Eq. (5) shows that increasing µ also reduces the m−fluctuations by making the effective potential sharper. The reduction in these fluctuations implies a decrease in the variance of the error σ2  . Taking into account the separation of time scales τa (cid:28) τy (cid:28) τad, the variance of the output y can be approximated as the sum of two variances σ2 ym and σ2 a. They respectively correspond to variation of y at time scale ∼ τy around its average a(m) for a fixed m, and the variation of a(m) due to variation of m at the adaptation time ∼ τad. We thus have a)/a2 ad  ≈ (σ2 σ2 ym + σ2 (8) . The variance σ2 ym of y is caused by the fast fluctuations of the binary variable A at timescale ∼ τa averaged over the output timescale τy (cid:29) τa (see SI for derivation): , (cid:82) 1 ym = (aad − a2 σ2 which clearly shows that σ2 averaging. ad)τa/(τy + τa) ym ∝ τa/τy is reduced by time- a = (cid:104)a(cid:105)2 − (cid:104)a2(cid:105), where (cid:104)an(cid:105) = The variance σ2 0 an(m)p(m)dm for n = 1, 2, is caused by the slow vari- ation of m. To obtain an analytical expression for σ2 we approximate p(m) by a Gaussian, valid for µ (cid:29) µc. a This results in σ2 ad. The variance m = (cid:104)m(cid:105)2 − (cid:104)m2(cid:105) within the same Gaussian approxi- σ2 m ≈ (µµc)−1. Defining now mation of p(m) is given by σ2 c = 4(1−aad)2a2 a characteristic variance as σ2 ad, we finally have: a ≈ (∂maad)2σ2 m/a2 a ≈ σ2 σ2 c µc/µ , (9) FIG. 4: Response and correlations in systems out of equilibrium. A) (top panel) Average output response to a signal decrease for several values of the chemical driving be- yond µc, see the color-code for µ in panel B. As the chemical driving µ increases, the maximal transient response (cid:104)y(cid:105)max in- creases, but the long time response ∆(cid:104)y(cid:105) = aad(cid:104)(cid:105) decreases. (bottom panel) The correlation function also decreases as the system is driven further away from equilibrium. B) The de- pendence of (cid:104)y(cid:105)max, ∆(cid:104)y(cid:105), and (cid:104)y(cid:105)max on the chemical driving µ. The long-time response (adaptation error) ∆(cid:104)y(cid:105) decreases quickly with µ. The decrease of the output fluctuation (noise) σy with µ is more gradual, and controls the increase in the maximal response (cid:104)y(cid:105)max for large µ. In this figure N = 15, and S = Sr. Discussion. Biochemical networks are non-equilibrium systems fueled by free energy dissipation to achieve their biological functions. Energy dissipation liberates the net- works from constraints such as the Fluctuation Dissipa- tion Theorem and Detailed Balance. Here, we show in a negative feedback network that the long-time output response ∆(cid:104)y(cid:105) = aad(cid:104)(cid:105) decreases with the free energy dissipation µ ≈ τad W exponentially, and its fluctuation  decreases as µ−1. Both these effects, espe- y = a2 σ2 cially the slower decay of σ2 y with µ, contribute to en- hance the short time response (cid:104)y(cid:105)max, see Fig. 4. adσ2 ym, where d = kσ2 Even though FDT is broken in the adaptive sys- tem studied here, fluctuations and long-time response y ≈ of the output are linked via a non-linear relation: σ2 c and yc = aad(cid:104)(cid:105). dµc/ log(yc/∆(cid:104)y(cid:105))+σ2 Unlike the linear non-equilibrium FDT derived by a change of observables [24–26], our non-linear relation links observables that are conjugate at equilibrium, mak- ing it particularly appealing. Another approach is taken in [27], where near equilibrium linear response is used to show that the dispersion of variables can be reduced by dissipation. Adaptation however is a far from equilib- rium phenomenon which requires a critical finite amount of free energy dissipation. As a result, the energy scale is set by the intrinsic energy µc instead of the thermal energy kBT in [27]. It remains a challenging question average error,error noise,N=5N=15N=50FPEN=5N=15N=5005010015010-210-410-6100chemical driving,A050100150chemical driving,0.00.20.60.4Btime, (s)02468100.000.100.050.00.40.2AB0.00.20.10.30.40102030405060chemical driving, signal decrease 5 whether these approaches can be combined to obtain a general relationship among response, fluctuation, and en- ergy dissipation for systems far from equilibrium. This work is partly supported by a NIH grant (R01GM081747 to YT). We thank Leo Granger and Jor- dan Horowitz for a critical reading of this manuscript. [1] H. Qian, Annu. Rev. Phys. Chem. 58, 113 (2007). [2] P. Mehta and D. J. Schwab, Proceedings of the National Academy of Sciences 109, 17978 (2012). [3] J. E. Niven and S. B. Laughlin, Journal of Experimental Biology 211, 1792 (2008). [4] C. H. Bennett, BioSystems 11, 85 (1979). [5] P. Martin, A. Hudspeth, and F. Julicher, Proceedings of the National Academy of Sciences 98, 14380 (2001). [6] H. C. Berg, D. A. Brown, et al., Nature 239, 500 (1972). [7] S. M. Block, J. E. Segall, and H. C. Berg, Journal of bacteriology 154, 312 (1983). [8] N. Barkai and S. Leibler, Nature 387, 913 (1997). [9] Y. Tu, Annual review of biophysics 42, 337 (2013). [10] D. E. Koshland, A. Goldbeter, and J. B. Stock, Science 217, 220 (1982). [11] G. Lan, P. Sartori, S. Neumann, V. Sourjik, and Y. Tu, Nature physics (2012). [12] P. Sartori and Y. Tu, Journal of statistical physics 142, 1206 (2011). [13] N. G. Van Kampen, Stochastic Processes in Physics and Chemistry (Elsevier Ltd., New York, 2007), 3rd ed. [14] Y. Kalinin, L. Jiang, Y. Tu, and M. Wu, Biophysical journal 96, 2439 (2009). [15] C. Gardiner, Applied Optics 25, 3145 (1986). [16] A. Parmeggiani, F. Julicher, A. Ajdari, and J. Prost, Physical Review E 60, 2127 (1999). [17] Y. Tu, T. S. Shimizu, and H. C. Berg, Proceedings of the National Academy of Sciences 105, 14855 (2008). [18] A. E. Allahverdyan and Q. A. Wang, Phys. Rev. E 87, 032139 (2013). [19] W. Buijsman and M. Sheinman, arXiv preprint arXiv:1212.5712 (2012). [20] G. De Palo and R. G. Endres, PLOS Computational Bi- ology 9, e1003300 (2013). [21] P. Sartori, L. Granger, C. F. Lee, and J. M. Horowitz, PLoS computational biology 10, e1003974 (2014). [22] L. F. Cugliandolo, J. Kurchan, and L. Peliti, Physical Review E 55, 3898 (1997). [23] F. Julicher and J. Prost, Physical review letters 78, 4510 (1997). [24] D. Bedeaux, S. Milosevic, and G. Paul, Journal of Sta- tistical Physics 3, 39 (1971). [25] J. Prost, J.-F. Joanny, and J. Parrondo, Physical review letters 103, 090601 (2009). [26] U. Seifert and T. Speck, EPL (Europhysics Letters) 89, 10007 (2010). [27] A. C. Barato and U. Seifert, Physical Review Letters 114, 158101 (2015). SUPPLEMENTARY MATERIAL Continuum limit of feedback adaptation. The out-of-equilibrium dynamics of the discrete model in its phase space A × M ∈ {[0, 1], [0, N ]} are governed by six sets of rates. The rates ω0(M ) govern the transi- tions from the inactive A = 0 states to the active A = 1 states, and the rates ω1(M ) the reciprocal inactivation transitions. The memory of active states is increased with a rate k+ 1 (M ). For inactive states, the memory can increase with a rate 0 (M ) and decrease with k− k+ Given the free energy from Eq. 1, we have that the rates of passive activity transitions satisfy detailed bal- ance, 1 (M ) and decreased with a rate k− 0 (M ). ω1(M ) ω0(M ) = eF1(M )−F0(M ) = e−[(M−Mr)E−(S−Sr)] . The memory transitions are driven out of equilibrium, and in general we have = eF0(M )−F0(M +1)+G = e−E/2+G = eF1(M )−F1(M +1)+G = eE/2−G, k+ 0 k− 0 k+ 1 k− 1 where the ratios are independent of M , and G is the free energy input in the reactions which keeps them out of equilibrium. When there is no free energy input the sys- tem satisfies detailed balance and its dynamic are simple equilibrium relaxation. For large values of G the memory only increases for inactive states and decreases for active ones, the chemotaxis limit [11]. The dynamics of this system are governed by the mas- ter equation. For the bulk states it is best written as two coupled equations ∂tP1(M ) = P0(M )ω0(M ) + P1(M − 1)k+ + P1(M + 1)k− 1 − P1(M )[ω1(M ) + k+ 1 + k− 1 ] ∂tP0(M ) = P0(M )ω0(M ) + P1(M − 1)k+ 1 1 , . + P1(M + 1)k− 1 − P1(M )[ω1(M ) + k+ 1 + k− 1 ] For the upper boundary M = N we have ∂tPA(N ) = P1−A(N )ω1−A(N ) + PA(N − 1)k+ A − PA(N )[ωA(N ) + k− A ] 6 To obtain the continuum theory we perform an expan- sion in the number of memory states N , and then take the limit N → ∞. This changes the discrete variable M ∈ [0, N ] to the continuous m ∈ [0, 1]. The proba- bility density is defined by pA(m) = P (A, M )/N , where m = M/N . Furthermore, for the continuum limit to ex- ist, the parameters µ = N G, e = N E and mr = Mr/N have to be kept constant as N → ∞. Using this, the continuous free energy is simply given by fA(m) = −(A − 1/2)[(m − mr)e − (S − Sr)] , which through detailed balance defines the continuous activation and inactivation rates ω0(m) and ω1(m) up to a time scale. Dividing now the bulk master equations by N and expanding to second order in 1/N results in two chemically coupled Fokker-Planck equations ∂tp1 = p0ω0 − p1ω1 − H1∂mp1 + D1∂2 ∂tp0 = p1ω1 − p0ω0 − H0∂mp0 + D0∂2 mp1 mp0 (10) , where the drift and diffusion coefficients are given by HA = lim N→∞ A − k− k+ N A ; DA = lim N→∞ A + k− k+ 2N 2 A . To calculate the first limit, we can use that A − k− k+ N A = 2N A + k− k+ 2N 2 A 1 − k− 1 + k− A A /k+ A /k+ A , and expand in 1/N the ratio of rates, which results in H0 = −(e/2 − µ)D0 and H1 = (e/2 − µ)D1. Note that H0 = D0(−∂mf0 + µ) and H1 = D1(−∂mf1 − µ), where the difference in sign in front of the chemical driving is due to the fact that memory transitions are driven in opposing directions for active and inactive states. Finally, proceeding in the same way on the discrete boundary equations gives at m = 1 and m = 0 the boundary condition 0 = HApA − DA∂mpA , which is nothing but a no-flux condition. A verification of the convergence of the discrete to the continuous solution on the steady state appears in Fig. 5 A, where the discrete model was solved numerically for increasing values of N . , Self-consistent choice of diffusion constants and analogously for the lower boundary M = 0, we have ∂tPA(0) = P1−A(0)ω1−A(N ) + PA(1)k− A − PA(0)[ωA(0) + k+ A] . A general solution of the master equation can be obtained using standard linear algebra [13]. The constants D0 and D1 with units of frequency are analogous to the diffusion constant of a random walk. They define the characteristic time scale of the m-dynamics, as well as the value of the adapted activ- ity. In fact, the adapted activity for the discrete model is reached when G → ∞ and k+ 0 , which results 1 = 0 = k− 7 as can be verified by taking the limit N → ∞ which de- fines DA. For this choice, we have that in the irreversible limit the "diffusion" term of the master equation for fi- nite N is G→∞ k+ lim A + k− A = 2N τad , ultimately determined by the adaptation time. The same is true for the "drift" of active and inactive memory states, which are given by G→∞ k+ lim 0 − k− 0 = 2N τad and G→∞ k+ lim 1 − k− 1 = − 2N τad . These choices of time scales, which together with the ratio of the rates uniquely defines all rates, ensures that the adaptation time takes a value of order τad in the discrete and continuum models alike. Indeed, as can be seen in Fig. 5 D, the response of the activity for a step change in ligand in the adaptive regimes is essentially the same for the continuous and discrete cases. Perturbative expansion for fast activity transitions Even at the steady state, Eqs. 10 are hard to solve analytically without any further assumption. In most sensory adaptive systems however there is a clear separa- tion between the adaptation time τad and the activation time τa (cid:28) τad, which we can use to obtain approximate solutions of pA(m). Consider that ωA = ¯ωA/τa, where from now on bars will denote dimensionless quantities, and ¯ω are of order unity. We can then define the param- eter δ = τa/τad, and write the dimensionless steady state equations p0 ¯ω0 − p1 ¯ω1 − δ p1 ¯ω1 − p0 ¯ω0 − δ e + µ e + µ ∂m ¯J1 = 0 ∂m ¯J0 = 0 , (12) where the dimensionless fluxes are defined using ¯DA. In addition to these equations, because of total flux conser- vation, we have that ∂m( ¯J0 + ¯J1) = 0, which together with the boundary conditions gives ¯J0 + ¯J1 = 0 . (13) Since typically in adaptive sensory systems δ (cid:28) 1, it is natural to expand the steady state probabilities as pA = p(0) A + δp(1) A + δ2p(2) A + . . . (14) We now provide the equations for the first two terms of this expansion. Zeroth order. We begin by using the definition of 1 = a(0)p(0) and conditional probability, which gives p(0) FIG. 5: A) Convergence of the steady state solution of the ME P (M ) to the FPE solution p(m). Parameters as elsewhere in the text, with µ = 20µc and S = Sr. B) Numerical solu- tion of the FPE (black) and two analytical approximations, the Gaussian (red) and the adiabatic using the potential h. Parameters as in B. C) In black, the equilibrium contribution to ∂mh, and in shades of blue the non-equilibrium contribu- tion for different values of µ. Both in units of e and as a function of m, with the range [0, 1] delimited by dashed lines. Sinc these terms have opposing sign, see Eq. 16, extrema cor- respond to points where a blue line crosses the black line, and if the slope of the blue curve is higher than that of the black one they are minima. As one can see, the maximum (in red) disappears through the left boundary, while the minimum (in green) comes through the right one. This feature makes the transition second order. 0 + k− in aad = k+ D0 and D1, this gives 0 /(k+ 1 ) [11]. Using the expressions for aad = D0 D0 + D1 . (11) The time scale of adaptation τad defines the rate at which the memory relaxes, which in the discrete case is given by 0 + k− k+ 1 . In the continuum limit however the relaxation rate of the memory is given by (e − µ)(D0 + D1). Since accurate adaptation is reached in the regime µ → ∞, which should not affect the adaptation time, we choose the diffusion-like constants as DA = ¯DA τad(e + µ) , with the dimensionless constants ¯DA of order one chosen such that Eq. 11 is satisfied. Note that this choice of diffusion constants limits the choice of rates in the discrete model, of which we have so far only specified the ratios. One choice of time scale compatible with the DA above is A + k− k+ A = 2N τad 1 + eE/2+G eE/2+G − 1 , BAC1/2-1/2100.00.20.40.60.81.0024680.00.20.40.60.81.002468FPEadiabticgaussmM/NFPEN=125N=25N=15N=5 0 = (1 − a(0))p(0). Here a(0) is the conditional activity p(0) distribution to zeroth order, which we call a(m) in the main text; and p(0) the marginal memory distribution to zeroth order, noted in the main text as p(m). To zeroth order, Eq. 12 establishes that ¯ω0p(0) 0 = ¯ω1p(0) 1 , which is a no-flow condition for the activity transitions. From it, we can derive that a(0) = ω0/(ω1+ω0). Inserting this in Eq. 13 gives to lowest order ( ¯H1a(0) + ¯H0(1 − a(0)))p(0) − ∂m[( ¯D1a(0) + ¯D0(1 − a(0)))p(0)] = 0 , This equation has a solution of the form p(0) = exp(−h)/Z, where Z is a normalization constant and h is an effective non-equilibrium potential. This potential can be integrated analytically, which results in h(m, S) = + log (cid:20) (cid:18) me (cid:18) me 2 (cid:21)(cid:19) 1 (cid:20) 1 + ω0/ω1 − µ e/2 + log 2 1 D0/D1 + ω0/ω1 (cid:21)(cid:19) . (15) The validity of this solution can be verified by substitu- tion in Eq. 15. (cid:82) 1 First order. The correction of first order in δ is p(1) A . Since we preserve the normalization of the zeroth order term, the first order term has as normalization condition 0 (p(1) 1 (m))dm = 0. From Eq. 12, the first order term directly gives 0 (m) + p(1) ¯ω0p(1) 0 − ¯ω1p(1) 1 − ∂m ¯J (0) 1 e + µ = 0 , which shows that to this order part of the m-flux is de- viated as probability currents on the activity transitions. The global flux balance in Eq. 13 becomes ¯H1p(1) 1 + ¯H0p(1) 0 − ∂m[ ¯D1p(1) 1 + ¯D0p(0) 0 ] = 0 . Unfortunately these two equations can not be solved an- alytically. In the following, analytical approximations are made using only the zeroth order contribution to the probability. Characterizing the adaptive transition To characterize the onset of adaptation, we study the extrema of the non-equilibrium potential. The condition for a point m to be an extrema is ∂mh = 0, where we have ∂mh = − e 2 e 1 + ω1/ω0 − µ e/2 − e 1 + D0ω1/ω0D1 (cid:18) e 2 (cid:19) . (16) 8 The second derivative of h characterizes whether the point is a maximum or a minimum. At equilibrium we have µ = 0, and the only extrema occurs at the value m∗ = mr + (S − S0)/e . Note that, for m∗ to be an extrema it must fall in the range [0, 1], something which we assume from now on. At m∗ we have that a(m∗) = 1/2, which corresponds to the maximal sensitivity of the activity to changes in the sig- nal. It is easy to show, however, that for µ = 0 this point is unstable, since ∂2 mh < 0. Because there is no other extrema, the memory will accumulate at the boundaries, which shows that adaptation to a high sensitivity state is not possible in equilibrium. When µ (cid:54)= 0, there can be two extrema of the potential h. At µ = 0 we still have m∗, however at µ → ∞ we have mad(µ → ∞) = mne ≡ mr + (S − Sr)/e + −1 log(D0/D1) (17) , where mne is the m value at the minimum of the nonequi- librium contribution to the effective free energy, i.e., the first term on the right hand side of Eq. (5) in the main text. At mad the activity is a(mad) = D0/(D0 + D1), which deviates from the maximum sensitivity point. Note that for mad to be a minima it must fall in the range [0, 1], which determines the range of signals to which the system can adapt. For the case in which mad ∈ [0, 1], there is a transition at intermediate values of µ from meq being an unstable point, and the probability accumulat- ing to the boundaries; to mad being a stable state, and the system being adaptive. By using a graphical con- struction of ∂mh, see Fig. 5 C, one can show that this transition is second order. We now describe the phase transition step by step. Near equilibrium and still far from the critical point the memory accumulates at the boundaries. Which bound- ary is more stable depends on the value of the signal. As µ increases the maximum (which is m∗ at equilibrium) is displaced towards m = 0, for D0/D1 > 1, or m = 1 , for D0/D1 < 1. Eventually the maximum leaves the range [0, 1], and there is just one stable boundary (the opposite to the one through which the maximum left) where the memory accumulates. Right at the critical point µc = e/2 the energy landscape has no extrema in m ∈ [−∞, +∞]. For values µ > µc a minima develops asymptotically on the side opposite to the one through which the maxima left. Eventually this minima comes in the range [0, 1] through the boundary where the mem- ory resides, and moves asymptotically to the value mad, which is reached for µ → ∞. Because the memory is con- tinuously taken from the boundary to the adapted point, this phase transition is second order. Fully irreversible limit A useful limit to study the continuous equations is µ → ∞ to all orders in δ. Note that in this limit Eq. 13 becomes simply −D1p1(m) + D0p0(m) = 0 . This equation is satisfied by the following expressions p1(m) = p0(m) = ω0(m) ω0(m) + ω1(m) ω1(m) ω0(m) + ω1(m) δ(m − mad) δ(m − mad) , which also satisfy the normalization condition. This ex- pressions indicate that in the fully irreversible limit the memory is exactly fixed to its adapted value, as the dif- fusive terms vanish. The average activity is thus clearly aad. In this limit the activity transitions are simply gov- erned by the two states master equation ∂tp1 = ω0p0 − ω1p1 . (18) Derivation of Eq. 7 To obtain an analytical expression of the adaptation error we use Eq. 13. Using this equation and defining the average activity as (cid:104)A(cid:105) =(cid:82) 1 0 p1(m)dm we get: ¯H1(cid:104)A(cid:105) + ¯H0(1 − (cid:104)A(cid:105)) = ¯D1p1(1) + ¯D0p0(1) − ( ¯D1p1(0) + ¯D0p0(0)) , which is valid to all orders. Note that when µ → ∞ the average activity converges to the adapted value aad in Eq. 11, in agreement with the solution derived above on the fully irreversible limit. The adaptation error of the output is defined as y − aad aad .  = Since (cid:104)y(cid:105) = (cid:104)A(cid:105), we can use the previous equation to show that (cid:104)(cid:105) = 1 − 0, with 1 = 0 = D1p1(1) + D0p0(1) D0(e/2 − µ) D1p1(0) + D0p0(0) D0(e/2 − µ) drop the equilibrium component of h and use a saddle point approximation with expansion parameter µ/µc (cid:29) 1 to evaluate Z: (cid:115) Z = e−h(m,S)dm ≈ (cid:90) 1 0 9 (cid:114) 2π µµc e−had , 2π mhad e−had = ∂2 (cid:114) µµc where the subindex indicates that h and its derivative are to be evaluated at mad. Inserting this in the expression for the errors we have b ≈ ∆ba(b) + ∆1−b(1 − a(b)) D0(µc − µ) 2π e−(h(b)−had) , (20) where b = 0 or b = 1 for each of the two boundary contributions to the error. While these expressions are analytical, their evaluation is not transparent due to the complicated form of the potential. A simpler expression can be obtained by evaluating h(1) and h(0) by an ex- pansion around mad. Truncating to second order we have b ≈ ∆ba(b) + ∆1−b(1 − a(b)) ∆0(µc − µ) e−(b−mad)2µµc/2 , (cid:114) µµc 2π We now define the characteristic error ∆ba(b) + ∆1−b(1 − a(b)) c = ∆0(µc − µ) (cid:114) µµc 2π , which has a weak dependence on µ. Introducing the con- stant k = (b − mad)2/2, we arrive at the following esti- mate for the adaptation error (cid:104)(cid:105) ≈ ce−kµµc , (21) large sig- where for simplicity we consider the case of nals, in which mad is also high and the error contribu- tion comes dominantly from the boundary b = 1. This expression is accurate up to order δ, and corresponds to Eq. 7 in the main text. It is important to note that the higher order terms do not imply saturation, since it was shown at the beginning of this section that when µ → ∞, then (cid:104)A(cid:105) → aad to all orders. Noise spectrum of output The output generated by the system activity is given by (19) y(t) = 1 τy (cid:90) ∞ −∞ Θ(t − t(cid:48))e−(t−t(cid:48))/τy A(t(cid:48))dt(cid:48) , Thus to obtain the error we just need to evaluate p at the boundaries. We calculate the zeroth order term, and for compact- ness note a(0)(m) as a(m), as done in the main text. Although to this order we have an exact form for the po- tential, we can not exactly integrate e−h to obtain the normalization Z. Deep into the adaptive phase, we can where Θ is the Heaviside step function. This corre- sponds to a relaxation of y towards A in a time scale τy. Using the definition of Fourier transform f (ω) = (2π)−1/2(cid:82) f (t) exp(−iωt)dt, the equation above trans- lates to y(ω) = τ−1 τ−1 y + iω y A(ω) , (22) where we have used that the time-domain convolution (f ∗ g)(t) is given by 2πf (ω)g(ω). √ To calculate the power spectrum of the output, we first note that (cid:104)A(ω)A(ω(cid:48))(cid:105) = √ 2πδ(ω + ω(cid:48))SA(ω) , where SA is the power spectrum of the activity, which can be calculated by standard methods [13, 15]. From this and Eq. 22 we obtain Sy(ω) = SA(ω)τ−2 y τ−2 y + ω2 , which we use to calculate the power-spectrum throughout this work. (cid:104)y2(cid:105) = The power spectrum is related to the autocorrelation function through a Fourier transform, that is e−iωtCy(t)dω ; Sy = (cid:90) ∞ −∞ 1√ 2π (cid:90) ∞ −∞ where the correlation function is defined as Cy(t) = (cid:104)y(t)y(0)(cid:105). The second moment relates to the correla- tion function as (cid:104)y2(cid:105) = Cy(0) = 1√ 2π Sy(ω)dω . Together with (cid:104)y(cid:105) = (cid:104)A(cid:105), this equation allows to calculate the variance of the output. For the particular case of µ → ∞, the memory state is fixed to mad. The activity however still fluctuates between one active and one inactive state according to Eq. 18. In this case, the power spectrum of the output takes a particularly simple form. To calculate it, we use that the correlation of the activity is generically given by (cid:104)A(t)A(0)(cid:105) = Aiqλ,iAjzλ,jpss j eλt where Ai = {1, 0} is the activity observable, pss j the sta- tionary probability distribution, λ the eigenvalues of the W−matrix, and qλ and zλ the corresponding right and left eigenvectors normalized by the condition zλ· qλ(cid:48) = δλλ(cid:48). For the particular case of the activity we thus have [13, 15] (cid:104)A(t)A(0)(cid:105) = qλ,1zλ,1ps 1eλt . (cid:88) (cid:88) λ i,j (cid:88) λ From Eq. 18 we have that the eigenvalues are 0 and −(ω0 + ω1), the corresponding right eigenvectors {ω0, ω1}/(ω1 + ω0) and {−1, 1}, and the left eigenvec- tors {1, 1} and {−ω1, ω0}/(ω1 + ω0). The component 1 of the steady state probability is simply the average ac- 1 = (cid:104)A(cid:105) = ω0/(ω0 +ω1). The correlation function tivity, ps is thus given by (cid:104)A(t)A(0)(cid:105) = (cid:104)A(cid:105)(cid:16)(cid:104)A(cid:105) + (1 − (cid:104)A(cid:105))e−(ω0+ω1)t(cid:17) 10 from which one can verify that (cid:104)A2(0)(cid:105) = (cid:104)A(cid:105) and (cid:104)A(∞)A(0)(cid:105) = (cid:104)A(cid:105)2. The power spectrum of the activity is then SA(ω) = √ 2π(cid:104)A(cid:105)2δ(ω) + (cid:104)A(cid:105)(1 − (cid:104)A(cid:105)) (cid:114) 2 ω0 + ω1 π (ω0 + ω1)2 + ω2 . Using this expression, the relationship between the power spectrum of y and A, and defining the activation time as τa = 1/(ω0 + ω1), we obtain the second moment of the output to be (cid:18) 1√ 2π = (cid:104)A(cid:105)2 + √ (cid:104)A(cid:105)2 τa a − τ−1 τ−1 a − τ−2 τ−2 y y 1 τy 2π + (cid:104)A(cid:105)(1 − (cid:104)A(cid:105)) τa + τy (cid:114) 2 (cid:19) π (cid:104)A(cid:105)(1 − (cid:104)A(cid:105))π From this, we obtain that the variance of the output in the fully irreversible regime is given by ym = (cid:104)A(cid:105)(1 − (cid:104)A(cid:105)) σ2 τa τa + τy , (23) which contains the intrinsic activity fluctuations, aver- aged by the CheY-P dynamics [12]. This defines the saturation error noise, which can be averaged out by re- ducing the ratio τa/τy. Derivation of Eq. 9 ym As described in the main text, there are two contribu- tions to the fluctuations of the output. The first comes from the fluctuation of y around a(m), we note it σ2 and was calculated before in the case µ → ∞. The second comes from the fluctuations of m itself, which make a(m) fluctuate, and we note it σ2 a. When there is a separation of time scales τa (cid:28) τy (cid:28) τad, we have that the total out- put variance is the sum of these two contributions. Thus, using σ2 ym)/aad. For finite values of µ the memory fluctuations dominate and a is the most relevant contribution, however as µ → ∞ σ2 the memory gets frozen and σ2 ym dominates. It is thus crucial to calculate the dependence of σ2 a in µ. The main steps on how to calculate σ2 a are given in the main text.  , we have that σ2  = aadσ2  = (σ2 a +σ2 DISSIPATED WORK To quantify how far from equilibrium the system is, we use the dissipated work W . At the steady state, and having set the thermal unit to one, we have (cid:90) 1 (cid:20) J 2 0 0 D0p0 , W = + J 2 1 D1p1 + (p0ω0 − p1ω1) log (cid:18) p0ω0 (cid:19)(cid:21) p1ω1 dm . We can estimate this quantity in the highly irreversible limit µ (cid:29) µc to zeroth order in δ. In that regime, we have (cid:21) + (µD1p1)2 D1p1 dm (cid:20) (µD0p0)2 (cid:90) 1 (cid:90) 1 D0p0 0 W ≈ = µ2 µ τad = [D0p0 + D1p1] dm 0 ( ¯D0(1 − aad) + ¯D1aad) . For the case considered in the main text in which D0 = D1, we have that W ≈ µ ¯D0/τad. 11 PARAMETERS Unless otherwise specified, the parameters used in the continuum limit are: e = 8, mr = 1/2, d = 1/e, Sr = 0, S = 2.5, D0 = D1, τa = 10−3s, τy = 10−1s, τad = 10s. The parameters for the discrete model are used taking these as references and using the appropriate N rescal- ing. For N = 5 as in E. Coli chemotaxis the resulting parameters are those in [17].
1705.08107
1
1705
2017-05-23T07:38:08
Investigating the inner structure of focal adhesions with single-molecule localization microscopy
[ "physics.bio-ph" ]
Cells rely on focal adhesions (FAs) to carry out a variety of important tasks, including motion, environmental sensing, and adhesion to the extracellular matrix. Although attaining a fundamental characterization of FAs is a compelling goal, their extensive complexity and small size, which can be below the diffraction limit, have hindered a full understanding. In this study we have used single-molecule localization microscopy (SMLM) to investigate integrin $\beta$3 and paxillin in rat embryonic fibroblasts growing on two different extracellular matrix-representing substrates (i.e. fibronectin-coated substrates and specifically bio-functionalized nano-patterned substrates). To quantify the substructure of FAs, we developed a method based on expectation maximization of a Gaussian mixture that accounts for localization uncertainty and background. Analysis of our SMLM data indicates that the structures within FAs, characterized as a Gaussian mixture, typically have areas between 0.01 and 1 $\mu$m$^2$, contain 10 to 100 localizations, and can exhibit substantial eccentricity. Our approach based on SMLM opens new avenues for studying structural and functional biology of molecular assemblies that display substantial varieties in size, shape, and density.
physics.bio-ph
physics
Investigating the inner structure of focal adhesions with single-molecule localization microscopy H. Deschout1, I. Platzman2, D. Sage3, L. Feletti1, J. P. Spatz2 and A. Radenovic1 1Laboratory of Nanoscale Biology, Institute of Bioengineering, School of Engineering, EPFL, Lausanne, Switzerland 2Max-Planck-Institute for Medical Research, Dept. of Cellular Biophysics, University of Heidelberg, Dept. of Biophysical Chemistry, Heidelberg, Germany 3Biomedical Imaging Group, School of Engineering, EPFL, Lausanne, Switzerland Abstract Cells rely on focal adhesions (FAs) to carry out a variety of important tasks, including motion, environmental sensing, and adhesion to the extracellular matrix. Although attaining a fundamental characterization of FAs is a compelling goal, their extensive complexity and small size, which can be below the diffraction limit, have hindered a full understanding. In this study we have used single- molecule localization microscopy (SMLM) to investigate integrin β3 and paxillin in rat embryonic fibroblasts growing on two different extracellular matrix-representing substrates (i.e. fibronectin- coated substrates and specifically bio-functionalized nano-patterned substrates). To quantify the substructure of FAs, we developed a method based on expectation maximization of a Gaussian mixture that accounts for localization uncertainty and background. Analysis of our SMLM data indicates that the structures within FAs, characterized as a Gaussian mixture, typically have areas between 0.01 and 1 µm2, contain 10 to 100 localizations, and can exhibit substantial eccentricity. Our approach based on SMLM opens new avenues for studying structural and functional biology of molecular assemblies that display substantial varieties in size, shape, and density. 1 1. Introduction Focal adhesions (FAs) are subcellular macromolecular assemblies consisting of dynamic protein complexes that are localized near the cell membrane. FAs affect nearly all aspects of a cell's life, including, but not limited to, adhesion, directional migration, cell proliferation, differentiation, survival, and gene expression (1). Despite having been studied for several decades, the inner architecture of FAs is still not completely understood. In part, this is due to the limitations of conventional fluorescence microscopy for FA analysis. FAs are molecularly diverse structures, containing a large number of proteins (2). Therefore, their investigation requires imaging techniques that offer sufficient multiplexing capabilities (3). Moreover, FAs have a size that is typically in the order of microns or less, and therefore their internal spatio-temporal organization is not fully resolvable with conventional microscopy. During the last decade, several super-resolution microscopy techniques have been employed to image FAs (4-9). An important insight from these studies was that FAs are not homogeneous spatial structures. Initially, photo-activated localization microscopy (PALM) was used to reveal that FAs can consist of patches of proteins with submicron dimensions (4, 9). Later on, Bayesian localization microscopy and structured illumination microscopy showed that many FAs exhibit discontinuous elongated (or fiber-like) substructures (5, 6). Moreover, single-particle tracking demonstrated that proteins can diffuse within FAs (7, 8), which again suggests that they have an internal spatial organization. However, these have all been qualitative observations, and a quantitative analysis of the FA substructure is still lacking. For quantitative analysis of the internal spatial organization of FAs, single-molecule localization microscopy (SMLM) can potentially be implemented (10). SMLM data consists of the localizations of individual photo-activatable or photo-switchable fluorescent molecules. Therefore, a variety of methods have been developed to identify and characterize clusters of such localizations (11, 12). These methods are often applied to investigate clusters of receptors in the cell membrane. Such clusters are usually radially symmetric, spatially well separated, and homogeneous in size and density. FA substructures, on the other hand, cannot be characterized as simply. Indeed, adhesions structures can vary from sub-diffraction entities composed of a couple of different proteins (e.g. focal complexes or nascent adhesions) to assemblies of many proteins measuring several microns (e.g. FAs) (13). Moreover, FA subunits are densely packed, since they cannot be resolved using a conventional microscope. Finally, FAs usually have an elongated shape, and the same is possibly true for their subcomponents. Therefore, it is not clear whether established SMLM clustering methods are suitable for the identification of FA substructures. In this study we have designed a novel approach to investigate the FA substructure. We used expectation maximization of a Gaussian mixture (EMGM) (14) to interpret SMLM data in terms of spatial probability distributions. EMGM allows to quantify the properties of closely packed localization patterns that exhibit substantial varieties in size, density, and shape, and is therefore well suited for studying the inner architecture of FAs. Importantly, we improved the classical EMGM framework to account for localization uncertainties and the presence of a localization background, both being ubiquitous in SMLM data. 2 The other goal of this study was to quantify the properties of the subunits of which FAs are composed. For this purpose we used PALM, an implementation of SMLM that is popular for imaging FAs (4, 9, 15- 17), since it makes use of photo-activatable fluorescent proteins that can be genetically expressed. More in particular, we used PALM to image integrin β3 and paxillin in fixed rat embryonic fibroblasts (REFs), a well-known cell line for FA investigation. Cell experiments were performed using fibronectin- coated substrates and specifically bio-functionalized nano-patterned substrates, on which ordered patterns of nanoscale adhesive spots were provided (18, 19). Such nano-patterned substrates have already been used to indirectly probe the behavior of FAs on the nanoscale (20). In this way, the spatial organization of integrin binding sites is precisely controlled, ensuring that the observed substructures are innate to FAs. Application of our improved version of EMGM on the PALM data allowed us to determine that FAs are composed of structures with areas between 0.01 and 1 µm2, containing 10 to 100 localizations, and exhibiting substantial eccentricities down to 0.1. 3 2. Materials and Methods 2.1 Microscope PALM imaging was carried out on a custom built microscope (21, 22). A 50 mW 405 nm laser (Cube, Coherent), a 100 mW 488 nm laser (Sapphire, Coherent), and a 100 mW 561 nm laser (Excelsior, Spectra Physics) were used for excitation/activation. The three lasers were focused into the back focal plane of the objective mounted on an inverted optical microscope (IX71, Olympus). We used a 100× objective (UApo N 100×, Olympus) with a numerical aperture of 1.49 configured for TIRF. A dichroic mirror (493/574 nm BrightLine, Semrock) and an emission filter (405/488/568 nm StopLine, Semrock) were used to separate fluorescence and illumination light. The fluorescence light was detected by an EMCCD camera (iXon DU-897, Andor). An adaptive optics system (Micao 3D-SR, Imagine Optics) and an optical system (DV2, Photometrics) equipped with a dichroic mirror (T565lpxr, Chroma) were placed in front of the EMCCD camera. 2.2 Imaging procedure Cells were imaged in PBS at room temperature. Prior to imaging, 100 nm gold fiducial markers (C-AU- 0.100, Corpuscular) were added to the sample for lateral drift monitoring. Axial drift correction was ensured by a nanometer positioning stage (Nano-Drive, Mad City Labs) driven by an optical feedback system (21). Excitation of the mEos2 done at 488 or 561 nm with ~10 mW power (as measured in the back focal plane of the objective). The mEos2 was activated at 405 nm with ~2 mW power. The gain of the EMCCD camera was set at 100 and the exposure time to 50 ms. For each experiment 10,000 camera frames were recorded. 2.3 Substrate preparation Quasi-hexagonal patterns of AuNPs were prepared on 25 mm diameter microscope cover slips (#1.5 Micro Coverglass, Electron Microscopy Sciences) by means of BCML as previously described (18, 19, 23) (Supporting Material). Fibronectin-coated cover slips were prepared by first cleaning with an oxygen plasma and then incubating with PBS containing 50 µg/ml fibronectin (Bovine Plasma Fibronectin, Invitrogen) for 30 minutes at 37 °C. To remove the excess of fibronectin, the cover slip was washed with PBS before seeding the cells. 2.4 Cell culture and fixation The REFs (CRL-1213, ATCC) were grown in DMEM supplemented with 10% fetal bovine serum, 1% penicillin-streptomycin, 1% non-essential amino acids and 1% glutamine, at 37 °C with 5% CO2. The cells were transfected by electroporation (Neon Transfection System, Invitrogen), which was performed on ~106 cells using 1 pulse of 1350 V lasting for 35 ms. The amount of DNA used for the transfection was 4 µg for both the mEos2-paxillin-22 vector and the mEos2-Integrin-β3-N-18 vector. Around 2.105 transfected cells were seeded on individual cover slips and grown in cell culture medium without penicillin-streptomycin, at 37 °C with 5% CO2. The cells were washed with PBS around 20h after transfection (Fig. S1 in the Supporting Material), and then incubated in PBS with 2.5% 4 paraformaldehyde at 37 °C for 10 minutes. After removing the fixative, the cells were again washed with PBS, and the cover slip was placed into a custom made holder. 2.5 PALM data analysis The recorded images were analyzed by a custom written algorithm (Matlab, The Mathworks) that was adapted from a previously published algorithm (4, 22). First, peaks were identified in each camera frame by filtering and applying an intensity threshold. Only peaks with an intensity at least 4 times the background were considered to be emitters. Subsequently, the peaks were fitted by maximum likelihood estimation of a 2D Gaussian distribution (24). Drift was corrected in each frame by subtracting the average position of the fiducial markers from the positions of the emitters in that frame. The localization uncertainty for each emitter was obtained from the Cramér-Rao lower bound of the maximum likelihood procedure (25). PALM images were generated by plotting a 2D Gaussian centered on each fitted position with a standard deviation equal to the corresponding localization uncertainty. Only positions with a localization uncertainty between 0 and 40 nm were used. 2.6 EMGM procedure The EMGM procedure (Supporting Material) was implemented in Matlab (The Mathworks). The initial values of the parameters that describe a mixture consisting of K components were estimated by deleting a component from the previously estimated mixture consisting of K-1 components and adding two new components that are generated from the deleted one (Supporting Material). Additionally, one new Gaussian component is generated from the background component of the previously estimated mixture. This is done 3 times for each of the original K-1 Gaussian components and the background component, resulting in a total of 3K initializations. In the case of K=1, the initialization is done randomly 3 times. The procedure is stopped when the null hypothesis that the previously estimated K-1 component mixture is the correct one is fulfilled (Supporting Material). For this purpose, we simulated the distribution of likelihood increments when comparing the K-1 and K component models under the null hypothesis. This distribution is obtained by simulating 100 datasets assuming the K-1 solution, and applying EMGM on each dataset, for both K-1 and K mixture components. If the real likelihood increment has a p-value below 0.01 under the null hypothesis, it is assumed that the K- 1 component solution is the correct one. Prior to analysis, the PALM data was split into overlapping 2×2 µm areas, and the EMGM analysis was performed on each area separately (Fig. S2 in the Supporting Material). Afterwards, identical mixture components in different EMGM results were combined according to a criterion based on the correlation between their posterior probabilities (Supporting Material). 5 3. Results 3.1 Expectation Maximization of a Gaussian Mixture (EMGM) FAs display a substantial variety in size, shape, and density, and their substructure potentially as well. Quantifying the properties of the FA substructure with SMLM clustering methods is therefore challenging. Clusters in SMLM data are often characterized using the pair correlation function (26) or Ripley's K(r) or L(r) function (27). These functions describe the density around a certain point as a function of the distance r from that point. As an illustration, we used PALM to image integrin β3 in a REF cell (Fig. 1A). We used Ripley's L(r)-r function (28) to analyze a subset of the data (Fig. 1, B-D). This function shows a peak around 0.2 µm, indicating that the degree of clustering is highest on this length scale. However, it is difficult to interpret this result in terms of FA substructure properties, especially considering the heterogeneity in size and shape of the FAs themselves. Such difficulties can be avoided by clustering methods that identify individual clusters based on criteria related to the local density of localizations, such as the nearest neighbor method (29) or density-based spatial clustering of applications with noise (DBSCAN) (30). We applied DBSCAN (31) to the same subset of the PALM data mentioned above (Fig. 1E). One value for the DBSCAN search radius identifies several substructures in the FA, while a larger value does not. However, the large search radius identifies two clusters that were considered to be background by the small search radius. It is clear that DBSCAN can handle the heterogeneity in size and shape of FAs, but identification of FA substructures largely depends on the values used for parameters that are related to a localization density threshold. Such a threshold is challenging to define, since FA substructures exhibit a variety of localization densities and can be closely packed (Fig. 1, A and B). The difficulties related to established SMLM clustering methods prompted us to develop an approach based on EMGM (14). The main assumption of EMGM is that FAs can be modeled by a mixture of bivariate Gaussian probability distributions (Supporting Material). After choosing initial values for the parameters of each Gaussian component, the posterior probability that a certain localization was generated from a certain Gaussian component is evaluated (i.e. the expectation step). The Gaussian component parameters are then re-estimated using the new posterior probabilities (i.e. the maximization step) and the likelihood of the updated Gaussian mixture is calculated and checked for convergence. In order to apply EMGM on SMLM data, we used a "greedy learning" approach (32) to initialize the parameters of the Gaussian components, and a model selection procedure based on hypothesis testing (33) to determine the number of components in the mixture (Supporting Material). However, the specific nature of SMLM data poses some additional challenges for EMGM. One problem is that not necessarily all localizations are part of the structure of interest, but can instead belong to a background. In case of a simple uniform background, the EMGM algorithm can readily be adjusted (Supporting Material). Moreover, the localizations in SMLM data contain measurement uncertainties (34). This localization uncertainty can be described by a spatial probability distribution that is usually modeled as a Gaussian. The EMGM can therefore be adapted by convolving the probability distributions that describe the mixture and the localization uncertainties (Supporting Material). 6 3.2 Evaluation of EMGM on simulations The performance of the EMGM algorithm adapted for SMLM data was evaluated and validated by applying it to simulated data. We simulated mixtures consisting of K closely spaced Gaussian components described by identical spatial probability distributions (i.e. 2D symmetric Gaussians with standard deviation σx = σy = 20 nm) and containing an identical number of positions (i.e. 100) (Fig. 2A, and Supporting Material). Such components have similar characteristics as nascent adhesions, or more speculatively, as the substructure of larger FAs. First, we verified the performance of our proposed initialization scheme and model selection procedure. The results show that the simulated mixtures are correctly identified, provided K is smaller than 10 (Fig. 2B, and Fig. S3 in the Supporting Material). We used 3K initializations for a mixture with K components (Supporting Material). Increasing the number of initializations does not substantially improve the EMGM performance (Fig. S4 in the Supporting Material). Next, we simulated the effect of a uniform localization background density bg and a localization uncertainty s. The results indicate that the adapted EMGM correctly predicts σx,y for values of bg up to 25,000 #/µm2 (Fig. 2C, and Fig. S5 in the Supporting Material). Our EMGM approach also captures the effect of the apparent increase in σx,y due to localization uncertainties for values of s up to 30 nm (Fig. 2D, Fig. S6 in the Supporting Material). Note that the largest values of s and bg included in these simulations are typically not encountered in "good quality" SMLM data. The adapted EMGM should be able to distinguish closely spaced substructures inside FAs. Towards this end, we simulated Gaussian mixtures with a decreasing spacing dx,y between the component centers (Fig. 2E, and Supporting Material). The adapted EMGM performs well when dx,y is larger than 70 nm. A smaller dx,y results in a significant overlap in the spatial probability distribution of two adjacent components. Since one cannot assume that the substructures of FAs are radially symmetric, the component shape should also be accounted for by the EMGM algorithm. We simulated mixture components with decreasing σx and simultaneously increasing σy (Supporting Material). The results (Fig. 2F) clearly show that the algorithm correctly predicts the changing eccentricity σx/σy. It should be noted that the results (Fig. 2) depend on the number of localizations that are contained by the components (Fig. S7 in the Supporting Material). 3.3 Application of EMGM on experimental data To demonstrate the application of our EMGM algorithm, we made use of the SMLM data of a REF cell expressing mEos2-labelled integrin β3 (Fig. 1B). Similar to DBSCAN applied with the small search radius (Fig. 1E), EMGM also finds several FA substructures. Moreover, EMGM identifies two structures on the right as well, as indicated by the DBSCAN result using the large search radius (Fig. 1E). We next proceeded to apply the EMGM algorithm on the whole PALM dataset (Fig. 1A). Since the simulation results (Fig. 2B) indicate that our algorithm works best for a small number of components, we reduce their number by applying a scanning procedure, consisting of splitting the original field of view into smaller overlapping areas, and by subsequently applying EMGM to each of these areas (Fig. S2). Afterwards, the results are combined, by merging identical Gaussian components in overlapping regions based on the correlation between their posterior probabilities, while excluding Gaussian 7 components that belong to structures that were clipped during the splitting procedure (Supporting Material). EMGM characterizes FA substructures in terms of bivariate Gaussian probability distributions. The properties of such a distribution can be translated into more intuitive properties using the error ellipse, i.e. the line of that describes a constant probability density. The major axis a and the minor axis b of the ellipse define its area and its shape (Fig. S8 in the Supporting Material). We therefore describe the FA substructure shape by the eccentricity b/a (similar to the definition above). To calculate the area, we choose the 2σ error ellipse, corresponding to twice the standard deviation of the Gaussian component. This error ellipse defines the area in which there is a probability to find ~95% of all localizations belonging to the component. We pooled the area and eccentricity values of all identified mixture components in our PALM data set (Fig. 1G). Most components have an area below 0.5 µm2 with a peak around 0.1 µm2, and many exhibit some degree of eccentricity, with most values lower than 0.8. The EMGM algorithm also returns the posterior probability of each localization belonging to a specific Gaussian distribution, which gives the total number of localizations of each FA substructure (Supporting Material). Making the simplifying assumption that the localizations are uniformly distributed within the 2σ error ellipse, this leads to a characteristic localization density. Most FA substructures have a localization density below 2000 #/µm2, and contain less than 100 localizations (Fig. 1G). 3.4 Integrin and paxillin Following the evaluation of the adapted EMGM, we applied our method to investigate the substructure of FAs in cells growing on often-used fibronectin-coated substrates. We used PALM to image fixed REF cells (n = 10) expressing paxillin or integrin β3 labelled with mEos2 (Fig. 3, A and B). To identify the FA substructure, we applied the adapted EMGM to each of these PALM datasets (Fig. 3C). As discussed above, the properties of individual mixture components, defined as bivariate Gaussians, can be described by three parameters: eccentricity, area, and number of localizations. We plotted these quantities as a function of each other, for both paxillin and integrin β3 (Fig. 3, D-F). Most mixture components contain between 10 and 100 localizations, and have an area between 0.01 and 1 µm2 (Fig. 3D). The paxillin case displays a slightly more pronounced tail towards components that contain more localizations (up to 1000 localizations). When plotting the eccentricity as a function of the number of localizations (Fig. 3E), it is again apparent that the paxillin FA substructures can contain more localizations than the integrin ones. Furthermore, the mixture components in both cases appear to be eccentric, with most values below 0.7. The FA substructures containing less localizations appear to be somewhat more eccentric, a tendency that is more apparent in the paxillin case. A similar observation can be made when plotting the eccentricity as a function of the area (Fig. 3F). The larger the FA substructure, the more eccentric it seems to be. Interestingly, both paxillin and integrin objects seem to have similar areas, with a pronounced peak around 0.1 µm2. 3.5 Nano-patterned substrates The FA substructure properties (Fig. 3) have been obtained from REF cells growing on fibronectin- coated substrates. It can therefore not be guaranteed that the observed FA substructure is innate, it might simply be reflecting how the integrin binding sites on the fibronectin-coated substrate are 8 organized on the nanoscale level. Such difficulties in interpretation of the data can be avoided by making use of a substrate where the integrin binding site locations are precisely controlled. We have therefore made use of block-copolymer micelle nanolithography (BCML) to pattern substrates with a quasi-hexagonal grid of 8 nm diameter gold nanoparticles (AuNPs) (18, 19) (Supporting Material). The AuNPs are functionalized with cyclic arginyl-glycyl-aspartic acid (cRGD) peptides, using a flexible polyethylene glycol (PEG) spacer. The area between the AuNPs is passivated with a PEG layer, ensuring that integrins can only adhere to the AuNPs. This enables a more unambiguous interpretation of the observed FA substructure. We chose a 56 nm spacing between the AuNPs, which was shown to result in normal cell adhesion (18). Furthermore, we also tested a 119 nm spacing, which poses more challenges for adhering cells (19). We again imaged fixed REFs (n = 10) expressing integrin β3 labelled with mEos2 (Fig. 4, A and B). The signal-to-noise ratio in the total internal reflection fluorescence (TIRF) images (Fig. 4A) appears to be lower than in the case of the fibronectin-coated substrates (Fig. 3A), suggesting that the grid of AuNPs affects the quality of the TIRF illumination. Next, we applied the adapted EMGM to each of the PALM datasets, in order to investigate the FA substructure (Fig. 4C). We plotted the number of localizations as a function of the area, for both the 56 and 119 nm AuNP spacings (Fig. 4, E and F). The fibronectin case (Fig. 4D) was added for comparison. It is clear that the objects on the fibronectin-coated substrate can contain up to 100 localizations, while the localization numbers on the 56 nm spacing substrate are generally below that level (Fig. 4, D and E). Interestingly, the FA substructure areas are very similar between both types of substrates, mostly between 0.01 and 1 µm2 (Fig. 4, E and F). The FA substructure observed on the nano-patterned substrates does not appear in contradiction with the results obtained from fibronectin-coated substrates. 3.6 Isolated and overlapping mixture components The interpretation of the EMGM results can be complicated (Fig. 3C, and Fig. 4C). Especially inside dense and large structures, which visually appear to be FAs, one can observe several components that overlap, based on their 2σ error ellipses. The isolated mixture components, on the other hand, seem to correspond with smaller structures that could be nascent adhesions or focal complexes. We, therefore, performed a post-analysis step on EMGM results (Fig. 5A, and Supporting Material). We split the mixture components into two categories: the ones whose 1σ error ellipse overlaps with at least one other 1σ error ellipse, called the "overlapping" components, and the ones whose 1σ error ellipse does not overlap with another one, called the "isolated" components. A new object can be calculated from a set of overlapping components, giving rise to a third category, called the "merged" components (Fig. 5A, and Supporting Material). Application of this merging procedure on a previously obtained EMGM result (Fig. 3C) shows that there are indeed several components that overlap (Fig. 5, B and C). We applied the merging procedure on the EMGM results of REFs (n = 10) expressing integrin β3 labelled with mEos2, growing on fibronectin-coated (Fig. 3D) and 56 nm spacing nano-patterned (Fig. 4D) substrates. As expected, on both types of substrate, the merged objects tend to have a larger area (up to 1 µm2) and contain more localizations (up to 1000 localizations) than the isolated and overlapping objects (Fig. 5D). The isolated components exhibit a similar behavior on both substrate types (Fig. 5E). Both cases exhibit FA substructures with an area between 0.01 and 0.1 µm2, containing less than 100 localizations. The overlapping components are also not showing much difference 9 between both substrate types, although the ones on the fibronectin-coated substrate can contain more localizations (Fig. 5F). Interestingly, the isolated and overlapping objects on the nano-patterned substrate also behave quite similarly (Fig. 5, E and F). The overlapping FA substructures are therefore not necessarily artifacts found by EMGM in a dense localization environment. 10 4. Discussion We propose a new way to explore the properties of unknown structures as observed by SMLM. Using EMGM, we interpret patterns in SMLM data as a mixture of bivariate Gaussians. This approach allows to describe densely packed structures that can display strong heterogeneities in size, shape, and density, and is therefore well suited for investigation of the substructure of FAs. However, application of EMGM to SMLM data is not without challenges. The result can be influenced by the choice of the initial values for the mixture component properties, and the number of components needs to be chosen as well. We identified an initialization procedure and a selection criterion for the number of components that gives good results for mixtures consisting of a small number of components (e.g. less than 10 for our simulated data). To allow analysis of larger numbers of components, we used a scanning procedure that essentially consists of splitting the SMLM data into smaller overlapping areas, and performing EMGM on each area separately. It is important to note that, unlike some SMLM clustering methods, the EMGM approach essentially does not depend on the choice of a free parameter (except for the area size of the scanning procedure). The properties of SMLM data pose challenges to the classic EMGM algorithm. One complication is the localization uncertainty, which leads to an overestimation of the standard deviation of the Gaussian mixture components. An important contribution of this work is that we improved the EMGM approach to account for this effect. For "reasonable" localization uncertainties (e.g. below 30 nm for our simulated data) we found that the adapted EMGM worked well. We would like to point out that the effect of localization uncertainties is ignored by most existing SMLM clustering methods. Besides localization uncertainty, we also adjusted the EMGM algorithm to account for the presence of a uniform localization background. The method was found to perform excellently for any realistic level of background (e.g. up to 25,000 #/µm2 for our simulated data). To investigate the inner architecture of FAs, we performed SMLM imaging of FAs in fixed REF cells. We first explored the use of points accumulation in nanoscale topography (PAINT) (35) for imaging integrin β3 (Supporting Material). Our PAINT data suggests that that not all integrins are accessible for antibodies (Fig. S9 in the Supporting Material). To avoid antibody labeling problems, we therefore opted for PALM. We imaged integrin β3 and paxillin in fixed REF cells on fibronectin-coated substrates. The EMGM algorithm allowed us to identify integrin β3 objects with a typical area in the range between 0.01 and 1 µm2, and containing between 10 and 100 localizations. Paxillin objects were found to have a similar area, but can contain more localizations, up to 1000. We attribute this difference to a tree-like organization of the FAs, rooting from isolated integrin islands, and expanding towards the actin filaments due to crosslinking and multivalent binding of paxillin and other proteins to their recruiting components. The equivalent diameter of the smallest objects was found to be around 100 nm (using the 2σ error ellipse area, which is 0.01 µm2 for the smaller objects). This indeed justifies the need for super-resolution microscopy to investigate the inner structure of FAs. Most objects were found to exhibit substantial eccentricity, with values down to 0.1. An algorithm that does not assume radial symmetry, such as EMGM, is therefore essential for the analysis of FA substructure. A fibronectin coating is often used to ensure good cell adhesion to the substrate. However, it is important to rule out that the observed FA substructure is a mere artifact of the binding sites presented by such fibronectin-coated substrates. We therefore repeated the experiments on 11 substrates that were patterned with a quasi-hexagonal grid of functionalized AuNPs. Our EMGM algorithm identified integrin β3 objects with areas in the same range as on fibronectin-coated substrates, while the number of localizations was lower, typically not exceeding 100. The FA substructure observed on the nano-patterned and fibronectin-coated substrates do not contradict each other. The EMGM results sometimes display strongly overlapping mixture components, which is mathematically perfectly possible, but difficult to interpret. One possibility is that the background within the FAs is more complex than a simple uniform distribution. This could lead to the background partially being characterized by some of the mixture components, while the others are actual FA substructures. Note that our scanning procedure already captures background heterogeneities on the scale of the scanned areas. Another possibility is that a bivariate Gaussian is not the most accurate model for the FA subunits. To a certain extent, a post-analysis step can provide more insight. We performed a merging procedure that describes FA substructures either as an isolated Gaussian component, or a combination of several overlapping components. We hypothesize that the isolated components (areas between 0.001 and 0.01 µm2, and number of localizations between 10 and 100) correspond to focal complexes or nascent adhesions. The overlapping mixture components, which appear to belong to FAs, have areas and localization numbers in the same range as the isolated components. This suggests that the observed objects are indicative of the real FA substructure. The merged components have a maximal area around 1 µm2 and contain up to 1000 localizations, which can be interpreted as an upper limit for the FA substructure. We envisage several ways in which our EMGM approach could be extended or adapted to allow a systematic and detailed study of the inner architecture of FAs. Several FA proteins could be investigated in multi-color mode to assess their spatial relationship. In this context, it could be of interest to develop an extension of EMGM that allows to investigate the "co-localization" of the mixture components. It would also be interesting to develop a 3D implementation of EMGM for the investigation of FA substructure in both the lateral and axial direction, as observed for instance by iPALM (17). It seems worthwhile to explore the possibility of incorporating other models than the Gaussian bivariate distribution, and other types of background besides the uniform one. Note that the effect of repetitive localizations on EMGM should be investigated, since photo-activatable fluorescent proteins can be localized more than once due to a phenomenon called "photoblinking" (36). Using transient transfection, a population of endogenous proteins will not fluorescently labeled, and the labelled proteins might be overexpressed. Techniques such as CRISPR/cas9 can bring solutions to this problem (37). 12 5. Conclusion We have used PALM to investigate FAs in REF cells growing on fibronectin-coated substrates and specifically bio-functionalized nano-patterned substrates, on which ordered patterns of nanoscale adhesive spots were provided. To quantify the FA subunit properties, we developed a method based on EMGM that accounts for localization uncertainty and background. Analysis of our PALM data indicates that integrin β3 and paxillin structures within FAs have areas between 0.01 and 1 µm2, contain 10 to 100 localizations, and can exhibit substantial eccentricities down to 0.1. We believe that our EMGM based approach is generic enough for the investigation of various other SMLM imaged nanoscale structures as well, especially for closely packed protein structures, or objects that display strong radial asymmetries and differences in size and density. 13 Acknowledgements The mEos2-paxillin-22 vector and the mEos2-Integrin-β3-N-18 vectors were kindly provided by Dr. Michael Davidson. H.D., J.P.S., and A.R. acknowledge the support of the Max Planck-EPFL Center for Molecular Nanoscience and Technology. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. Zamir, E., and B. Geiger. 2001. Molecular complexity and dynamics of cell-matrix adhesions. J Cell Sci 114:3583-3590. Zaidel-Bar, R., S. Itzkovitz, A. Ma'ayan, R. Iyengar, and B. Geiger. 2007. Functional atlas of the integrin adhesome. Nat Cell Biol 9:858-868. Harizanova, J., Y. Fermin, R. S. Malik-Sheriff, J. Wieczorek, K. Ickstadt, H. E. Grecco, and E. Zamir. 2016. Highly Multiplexed Imaging Uncovers Changes in Compositional Noise within Assembling Focal Adhesions. Plos One 11:e0160591. Betzig, E., G. H. Patterson, R. Sougrat, O. W. Lindwasser, S. Olenych, J. S. Bonifacino, M. W. Davidson, J. Lippincott-Schwartz, and H. F. Hess. 2006. Imaging intracellular fluorescent proteins at nanometer resolution. Science 313:1642-1645. Hu, S. Q., Y. H. Tee, A. Kabla, R. Zaidel-Bar, A. Bershadsky, and P. Hersen. 2015. Structured illumination microscopy reveals focal adhesions are composed of linear subunits. Cytoskeleton 72:235-245. Morimatsu, M., A. H. Mekhdjian, A. C. Chang, S. J. Tan, and A. R. Dunn. 2015. Visualizing the Interior Architecture of Focal Adhesions with High-Resolution Traction Maps. Nano Lett 15:2220-2228. Rossier, O., V. Octeau, J. B. Sibarita, C. Leduc, B. Tessier, D. Nair, V. Gatterdam, O. Destaing, C. Albiges-Rizo, R. Tampe, L. Cognet, D. Choquet, B. Lounis, and G. Giannone. 2012. Integrins beta(1) and beta(3) exhibit distinct dynamic nanoscale organizations inside focal adhesions. Nat Cell Biol 14:1057-1067. Shibata, A. C. E., T. K. Fujiwara, L. M. Chen, K. G. N. Suzuki, Y. Ishikawa, Y. L. Nemoto, Y. Miwa, Z. Kalay, R. Chadda, K. Naruse, and A. Kusumi. 2012. Archipelago architecture of the focal adhesion: Membrane molecules freely enter and exit from the focal adhesion zone. Cytoskeleton 69:380-392. Shroff, H., C. G. Galbraith, J. A. Galbraith, H. White, J. Gillette, S. Olenych, M. W. Davidson, and E. Betzig. 2007. Dual-color superresolution imaging of genetically expressed probes within individual adhesion complexes. P Natl Acad Sci USA 104:20308-20313. Tabarin, T., S. V. Pageon, C. T. T. Bach, Y. Lu, G. M. O'Neill, J. J. Gooding, and K. Gaus. 2014. Insights into Adhesion Biology Using Single-Molecule Localization Microscopy. Chemphyschem 15:606-618. Deschout, H., A. Shivanandan, P. Annibale, M. Scarselli, and A. Radenovic. 2014. Progress in quantitative single-molecule localization microscopy. Histochem Cell Biol 142:5-17. Nicovich, P. R., D. M. Owen, and K. Gaus. 2017. Turning single-molecule localization microscopy into a quantitative bioanalytical tool. Nat Protoc 12:435-461. Gardel, M. L., I. C. Schneider, Y. Aratyn-Schaus, and C. M. Waterman. 2010. Mechanical Integration of Actin and Adhesion Dynamics in Cell Migration. Annu Rev Cell Dev Bi 26:315- 333. Bishop, C. M. 2006. Pattern recognition and machine learning. Springer. 14 Shroff, H., C. G. Galbraith, J. A. Galbraith, and E. Betzig. 2008. Live-cell photoactivated localization microscopy of nanoscale adhesion dynamics. Nat Methods 5:417-423. Fuchs, J., S. Bohme, F. Oswald, P. N. Hedde, M. Krause, J. Wiedenmann, and G. U. Nienhaus. 2010. A photoactivatable marker protein for pulse-chase imaging with superresolution. Nat Methods 7:627-630. Kanchanawong, P., G. Shtengel, A. M. Pasapera, E. B. Ramko, M. W. Davidson, H. F. Hess, and C. M. Waterman. 2010. Nanoscale architecture of integrin-based cell adhesions. Nature 468:580-584. Arnold, M., E. A. Cavalcanti-Adam, R. Glass, J. Blummel, W. Eck, M. Kantlehner, H. Kessler, and J. P. Spatz. 2004. Activation of integrin function by nanopatterned adhesive interfaces. Chemphyschem 5:383-388. Platzman, I., C. A. Muth, C. Lee-Thedieck, D. Pallarola, R. Atanasova, I. Louban, E. Altrock, and J. P. Spatz. 2013. Surface properties of nanostructured bio-active interfaces: impacts of surface stiffness and topography on cell-surface interactions. Rsc Adv 3:13293-13303. Geiger, B., J. P. Spatz, and A. D. Bershadsky. 2009. Environmental sensing through focal adhesions. Nat Rev Mol Cell Bio 10:21-33. Annibale, P., M. Scarselli, M. Greco, and A. Radenovic. 2012. Identification of the factors affecting co-localization precision for quantitative multicolor localization microscopy. Optical Nanoscopy 1:9. Deschout, H., T. Lukes, A. Sharipov, D. Szlag, L. Feletti, W. Vandenberg, P. Dedecker, J. Hofkens, M. Leutenegger, T. Lasser, and A. Radenovic. 2016. Complementarity of PALM and SOFI for super-resolution live-cell imaging of focal adhesions. Nat Commun 7:13693. Pallarola, D., I. Platzman, A. Bochen, E. A. Cavalcanti-Adam, H. Axmann, H. Kessler, B. Geiger, and J. P. Spatz. 2017. Focal adhesion stabilization by enhanced integrin-cRGD binding affinity. BioNanoMaterials https://doi.org/10.1515/bnm-2016-0014. Mortensen, K. I., L. S. Churchman, J. A. Spudich, and H. Flyvbjerg. 2010. Optimized localization analysis for single-molecule tracking and super-resolution microscopy. Nat Methods 7:377-381. Ober, R. J., S. Ram, and E. S. Ward. 2004. Localization accuracy in single-molecule microscopy. Biophys J 86:1185-1200. Sengupta, P., T. Jovanovic-Talisman, D. Skoko, M. Renz, S. L. Veatch, and J. Lippincott- Schwartz. 2011. Probing protein heterogeneity in the plasma membrane using PALM and pair correlation analysis. Nat Methods 8:969-975. Owen, D. M., C. Rentero, J. Rossy, A. Magenau, D. Williamson, M. Rodriguez, and K. Gaus. 2010. PALM imaging and cluster analysis of protein heterogeneity at the cell surface. J Biophotonics 3:446-454. Kiskowski, M. A., J. F. Hancock, and A. K. Kenworthy. 2009. On the Use of Ripley's K-Function and Its Derivatives to Analyze Domain Size. Biophys J 97:1095-1103. Baddeley, D., I. D. Jayasinghe, L. Lam, S. Rossberger, M. B. Cannell, and C. Soeller. 2009. Optical single-channel resolution imaging of the ryanodine receptor distribution in rat cardiac myocytes. P Natl Acad Sci USA 106:22275-22280. Endesfelder, U., K. Finan, S. J. Holden, P. R. Cook, A. N. Kapanidis, and M. Heilemann. 2013. Multiscale Spatial Organization of RNA Polymerase in Escherichia coli. Biophys J 105:172- 181. Ester, M., H. P. Kriegel, J. Sander, and X. Xu. 1996. A density-based algorithm for discovering clusters in large spatial databases with noise. Proceedings of the Second International Conference on Knowledge Discovery and Data Mining (KDD-96):226-231. Verbeek, J. J., N. Vlassis, and B. Krose. 2003. Efficient greedy learning of Gaussian mixture models. Neural Comput 15:469-485. Punzo, A., R. P. Browne, and P. D. McNicholas. 2014. Hypothesis testing for parsimonious Gaussian mixture models. arXiv 1405.0377. 15 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. Deschout, H., F. C. Zanacchi, M. Mlodzianoski, A. Diaspro, J. Bewersdorf, S. T. Hess, and K. Braeckmans. 2014. Precisely and accurately localizing single emitters in fluorescence microscopy. Nat Methods 11:253-266. Sharonov, A., and R. M. Hochstrasser. 2006. Wide-field subdiffraction imaging by accumulated binding of diffusing probes. P Natl Acad Sci USA 103:18911-18916. Annibale, P., S. Vanni, M. Scarselli, U. Rothlisberger, and A. Radenovic. 2011. Identification of clustering artifacts in photoactivated localization microscopy. Nat Methods 8:527-528. Ratz, M., I. Testa, S. W. Hell, and S. Jakobs. 2015. CRISPR/Cas9-mediated endogenous protein tagging for RESOLFT super-resolution microscopy of living human cells. Sci Rep-Uk 5:9592. 34. 35. 36. 37. 16 Figures Figure 1. Application of SMLM clustering algorithms to PALM data of focal adhesions. (A) PALM image of a fixed REF cell expressing integrin β3 labelled with mEos2. (B) Zoom-in PALM image corresponding to the green rectangle in (A). (C) Scatter plot of the mEos2 localizations corresponding to the green rectangle in (A). (D) Ripley's L(r)-r as a function of r, obtained from the localizations in (C). (E) Clusters obtained from the localizations in (C) by DBSCAN. The minimum number of localizations was set to 10, and two values were chosen for the maximum search radius rmax: 0.05 and 0.10 µm. The different colors of the localizations indicate to which cluster they belong, the background localizations are red. (F) Result of EMGM analysis of the localizations in (C). The red dots symbolize the localizations, and 17 the blue ellipses the 2σ error ellipses of the components. (G) Histograms showing the eccentricity b/a, localization density, number of localizations, and area πab of the 2σ error ellipses of the components obtained by EMGM from the complete PALM data set in (A). The rightmost bins contains all values within that bin and larger. 18 Figure 2. Evaluation of EMGM using simulated data. (A) On the left, an example of a simulated Gaussian mixture consisting of K = 4 components, each containing 100 localizations, described by a symmetric 2D Gaussian distribution with a standard deviation σx = σy = 20 nm. The Gaussian centers are placed in a square grid with spacing dx,y = 100 nm. On the right, the EMGM result. The red dots symbolize the localizations. The blue dots symbolize the center positions and the blue ellipses the 2σ error ellipses of the components. (B) On the right, the average number of mixture components correctly identified by EMGM as a function of the simulated K. On the left, an example EMGM result for K = 16. (C) On the right, the average standard deviation σx,y of the mixture components calculated by EMGM as a function of the simulated localization background density bg. On the left, an example EMGM result for bg = 40,000 #/µm2. (D) On the right, the average σx,y calculated by EMGM as a function of the simulated localization uncertainty s. On the left, an example EMGM result for s = 30 nm. (E) On the right, the average number of mixture components correctly identified by EMGM as a function of the simulated spacing dx,y. On the left, an example EMGM result for dx,y = 60 nm. (F) On the right, the average eccentricity σx/σy of the mixture components calculated by EMGM as a function of the simulated σx/σy. On the left, an example EMGM result for σx/σy = 0.2. The simulated Gaussian mixtures in (C-F) consist of K = 4 components, similar to (A). The dashed lines in (B-F) represent the ground truth (GT) and the shaded areas the standard deviation (n = 100). 19 Figure 3. EMGM analysis of PALM data of integrin β3 or paxillin on fibronectin-coated substrates. (A) Summed TIRF images of the mEos2 off-state of fixed REF cells expressing integrin β3 or paxillin labelled with mEos2, growing on fibronectin-coated substrates. (B) Zoom-in PALM images corresponding to the green rectangles in (A). (C) Result of the EMGM analysis of the PALM data shown in (B). The red dots symbolize the localizations, and the blue ellipses the 2σ error ellipses of the mixture components. (D-F) Result of the EMGM analysis of PALM data corresponding to different REF cells (n = 10): (D) number of localizations in each mixture component as a function of the area of its 2σ error ellipse, (E) eccentricity of the 2σ error ellipse of each mixture component as a function of its number of localizations, (F) eccentricity of the 2σ error ellipse of each mixture component as a function of its area. The dashed white rounded rectangles in (D) and (E) are visual guides. 20 Figure 4. EMGM analysis of PALM data of integrin β3 on nano-patterned substrates. (A) Summed TIRF images of the mEos2 off-state of fixed REF cells expressing integrin β3 labelled with mEos2, growing on nano-patterned substrates with 56 nm or 119 nm spacing between the AuNPs. (B) Zoom-in PALM images corresponding to the green rectangles in (A). (C) Result of the EMGM analysis of the PALM data shown in (B). The red dots symbolize the localizations, and the blue ellipses the 2σ error ellipses of the mixture components. (D-F) Result of the EMGM analysis of PALM data corresponding to different REF cells (n = 10). The number of localizations in each mixture component is shown as a function of the area of its 2σ error ellipse, for (D) fibronectin coated substrates (Fig. 3D), (E) nano- patterned substrates with 56 nm spacing, (F) nano-patterned substrates with 119 nm spacing. The dashed white rounded rectangles in (D) and (E) are visual guides. 21 Figure 5. Merging procedure applied on EMGM results for integrin β3. (A) Illustration of the concept of merging overlapping mixture components based on overlapping error ellipses. The red dots symbolize the localizations. The black/green/blue ellipses represent the 2σ error ellipses of the merged/isolated/overlapping mixture components. (B) EMGM result for PALM data of a fixed REF cell growing on a fibronectin-coated substrate expressing integrin β3 labelled with mEos2 (Fig. 3C). (C) Result of the merging procedure applied on the EMGM result in (B). (D-F) Result of the merging procedure applied on EMGM results for integrin β3 (Fig. 4D, and Fig. 5D). The number of localizations in each mixture component is shown as a function of the area of its 2σ error ellipse, for (D) the merged components, (E) the isolated components, and (F) the overlapping components. The dashed white rounded rectangles in (F) are visual guides. 22 Supporting Material Supporting Text ...................................................................................................................................... 2 1. Expectation maximization of a Gaussian mixture (EMGM) .......................................................... 2 1.1 Classic algorithm ....................................................................................................................... 2 1.2 Initialization by greedy learning .............................................................................................. 2 1.3 Model selection by hypothesis testing .................................................................................... 3 1.4 Localization background .......................................................................................................... 4 1.5 Localization uncertainty ........................................................................................................... 4 2. Simulations ..................................................................................................................................... 6 2.1 Simulation details ..................................................................................................................... 6 2.2 Number of mixture components ............................................................................................. 6 2.3 Number of initializations.......................................................................................................... 6 2.4 Localization background .......................................................................................................... 7 2.5 Localization uncertainty ........................................................................................................... 7 2.6 Number of localizations ........................................................................................................... 7 3. Applying EMGM on experimental data ......................................................................................... 8 3.1 Scanning procedure .................................................................................................................. 8 3.2 Combining procedure ............................................................................................................... 8 4. Merging procedure ....................................................................................................................... 10 5. PAINT imaging of integrin β3 ....................................................................................................... 11 5.1 Sample preparation ................................................................................................................ 11 5.2 Imaging procedure ................................................................................................................. 11 5.3 Discussion ............................................................................................................................... 11 6. Production of nano-patterned substrates ................................................................................... 12 References ........................................................................................................................................ 13 Supporting Figures ............................................................................................................................... 14 Supporting Tables ................................................................................................................................. 23 1 Supporting Material Supporting Text 1. Expectation maximization of a Gaussian mixture (EMGM) 1.1 Classic algorithm We apply expectation maximization of a Gaussian mixture (EMGM) [1] on single-molecule localization (SMLM) data to investigate the substructure of focal adhesions (FAs). The main assumption is that the FA subunits can be described as bivariate Gaussians. The spatial probability distribution of an FA subunit is thus given by: 1 𝑁(𝒓𝝁, 𝚺) = exp (− (𝒓 − 𝝁)T ∙ 𝚺−1 ∙ (𝒓 − 𝝁)) (1) 2𝜋√𝚺 1 2 where 𝒓 is the position in which the Gaussian is being evaluated, 𝝁 the center position of the Gaussian, and 𝚺 the covariance matrix of the Gaussian. Assume one or more FAs consisting out of 𝑁 positions 𝒓𝑛. According to our assumption, these FAs can be modeled by a mixture of bivariate Gaussians. Assume that that this mixture consists of 𝐾 components with the weight of component 𝑘 described by the mixing coefficient 𝜋𝑘. These mixing coefficients fulfil the condition: 𝐾 ∑ 𝜋𝑘 = 1 𝑘=1 (2) (3) Expectation maximization is a popular algorithm to identify the properties 𝝁𝑘, 𝚺𝑘 and 𝜋𝑘 of each component the Gaussian mixture. After choosing initial values, the expectation step consists of evaluating the posterior probability that localization 𝒓𝑛 was generated from component 𝑘: 𝛾𝑛𝑘 = 𝜋𝑘𝑁(𝒓𝑛𝝁𝑘, 𝚺𝑘) 𝐾 𝑗=1 𝜋𝑗𝑁(𝒓𝑛𝝁𝑗, 𝚺𝑗) ∑ In the maximization step, the parameters are re-estimated using the posterior probabilities: 𝑁 new = 𝝁𝑘 1 𝑁𝑘 ∑ 𝛾𝑛𝑘𝒓𝑛 𝑛=1 new = 𝚺𝑘 1 𝑁𝑘 ∑ 𝛾𝑛𝑘(𝒓𝑛 − 𝝁𝑘 new) ∙ (𝒓𝑛 − 𝝁𝑘 new)T 𝑁 𝑛=1 new = 𝜋𝑘 𝑁𝑘 𝑁 Where 𝑁𝑘 is defined as the number of localizations that belong to component 𝑘: 𝑁 𝑁𝑘 = ∑ 𝛾𝑛𝑘 𝑛=1 (4) (5) Finally, the likelihood of the updated Gaussian mixture is calculated and checked for convergence: 𝑁 𝐾 ℒ = ∏ ∑ 𝜋𝑘 new𝑁(𝒓𝑛𝝁𝑗 new, 𝚺𝑗 new) 𝑛=1 𝑗=1 (6) If the convergence criterion is not satisfied, the expectation and maximization steps described in Eqs. (3) and (4) are repeated. 1.2 Initialization by greedy learning EMGM is known to be sensitive to local maxima. To avoid finding such a solution, initial values of the parameters 𝝁𝑘, 𝚺𝑘 and 𝜋𝑘 (see Supporting Text, Section 1.1) need to be chosen sufficiently close to the real values. In the context of SMLM, these values are not known. Although several approaches have been reported in order to initialize the model parameters for EMGM, there is no widely accepted 2 Supporting Material method. Popular approaches are randomly generating the initial parameter values, or estimating them using the k-means clustering algorithm [1]. An interesting alternative to these initialization methods is the so-called "greedy learning" approach [2], based on repeating the EMGM by starting from a trivial Gaussian mixture consisting of one component, and each time adding an extra component. The EMGM solution obtained for a 𝑃-1 component mixture is used as initialization for the 𝑃 component mixture, by deleting one component and inserting two random components, based on the deleted one. This can be done 𝑃-1 times, for each component of the old mixture, and the solution with the highest likelihood is retained. By doing so, one proceeds until a desired number of components 𝐾 is attained. Additionally, each step consisting of 𝑃-1 initializations can be repeated 𝑄 times to increase the accuracy of the result. The total number of EMGM repeats to obtain the correct solution of 𝐾 components is thus given by 𝑄(1 + ∑ 𝑖𝐾 This shows that the initialization procedure becomes computationally more expensive for datasets containing more components. The computation time on a mid-range personal computer for the simulations shown in Fig. 2 ranged from ~3 s (for 𝐾 = 1 and 𝑄 = 3) to ~1000 s (for 𝐾 = 20 and 𝑄 = 3). Note that we actually used 𝑄(1 + ∑ [𝑖 + 1] ) initializations due to an extra background "component", see Supporting Text, Section 1.4. 1.3 Model selection by hypothesis testing 𝑖=1 ). 𝐾 𝑖=1 When applying EMGM, the number of components 𝐾 for the Gaussian mixture needs to be chosen. In the context of SMLM, this number is unknown. In order to select the most appropriate number of components, one can repeat the EMGM procedure for a range of 𝐾 values. The likelihood value is not a good selection criterion, as increasing the number of components increases the likelihood monotonously. A solution provided by information theory is the Akaike or Bayes information criterion [1], which penalizes an increasing number of components and therefore leads to a maximum value for a certain 𝐾 value. However, this value has been reported to typically overestimate the real number of components [3]. Hypothesis testing can provide a more conservative approach towards selecting to right mixture model [4]. Assume two mixtures calculated by EMGM, one containing 𝐾-1 components and the other containing 𝐾 components. The 𝐾 component model will have a larger likelihood than the 𝐾-1 component model. Consider the null hypothesis that the 𝐾-1 component model is the correct one, which will correspond to a specific distribution of likelihood increments. If the real model consists of more than 𝐾-1 components, the likelihood increment can be expected to be larger than the values described by the null hypothesis distribution. This distribution, however, is unknown, but can be simulated from the identified 𝐾-1 component model, i.e. a number of bootstrapped data sets are generated assuming the null hypothesis and the increments in likelihood are obtained by applying EMGM for both 𝐾-1 and 𝐾 components. Comparing the real likelihood increment with the bootstrap null hypothesis distribution allows to determine the p-value, in turn allowing to accept or reject the null hypothesis. Choosing the maximum allowed p-value sufficiently small, e.g. equal to 0.01, means that there is only a 1% chance to select a mixture model that contains too many components, preventing overestimation of the number of components. 3 1.4 Localization background Supporting Material While initialization and model selection issues are inherent to EMGM, other problems arise because of the nature of SMLM data. One important problem is that not necessarily all localizations are part of FAs, but instead can belong to a background. Consider an SMLM dataset consisting of 𝑁 positions that belong to a mixture of multivariate Gaussians, and an extra 𝑁b positions that belong to a background, within an area 𝐴. In case of a simple uniform background, the probability distribution of the background localizations is given by: 𝐵 = 1 𝐴 The algorithm can readily be adjusted to incorporate the background described by 𝐵. First of all, the posterior probability that localization 𝒓𝑛 was generated from component 𝑘 (see Eq. (3)) is now given by: 𝛾𝑛𝑘 = 𝜋𝑘𝑁(𝒓𝑛𝝁𝑘, 𝚺𝑘) 𝜋𝑗𝑁(𝒓𝑛𝝁𝑗, 𝚺𝑗) + 𝐵 ∑ 𝐾 𝑗=1 (7) (8) (9) (10) And an equivalent posterior probability for the background can be defined as: 𝐵 𝛿𝑛 = ∑ 𝐾 𝑗=1 𝜋𝑗𝑁(𝒓𝑛𝝁𝑗, 𝚺𝑗) + 𝐵 The re-estimation of the parameters 𝝁𝑘 and 𝚺𝑘 can be done as before, while the re-estimation of the mixing coefficients (see Eq. (4)) has to be adjusted as follows: 𝑁𝑘 new = 𝜋𝑘 𝑁 + 𝑁b where 𝑁b can be calculated using the background posterior probabilities: 𝑁 𝑁b = ∑ 𝛿𝑛 𝑛=1 (11) Finally, the calculation of the likelihood of the updated Gaussian mixture (see Eq. (6)) is adjusted as follows: 𝑁+𝑁b 𝐾 ℒ = ∏ {∑ 𝜋𝑘 new𝑁(𝒓𝑛𝝁𝑗 new, 𝚺𝑗 new) + 𝐵} 𝑛=1 𝑗=1 (12) The background can effectively be considered as an extra component of the Gaussian mixture, requiring an adaptation of the initialization procedure described in Supporting Text, Section 1.2. Initialization of a 𝑃 component Gaussian mixture is done 𝑃 times instead of 𝑃 – 1 times (i.e. 𝑃 – 1 initializations corresponding to each component of the previous solution, and 1 initialization corresponding to the background of the previous solution). 1.5 Localization uncertainty The localizations in SMLM data contain measurement uncertainties [5]. The localization uncertainty can be described as an extra contribution 𝜺 to the real position of the molecule. This contribution is described by a spatial probability distribution that is usually modeled as a Gaussian: 1 𝐸(𝜺𝑠) = exp (− 2𝜋𝑠 𝜺2 2𝑠2) (13) The standard deviation 𝑠 is often termed as the localization uncertainty or precision. An observed localization 𝒓 belonging to component 𝑘 is described by the sum of 𝜺 and the real emitter position. Since both variables are independent, the spatial probability distribution of their sum is given by the convolution of their corresponding spatial probability distributions (see Eqs. (1) and (13)): 4 +∞ 𝑁(𝒓𝝁𝑘, 𝚺𝑘, 𝑠) = ∫ 𝐸(𝒓 − 𝒓′𝑠)𝑁(𝒓′𝝁𝑘, 𝚺𝑘) 𝑑𝒓′ −∞ This is the convolution of two bivariate Gaussians, which can be solved as [6]: 1 𝑁(𝒓𝝁𝑘, 𝚺𝑘, 𝑠) = exp (− (𝒓 − 𝝁𝑘)T ∙ (𝚺𝑘 + 𝑠2𝑰)−1 ∙ (𝒓 − 𝝁𝑘)) 2𝜋√𝚺𝑘 + 𝑠2𝑰 1 2 Supporting Material (14) (15) (16) where 𝑰 is the identity matrix. This expression describes the observed spatial probability distribution of component 𝑘. In order to incorporate the effect of the localization uncertainty in EMGM, we need to adjust the algorithm in two ways. First of all, the expectation step needs to be adjusted, since the expression for the posterior probability 𝛾𝑛𝑘 of molecule position 𝒓𝑛 of component 𝑘 contains the spatial probability distribution of that component (see Eq. (3)). Substitution of Eq. (15) in Eq. (3) yields the adjusted posterior probability: 𝛾𝑛𝑘 = 𝜋𝑘𝑁(𝒓𝑛𝝁𝑘, 𝚺𝑘, 𝑠𝑛) 𝐾 𝑗=1 𝜋𝑗𝑁(𝒓𝑛𝝁𝑗, 𝚺𝑗, 𝑠𝑛) ∑ where 𝑠𝑛 is the localization uncertainty corresponding to localization 𝒓𝑛. Secondly, the maximization step needs to be adjusted, because the apparent spatial probability distribution is a bivariate Gaussian with a covariance matrix equal to 𝚺𝑘 + 𝑠2𝑰 (see Eq. (15)). This means that the presence of localization uncertainties affects both the shape and size of the observed component 𝑘. The re-estimation of the covariance matrix (see Eq. (4)) should be adjusted as follows: 𝑁 new = 𝚺𝑘 ∑ 𝛾𝑛𝑘{(𝒓𝑛 − 𝝁𝑘 new) ∙ (𝒓𝑛 − 𝝁𝑘 new)T − 𝑠𝑛 2𝑰} (17) 1 𝑁𝑘 𝑛=1 The contribution coming from the localization uncertainty is included within the sum, since the value of the localization uncertainty can change for different localizations. Note that Eq. (17) suggests that the covariance matrix values of certain mixture components can possibly become negative during the EMGM procedure. If this occurs during EMGM, the covariance matrix is not updated, and the value of the previous iteration is retained. 5 Supporting Material 2. Simulations 2.1 Simulation details The simulations shown in Fig. 2 were performed in Matlab (The Mathworks). Briefly, Gaussian mixtures consisting of 𝐾 components were simulated. The localizations in each component were obtained from a Gaussian probability distribution, using the Matlab function mvnrnd. The Gaussian standard deviation was 𝜎𝑥 = 𝜎𝑦 = 20 nm (except for Fig. 2F), and the number of localizations for each component was 𝑁𝑘 = 100. The number of mixture components 𝐾 was varied between 1 and 20 in Fig. 2B, and fixed at 4 in Fig. 2C-F. The centers of the mixture components were placed in a square grid with a spacing 𝑑𝑥,𝑦 equal to five times 𝜎𝑥,𝑦 (except for Fig. 2D). A uniform localization background was added in Fig. 2C by randomly generating a number of localizations from a uniform distribution, using the Matlab function rand. The number of background localizations was determined from the localization background density 𝑏𝑔, which was varied between 0 and 50,000 #/µm2, in steps of 1000 #/µm2. The effect of the localization uncertainty shown in Fig. 2D was simulated by adding to each localization coordinate a value randomly generated from a Gaussian distribution with standard deviation 𝑠, using the Matlab function randn. The value of the localization uncertainty 𝑠 was varied from 0 to a 40 nm, in steps of 1 nm. To account for the apparent increase in component size, the spacing between the component centers was adjusted to five times √𝜎𝑥,𝑦 2 + 𝑠2. In Fig. 2E, the spacing 𝑑𝑥,𝑦 between the component centers was increased from 0 to 200 nm, in steps of 5 nm. The changing component eccentricity shown in Fig. 2F was simulated by increasing the component standard deviation 𝜎𝑥 from 2.8 to 20 nm, and simultaneously decreasing the standard deviation 𝜎𝑦 from 140 to 20 nm, resulting in eccentricities 𝜎𝑥 𝜎𝑦⁄ increasing from 0.02 to 1. For each case, 100 simulations were performed. 2.2 Number of mixture components The simulation results in Fig. 2B show that EMGM increasingly underestimates the number of mixture components for an increasing value of 𝐾. Additionally, the number of non-existing components (i.e. false positives) identified by EMGM also increases with 𝐾, as illustrated in Fig. S3, A and B. We define 𝐾id as the number of mixture components correctly identified by EMGM, and 𝐾fp as the number of false positive components found by EMGM. Using the simulated data from Fig. 2B, we calculated the probability of obtaining a completely correct EMGM result (i.e. 𝐾id = 𝐾 and 𝐾fp = 0) as a function of 𝐾. The results are shown in Fig. S3C. For mixtures with 𝐾 < 10, this probability is on average equal to 94%. For larger numbers, the method starts to underestimate 𝐾, most likely because the contribution of correctly fitting individual components to the total likelihood becomes smaller with an increasing number. Fig. S3D shows the average values of 𝐾id and 𝐾fp as a function of 𝐾. The average number of false positives is smaller than 1 for mixtures with 𝐾 < 10. 2.3 Number of initializations The initialization procedure described in Supporting Text, Section 1.2 consists of 𝑃-1 separate initializations for a 𝑃 component Gaussian mixture. If the localization background is considered as an extra component, the procedure actually consists of 𝑃 separate initializations for a 𝑃 component 6 Supporting Material mixture (see Supporting Text, Section 1.4). This procedure can be repeated several times 𝑄 to improve the accuracy of the EMGM result, resulting in a total of 𝑄𝑃 initializations for a 𝑃 component Gaussian mixture. In order to investigate the effect of the value of 𝑄 on the EMGM performance, we performed simulations similar to the ones shown in Fig. 2B, for different values of 𝑄. Fig. S4A shows that an increasing 𝑄 results in less underestimation of 𝐾, although the improvement is small for 𝑄 > 3. The number of false positive components 𝐾fp does not seem to be affected by the value of 𝑄, see Fig. S4B. 2.4 Localization background The adapted EMGM performs excellently in the presence of a uniform localization background (see Fig. 2C and Fig. S5, A and B). Only for values of the localization background density that are not representative for our experimental conditions (e.g. 𝑏𝑔 = 50,000 #/µm2 in Fig S5C), the algorithm starts to underestimate the true amount of mixture components and to find false positive components. Using the simulated data from Fig. 2C, we calculated the probability of obtaining a completely correct EMGM result (i.e. 𝐾id = 𝐾 and 𝐾fp = 0) as a function of 𝑏𝑔 (see Fig. S5C). For mixtures with 𝑏𝑔 < 25,000 #/µm2, this probability is on average equal to 93%. Fig. S5D shows the average values of 𝐾id and 𝐾fp as a function of 𝑏𝑔, confirming that the EMGM performance deteriorates for values larger than 25,000 #/µm2. This is not a surprise, since the characteristic localization density of the component mixtures themselves is lower (each component counts 100 localization and has a standard deviation of 𝜎𝑥,𝑦 = 20 nm, resulting in a 2𝜎 ellipse area of 0.016 µm2, which yields a characteristic localization density around 20,000 #/µm2). 2.5 Localization uncertainty The simulation results in Fig. 2D show that the estimated standard deviation 𝜎𝑥,𝑦 of the mixture components is slightly affected by an increasing localization uncertainty 𝑠. However, as illustrated in Fig. S6, A-C, an increasing value of 𝑠 can have an important impact on the values of 𝐾id and 𝐾fp. We assessed the probability of obtaining a completely correct EMGM result (i.e. 𝐾id = 𝐾 and 𝐾fp = 0) as a function of 𝑠, using the simulated data shown in Fig. 2D. The results are shown in Fig. S6D, indicating that the probability decreases strongly when 𝑠 becomes larger than 30 nm. This is to be expected, since the localization uncertainty is larger than the standard deviation 𝜎𝑥,𝑦 = 20 nm of the mixture components itself. Fig. S6E shows 𝐾id and 𝐾fp as a function of 𝑠. For localization uncertainties larger than 30 nm, the average number of correctly identified components slightly decreases, while the average number of false positives increases more strongly. 2.6 Number of localizations The simulations presented in Fig. 2 describe Gaussian mixtures with components that each consist of 𝑁𝑘 = 100 localizations. However, as illustrated in Fig. S7, A-C, the performance of the EMGM algorithm can depend on the value of 𝑁𝑘. We assessed the probability of obtaining a completely correct EMGM result (i.e. 𝐾id = 𝐾 and 𝐾fp = 0) as a function of 𝑁𝑘, using simulations similar to Fig. 2A. The results are shown in Fig. S7D, indicating that the probability decreases strongly when 𝑁𝑘 becomes smaller than 50. Fig. S7E shows 𝐾id and 𝐾fp as a function of 𝑁𝑘, indicating that this low probability is mainly due to EMGM not detecting all mixture components for such low number of localizations. 7 Supporting Material 3. Applying EMGM on experimental data 3.1 Scanning procedure The number of FA substructures present in a typical SMLM dataset is not known, and can be assumed to be larger than 10. However, the simulation results in Fig. 2B indicate that the EMGM analysis is optimal when the Gaussian mixture consists of a smaller number of components. We therefore split the SMLM dataset into smaller subsets and perform the EMGM analysis on each subset separately. This can be done simply by scanning the original region of interest along non-overlapping square subregions with side length 𝐿, as illustrated in Fig. S2, A and B. However, this scanning procedure clips Gaussian mixture components that are not completely contained in a single subregion. A solution is repeating the scan with subregions that are shifted over a distance equal to 𝐿 2⁄ . If this shift is done in three different directions, as shown in Fig. S2, B-E, each component with dimensions below 𝐿 2⁄ is completely included in at least one subregion of at least one scan. Considering that the FA substructures of interest have sizes below the diffraction limit, we choose 𝐿 = 2 µm. 3.2 Combining procedure Combining the EMGM results obtained from the scanning procedure described in Supporting Text, Section 3.1 consists of two steps: (1) the EMGM results of the subregions within each separate scan need to be combined, resulting in four different EMGM descriptions of the same original dataset, and (2) combining these four results yields the final EMGM result. For the first step, we make the approximation that all components identified in a subregion are completely described by the localizations within that subregion. The posterior probability (see Eq. (3)) of a localization within a certain subregion belonging to a component identified in another subregion will therefore be zero. This means that the posterior probabilities of all 𝑀 subregions of a single scan can be assembled into a sparse matrix 𝜸scan to describe the posterior probabilities of the full dataset: 𝜸scan = [ ] (18) 𝜸1 ⋯ 𝟎 ⋮ ⋮ 𝟎 ⋯ 𝜸𝑀 ⋱ where the matrices 𝜸𝑖 describe the posterior probabilities of the localizations inside subregion 𝑖, with 𝑖 = 1, …, 𝑀. The posterior probabilities corresponding to the full dataset for a localization to belong to the background (see Eq. (9)) are similarly given by: 𝜹scan = [ ] (19) 𝜹1 ⋮ 𝜹𝑀 This approximation is not optimal for the case of Gaussian mixture components being clipped, as described in Supporting Text, Section 3.1. The column in 𝜸scan corresponding to such a clipped component is therefore deleted, and its values are added to 𝜹scan. The criterion for determining whether a component is clipped is chosen as whether its 2𝜎 error ellipse (containing around ~95% of localizations) is completely inside the subregion or not. The resulting 𝜸scan does not provide a complete description of the Gaussian mixture in the full dataset due to the deletion of components that are clipped during the scanning procedure. However, each clipped component that is deleted from a certain scan is, in theory, identified in at least one of the three other scans (see Fig. S2). The second step therefore consists of merging the 𝜸scan matrices of the four different scans. For this purpose, the Pearson correlation between the posterior probabilities of each pair of components 𝑖 and 𝑗 belonging to different scans is calculated: 8 𝜌𝑖𝑗 = ∑ (𝛾𝑖𝑛 − 𝛾𝑖𝑛) 𝑛 (𝛾𝑗𝑛 − 𝛾𝑗𝑛) √∑ (𝛾𝑖𝑛 − 𝛾𝑖𝑛)2 𝑛 2 ∑ (𝛾𝑗𝑛 − 𝛾𝑗𝑛) 𝑘 Supporting Material (20) The sum runs over all localizations 𝑛 that have a non-zero posterior probability (e.g. excluding all localizations outside subregion 𝑖 and 𝑗). The correlation is tested against the null hypothesis that the posterior probabilities of components 𝑖 and 𝑗 are not correlated (i.e. it is verified that the correlation is larger than the values described by a simulated null hypothesis distribution). Two components identified in two different scans are considered to be identical if their correlation is significant according to the null hypothesis and if the correlation is larger than any other significant correlation involving either 𝑖 or 𝑗. After identifying all identical components, their posterior probabilities are combined by averaging, while the posterior probabilities of components identified in only one scan are retained. This results in a final 𝜸 that describes the full dataset without clipped components. The background posterior probabilities are combined similarly into a final 𝜹. 9 Supporting Material 4. Merging procedure The merging procedure illustrated in Fig. 5A is performed by splitting the mixture components obtained by EMGM into two categories: the ones whose 1𝜎 error ellipse intersects with at least one other error ellipse, called the "overlapping" components, and the ones whose 1𝜎 error ellipse does not intersect with another one, called the "isolated" components. The 1𝜎 error ellipse is chosen because it corresponds to the probability of containing ~40% of all localizations. This means that localizations on the intersection between two such error ellipses have approximately an equal probability to belong to both corresponding components, therefore suggesting that they can be viewed as a single merged object. Once a set of 𝐾overlap overlapping components have been verified, a new merged object can be calculated by summing their posterior probabilities 𝛾𝑛𝑘 (see Eq. (3)): 𝐾overlap The properties of the merged object can then be calculated using Eq. (4). This gives rise to a third category, called the "merged" components. 𝛾𝑛,merged = ∑ 𝛾𝑛𝑘 𝑘=1 (21) 10 5. PAINT imaging of integrin β3 5.1 Sample preparation Supporting Material to [7] experiments. The sample was prepared according We used a commercial kit (Ultivue-2, Ultivue) for our points accumulation in nanoscale topography (PAINT) the manufacturer's recommendations. Briefly, we seeded around 105 REF cells on a fibronectin-coated 25 mm diameter cover slip, incubated them at 37° C in cell culture medium, washed them with PBS after 24h, and fixed them with 2.5% paraformaldehyde at 37° C for 10 minutes (see Materials and Methods). After removing the fixative, the cells were washed three times with PBS, the cover slip was placed into a custom made holder, and they were incubated in PBS for 10 minutes at 37° C. The cells were subsequently reduced by incubating them for 10 minutes in a freshly prepared 0.1% sodium borohydride solution at room temperature. Afterwards, the cells were washed three times with PBS, and incubated in PBS for 10 minutes at room temperature. Next, the cells were incubated for 1.5h at room temperature in a blocking and permeabilization buffer consisting of PBS with 3% bovine serum albumin and 0.2% Triton X-100. The primary antibody staining was carried out by incubating the cells overnight at 4 °C with integrin β3 mouse monoclonal antibodies (sc-7311, Santa Cruz Biotechnology) diluted 100 times in staining buffer composed of PBS with 1% bovine serum albumin and 0.2% Triton X-100. Next, the cells were washed four times with PBS, and incubated in PBS for 10 minutes at room temperature. The secondary antibody staining was carried out by first incubating the cells in Antibody Dilution Buffer (Ultivue-2, Ultivue) for 10 minutes at room temperature, and then for 2h with Goat-anti-Mouse-D1 antibodies (Ultivue-2, Ultivue) diluted 100 times in Antibody Dilution Buffer. Next, the cells were washed four times with PBS, and incubated in PBS for 10 minutes at room temperature. 5.2 Imaging procedure Prior to imaging, 100 nm gold nanospheres (C-AU-0.100, Corpuscular) were added to the sample for lateral drift correction (see Materials and Methods). Imaging was performed using image strand I1- 560 (Ultivue-2, Ultivue) diluted in Image Buffer (Ultivue-2, Ultivue) at a concentration of 1 nM. The imaging procedure was similar as for the PALM measurements (see Materials and Methods). 5.3 Discussion We used PAINT to image fixed rat embryonic fibroblast (REF) cells where integrin β3 was antibody stained. The resulting PAINT images show FAs as patchy structures (Fig. S9). We hypothesize that this is caused by difficulties in labeling integrin with antibodies, for instance due to cell membrane areas that are curved inwards, resulting in an integrin epitope that is more difficult to access. We also noticed that mostly the cell periphery was labelled, again suggesting that not all integrins are accessible for the antibodies. 11 Supporting Material 6. Production of nano-patterned substrates Nano-patterned substrates were prepared by means of block-copolymer micelle nanolithography (BCML) as previously described [8-10]. Briefly, quasi-hexagonally ordered gold nanoparticles arrays on cleaned 25 mm diameter microscope cover slips (#1.5 Micro Coverglass, Electron Microscopy Sciences) were fabricated using toluene solution of poly(styrene)-block-poly(2-vinyl pyridine) (PS-b- P2VP, Polymer Source Inc.) [9, 10]. The PS-b-P2VP toluene solution was treated with HAuCl4 (Sigma Aldrich) at a stoichiometric loading of (P2VP/HAuCl4) = 0.5 and stirred at least for 24h in order to obtain gold nanoparticles (AuNPs) with a diameter between 6-8 nm. The lateral distance between the individual AuNPs was adjusted by varying the micellar coating process (spinning speed). Details concerning the applied block polymers and the spin casting processes are included in Table S1. The area between the AuNPs was passivated with PLL-g-PEG (PLL(20kDa)-g[3.5]-PEG(2kDa), Susos AG) to prevent non-specific adhesion. The substrates were first activated in an oxygen plasma at 0.4 mbar and 150 W for 10 minutes. The PLL-g-PEG was diluted to a concentration of 0.25 mg/ml in a 10 mM HEPES buffer at pH 7.4. The freshly activated substrates were incubated upside down for 45 minutes at room temperature on a 60 µl drop of the PLL-g-PEG solution on parafilm in a moist chamber. Afterwards the substrates are washed once with milli-Q water. Following passivation, each surface was functionalized with cRGD pentapeptide (Peptide Specialty Laboratories GmbH) at a concentration of 25 μM in MilliQ water for 2h at room temperature. The cRGD pentapeptide was conjugated with a PEG spacer (6 units) that serves as a breach between the peptide and the cysteine. The physisorbed material was removed by thorough rinsing with MilliQ water and PBS. 12 Supporting Material References 1. 2. Bishop, C.M., Pattern recognition and machine learning. 2006: Springer. Verbeek, J.J., N. Vlassis, and B. Krose, Efficient greedy learning of Gaussian mixture models. Neural Computation, 2003. 15(2): p. 469-485. Busemeyer, J.R. and Y.M. Wang, Model comparisons and model selections based on generalization criterion methodology. Journal of Mathematical Psychology, 2000. 44(1): p. 171-189. Punzo, A., R.P. Browne, and P.D. McNicholas, Hypothesis testing for parsimonious Gaussian mixture models. arXiv, 2014. 1405.0377. Deschout, H., et al., Precisely and accurately localizing single emitters in fluorescence microscopy. Nature Methods, 2014. 11(3): p. 253-266. Vinga, S. and J.S. Almeida, Renyi continuous entropy of DNA sequences. Journal of Theoretical Biology, 2004. 231(3): p. 377-388. Sharonov, A. and R.M. Hochstrasser, Wide-field subdiffraction imaging by accumulated binding of diffusing probes. Proceedings of the National Academy of Sciences of the United States of America, 2006. 103(50): p. 18911-18916. Arnold, M., et al., Activation of integrin function by nanopatterned adhesive interfaces. Chemphyschem, 2004. 5(3): p. 383-388. Platzman, I., et al., Surface properties of nanostructured bio-active interfaces: impacts of surface stiffness and topography on cell-surface interactions. Rsc Advances, 2013. 3(32): p. 13293-13303. Pallarola, D., et al., Focal adhesion stabilization by enhanced integrin-cRGD binding affinity. BioNanoMaterials, 2017. https://doi.org/10.1515/bnm-2016-0014. 3. 4. 5. 6. 7. 8. 9. 10. 13 Supporting Figures Supporting Material Figure S1. Dark field microscopy imaging of REF cells. (A-C) The REF cells were growing on (A) a fibronectin coated substrate, (b) a nano-patterned substrate with 56 nm spacing between AuNPs, or (C) a nano-patterned substrate with 119 nm spacing between AuNPs. The images were recorded 24h after transection with the integrin β3 vector. 14 Supporting Material Figure S2. Scanning procedure for EMGM analysis of SMLM data. (A) Illustration of a Gaussian mixture with components represented by black ellipses. (B-E) Scanning procedure consisting of 4 different scans. During each scan, the EMGM analysis is performed on separate square subregions with a side length 𝐿, indicated by the colored squares. The Gaussian mixture components that can be correctly identified in a certain scan are indicated by the ellipses that have the same color as the squares. In between scans, the subregions are shifted over a distance 𝐿 2⁄ in one of the following directions: left, right, up, or down. 15 Supporting Material Figure S3. Influence of the number of mixture components 𝐾 on the EMGM performance. (A-B) Example EMGM results for simulated Gaussian mixtures with 𝐾 = 20 components. EMGM correctly identified 𝐾id = 20 components and found 𝐾fp = 0 false positive components for (A). EMGM correctly identified 𝐾id = 18 components and found 𝐾fp = 1 false positive components for (B). The red dots symbolize the simulated localizations. The blue/green dots symbolize the center positions of the correct/false positive components, the blue/green ellipses symbolize the 2𝜎 error ellipses of the correct/false positive components. (D) The simulated probability of obtaining a completely correct EMGM result (i.e. 𝐾id = 𝐾 and 𝐾fp = 0) as a function of 𝐾. (E) The simulated average values of 𝐾id and 𝐾fp as a function of 𝐾. The dashed line represents the ground truth (GT) and the shaded areas the standard deviation (𝑛 = 100). 16 Supporting Material Figure S4. Influence of the number of initialization procedures 𝑄 on the EMGM performance. Gaussian mixtures with different values of 𝐾 were simulated and analyzed by EMGM. (A) The average value of the number of correctly identified compenents 𝐾id as a function of 𝐾, for different values of 𝑄. (B) The average value of number of false positive components 𝐾fp as a function of 𝐾, for different values of 𝑄. The dashed line represents the ground truth (GT) and the shaded areas the standard deviation (𝑛 = 100). 17 Supporting Material Figure S5. Influence of the localization background on the EMGM performance. (A-C) Example EMGM results for simulated Gaussian mixtures. Each mixture consists of 𝐾 = 4 components with localization background density (A) 𝑏𝑔 = 0, (B) 𝑏𝑔 = 25,000 #/µm2, or (C) 𝑏𝑔 = 50,000 #/µm2. EMGM correctly identified 𝐾id = 4 components and found 𝐾fp = 0 false positive components for (A) and (B). EMGM correctly identified 𝐾id = 2 components and found 𝐾fp = 1 false positive component for (C). The red dots symbolize the simulated localizations. The blue/green dots symbolize the center positions of the correct/false positive components, the blue/green ellipses symbolize the 2𝜎 error ellipses of the correct/false positive components. (D) The simulated probability of obtaining a completely correct EMGM result (i.e. 𝐾id = 4 and 𝐾fp = 0) as a function of 𝑏𝑔. (E) The simulated average values of 𝐾id and 𝐾fp as a function of 𝑏𝑔. The dashed line represents the ground truth (GT) and the shaded areas the standard deviation (𝑛 = 100). 18 Supporting Material Figure S6. Influence of the localization uncertainty on the EMGM performance. (A-C) Example EMGM results for simulated Gaussian mixtures. Each mixture consists of 𝐾 = 4 components with localization uncertainty (A) 𝑠 = 0, (B) 𝑠 = 20 nm, or (C) 𝑠 = 40 nm. EMGM correctly identified 𝐾id = 4 components and found 𝐾fp = 0 false positive components for (A) and (B). EMGM correctly identified 𝐾id = 4 components and found 𝐾fp = 1 false positive component for (C). The red dots symbolize the simulated localizations. The blue/green dots symbolize the center positions of the correct/false positive components, the blue/green ellipses symbolize the 2𝜎 error ellipses of the correct/false positive components. (D) The simulated probability of obtaining a completely correct EMGM result (i.e. 𝐾id = 4 and 𝐾fp = 0) as a function of 𝑠. (E) The simulated average values of 𝐾id and 𝐾fp as a function of 𝑠. The dashed line represents the ground truth (GT) and the shaded areas the standard deviation (𝑛 = 100). 19 Supporting Material Figure S7. Influence of the number of localizations on the EMGM performance. (A-C) Example EMGM results for simulated Gaussian mixtures. Each mixture consists of 𝐾 = 4 components with localization number (A) 𝑁𝑘 = 30, (B) 𝑁𝑘 = 100, or (C) 𝑁𝑘 = 200. EMGM correctly identified 𝐾id = 2 components and found 𝐾fp = 1 false positive components for (A). EMGM correctly identified 𝐾id = 4 components and found 𝐾fp = 0 false positive component for (B) and (C). The red dots symbolize the simulated localizations. The blue/green dots symbolize the center positions of the correct/false positive components, the blue/green ellipses symbolize the 2𝜎 error ellipses of the correct/false positive components. (D) The simulated probability of obtaining a completely correct EMGM result (i.e. 𝐾id = 4 and 𝐾fp = 0) as a function of 𝑁𝑘. (E) The simulated average values of 𝐾id and 𝐾fp as a function of 𝑁𝑘. The dashed line represents the ground truth (GT) and the shaded areas the standard deviation (𝑛 = 100). 20 Supporting Material Figure S8. Illustration of a Gaussian component with standard deviation 𝜎𝑥 and 𝜎𝑦, together with the corresponding 2𝜎 error ellipse with major axis 𝑎 and minor axis 𝑏. 21 Supporting Material Figure S9. PAINT imaging of focal adhesions. (A) PAINT image of a fixed REF cell where integrin β3 was antibody stained. (B) Zoom-in of the region in (A) indicated by the white rectangle. 22 Supporting Material Supporting Tables Polymer PS(units)-b-P2VP(units) PDI Polymer concentration [mg/ml] PS1056-b-P2VP671 1.09 5 2.5 Spinning speed 2000 rpm 6000 rpm Distance on glass [nm] 56  9 119  11 Table S1. Details concerning the block polymers and the spin casting processes used for the fabrication of the nano-patterned substrates. 23
1512.05431
4
1512
2016-11-18T19:51:43
A Path Integral Approach to Age Dependent Branching Processes
[ "physics.bio-ph", "math.PR", "q-bio.PE" ]
Age dependent population dynamics are frequently modeled with generalizations of the classic McKendrick-von Foerster equation. These are deterministic systems, and a stochastic generalization was recently reported in [1,2]. Here we develop a fully stochastic theory for age-structured populations via quantum field theoretical Doi-Peliti techniques. This results in a path integral formulation where birth and death events correspond to cubic and quadratic interaction terms. This formalism allows us to efficiently recapitulate the results in [1,2], exemplifying the utility of Doi-Peliti methods. Furthermore, we find that the path integral formulation for age-structured moments has an exact perturbative expansion that explicitly relates to the hereditary structure between correlated individuals. These methods are then generalized with a binary fission model of cell division.
physics.bio-ph
physics
A Path Integral Approach to Age Dependent Branching Processes Chris D. Greenman School of Computing Sciences, University of East Anglia, Norwich, UK, NR4 7TJ Abstract. Age dependent population dynamics are frequently modeled with generalizations of the classic McKendrick-von Foerster equation. These are deterministic systems, and a stochastic generalization was recently reported in [1, 2]. Here we develop a fully stochastic theory for age-structured populations via quantum field theoretical Doi-Peliti techniques. This results in a path integral formulation where birth and death events correspond to cubic and quadratic interaction terms. This formalism allows us to efficiently recapitulate the results in [1, 2], exemplifying the utility of Doi-Peliti methods. Furthermore, we find that the path integral formulation for age-structured moments has an exact perturbative expansion that explicitly relates to the hereditary structure between correlated individuals. These methods are then generalized with a binary fission model of cell division. Keywords: Doi-Peliti, Second Quantization, Path Integral, Birth-Death Process, Age-Structure 6 1 0 2 v o N 8 1 ] h p - o i b . s c i s y h p [ 4 v 1 3 4 5 0 . 2 1 5 1 : v i X r a Contents 1 Introduction 2 Age-Structured Dynamics 2.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Simple Budding Birth and Death Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Numerics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 Doi Peliti Second Quantization Methods 3.1 Machinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Kinetic Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Age-Structured Moments 4 Path Integral Representation 4.1 Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Perturbative Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 Binary Fission and Death Processes 5.1 Kinetic Equations 5.2 Path Integrals and Correlation Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 Conclusions Appendix A: Resolution of the Identity Appendix B: Alternative Path Integral Formulation Appendix C: Correlation Functions up to Second Order Appendix D: Correlation Function Equations for Fission Appendix E: Feynman Diagram Summary Appendix F: Exact Approach to Correlation Functions 2 3 3 4 7 7 7 9 10 11 11 12 15 15 17 20 21 22 23 24 25 27 1. Introduction Populations are dynamic entities that exhibit stochastic variation through time; the number of individuals in the population fluctuates. Examples range from microscopic populations of cells to evolutionary processes of large multicellular organisms [3], and models of the underlying processes find a range of application. Historically these models have been deterministic, such as with Malthusian exponential growth. However, for populations of smaller volume, or for those going through a bottleneck, fluctuations can have significant effects on the dynamics and stochastic properties need to be considered. These can be modeled with branching processes, such as portrayed in Fig. 1A, where we see a population growing through time, starting from a single individual. The differing nature of branching processes can be seen to significantly influence dynamic effects. At I, for example, we have a simple (or budding) mode of birth, such as that observed in yeast cell division [4, 5], where the parental cell survives the creation of an offspring cell. At II we have binary fission, more commonly associated with cell division, where the parental cell effectively terminates at the moment two identical twin daughter cells are created. Fig. 1B contains a full example of this, which will later be examined in detail. At III we see a more general fission process, with 2 four daughter cells. Along with such birth processes, there are also death processes, such as at IV, which can also be viewed as fission with zero progeny. Simpler models of branching processes generally assume dynamics only depend upon the current population size. However, the age of individuals is an important factor in population growth, with fecundity often a function of age. Cell division is most likely to occur after the cell has had sufficient time to mature and enter the mitosis phase, for example. In terms of Fig. 1, incorporating age effects is equivalent to assuming the lengths of the branches affect dynamics. The incorporation of age into population dynamics was initially described by the McKendrick-von Foerster equation [6, 7], where a partial differential equation approach was used to describe the evolution of a continuous age structure. More recent developments utilize the machinery of semi-groups and have included other effects such as spatial dependence and cell division processes that depend upon the cells size. More details of these methods and their applications can be found in [8]. Leslie matrices [9, 10] are a discretized version of the process modeled by the McKendrick-von Foerster equation. These linear models have certain algebraic advantages and have proved popular in ecological modeling [11]. However, these models ignore the stochastic fluctuations in population size. The Bellman-Harris process went some way in addressing this problem [12], providing an integral equation for the sample size generating function. Although this has an implicit dependence on the underlying age dependence, solving this equation does not reveal the age-structure for the system. It is also designed for fission events (such as events II, III or IV in Fig. 1B) and does not cater for simple birth (e.g., event I ). More recent fully stochastic approaches have emerged that utilize Martingale techniques [13–16] and adapt methods from gas kinetics [1, 2]. Temporal fluctuations in population number have parallels with particle physics, where the number of subatomic particles can change due to interactions arising from fundamental forces. This parallel was first noted by Doi [17, 18], who adapted time-ordered perturbative expansion techniques to chemical reaction processes. An extensive review of this approach applied to a range of problems, including spatial effects, can be found in [19]. Peliti constructed a path integral formulation, developing lattice methods for birth death processes [20]. Although these methods are generally bosonic, where lattice sites can have multiple occupation, they have also been extended for exclusive occupation (fermionic) processes [21]. A recent review of path integral techniques for master equations can be found in [22]. Significant work in the development of renormalization techniques has also taken place [23–25]. Recent pedagogic expositions can be found in [26, 27]. To date there has been no application of these methods to age dependent processes. In this work we address this, and show that Doi-Peliti methods offer an efficient approach to such problems. In the next section we give an outline of age structured population dynamics. Section 3 details how Doi-Peliti formalism can be adapted to age-structured problems. In particular, it is used to derive kinetic equations for correlation functions, and full age-structure-population-size probability densities. Section 4 derives path integral solutions to these equations, obtaining an exact perturbative expansion for correlation functions. This expansion can be directly related to the hereditary structure of correlated individuals and is our main result. Section 5 generalizes these methods to binary fission models of cell population dynamics. Conclusions complete the study. 2. Age-Structured Dynamics 2.1. Background X(q, t), where(cid:82) The majority of methods analyzing the age structure of a population consider a distribution of the form Ω X(q, t)dq represents the number of individuals from the population of interest with age in the set Ω. The backbone of these modeling techniques is the McKendrick-von Foerster equation [6–8] ∂X(q, t) ∂t + ∂X(q, t) ∂q = −µ(q)X(q, t), (1) where µ(q) represents the death rate per individual of age q. In survival statistics this is also referred to as the hazard function, where µ(q)dt represents the probability that an individual with current age q will subsequently die within an infinitesimal amount of time dt. The death rate can also be written as µ(q) = m(q) 1−M (q) , where m(q) is the probability density function for the waiting time until death occurs, with cumulative density function M (q). The survival function can thus be written as S(q) = 1 − M (q) = 3 Birth Death Time t1 t2 Time Figure 1. (A) A generalized branching process. At I we have a budding (or simple) mode of birth, where the parental individual continues to exist after birth. At II we have binary fission, where the parental individual is terminated at the moment two identical twin daughters are created. The number of progeny in fission can vary, such as at III where four daughters are born, or the death process at IV, which can be viewed as fission with no progeny. (B) A binary fission process typically seen in cell division. At time t1 we have two sets of twins, each pair having identical ages. At time t2 one has died and one has divided, leaving a single pair of twins, and two singletons with distinct ages. e−(cid:82) q 0 µ(u)du, representing the probability that an individuals time of death is greater than q. More details of the relationships between these terms can be found in [28]. For the McKendrick-von Foerster equation there is also an associated boundary condition representing the arrival of newborn individuals of age q = 0, (cid:90) ∞ X(q = 0, t) = β(u)X(u, t)du, (2) 0 where β(q) represents the birth rate per parental individual of age q. The birth rate β(q) = b(q) 1−B(q) also has corresponding waiting time distributions b(q) and B(q). This formulation of the McKendrick- von Foerster equation is for the simple mode of birth, such as the case labelled I in Fig. 1. An initial condition X(q, t = 0) = g(q) describing the age distribution of founder individuals completely specifies the mathematical model. However, from the McKendrick-von Foerster equation we find the total population(cid:82) ∞ 0 X(q, t)dq grows deterministically and the stochastic fluctuations in population size are not incorporated. The natural way to model such variation is with the standard forward continuous-time master equation [29, 30]: ∂ρn(t) ∂t = −n [βn(t) + µn(t)] ρn(t) + (n − 1)βn−1(t)ρn−1(t) + (n + 1)µn+1(t)ρn+1(t), (3) where ρn(t) is the probability of population size n at time t, and βn(t) and µn(t) are the birth and death rates, per individual, respectively. Each of these rates can be population size and time dependent. This master equation does not consider age effects, however, and methods that incorporate both stochastic fluctuations and age are required. The models to do this will differ depending on the microscopic behavior of the corresponding branching process. We next develop the theory for simple budding birth such as at event I in Fig. 1A. More complex modes of birth such as binary fission at event II shall later be developed in Section 5. 2.2. Simple Budding Birth and Death Processes We consider the following stochastic process describing a population of n individuals at time t. Each individual of age p is represented by A(p). We have a death process A(p) → φ occurring at rate µn(p) that 4 AIIIIIIIVB can depend upon age p and population size n. We also have a budding birth process A(p) → A(p) + A(0) giving rise to a new individual of age zero, along with the surviving parent, occurring at a rate βn(p). To represent this process requires some terminology. The vector qn = (q1, q2, . . . , qn) denotes the ages of the population of size n. The term ρn(qn; t)dqn represents the probability that the population is of size n and, if the n individuals are randomly labeled 1, 2, . . . , n, the ith individual has age in the interval [qi, qi + dqi], at time t. The probability density ρn is thus a symmetric function of the age arguments qn. The marginal probability densities are defined as ρ(m) We also introduce the term fn(qn; t) as the probability density for the case where the ages are ordered, q1 ≤ q2 ≤ . . . ≤ qn. The age domain can be extended to Rn by defining fn(qn; t) = fn(π(qn); t), where π is the permutation that orders the components of the vector qn. Note that fn ≡ n!ρn. With this new domain, fn is no longer a probability density, however, many expressions are simpler to express with fn and we shall interchange between the two functions. n (qm; t) =(cid:82) dq(cid:48) n−m ρn(qm, q(cid:48) n−m; t). n≥m(n)mρ(m) The final term of interest considered is the correlation functions Xm(qm; t), where Xm(qm; t)dqm = n (qm; t) represents the probability that we can find m individuals in the population and label them 1, 2, . . . , m such that the ith has age in [qi, qi + dqi]. Here (n)m = n(n − 1) . . . (n − (m − 1)) denotes the Pochhammer symbol. (cid:80) More details about this representation can be found in [1, 2], where the probability density function ρn(qn; t) and correlation function Xn(qn; t) are examined in greater depth, which we summarize below. These results shall later be rederived using Doi-Peliti methods in Section 3. For the density ρn(qn; t) the following hierarchy of equations was established: n(cid:88) ∂ρn(qn; t) ∂t + ∂ρn(qn; t) = −ρn(qn; t) ∂qj j=1 n(cid:88) (cid:90) ∞ γn(qi) i=1 + (n + 1) along with boundary condition nρn(qn−1, 0; t) = ρn−1(qn−1; t) 0 n−1(cid:88) i=1 βn−1(qi). µn+1(y)ρn+1(qn, y; t)dy, (4) (5) (6) (7) Note that the birth and death rates βn(q) and µn(q) can be functions of both population size n and age q, and we have used the event rate γn(q) = βn(q) + µn(q). This system was solved analytically in [1, 2] for the cases of either pure birth (µn(q) = 0) or pure death (βn(q) = 0). Furthermore, provided the rates βn(q) = βn and µn(q) = µn are independent of age, the population size density ρ(0) shown to reduce to master Eq. 3. n (t) =(cid:82) dqnρn(q; t) was For the mth order correlation function Xm(qm; t), provided the birth and death rates rates βn(q) = β(q) and µn(q) = µ(q) are independent of population size, the following equation was obtained: m(cid:88) j=1 ∂Xm ∂t + with boundary condition m(cid:88) i=1 ∂Xm ∂qj + Xm µn(qi) = 0, Xm(qm−1, 0; t) = Xm−1(qm−1; t) m−1(cid:88) i=1 β(qi) + (cid:90) ∞ 0 Xm(qm−1, y; t)β(y)dy. For the mean-field case (m = 1), this equation is precisely the McKendrick-von Foerster Eq. 1, and we find the correlation function satisfies a natural generalization. Now, the hierarchy for the full probability densities given by Eq. 4 and 5 express ρn in terms of both lower and higher order terms ρn−1 and ρn+1. This mirrors the BBGKY hierarchies seen in gas kinetics [31, 32], which are notoriously difficult to solve, and approximate closure schemes are often employed [33, 34]. 5 Figure 2. An age structured budding birth death process. (Ai) Poisson distributed initial population size with rate parameter 5. (Aii) Birth and death rates dependency upon age (see Section 2.3 for specifics). (Bi-iv) Population densities for a birth death process at times T = 0, 4, 8, 14. Distributions for individual generations are highlighted. (Ci) Plots of N[0,q] (the number of individuals younger than age q) at time T = 4 are plotted for 1000 simulated birth death processes. Each color is one simulation. The two solid thick lines are the simulation mean (blue) and E(N[0,q]) (black). (Cii) Standard deviation of simulations in Ci and STD(N[0,q]) (see Eq. 9). 6 BiiiiiiivAiiiCiii However, we note that the hierarchy for the correlation functions given in Eq. 6 and 7 expresses Xm just in terms of the lower order term Xm−1 suggesting that this hierarchy is easier to analyze (by back substitution for example, starting from the solution X1 to the McKendrick-von Foerster equation). This analytic simplification is something that we will also see reflected in the path integral techniques applied in Section 4, where we will find Xm easier to analyze than ρn. Finally, we consider age-structured population-size variance. We first let NΩ denote the random variable counting the number of individuals with age inside Ω. We can then relate Eq. 6 to the variance Var(NΩ) as follows. Var(NΩ) = Var Ndu = Cov(Ndu, Ndv) (cid:32)(cid:88) du∈Ω (cid:88) du∈Ω (cid:90) (cid:88) (cid:88) du,dv∈Ω du(cid:54)=dv∈Ω (cid:33) (cid:90) = Var(Ndu) + Cov(Ndu, Ndv), (8) Bernoulli variable indexing whether an individual exists in du. Thus N 2 X1(u)du. Also, E(Ndu)2 is vanishingly small, and we find (cid:80) Now, a sufficiently small interval du will contain the age of at most one individual, and so Ndu is simply a du) = E(Ndu) = Ω X1(u)du. We similarly note that E(NduNdv) = X2(u, v)dudv is the probability of finding a pair of individuals with ages in intervals du, dv. Thus we finally obtain the following, correcting an error in [2]: du∈Ω Var(Ndu) = (cid:82) du = Ndu and so E(N 2 (X2(u, v) − X1(u)X1(v)) dudv. (9) Var(NΩ) = 2.3. Numerics X1(u)du + Ω Ω2 We next exemplify the simple budding birth death process of Fig. 2 with a comparison of simulations to analytic expression for the mean-field and variance of the process. We suppose the death waiting time probability density function m(q) = Γ(k)−1θ−kqk−1e−q/θ is gamma distributed with mean 4 and variance 1 (k = 16, θ = 1 4 ). We suppose the birth rate β(q) = cqzµ(q), where c = 1.2 and z = 0.2. The resultant densities b(q) and m(q) are plotted in Fig. 2Aii. Note that the birth rate is higher than the death rate and a mean increase in population is expected. We suppose the initial population size at time t = 0 is Poisson distributed with mean 5 (Fig. 2Ai), where each founder individual has an age that is gamma distributed with mean 1 and variance 1 4 ; Fig. 2Bi). In Fig. 2B we see the mean field solution X1(q; t) to the McKendrick-von Foerster equation. We have the age distribution of the generation zero founder individuals (2Bi). After a time t = 4 (2Bii) this peak has significantly reduced due to death, and moved to the right, due to aging. The birth process has resulted in a new peak of individuals from the next generation. By time t = 14 (2Biv)we see that the majority of individuals are from generations three and four, which significantly overlap in age. Note that the total population (area under black curve) has significantly increased from the initial case t = 0. We see from the simulations in 2C that stochastic fluctuations lead to significant differences from the mean field, although the mean and variance of the simulations can be seen to match the solutions to Eq.s 1 and 9 (2Cii). 4 (k = 4, θ = 1 This leaves open a two-fold problem. Firstly, how to solve Eq. 6 for the correlation functions Xm(qm; t) (the boundary conditions in Eq. 7 complicate the derivation of an explicit solution; see [2]). Secondly, how to derive the density fn(qn; t). To attack these problems we will use quantum field theoretic techniques, where we find that these two terms have very similar path integral formulations. In particular, we shall show that correlation functions restricted by generation numbers (such as the mean field densities in Fig. 2B) and hereditary structure arise naturally in path integral perturbative expansions. 3. Doi Peliti Second Quantization Methods 3.1. Machinery Here, we introduce a Doi-Peliti operator formalism tailored to age-dependent birth-death processes. The [17, 18]. This was based upon a state vector use of field theoretic techniques was initiated by Doi 7 qn(cid:105) ≡ q1, q2,··· , qn(cid:105) used to represent a set of n objects, with corresponding properties q1, q2,··· , qn, such as the coordinates of n molecules for example. However, we interpret them as the ages of n individuals. In this representation individuals are indistinguishable and the order of the components qi is immaterial. We use φ(cid:105) to represent the 'vacuum' state consisting of an empty population. Next, we introduce annihilation and creation operators ψq and ψ† q which satisfy standard commutation relations: [ψq, ψ† [ψq, ψp] = [ψ† q, ψ† States can be constructed using creation operators: commutation relations, we obtain the normalization δ(qi − pπ(i)), p] = δ(q − p), (cid:88) (cid:104)qmpn(cid:105) = δmn n(cid:89) π∈Sm i=1 p] = 0. qn(cid:105) = ψ† q1 ψ† q2 ··· ψ† qn (10) φ(cid:105). From these states and (11) where (cid:104)φφ(cid:105) ≡ 1 and Sm is the symmetry group of permutations on m symbols. The annihilation operators are assumed to kill the vacuum state; ψq φ(cid:105) = 0. Note that although we will use this formalism to model (positive) ages, we place no restrictions on the states qn(cid:105) which can contain negative entries. It is relatively straightforward to use Eq. 11 to verify the following resolution of the identity operator: Next, probabilty is introduced into the system by defining the following superposition of states: qm(cid:105)(cid:104)qm . ∞(cid:88) m=0 I = (cid:90) dqm (cid:90) ∞(cid:88) m! n=0 f (t)(cid:105) = fn(qn; t)qn(cid:105) . dqn n! Rn (12) (13) This is the representation used in Doi [17, 18], where the distribution fn(qn; t) ≡ n!ρn(qn; t) is equivalent to the one mentioned in the previous section. The utility of this representation is manifest when the evolution of the state can be described linearly as, f (t)(cid:105) = ζ f (t)(cid:105) , ∂ ∂t (14) (cid:90) (cid:90) with a suitable operator ζ. For our application ζ = ζ0 + ζb + ζd can be decomposed into three parts. The term ζ0 describes the increase of all age variables in time, the term ζb represents the increase in population size due to birth and ζd represents the decrease in population size due to death. Then we have, following Doi methodology [17, 18], the following expressions: ζ0 = dq ψ† q ∂ ∂q ψq, ζb = dq β(q)(ψ† qψq − ψ† qψ † 0ψq), ζd = dq µ(q)(ψ† qψq − ψq). (15) † 0, representing the birth of a new Thus for example, the second term in ζb contains creation operator ψ individual of age zero, and the annihilation and creation operators ψq and ψ† q are a bookkeeping measure that preserves the parental individual of age q. These states and operators are given in the Schrodinger representation where the operators are constant and the states vary in time, with Eq. 14 having a formal solution of the form f (t)(cid:105) = e−ζt f (0)(cid:105) . (16) (cid:90) We complete this section by introducing functional coherent states. If we take any complex function u(q) of the real value q, with complex conjugate u∗(q), we can construct the the following state superposition: u(q1)··· u(qn)ψ† q1 ··· ψ† qn φ(cid:105) . (17) (cid:82) dq u(q)ψ† q φ(cid:105) = u(cid:105) = e ∞(cid:88) (cid:90) dqn n! n=0 These coherent states satisfy the following eigenstate property: 8 (cid:90) (cid:90) ∞(cid:88) ∞(cid:88) n=0 n=0 1 n! 1 n! ψp u(cid:105) = = dqn u(q1)··· u(qn)ψpψ† ··· ψ† qn q1 dqn u(q1)··· u(qn) δ(p − qi) n(cid:88) i=1 φ(cid:105) (cid:89) j(cid:54)=i ψ† qj Coherent states also satisfy the following normalization property, (cid:82) dq u∗ψq e (cid:82) dq u(cid:48)ψ† q φ(cid:105) = (cid:104)φ e (cid:82) dq u∗u(cid:48) (cid:82) dq u(cid:48)ψ† e (cid:104)uu(cid:48)(cid:105) = (cid:104)φ e φ(cid:105) = u(p)u(cid:105) . (cid:82) dq u∗ψq φ(cid:105) = e q e (18) (cid:82) dq u∗u(cid:48) , (19) where we have used the Baker-Campbell-Hausdorff theorem to commute operators [35, 36]. The functional coherent states u(q)(cid:105) generalize the constant coherent states z(cid:105) used by Doi [17, 18], which can be recovered by setting the function u(q) ≡ z to be constant. Doi noticed the form z(cid:105) allows many summary statistics of interest to be simply expressed. In particular, the correlation density Xm(qm; t) is given by the following expectation of the number operator ψ† ··· ψ† ψq1 ··· ψqn : q1 qn q1 qm ··· ψ† Xm(qm; t) = (cid:104)1ψ† ψq1 ··· ψqmf (t)(cid:105) = (cid:104)1ψq1 ··· ψqm e−ζtf (0)(cid:105) . (20) Note that in this expression (cid:104)1 represents a coherent state with function u(q) ≡ 1 (rather than a state representing a single individual of age 1). We have also used the fact that (cid:104)1 is a left eigenstate of ψ† with eigenvalue 1, as seen in Eq. 18. Peliti [20] adapted these methods for birth death processes, developing path integral techniques rather than the time-ordered perturbative methods of Doi [17, 18]. In particular, a state vector n(cid:105) was used to represent a population of n individuals, and a functional mapping n(cid:105) ≡ zn utilized to construct a path integral formulation for properties of interest, with lattice methods used to incorporate spatial effects. The more general coherent states we have introduced will allow us to directly construct a path integral formulation for our problem without recourse to lattice discretization methods. qi 3.2. Kinetic Equations We now use the formalism outlined above to derive equations for the distribution fn(qn; t). Specifically, we utilize the field theoretic framework to derive the hierarchy given in Eq.s 4 and 5. If we condition upon an initial distribution fn(qn; 0), then from Eq.s 11 and 13 we find that fn(qn; t) is the projection of the state f (t)(cid:105) onto the fundamental state qn(cid:105): fn(qn; t) = (cid:104)qnf (t)(cid:105) = (cid:104)qne−ζtf (0)(cid:105) . (21) Since both (cid:104)qn and f (0)(cid:105) are constant in time, we can differentiate Eq. 21 with respect to time to find ∂fn(qn; t) ∂t = −(cid:104)qnζf (t)(cid:105) = −(cid:104)qnζ0f (t)(cid:105) − (cid:104)qnζbf (t)(cid:105) − (cid:104)qnζdf (t)(cid:105) . (22) Next we use the commutation relations to calculate the left action of the operators ζ0, ζb and ζd upon (cid:104)qn. The first term gives i=1 9 (cid:104)qnζ0f (t)(cid:105) = = (cid:90) dp(cid:104)φψq1ψq2 ··· ψqn ψ† n(cid:88) (cid:104)φ(cid:89) p ∂ ∂qi i=1 j(cid:54)=i ψqj ψqif (t)(cid:105) = ψpf (t)(cid:105) = ∂ ∂p n(cid:88) (cid:90) dp(cid:104)φ n(cid:88) n(cid:88) i=1 (cid:89) j(cid:54)=i ψqj ∂ ∂p ψpf (t)(cid:105) δ(p − qi) (cid:104)qnf (t)(cid:105) = ∂ ∂qi ∂ ∂qi fn(qn; t). (23) i=1 For the birth term, a similar derivation yields (cid:104)qnζbf (t)(cid:105) = = (cid:90) dp β(p)(cid:104)φψq1ψq2 ··· ψqn (ψ† n(cid:88) β(qi) fn(qn; t) −(cid:88) (cid:90) j(cid:54)=i i=1 n(cid:88) pψp − ψ† pψ † 0ψp)f (t)(cid:105) δ(qj)fn−1(q(−j) n ; t)  . where q(−j) is omitted. Finally, the death term yields n = (q1,··· , qj−1, qj+1,··· , qn) represents the age-chart with all n ages except the jth one, which (cid:104)qnζdf (t)(cid:105) = µ(qi)fn(qn; t) − dp µ(p)fn+1(qn, p; t). i=1 Upon combining these results for any strictly positive age chart qn we lose the delta functions in Eq. 24 and (using ρn ≡ fn/n!) rediscover Eq. 4. Furthermore, integrating Eq. 22 with respect to the variable qj over a small interval containing the boundary qj = 0 captures a delta function from Eq. 24 and recovers the boundary condition in Eq. 5. The field theoretic methods thus provide an efficient means of deriving kinetic equations for the population distribution function fn(qn; t), and thus, a complete stochastic description of the population size and age structure for the entire population. 3.3. Age-Structured Moments We now derive kinetic equations for the age structured correlation densities Xm(qm; t) of the age-structure. Consider first the case X1(q; t) ≡ X(q; t). Then differentiating Eq. 20 with respect to time, with m = 1, we find = −(cid:104)1ψqζf (t)(cid:105) = −(cid:104)1ψq(ζ0 + ζb + ζd)f (t)(cid:105) . ∂X ∂t This yields three terms on the right-hand side that can be written in the forms (cid:90) (cid:90) (cid:104)1ψqζ0f (t)(cid:105) = (cid:104)1ψq ψpf (t)(cid:105) = p dpψ† ∂ ∂p (cid:104)1ψqf (t)(cid:105) + (cid:90) = ∂ ∂q (cid:104)1ψqζbf (t)(cid:105) = (cid:104)1ψq dp β(p) ψpf (t)(cid:105) + (cid:104)1ψqψpf (t)(cid:105) = dp (cid:104)1δ(q − p) ∂ ∂p (cid:90) (cid:105)f (t)(cid:105) = β(q)X(q) + ∂X(q) ∂q + dp ∂ ∂p (cid:104) (cid:90) pψp − ψ† ψ† pψ † 0ψp (cid:90) dp (cid:104)1ψ† pψq ψpf (t)(cid:105) ∂ ∂p dp (cid:90) ∂ ∂p X2(q, p), dp β(p)(cid:104)1ψqψpf (t)(cid:105) − β(q)X(q) − δ(q) dp β(p)(cid:104)1ψpf (t)(cid:105) − dp β(p)(cid:104)1ψqψpf (t)(cid:105) (24) (25) (26) (27) (28) (cid:90) (cid:90) (cid:90) (cid:90) and (cid:104)1ψqζdf (t)(cid:105) = (cid:104)1ψq = −δ(q) dp β(p)X(p), (cid:3)f (t)(cid:105) pψp − ψp dp µ(p)(cid:2)ψ† (cid:90) = µ(q)X(q) + dp µ(p)(cid:104)1ψqψpf (t)(cid:105) − (cid:90) dp µ(p)(cid:104)1ψqψpf (t)(cid:105) = µ(q)X(q). (29) Then assuming X2(q, p) vanishes for the extreme values of p, the last term in Eq. 27 vanishes, yielding the PDE (cid:90) ∂X ∂t + ∂X ∂q + µ(q)X = δ(q) dp β(p)X(p). 10 (30) For any q > 0 we lose the delta function and recover the McKendrick-von Foerster Eq. 1, as expected. If we integrate across a vanishing small interval containing q = 0, we also recover the McKendrick-von Foerster boundary condition of Eq. 2. Higher order correlations Xm(qm; t) obey the following equation, which can be derived in much the same way as Eq. 30; by differentiating Eq. 20 with respect to time and calculating the resulting matrix elements, the details of which are left to the reader: m(cid:88) n(cid:88) (cid:90) ∂Xm ∂t + ∂Xm ∂qi + Xm µ(qi) = δ(qi) i=1 i=1 m(cid:88) i=1 dp β(p)Xm(q(−i) m , p; t) + Xm−1(q(−i) m ; t) β(qj) (31)  . (cid:88) j(cid:54)=i This equation is equivalent to Eq. 6 and its associated boundary condition, and we find field theoretic methods provide a natural way to describe the correlation functions for age-structured populations. 4. Path Integral Representation The path integral formulation of quantum mechanics works well in part because the fundamental position and momentum states are eigenstates of terms in Hamiltonians corresponding to many fields of interest. These fundamental states are then use to construct resolutions of the identity, which are applied between time slices across the time period of interest, resulting in path integrals [37, 38]. This technique will not work for the systems we consider. Specifically, we have fundamental states qn(cid:105) that represent the age- charts of populations of size n. However, the 'Hamiltonians' ζ we consider are functions of creation and annihilation operators, and the states qn(cid:105) are not eigenstates for these operators. Creation operators increase the minimum occupation number for any state superposition indicating that an eigenstate will not exist. However, eigenstates exist for annihilation operators; the coherent states we have seen in Eq. 18. To construct a path integral, we thus need a resolution of the identity in terms of coherent states. There are two possible approaches. One generalizes that of Peliti [20] and is the approach we take, as detailed in Appendix A. The other approach adapts techniques more commonly found in quantum field theoretic applications, as detailed in Appendix B, along with an explanation why two path integral formulations are possible. 4.1. Construction We have, then, the following path integral resolution of the identity, (cid:90) I = (cid:90) DuDv e−i(cid:82) dq u(q)v(q) iv(cid:105)(cid:104)u , (cid:90)  Q(cid:89) DuDv ≡ lim  → 0 Q → ∞ 2π q = −Q ∆q =  d[u(q)]d[v(q)]. where the functional integrals over u and v are over real functions such that (32) (33) We next construct a path integral representation of the amplitude between two coherent states, using the resolution of the identity from Eq. 32 at the time slices. To do this we first obtain matrix elements for the evolution operators ζ0, ζb and ζd between coherent states (cid:104)u and iv(cid:105). We find, using the eigenvalue properties of Eq. 18, that (cid:90) (cid:90) dq u (cid:104)uζ0iv(cid:105) = (cid:104)uiv(cid:105) i ∂v ∂q (cid:104)uζbiv(cid:105) = (cid:104)uiv(cid:105) i(1 − u(0)) (cid:104)uζdiv(cid:105) = (cid:104)uiv(cid:105) i dq µv(u − 1). , (cid:90) dq βuv, 11 (34) (cid:21)(cid:27) . (35) (cid:21) (cid:90) (cid:27) For small time interval , and using the normalization property of Eq. 19, we thus find + (1 − u(0))βuv + µv(u − 1) (cid:104)ue−ζiv(cid:105) = exp −i + u dq (cid:26) (cid:90) (cid:20) − uv  (cid:90) (cid:90) (cid:26) −i (cid:20) (cid:18) ∂v ∂q + ∂v ∂t ∂v ∂q (cid:19) Taking a product of such time slices over a time interval [0, T ], we obtain the following path integral formulation, where we have used integration by parts on the time differential term: (cid:104)uTe−ζTiv0(cid:105) = DuDv exp dqdt u + µv(u − 1) + βuv (1 − u0) + i dq uT vT .(36) Here u(q, t) and v(q, t) are now real functions of age and time, and we introduce shorthand notation u0 = u(0, t), u0 = u(q, 0), uT = u(q, T ), v0 = v(q, 0) and vT = v(q, T ). Now the two features of interest we investigate with this framework are the correlation density function Xm(qm; T ) = (cid:104)1ψq1 . . . ψqme−ζTf (0)(cid:105) and the density function fm(qm; t) = (cid:104)qme−ζTf (0)(cid:105) = (cid:104)φψq1 . . . ψqm e−ζTf (0)(cid:105). In both cases we need to specify the initial population distribution state f (0)(cid:105). The formalism is most compact when this is a coherent state. Specifically, we assume that the initial population size is Poisson distributed with parameter α, such that each individual has an initial age distribution ω(q). We then find that f (0)(cid:105) = e−α αω(cid:105). (cid:104)1ivT(cid:105) = ei(cid:82) dq vT , (cid:104)u0f (0)(cid:105) = eα(cid:82) dq ω(u0−1) and (cid:104)φivT(cid:105) = 1. Now using Eqs 17 and 19, we obtain the following identities (37) information in Eq. 37, the eigenstate property ψq1 . . . ψqm ivT(cid:105) =(cid:81)m Then using resolutions of the identity at time points 0 and T , placed at the right and left sides of operator e−ζT in Eq. 36, respectively, we obtain the following expressions, where we have utilized the boundary j=1 [ivT (qj)]ivT(cid:105), and finally the field shift u → u + 1 to transform the integrals, giving: (cid:21)(cid:27) (cid:20) (cid:26) (cid:90) (cid:90) fm(qm; T ) = DuDv [ivT (qj)] exp dqdt u + + µv − β(u + 1)u0v · and (cid:90) Xm(qm; T ) = m(cid:89) j=1 DuDv dq u0v0 + α dq ωu0 − i dq vT (cid:18) ∂v ∂q ∂v ∂t (cid:90) (cid:20) (cid:18) ∂v (cid:27) ∂q −i (cid:90) (cid:26) −i (cid:90) (cid:90) (cid:19) (cid:27) (cid:19) , [ivT (qj)] exp dqdt u + ∂v ∂t + µv m(cid:89) (cid:26) (cid:90) j=1 exp −i (cid:26) (cid:90) exp −i (38) (cid:21)(cid:27) · − β(u + 1)u0v dq u0v0 + α dq ωu0 . (39) Although the two expressions are almost identical and both fm(qm; t) and Xm(qm; t) can be examined by perturbative methods, it is only the latter expansion that readily combines into a simpler form. 4.2. Perturbative Expansion The path integral in Eq. 39 contains two interaction terms, the quadratic death term µuv and the cubic birth term βv(u + 1)u(0, t). We calculate this by expanding the birth term and the initial term α(cid:82) dq ωu0. (cid:27) The death term is quadratic and so can be dealt with directly. We are then in a position to use standard methods, and require the following propagator G(p, τ ; p(cid:48), τ(cid:48)), which we derive using generating functional techniques (see [39] for more discussion of such methods): (cid:19)(cid:21) (cid:26) (cid:20) (cid:90) (cid:90) (cid:90) G(p, τ ; p(cid:48), τ(cid:48)) = DuDv u(p, τ )iv(p(cid:48), τ(cid:48)) exp −i dqdt u + + µv − i (cid:18) ∂v ∂q ∂v ∂t dq u0v0 (40) = iδ2Z(J, K) δJ(p; τ )δK(p(cid:48); τ(cid:48)) (cid:12)(cid:12)(cid:12)(cid:12)J≡K≡0 , 12 where we introduce indeterminate functions J = J(q, t) and K = K(q, t) along with generating functional Z(J, K) = DuDv exp dq dt u + µv + J (cid:90) (cid:26) −i (cid:20) (cid:18) ∂v (cid:26) ∂q −i + (cid:90) ∂v ∂t (cid:19)(cid:21)(cid:27) (cid:90) · exp dq u0v0 + i dq dt vK (cid:27) . (41) (cid:90) (cid:90) t 0 Now the functional integral over variable u forces the constraint ∂v integrating over u0 forces the constraint v0 = 0, which we solve to give, ∂q + ∂v ∂t + µv + J = 0 and functionally v(q, t) = − s µ(q−t+x)dxds, J(q − t + s, s)e−(cid:82) t (cid:26) (cid:90) (cid:27) Z(J, K) = exp i dq dt v(q, t)K(q, t) . so the generating functional is given by (42) (43) Performing the functional differentiations of Eq. 40 then results in propagator G(p, τ ; p(cid:48), τ(cid:48)) = δ((p(cid:48) − τ(cid:48)) − (p − τ ))S(p, p(cid:48))θ(τ(cid:48) − τ ). (44) The delta function in this expression ensures an individual of age p at time τ has age p(cid:48) at time τ(cid:48), the survival function S(p, p(cid:48)) = exp ensures the individual does not die between the two times, and the theta function ensures end point τ(cid:48) is later then the initial time τ of the interval being propagated across. (cid:110)−(cid:82) p(cid:48) p µ(x)dx (cid:111) Finally we observe that the birth term β(u + 1)u0v can be decomposed into two possible terms. Firstly, we can have βuu0v, where the terminal variable v at the end of one propagator gives rise to variables u (representing parental individuals) and u0 (representing new born individuals) at the beginning of two propagators. Alternatively, we can have term βu0v, meaning we only get a single new propagator with initiating variable u0. This means that there are two possible types of internal vertex in the corresponding Feynman diagram, of degrees three and two, respectively, as indicated in Fig. 3A. In summary, then, to any Feynman diagram associated with the expansion of Xm(qm; t), we associate the following. We have a certain number of initiating nodes, each associated with variable αω. We have internal nodes from the two types mentioned above, each associated with a birth rate function β(p), where p is the age of the parent giving birth. We have m terminating nodes. Propagators bridging the nodes then complete the picture. More explicit details of the terms and resulting integral can be found in Appendix E. These integrals can be simplified with the aid of Laplace transforms, as we now demonstrate with an example.Consider the pair correlation example in Fig. 3. Here we have a Feynman diagram (3B) containing five birth terms, occurring at times t1, T − q2, τ1, τ2 and T − q1, and one initiating event at age r. We also have a corresponding Age-Time picture (3C). Although the Feynman diagram is a connected graph, the Age-Time representation is not; any propagated edge associated with a variable u0 is representing a newborn individual, and the difference in age between parent and offspring results in a corresponding discontinuity. However, the advantage of this representation is that we can now see the hereditary structure between correlated individuals; the pair correlation in Fig. 3C is between ages of individuals in generations five and two, such that the generation five individuals are descendants of generation two individuals. This leaves us with an integral of the following form, where we define Y (q) = β(q)S(q), with S(q) shorthand for S(0, q): X (2,0,3) 2 (q1, q2; T ) = αS(q1)S(q2) dr dt1 Y (r + t1)Y (T − q2 − t1) · 0 0 dτ1dτ2 β(τ1 − [T − q1])Y (τ2 − τ1)Y ([T − q1] − τ2) (45) ∞(cid:90) T−q2(cid:90) ω(r) S(r) (cid:90) T−q2<τ1<τ2<T−q1 13 u αω(p) v β(p) u0 v β(p) uu0 v Time q1 T q2 T − q1 τ2 τ1 T − q2 t1 0 r Age Figure 3. Feynman diagram and associated Age-Time diagram. (A) Classes of node in Feynman diagram; one initiating, one terminating, and two internal nodes. (B) Feynman diagram consisting of one initiating node, two terminating nodes, four degree two nodes and one degree three node. (C) The associated Age-Time diagram. Individuals of age q1 are generation five individuals, descending from generation two individuals of age q2. Now both the t1 and τ1, τ2 integrals are convolutions, so can be calculated with the aid of Laplace transformations (L) and their inverses (L−1 t ) to give the following: ∞(cid:90) X (2,0,3) 2 (q1, q2) = αS(q1)S(q2)L−1 q2−q1 (L(β)L(Y )2) dr ω(r) S(r) L−1 T−q2 (L(Y (t + r))L(Y )) (46) 0 2 Thus we have a pair correlation function corresponding to a specific hereditary structure (the superscript of X (2,0,3) (q1, q2; T ) labels the hereditary structure, which is explained in Appendix C). If the larger generation number is modified, say from five to seven, this is equivalent to adding another two birth events between times T − q2 and T − q1 in Fig. 3C. This will induce the change of L(Y )2 to L(Y )4 in Eq. 46, the rest of the solution taking the same form. Similarly, adding more birth events between times 0 and T − q2 will increase the power of the rightmost L(Y ) term, which again can be dealt with by Laplace transforms. To extend the correlation to all generations such that individuals with age q1 descend from individuals with age q2 (irrespective of generation number) requires summing over all possible generation numbers. This is 14 BCA achieved by replacing the L(Y ) terms with(cid:80)∞ L(Y )k = (1− L(Y ))−1. Provided the Laplace convolutions are possible, the pair correlation is still a single integral over variable r and the high dimensionality of the problem can be avoided. k=0 For more general hereditary structures, the same techniques can be applied, and it is the complexity of the hereditary structure that dictates the degree of integration involved, not the number of generations. Note that the number of different hereditary structures is limited; the internal nodes either preserve or increase the number of propagators through time. Thus any diagram for a paired correlation can only have one or two initiating nodes, for example. This summability and restricted choice of diagram is somewhat analogous to the summability of the 'parquet' diagrams observed in [23]. The full set of correlation functions up to second order can be found in Appendix C, where the results have also been implemented for the example given in Fig. 2. function by the single exponent i(cid:82) dp vT . Although one can expand this term and perform a similar series We note that the path integral for the density function fm(qm; t) in Eq. 38 differs from the correlation expansion to above, this results in a significant increase in the number of terms and it is not clear whether the resulting series can be summed exactly or approximated easily, and finding efficient means of determining the probability density function remains an open problem. Finally, we note that the action in Eq. 38 and 39 are identical, and linear in v. This suggests a direct functional integration with respect to v might yield a direct evaluation for fm(qm; t) or Xm(qm; t) without recourse to perturbative expansion. Although integration with respect to v can be done, resulting in a constraint upon u, which can furthermore be solved, the result requires mth order functional derivatives which soon become intractable and an exact approach without perturbative expansion is still lacking. An outline of this approach is given in Appendix F. 5. Binary Fission and Death Processes We now extend these methods to binary fission, a more general process where two individuals simultaneously arise at the moment the parental individual terminates, a process exemplified by cell division. This process is represented in Fig. 1B, where a single cell undergoes a process of cell division and death. Note that at time t1, after three cell divisions, we have two pairs of twins, each pair having identical ages. By time t2, one twin has died and one has divided, leaving one pair of twins, and two lone individuals with distinct ages, which we refer to as singletons. Fission is thus a degenerate process because a subset of ages are duplicated, and we will show that the ages of singletons and twins need to be tracked separately. We shall first consider mean-field behavior, then the full probability density for the population-size-age- chart, and finally use path integrals to determine the correlation functions. 5.1. Kinetic Equations We first consider a single species model, before establishing that the age degeneracy is better handled with a two species model. We thus have a death process A(p) → φ occurring at a rate µ(p) and a fission process of the form A(p) → A(0) + A(0) occurring at a rate β(p), where A(p) represents a single cell of age p. apply the number operator ψ† the density (cid:104)1ψ† Note that age degeneracy is automatically handled in field theoretic formalism. For example, if we rψr to the degenerate state p, p, q(cid:105), where one age p is duplicated, we obtain rψrp, p, q(cid:105) = 2δ(r − p) + δ(r − q) and find the duplicated age is correctly counted. of twins, and ψp represents termination of the parent. For the fission process, the operators ζ0 and ζd are identical to those in Eq. 15, but the birth operator † 0 operators in the latter term account for the birth To study the mean field behavior, we define the mean density X(p) = (cid:104)1ψpe−ζtf (0)(cid:105) and differentiate with respect to time, following the methodology used in earlier sections. The derivation is largely the same as that for Eq. 30, resulting in a McKendrick-von Foerster-like equation, † 0ψp], where the two ψ becomes ζb =(cid:82) dpβ(p)[ψ† pψp − ψ † 0ψ ∂X ∂t + ∂X ∂p = −γ(p)X. (47) 15 Note that the difference between this microscopic model and that encapsulated in Eq. 30 is that in fission both birth from, and death to, an individual results in their termination. This is reflected in the right-hand side of Eq. 47, where the event rate γ(p) = β(p) + µ(p) is used. The boundary condition is also modified to account for the duplicated offspring: (cid:90) X(0) = 2 dpβ(p)X(p). (48) Although this formalism works for the mean-field, if we attempt to take the temporal derivative of the pair correlation X(p, q) = (cid:104)1ψpψqe−ζtf (0)(cid:105), the following PDE and boundary condition results, ∂X ∂t + ∂X ∂p ∂X ∂q + (cid:90) = −(γ(p) + γ(q))X, (cid:90) X(0, q) = 2 drβ(r)X(q, r) + 2δ(q) drβ(r)X(0, q, r). (49) Note that the boundary condition contains an extra term (compared to Eq. 48) implicating the third correlation function X(0, p, r) and a delta function δ(q). These terms arise due to the fact that pairs of individuals can have the same age, and are difficult to deal with analytically. To combat this degeneracy, we adopt a multi-species Doi-Peliti paradigm. We treat the two classes (singletons and twins) as separate species, where A(p) represents a singleton cell of age p and B(p) represents a pair of twin cells with identical ages p. Then a singleton undergoes a death process A(p) → φ at rate µ(p) and a fission process A(p) → B(0) at rate β(p), whereas twins undergo a death process B(p) → A(p) at a rate 2µ(p) and a fission process B(p) → A(p) + B(0) at a rate 2β(p). We thus have two pairs of creation and annihilation operators; ψ† p, χp, respectively. We then represent m singletons by age-chart pm and n sets of twins by age-chart qn, resulting in a general state of the form p, ψp and χ† pm; qn(cid:105) ≡ ψ† p1 ··· ψ† pm χ† q1 ··· χ† qn φ(cid:105) . (50) Here the action of the singleton (resp. twin) operator is the same, irrespective of the current twin (singleton) p pm; qn(cid:105), q p, pm; qn(cid:105) = χ† p pm; q, qn(cid:105) = p, pm; q, qn(cid:105) = χ† state. Thus, for example, ψ† p, χ† and we find the two classes of operators commute with each other (e.g., [ψ† q] = 0). Within each operator † † q(cid:48)] = δ(q − q(cid:48)). p(cid:48)] = δ(p − p(cid:48)) and [χq, χ class, the usual commutation relations apply; [ψp, ψ q pm; qn(cid:105) = ψ† qψ† pχ† The evolution operators take on the following form: (cid:90) (cid:90) (cid:90) (cid:90) ψp + pψp − χ ψ† pψp − ψp dp χ† p ∂ ∂p (cid:105) (cid:3) + 2 † 0ψp p ∂ ∂p dp ψ† (cid:104) dp µ(p)(cid:2)ψ† dp β(p) χp, + 2 (cid:90) ζ0 = ζb = ζd = (cid:90) (cid:104) pχp − χ χ† dp µ(p)(cid:2)χ† pχp − ψ† dp β(p) pχp † 0ψ† (cid:3) . pχp (cid:105) , (51) These operators reflect the microscopic fission process, generalizing the operators in Eq. 15 that represent † 0ψ† pχp in ζb represents the simpler, non-fission budding birth-death process. For example, the last term χ the event that one individual from a pair of twins divides into a newborn pair of twins; the operator χp † 0 represents the creation of newborn represents the annihilation of the pair of twins of age p, the term χ twins of age zero, and the creation operator ψ† p represents the surviving singleton of age p. The coefficient 2 reflects the fact that either twin of age p can be annihilated. singleton density A(p) = (cid:104)1ψ† (cid:82) dpψ† We next derive the singleton-twin specific mean field equations given in [2] as follows. First we define pχpf (t)(cid:105), along with the coherent state pψpf (t)(cid:105) and twin density B(p) = (cid:104)1χ† (cid:82) dqχ† (52) Then differentiating A(p) = (cid:104)1ψpe−ζtf (0)(cid:105) and B(p) = (cid:104)1χpe−ζtf (0)(cid:105) with respect to time, using the commutator relations and eigenstate properties of 1(cid:105), we find 1(cid:105) ≡ e q φ(cid:105) . p e 16 , ∂A ∂r (cid:104)1ψrζ0f (t)(cid:105) = (cid:104)1ψrζdf (t)(cid:105) = µ(r)(A − 2B), (cid:104)1ψrζbf (t)(cid:105) = β(r)(A − 2B), , ∂B ∂r (cid:104)1χrζ0f (t)(cid:105) = (cid:104)1χrζdf (t)(cid:105) = 2µ(r)B, (cid:104)1χrζbf (t)(cid:105) = 2β(r)B − δ(r) (cid:90) dp β(p)(A + 2B), (53) which result in the following equations and boundary conditions: ∂A ∂t + ∂A ∂r A(0) = 0, = −(β(r) + µ(r))(A − 2B), ∂B ∂t + ∂B ∂r (cid:90) = −2(β(r) + µ(r))B, dp (A + 2B). (54) B(0) = This approach can also be used to derive higher order correlation functions, which can be found in Appendix D. Note that unlike Eq. 49, in this multispecies formulation there are no awkward delta functions in the boundary condition (viz. Eq. D.2). These mean-field equations agree with the system in [2] (where solutions can also be found), although unlike [2], the derivation above does not require the kinetic equation of the full probability density (Eq. 56), which we derive next. These two formulations of mean field analysis also serve to show that different Doi-Peliti models can reveal different levels of detail from the same underlying stochastic process. Now 1 m!n! fm,n(pm; qn; t) is the associated full probability density which has the state representation f (t)(cid:105) = fm,n(pm; qn; t)pm; qn(cid:105) . (55) ∞(cid:88) (cid:90) dpm dqn m,n=0 m!n! We differentiate fm,n(pm; qn; t) = (cid:104)pm; qne−ζtf (0)(cid:105) with respect to time and derive bulk equations in the much same manner as Section 3.2 to yield (β(pi) + µ(pi)) + 2 n(cid:88) j=1 (β(qj) + µ(qj))   m(cid:88) m(cid:88) i=1 i=1 dr µ(r)fm+1,n(pm, r; qn; t) + 2 µ(pi)fm−1,n+1(p(−i) m ; qn, pi; t), (56) fm+1,n(pm, 0; qn; t) = 0, fm,n+1(pm; qn, 0; t) = dr β(r)fm+1,n−1(pm, r; qn; t) m(cid:88) + 2 β(pi)fm−1,n+1(p(−i) m ; qn, pi; t). (57) These results agree with the equations obtained via a probabilistic derivation in [2]. j=1 5.2. Path Integrals and Correlation Functions We next define correlation function Xm,n(qm; ¯qn; t)dqmd¯qn to be the probability that there are m singletons that can be labeled such that the ith has age in interval [qi, qi +dqi], and n pairs of twins that can be arranged to have ages in intervals [¯qi, ¯qi + d¯qi]. In all that follows, a variable covered by a bar corresponds to a twin property, whereas an unbarred variable refers to singletons. There are two approaches to investigate the correlation functions. Firstly, we can derive kinetic equations for correlation functions in much the same way that Eq. 54, D.1 and D.2 were derived, which we can then attempt to solve. However, the equations 17 m(cid:88) (cid:90) i=1 ∂ ∂t fm,n + + ∂ ∂pi fm,n + fm,n = −fm,n ∂ ∂qj n(cid:88) j=1 (cid:90) for higher order correlations quickly become complicated (see Appendix D). The second approach develops a suitable path integral representation, which we now consider. To do this we require functional coherent states; along with a resolution of the identity; u, ¯u(cid:105) ≡ e p e (cid:82) dq ¯u(q)χ† (cid:82) dp u(p)ψ† DuD¯uDvD¯v e−i(cid:82) dq (uv+¯u¯v) iv, i¯v(cid:105)(cid:104)u, ¯u . q φ(cid:105) , (cid:90) I = Now if we assume that the initial number of singletons and pairs of twins are Poisson distributed with means α and ¯α, and the ages of these individuals have probability densities ω(q) and ¯ω(q), respectively, then by using the resolution of the identity between time slices, the correlation function can be written as the following path integral, in much the same way as Eq. 39 to give the following: Xm,n(qm; ¯qn; T ) = DuD¯uDvD¯v ivT (qi) m(cid:89) (cid:20) (cid:20) u j=1 (cid:18) ∂v (cid:18) ∂¯v ∂q ∂q n(cid:89) k=1 ∂v ∂t ∂¯v ∂t + + dqdt dqdt ¯u (cid:90) (cid:90) (cid:90) (cid:90) (cid:26) (cid:26) (cid:26) −i −i −i exp exp exp i¯vT (¯qi) · + γ(v − 2¯v) (cid:19) (cid:19) − β ¯u0v + 2γ¯v − 2β(u + 1)¯u0¯v · (cid:21)(cid:27) (cid:21)(cid:27) (cid:27) · (cid:90) (cid:90) (58) (59) (62) (63) terms ¯α(cid:82) dq ωu0 and ¯α(cid:82) dq ¯ω¯u0. We again split the (latter) birth term giving three types of birth events, Now, as before, we expand the birth terms β ¯u0v and 2β(u + 1)¯u0¯v, along with the initial distribution dq (u0v0 + ¯u0¯v0) + α dq ωu0 + ¯α dq ¯ω¯u0 . (60) meaning we have three types of internal nodes in the associated Feynman diagram, as indicated in Fig. 4B. The two initial distribution terms mean we have two classes of initiating nodes, and the singleton and twin class also mean there are two types of terminating node. We shall construct propagators in terms of functional derivatives of the following generating functional: Z(J, K, ¯J, ¯K) = DuD¯uDvD¯v exp dqdt u + γ(v − 2¯v) + J (cid:90) (cid:90) (cid:90) (cid:26) (cid:26) −i −i exp exp (cid:90) (cid:26) (cid:18) ∂¯v −i ∂q (cid:20) (cid:18) ∂v ∂q (cid:20) (cid:90) + ∂v ∂t (cid:19)(cid:21)(cid:27) · (cid:27) dqdt ¯u + ∂¯v ∂t + 2γ¯v + ¯J (cid:19)(cid:21)(cid:27) · dq (u0v0 + ¯u0¯v0) + i dqdt (vK + ¯v ¯K) . (61) ∂t + γ(v − 2¯v) + J = 0 and Functionally integrating over u and ¯u results in the conditions ∂v ∂t + 2γ¯v + ¯J = 0. Furthermore, integration over u0 and ¯u0 results in boundary conditions v0 = 0 and ∂q + ∂v ∂ ¯v ∂q + ∂ ¯v ¯v0 = 0, resulting in solutions: (cid:90) t (cid:90) t 0 ¯v(q, t) = − v(q, t) = − ¯J(q − t + s, s)e−(cid:82) t [J(q − t + s, s) − 2¯v(q − t + s, t)γ(q − t + s)] e−(cid:82) t s 2γ(q−t+x)dxds, s γ(q−t+x)dxds, with generating functional 0 Z(J, K, ¯J, ¯K) = exp (cid:26) (cid:90) i dqdt (vK + ¯v ¯K) . (cid:27) 18 Guv ∼ S β(q) ¯u0 v 2β(q) ¯u0 u ¯v 2β(q) ¯u0 ¯v G¯uv ∼ 2 S(1 − S) G¯u¯v ∼ S2 v ¯v u αω(q) ¯u ¯α¯ω(q) Time ¯q2 T q1 T − ¯q2 τ2 τ1 T − q1 0 r Age Figure 4. Feynman diagram and associated Age-Time diagram for fission process. (A) Three classes of propagator. (B) Three classes of internal nodes, two initiating and two terminating nodes. (C) Feynman diagram consisting of one initiating node, two terminating nodes, three degree two nodes and one degree three node. (D) The associated Age-Time diagram. Individuals of age ¯q2 are twins with generation number 4, descending from generation 1, which contain singletons of age q1. Guv(p, τ ; p(cid:48), τ(cid:48)) = Gu¯v(p, τ ; p(cid:48), τ(cid:48)) = G¯uv(p, τ ; p(cid:48), τ(cid:48)) = G¯u¯v(p, τ ; p(cid:48), τ(cid:48)) = iδ2Z(J, K, ¯J, ¯K) δJ(p; τ )δK(p(cid:48); τ(cid:48)) iδ2Z(J, K, ¯J, ¯K) δJ(p; τ )δ ¯K(p(cid:48); τ(cid:48)) iδ2Z(J, K, ¯J, ¯K) δ ¯J(p; τ )δK(p(cid:48); τ(cid:48)) iδ2Z(J, K, ¯J, ¯K) δ ¯J(p; τ )δ ¯K(p(cid:48); τ(cid:48)) (cid:12)(cid:12)(cid:12)(cid:12)J, ¯J,K, ¯K≡0 (cid:12)(cid:12)(cid:12)(cid:12)J, ¯J,K, ¯K≡0 (cid:12)(cid:12)(cid:12)(cid:12)J, ¯J,K, ¯K≡0 (cid:12)(cid:12)(cid:12)(cid:12)J, ¯J,K, ¯K≡0 We then find that the four possible propagators are: = δ((p(cid:48) − τ(cid:48)) − (p − τ )) S(p, p(cid:48))θ(τ − t), = 0, = δ((p(cid:48) − τ(cid:48)) − (p − τ ))2 S(p, p(cid:48))(1 − S(p, p(cid:48)))θ(τ − t), = δ((p(cid:48) − τ(cid:48)) − (p − τ )) S(p, p(cid:48))2θ(τ − t), (64) 19 ABCD (cid:110)−(cid:82) p(cid:48) (cid:111) (cid:110)−(cid:82) p(cid:48) (cid:111) p γ(x)dx p µ(x)dx where we have used the survival function S(p, p(cid:48)) = exp survival function S(p, p(cid:48)) = exp employed in previous sections. This is to be expected; in the simple mode of birth, where each parent can have multiple birth events, survival is just predicated on death not occurring over the interval [p, p(cid:48)], encapsulated by the death rate µ(x). In the fission process, however, birth induces death of the parent, so survival requires that both birth and death does not occur across the interval [p, p(cid:48)] prior to the birth event at time t(cid:48), hence the event rate γ(x) = β(x) + µ(x) is employed. The propagator function S2 thus represents the probability that a pair of twins do not die or give birth across an interval. The function 2 S(1 − S) represents the Binomial probability that one of a pair of twins dies or gives birth and the remaining singleton survives. . Note the distinction from the Thus we have three non-trivial propagators, resulting in three types of edges in the associated Feynman diagram, as highlighted in Fig. 4A (see also Appendix E). Consider then the Feynman diagram of Fig. 4C, which is directly mapped onto the Age-Time chart in Fig. 4D. We have a founder generation (generation number 0) of twins of age r, giving rise to generation number 1 twins at time T − q1. At time τ1 one of the twins gives birth and dies, leaving the singleton of age q1 at time T . We have two more generations giving birth; one a twin giving birth at time τ2 and one a singleton at time T − ¯q2, resulting in twins of age ¯q2 at time T . The integral corresponding to this diagram thus contributes to the paired correlation function X1,1(q1; ¯q2; t) with a term of the form: (cid:90) ∞ 0 (cid:90) (cid:90) ∞ X1,1(q1; ¯q2; T ) = ¯α S(¯q2)2 S(q1) dr ¯ω(r) S2(r, r + [T − q1])β(r + [T − q1]) · dτ1dτ2Y1(τ1 − [T − q1])Y2(τ2 − τ1)Y3([T − ¯q2] − τ2) = ¯α S(¯q2)2 S(q1)L−1 T−q1<τ1<τ2<T−¯q2 q1−¯q2 dr ω(r) S2(r, r + [T − q1])β(r + [T − q1]), (L(Y1)L(Y2)L(Y3)) · (65) 0 where we have used the shorthand notation S2 = 2 S(1 − S), Y1 = 2β S, Y2 = 2β S2 and Y3 = 2β S(1 − S). We have also used Laplace transforms to calculate the τ1, τ2 convolution integral. Note that we can place as many generations as we like between times T − q1 and T − ¯q2. These will give rise to a multiple convolution of terms where L(Y2)L(Y3) is replaced with powers of L(Y2) and L(Y3). If we t (L(Y1)(1 − (L(Y2) + L(Y3)))−1) sum over all possibilities, we just replace L−1 and find that the correlation functions can be entirely summed. Furthermore, just like the simple budding mode of birth, we find that the internal nodes of Feyman diagrams either preserve or increase the number of propagators through time, meaning that the number of initiating nodes is limited by the degrees of the correlation function, limiting the number of possible hereditary structures that need to be considered. t (L(Y1)L(Y2)L(Y3)) with L−1 6. Conclusions In this work we have used age-structured population modeling to highlight Doi-Peliti methods of population inference. In particular, we have generalized the formulation of coherent states used in [17, 18] to construct a path integral formulation similar to that of [20] without recourse to lattice methods. This technique was applied to age structured systems, firstly providing an efficient means of deriving kinetic equations for features of interest, and secondly demonstrated that perturbative expansions of path integrals can be calculated exactly, providing a means to construct age-structured correlation functions. These expansion methods do not so easily sum for probability density functions, which generally need to be approximated by truncation methods [33, 34]; finding efficient means of solving hierarchical equations such as Eq. 4 remains an open problem. it involves calculating terms such as (L(β)(1 − L(Y ))−1), which will appear in Eq. 46 once it is modified to sum over all generations. L−1 q2−q1 The Laplace inversion here will likely require a Bromwich integral and the resulting expressions can be Although summing the perturbation series is possible, 20 complicated (an example of this kind of calculation can be found in [2]). However, we see from Fig. 2B that for any given time the number of generations contributing to correlation functions are limited, and summing over a subset of generations can prove easier and just as effective. The path integral construction presented assumes that the initial population size is Poisson distributed. This naturally fits with the coherent state framework that was used and is the most efficient formulation. However, the machinery can be adapted to enable more general initial conditions, although some of the efficiency will be lost, resulting in integrals over a range of initial conditions. The age structured systems we have introduced can be considered as spatial systems where age is a convective term increasing with time. Other spatial properties, such as the molecular diffusive phenotype commonly analyzed in reaction-diffusion systems, can readily be incorporated into the models we have considered by adding suitable diffusive terms into the evolution equations, although this was not explored in this work. There is a need to develop the Doi-Peliti approach presented above to incorporate non-linear effects. For example, more realistic patterns of growth often involve population-size constraints such as a carrying capacity. This was noticed early in the development of population dynamics by Verhulst who modeled the growth deterministically with logistic functions [40]. In [1, 2] population-size (and age-dependent) birth and death rates, βn(q) and µn(q), were analyzed for a fully stochastic model, although the corresponding kinetic equations are significantly complicated by population-size dependency. For example, the mean-field equations are only equivalent to the McKendrick-von Foerster equation when the rates are independent of population size. Another non-linear effect of interest arises when fission processes are mediated by cell size rather than age [8], for which the non-linearity can manifest as chaos [41]. The second quantization techniques we have introduced need to be generalized to cater for such effects. The linear action of operators in field theory suggest that dealing with non-linear effects will be difficult, although the success of approaches in the area of quantum chaos give some cause for optimism [42]. The models we have introduced model interactions at the level of a single cell; population-size effects could be modeled by requiring that all cells interact with each other, for example. It seems unlikely that corresponding path integrals will be either simple or exact, and perturbative methods of approximation or semi-classical techniques will likely be required [43]. Exploring such avenues will be required if we are to further develop these methods in more realistic situations. However, the machinery we have introduced provides an initial framework that can efficiently deal with many age-structured models, offering an approach that avoids some of the complications presented when more classical probabilistic techniques are employed [1, 2]. Furthermore, path integral formulations of this machinery enabled explicit relationships between correlation functions and inheritance structure to be identified, results that are not obvious when applying more probabilistic approaches. Appendix A: Resolution of the Identity We outline why Eq. 32 is equivalent to the identity operator. More information about using coherent states and resolutions of the identity can be found in [44]. The path integral in Eq. 32 is understood in the following sense: P/(cid:89) (cid:90)  I = lim P → ∞  → 0 i=−P/ 2π d[u(pi)] d[v(pi)]e i=−P / u(pi)v(pi) iv(cid:105)(cid:104)u , (A.1) −i(cid:80)P / where age has been discretized over the interval [−P, P ] such that pi = i. To interpret the coherent states (cid:104)u and iv(cid:105) we need to use a discretized form of fundamental states pn(cid:105). To do this we use an occupation number state representation. For example, suppose we have a coarse grain resolution of age so that there are five possible values, p−2, p−1, p0, p1 and p2. Suppose, furthermore, that we have a discretized state p4(cid:105) = p0, p−2, p0, p1(cid:105) comprising the ages of four individuals. Then we write p4(cid:105) ≡ 1, 0, 2, 1, 0(cid:105)0, where the subscript 0 indicates an occupation number state representation. Two of the four individuals have middle age p0, and so the middle occupation number is 2 in the occupation state. Then we expand and discretize a general coherent state g(cid:105) as: 21  P/(cid:88) ∞(cid:88) n n! n φ(cid:105) g(cid:105) = e (cid:82) dp g(p)ψp φ(cid:105) = e(cid:80) (cid:18)n ∞(cid:88) where we use the notation(cid:0)n n n! n=0 = n! i g(pi)ψpi φ(cid:105) = g(pi)ψpi (cid:19) P/(cid:89) x n=0 i=−P/ i=−P/ g(pi)xi x(cid:105)0 , (cid:88) {x:(cid:80) (cid:1) ≡ i xi=n} x−P /!···xP /! . Then if we similarly define x! ≡ x−P/!··· xP/! and use Eq. ∞(cid:88) (A.2) 1 A.2 to obtain a discretized version of iv(cid:105)(cid:104)u, Eq. A.1 can be written as x I = lim P → ∞  → 0 · x!y!  i xi = n j yj = m d[u(pi)] d[v(pi)]e−iu(pi)v(pi)u(pi)xi(iv(p))yi (cid:21) y(cid:105)0 (cid:104)x0 . (A.3) δ(u) = m! δmn, (A.4) (A.5) (cid:1) possible states pm(cid:105). After (A.6) m,n=0 P/(cid:89) i=−P/ 2π m+n (cid:88)  x, y :(cid:80) (cid:80) (cid:20)(cid:90)  (cid:90) m (cid:88) {x:(cid:80) du um ∞(cid:88) Now, integration by parts establishes the following identity [20, 45]: (cid:90) du dv 2π e−iuvum(iv)n = (cid:19)n (cid:18) − ∂ ∂u which can be used to simplify Eq. A.3, giving I = lim P → ∞  → 0 Each occupancy vector x(cid:105)0 with total occupation number m corresponds to(cid:0)m i xi=m} m=0 x x(cid:105)0 (cid:104)x0 . 1 x! taking the continuum limit we find ∞(cid:88) (cid:90) dpm m=0 m! I = pm(cid:105)(cid:104)pm . Thus we have recovered Eq. 12; a standard resolution of the identity, as required. Appendix B: Alternative Path Integral Formulation The fundamental property that enables the path integral formulation given in Eq. 36 is the resolution of the identity described in Appendix A. This relies on the fundamental integral representation of the Kronecker delta function given in Eq. A.4, which was also the formulation used by Peliti [20]. However, there also exists the following expression, which can be established by converting to polar coordinates and integrating the subsequent gamma distribution: e−zz∗ zmz∗n = m!δmn. (B.1) In a manner largely identical to that in Appendix A, this can be used to construct a resolution of the identity more commonly seen in quantum mechanics: (cid:90) dz dz∗ (cid:90) 2πi I = (B.2) Df Df∗e−(cid:82) dpf∗(p)f (p) f(cid:105)(cid:104)f . 22 (cid:104)fζ0g(cid:105) = (cid:104)fg(cid:105) dp f∗(p) (cid:104)fζbg(cid:105) = (cid:104)fg(cid:105) (1 − f∗(0)) (cid:104)fζdg(cid:105) = (cid:104)fg(cid:105) ∂g ∂p (cid:90) dp µ(p)g(p)(f∗(p) − 1). (p), dp β(p)f∗(p)g(p), (cid:90) (cid:90) (cid:26) (cid:90) (cid:90) (cid:20) − f∗g (cid:26)  (cid:90) (cid:26)(cid:90) (cid:21)(cid:27) (B.3) , (B.4) (cid:27) · In much the same way as the derivation of Eq. 34, we can use eigenstate properties to determine the action of operators ζ0, ζb and ζd upon coherent states f(cid:105) and g(cid:105), where f and g are possibly complex functions, to give the following: We then use these to construct the incremental term (cid:104)fe−ζg(cid:105) = exp − dp + f∗ ∂g ∂p + (1 − f∗(0))βf∗g + µg(f∗ − 1) and thus arrive at the following path integral formulation: (cid:104)fTe−ζTf0(cid:105) = Df Df∗ exp − dpdt [f∗(ft + fp) + β(1 − f∗(0, t))f f∗ + µf (f∗ − 1)] exp dpf∗(p, T )f (p, T ) . (B.5) (cid:27) This provides an alternative path integral formulation to Eq. representation any further. 36, although we did not explore this Appendix C: Correlation Functions up to Second Order Here we use the path integral expansion to construct correlation functions up to order 2 for the process described in Section 2.3. We let g(t, k, θ) = θ−kΓ(k)−1tk−1e−t/θ denote the gamma probability density function and G(t, k, θ) the 1−G(q,k,θ) and birth rate β(q) = cqzµ(q). 4 . We also had Poisson 0 xa−1(1 − x)b−1dx denote the incomplete beta function, we let Y (t) = β(t)S(0, t), corresponding cumulative function. We assume death rate µ(q) = g(q,k,θ) For the implementation in Fig. 2 we used values c = 1.2, z = 0.2, k = 16 and θ = 1 distributed initial population size with mean 5, and initial age distribution ω(q) = g(q, 4, 1 We let I(t, a, b) =(cid:82) t 4 ). and define C = cΓ(z+k) θzΓ(k) . Then we obtain the following convolution: (cid:26) C ng(t + r, (z + k)n, θ)I( t Y (t + r), t+r , (z + k)(n − 1), z + k), n > 1, (C.1) n = 1. Rn(r, t) = L−1 t (L(Y (t+r))L(Y (t))n−1) = Note also that R0(0, t) = δ(t). Then we find that: We let X ((cid:96)) 1 (q, t) denote the mean density of the (cid:96)th generation individuals (as plotted in Fig. 2B). X ((cid:96)) 1 (q, T ) = α(cid:82) ∞ αω(q − T )S(q − T, q), 0 dr ω(r) S(0,r) R(cid:96)(r, T − q), (cid:96) = 0, (cid:96) ≥ 1, (C.2) (cid:40) Now if we let X ((cid:96),m) (q1, q2; T ) denote the pair correlation function such that ages q1 and q2 correspond to individuals in generations (cid:96) and m, respectively, such that both descend from distinct founder individuals (i.e. generation 0), then the Feynman diagram consists of two disconnected components and we find the simple relationship 2 X ((cid:96),m) 2 (q1, q2; T ) = X ((cid:96)) 1 (q1; T )X (m) 1 (q2; T ). 23 (C.3) Finally, we let X (n,(cid:96),m) (q1, q2; T ) denote the pair correlation such that q1 and q2 denote the ages of a pair of individuals that descend from the same individual in generation n, such that the two correlated individuals belong to generations n + (cid:96) and n + m, respectively. Then we have the following cases. For n, (cid:96), m ≥ 1 the two individuals lie on distinct branches of a hereditary tree, neither being a founder 2 individual, and we find that: ∞(cid:90) (cid:90) min(T−q1,T−q2) [T−q1]−t (cid:90) (cid:90) [T−q2]−t dv · du X (n,(cid:96),m) 2 (q1, q2; T ) = αS(q1)S(q2) ω(r) S(r) dr dt Rn(r, t) Y ([T − q1] − t − u)Y ([T − q2] − t − v) S(min([T − q1] − t − u, [T − q2] − t − v)) 0 0 0 0 R(cid:96)−1(0, u)Rm−1(0, v). (C.4) For the case X (n,0,m) (q1, q2; T ) we find that the generation n + m individual with age q2 is a direct descendant of the generation n individual with age q1(> q2), where we obtain the following (the case for X (n,m,0) (q1, q2; T ) is symmetrically similar): 2 2 X (n,0,m) 2 (q1, q2; T ) = αS(q1)S(q2)L−1 q1−q2 (L(β)L(Rm−1(0, t))) dr ω(r) S(r) Rn(r, T − q1). (C.5) (q1, q2; T ) that the individuals with ages q1, q2 descend from a founder individual 0 For the case X (0,(cid:96),m) 2 (generation 0), we find that: X (0,(cid:96),m) 2 (q1, q2; T ) = αS(q1)S(q2) ∞(cid:90) dr ω(r) S(r)2 T−q1(cid:90) du T−q2(cid:90) dv Y (r + u)Y (r + v) S(min(r + u, r + v)) · 0 R(cid:96)−1(0, [T − q1] − u)Rm−1(0, [T − q2] − v). 0 0 (C.6) Finally, for the case X (0,0,m) generation m individual (so q1 > T > q2), we get the following (the case of X (0,(cid:96),0) similar): 2 (q1, q2; T ) that q1 is the age of a founder individual, and q2 is the age of a (q1, q2; T ) is symmetrically 2 X (0,0,m) 2 (q1, q2; T ) = αS(q1)S(q2) ω(q1 − T ) S(q1 − T ) L−1 T−q2 (L(β(q1 − T + t))L(Rm−1(0, t))). (C.7) The example described in more detail in Eq. 46 is precisely X (3,3,0) (q1, q2; T ). To obtain the analytic variance given in Fig. 2Cii, the expressions above were summed for all generations up to 5 and substituted into Eq. 9. 2 Appendix D: Correlation Function Equations for Fission We have the following equation for the correlation function obtained by differentiating Xm,n(pm; qn; t) = (cid:104)1ψp1 . . . ψpm χq1 . . . χqn e−ζtf (0)(cid:105) with respect to time and using commutation relations in much the same way as the derivation of Eq. 54: ∞(cid:90) m(cid:88) i=1 n(cid:88) j=1 ∂Xm,n ∂t + ∂Xm,n ∂pi + ∂Xm,n ∂qj + n(cid:88)  Xm,n = γ(pi) + 2 γ(qj)  m(cid:88) m(cid:88) i=1 2 i=1 24 j=1 γ(pi)Xm−1,n+1(p(−i) m ; qn, pi; t), (D.1) with boundary condition Xm,n(pm; qn−1, 0; t) = (cid:90) (cid:90) dr β(r)Xm,n(pm; qn−1, r; t) dr β(r)Xm+1,n−1(pm, r; qn−1; t) + 2 m(cid:88) + 2 β(pi)Xm−1,n(p(−i) m ; qn−1, pi; t). (D.2) i=1 Appendix E: Feynman Diagram Summary Budding Birth The construction of Feynman diagrams and corresponding Age-Time diagrams is required to represent terms arising in the expansion of Xm(qm; t) (see also Fig. 3), each term begin an integral over a product of factors. These are outlined below: • There are a number I ≤ m of initiating nodes, each associated with factor αω(pi), i = 1, ..., I, where pi • There are a number J of internal nodes, of degrees 2 or 3. Collectively, they are associated with times represents the age of a founder individual. 0 ≤ t1 ≤ t2 ≤ . . . ≤ tJ ≤ T . • There are m terminating nodes, all associated with time T . The (cid:96)th is also associated with age q(cid:96), (cid:96) = 1, 2, . . . , m. We take the first node connected below the (cid:96)th terminating node that does not involve any edge to the right of a node of degree 3. – If it is the jth internal node, we associate factor δ(q(cid:96) − (T − tj)). – If it is the ith initiating node, we associate factor δ(q(cid:96) − (T + pi)). • The Age-Time diagram is constructed as follows. Each edge connecting a pair of nodes in the Feynman diagram is associated with initial and final age-time pairings (p, t) and (p(cid:48), t(cid:48)), respectively. There is a corresponding line from (p, t) to (p(cid:48), t(cid:48)) in the Age-Time diagram (Fig. 3C). The pairings can be calculated by working from the initiating nodes, upwards, with the following observations. – Edges connected to the ith initial node are given an initial age-time pairing of (pi, 0). – If the starting node of an edge is the jth internal node, which also has degree 2, the initial age-time pairing is (0, tj). – If the starting node of an edge is the jth internal node, which also has degree 3, and the edge extends up and to the left from the node, the initial age-time pairing is (0, tj) (representing birth of a new individual). – If the starting node of an edge is the jth internal node, which also has degree 3, and the edge extends up and to the right from the node, the initial age-time pairing is (p(cid:48)(cid:48), tj), where p(cid:48)(cid:48) is the final age of the edge below the node (representing the age of the parent who gave birth to a new individual). – The final age-time pairing for an edge where the final node is the jth internal node is (p− t + tj, tj), where (p, t) is the initial age-time pairing. – Edges connected to the (cid:96)th terminating node are given a final age-time pairing (q(cid:96), T ). (cid:110)−(cid:82) p(cid:48) (cid:111) p µ(x) dx , representing survival • Each edge is associated with a propagator factor S(p, p(cid:48)) = exp from age p at the beginning of the edge, to age p(cid:48) at the end. • The jth internal node is associated with birth factor β(p(cid:48)), where p(cid:48) is the age associated with the end of the edge below the node. • All factors are multiplied and integrated with respect to the measure represents the simplex region 0 ≤ t1 ≤ t2 ≤ ··· ≤ tJ ≤ T . 25 (cid:90) (cid:90) dpI dtJ , where ∆ (R+)I ∆ Fission Birth The construction of Feynman diagrams and corresponding Age-Time diagrams is required to represent terms arising in the expansion of Xm,n(qm; ¯qn; t) (see also Fig. 4), each term begin an integral over a product of factors. These are outlined below: • There are two types of initiating nodes (I1 + I2 < m + n): – A number I1 of circular initiating nodes, , each associated with factor αω(pi), i = 1, ..., I1, where pi represents the age of a founder singleton. – A number I2 of square initiating nodes, , each associated with factor ¯α¯ω(¯pi), i = 1, ..., I2, where degree 3 (hexagons, ¯pi represents the age of a founder pair of twins. • There are a number J of internal nodes, two types of degrees 2 (squares, ). Collectively, they are associated with times 0 ≤ t1 ≤ t2 ≤ . . . ≤ tJ ≤ T . , and circles, ) and one of • There are two types of terminating nodes, all associated with time T . There are m circular nodes, associated with singletons, where the (cid:96)th is associated with age q(cid:96), (cid:96) = 1, 2, . . . , m. There are n square associated with singletons, where the (cid:96)th is associated with age ¯q(cid:96), (cid:96) = 1, 2, . . . , n. For any nodes, terminating node with corresponding age q(cid:48) we take the first node connected below that is that does not involve any edge to the right of a node of degree 3. – If it is the jth internal node, we associate factor δ(q(cid:48) − (T − tj)). – If it is an initiating node with associated age p(cid:48)(cid:48), we associate factor δ(q(cid:48) − (T + p(cid:48)(cid:48))). • There are three classes of edges bridging nodes with corresponding initial and final times t and t(cid:48) where: – Edges of the form – Edges of the form twin dies). – Edges of the form represent a pair of twins at time t that survive through to time t(cid:48). represent a pair of twins at time t that become a singleton by time t(cid:48) (one represent a singleton at time t that is a singleton at time t(cid:48) (the singleton survives). • The nodes that can be connected by these edges are restricted, depending upon whether they represent singeltons or twins. – Edges of the form or can start from any node representing the birth of twins: , , , or (note the edge is from the left side of a hexagonal node). – Edges of the form can start from any node representing singletons: , or (note the edge is from the right side of a hexagonal node). – Edges of the form – Edges of the form or can end at any node that is terminal or representing the birth from a can end at any node representing singletons: , or . twin parent: , , or • The Age-Time diagram is constructed as follows. Each edge connecting a pair of nodes in the Feynman diagram is associated with initial and final age-time pairings (p, t) and (p(cid:48), t(cid:48)), respectively. There is a corresponding line from (p, t) to (p(cid:48), t(cid:48)) in the Age-Time diagram (Fig. 4E). The pairings can be calculated by working from the initiating nodes, upwards, with the following observations. – Edges connected to the ith initial singleton or twin node are given an initial age-time pairing of (pi, 0) or (¯pi, 0), respectively. – If the starting node of an edge is the jth internal node, which also has degree 2, the initial age-time pairing is (0, tj). – If the starting node of an edge is the jth internal node, which also has degree 3, and the edge extends to the left from the node, the initial age-time pairing is (0, tj) (representing birth of a new individual). – If the starting node of an edge is the jth internal node, which also has degree 3, and the edge extends to the right from the node, the initial age-time pairing is (p(cid:48)(cid:48), tj), where p(cid:48)(cid:48) is the final age of the edge below the node (representing the age of the parent who gave birth to a new individual). 26 – The final age-time pairing for an edge where the final node is the jth internal node is (p− t + tj, tj), where (p, t) is the initial age-time pairing. – Edges connected to the (cid:96)th terminating singleton or twin node are given a final age-time pairing of (q(cid:96), T ) or (¯q(cid:96), T ), respectively. (cid:110)−(cid:82) p(cid:48) (cid:111) • Each edge is associated with a propagator factor based upon the survival function S(p, p(cid:48)) = , where p and p(cid:48) are the initial and final ages associated with the edge, as follows: exp p µ(x) dx – Edges of the form – Edges of the form a pair of twins. – Edges of the form have a propagator S(p, p(cid:48))2 representing survival of a singleton. have a propagator 2S(p, p(cid:48))(1 − S(p, p(cid:48))), representing survival of one of have a propagator S(p, p(cid:48)), representing survival of a pair of twins. • We have the following birth factors: – If the jth internal node is of the form it is associated with birth factor β(p(cid:48)), where p(cid:48) is the age associated with the end of the edge below the node. – If the jth internal node is of the form or it is associated with birth factor 2β(p(cid:48)), where p(cid:48) is (cid:90) ∆ dpI1 dpI2 dtJ , where (R+)I1+I2 the age associated with the end of the edge below the node. • All factors are multiplied and integrated with respect to the measure ∆ represents the simplex region 0 ≤ t1 ≤ t2 ≤ ··· ≤ tJ ≤ T . Appendix F: Exact Approach to Correlation Functions The path integral in Eq. 39 can be written (using integration by parts) as where Xm(qm; T ) = (cid:90) (cid:81)m δmZ(K) i=1 δK(qi; T ) (cid:26) (cid:90) (cid:26) , (cid:12)(cid:12)(cid:12)(cid:12)K≡0 (cid:20)(cid:18) ∂u (cid:90) ∂q (cid:90) (cid:19) (cid:21)(cid:27) · (F.1) (F.2) i dq dt v Z(K) = DuDv exp (cid:27) of characteristics to give an expression of the form Z(K) = exp(cid:8)α(cid:82) dq ωu0 Now functional integration over the v variable results in the constraint ∂u ∂t −µv +β(u+1)u0 +K = 0 and integration over vT results in the boundary condition uT = 0. Now this can be solved with the method + β(u + 1)u0 + K ∂q + ∂u (cid:9) where u0(q) = u(q, 0) is found dq uT vT + α − µu ∂u ∂t (cid:90) dq ωu0 . −i exp + from, u(q, t) = (cid:90) T−t (cid:40) 0 exp − (cid:90) T−t ds [K(q + T − t, s) + β(q + T − t − s)B(s)] · (cid:41) ds(cid:48) [µ(q + T − t − s(cid:48)) − β(q + T − t − s(cid:48))B(s(cid:48))] , (F.3) s and the boundary term B(t) = u(0, T − t) = u0(T − t) satisfies the integral equation obtained when substituting q → 0, t → T − t. The fact that u(q, t) and B(t) have functional dependence upon K makes the functional derivative of Eq. F.1 difficult to implement, and finding a useful way to analyze Xm(qm; t) or fm(qm; t) without perturbation remains an open problem. 27 References [1] Greenman C D and Chou T 2016 Physical Review E 93 012112 [2] Chou T and Greenman C D 2016 J. Stat. Phys. 164 [3] Charlesworth B et al. 1994 Evolution in age-structured populations (Cambridge University Press Cambridge) [4] Shcheprova Z, Baldi S, Frei S B, Gonnet G and Barral Y 2008 Nature 454 728–734 [5] Lippuner A D, Julou T and Barral Y 2014 FEMS Microbiology Reviews 38 300–325 [6] McKendrick A G 1926 Proc. Edinburgh Math. Soc. 44 98–130 [7] von Foerster H 1959 Some remarks on changing populations in The Kinetics of Cell Proliferation (Springer) [8] Webb G F 2008 Population models structured by age, size, and spatial position Structured population models in biology and epidemiology ed Magal P and Ruan S (Berlin, Heidelberg: Springer) pp 1–49 [9] Leslie P H 1945 Biometrika 33 183–212 [10] Leslie P H 1948 Biometrika 35 213–245 [11] Caswell H 2001 Matrix population models (Wiley Online Library) [12] Bellman R and Harris T E 1948 Proceedings of the National Academy of Sciences 34 601–604 [13] Jagers P and Klebaner F C 2000 Stochastic Processes and their Applications 87 235–254 [14] Hamza K, Jagers P and Klebaner F C 2016 Journal of mathematical biology 72 797–820 [15] Hong J 2011 Coalescence in Bellman-Harris and multi-type branching processes Ph.D. thesis Iowa State University [16] Hong J et al. 2013 Journal of Applied Probability 50 576–591 [17] Doi M 1976 Journal of Physics A: Mathematical and General 9 1465 [18] Doi M 1976 Journal of Physics A: Mathematical and General 9 1479 [19] Mattis D C and Glasser M L 1998 Reviews of Modern Physics 70 979 [20] Peliti L 1985 Journal de Physique 46 1469–1483 [21] Schulz M and Reineker P 2005 New Journal of Physics 7 31 [22] Weber M F and Frey E 2016 arXiv preprint arXiv:1609.02849 [23] Peliti L 1986 Journal of Physics A: Mathematical and General 19 L365 [24] Lee B P and Cardy J 1994 Physical Review E 50 R3287 [25] Tauber U C, Howard M and Vollmayr-Lee B P 2005 Journal of Physics A: Mathematical and General 38 R79 [26] Tauber U C 2012 Field theoretic methods Computational Complexity (Springer) pp 1080–1093 [27] Tauber U C 2014 Critical dynamics: a field theory approach to equilibrium and non-equilibrium scaling behavior (Cambridge University Press) [28] Collett D 2015 Modelling survival data in medical research (CRC press) [29] Kampen N G V 2011 Stochastic Processes in Physics and Chemistry North-Holland Personal Library (Elsevier Science) [30] Chou T and D'Orsogna M R 2014 First Passage Problems in Biology First-Passage Phenomena and Their Applications ed Metzler R, Oshanin G and Redner S (Singapore: World Scientific) pp 306–345 [31] McQuarrie D A 2000 Statistical Mechanics (University Science Books) [32] Zanette D H 1990 Physica A 162 414–426 [33] Raghib M, Hill N A and Dieckmann U 2011 Journal of mathematical biology 62 605–653 [34] Rogers T 2011 Journal of Statistical Mechanics: Theory and Experiment 2011 P05007 28 [35] Maggiore M 2004 A modern introduction to quantum field theory (Oxford University Press) [36] Peskin M E and Schroeder D V 1995 An introduction to quantum field theory (Westview) [37] Srednicki M 2007 Quantum field theory (Cambridge University Press) [38] Feynman R P and Hibbs A R 1965 Quantum mechanics and path integrals vol 2 (McGraw-Hill New York) [39] Mandl F and Shaw G 2010 Quantum field theory (John Wiley & Sons) [40] Verhulst P F 1838 Quetelet 10 113–121 [41] Brze´zniak Z and Dawidowicz A L 2014 On the chaotic properties of the von foerster-lasota equation Semigroup Forum vol 88 (Springer) pp 287–299 [42] Casati G and Chirikov B 2006 Quantum chaos: between order and disorder (Cambridge University Press) [43] Assaf M and Meerson B 2010 Physical Review E 81 021116 [44] Fradkin E 1991 Field theories of condensed matter systems 82 (Addison Wesley Publishing Company) [45] Gardiner C W 1985 Handbook of stochastic methods (Springer) 29
1803.10019
1
1803
2018-03-27T11:37:33
Statistical Physics of Medical Diagnostics: Study of a Probabilistic Model
[ "physics.bio-ph" ]
We study a diagnostic strategy which is based on the anticipation of the diagnostic process by simulation of the dynamical process starting from the initial findings. We show that such a strategy could result in more accurate diagnoses compared to a strategy that is solely based on the direct implications of the initial observations. We demonstrate this by employing the mean-field approximation of statistical physics to compute the posterior disease probabilities for a given subset of observed signs (symptoms) in a probabilistic model of signs and diseases. A Monte Carlo optimization algorithm is then used to maximize an objective function of the sequence of observations, which favors the more decisive observations resulting in more polarized disease probabilities. We see how the observed signs change the nature of the macroscopic (Gibbs) states of the sign and disease probability distributions. The structure of these macroscopic states in the configuration space of the variables affects the quality of any approximate inference algorithm (so the diagnostic performance) which tries to estimate the sign/disease marginal probabilities. In particular, we find that the simulation (or extrapolation) of the diagnostic process is helpful when the disease landscape is not trivial and the system undergoes a phase transition to an ordered phase.
physics.bio-ph
physics
Statistical Physics of Medical Diagnostics: Study of a Probabilistic Model Alireza Mashaghia,b,∗, Abolfazl Ramezanpoura,c aLeiden Academic Centre for Drug Research, Faculty of Mathematics and Natural Sciences, Leiden University, Leiden, The Netherlands bHarvard Medical School, Harvard University, Boston, Massachusetts, USA cDepartment of Physics, University of Neyshabur, Neyshabur, Iran and ∗[email protected] (Dated: March 28, 2018) Abstract We study a diagnostic strategy which is based on the anticipation of the diagnostic process by simulation of the dynamical process starting from the initial findings. We show that such a strategy could result in more accurate diagnoses compared to a strategy that is solely based on the direct implications of the initial observations. We demonstrate this by employing the mean- field approximation of statistical physics to compute the posterior disease probabilities for a given subset of observed signs (symptoms) in a probabilistic model of signs and diseases. A Monte Carlo optimization algorithm is then used to maximize an objective function of the sequence of observations, which favors the more decisive observations resulting in more polarized disease probabilities. We see how the observed signs change the nature of the macroscopic (Gibbs) states of the sign and disease probability distributions. The structure of these macroscopic states in the configuration space of the variables affects the quality of any approximate inference algorithm (so the diagnostic performance) which tries to estimate the sign/disease marginal probabilities. In particular, we find that the simulation (or extrapolation) of the diagnostic process is helpful when the disease landscape is not trivial and the system undergoes a phase transition to an ordered phase. 8 1 0 2 r a M 7 2 ] h p - o i b . s c i s y h p [ 1 v 9 1 0 0 1 . 3 0 8 1 : v i X r a 1 I. INTRODUCTION Statistical physics has been widely used to extract macroscopic properties of a wide range of systems from their microscopic interaction models, yet it has not been employed to medical diagnostics. Given an initial subset of observed signs (symptoms, clinical and laboratory findings) with some prior knowledge about the patient (or a complex system like a biological cell), a diagnosis problem simply asks for the most probable diseases (or macrostates with specific phenotypes) [1–3]. An efficient and accurate diagnostic procedure is important specially in the early stages of diseases, where the number and quality of medical evidences are often insufficient to reach a definite diagnosis. Here, we use approximate inference and optimization algorithms of statistical physics [4, 5] to show that a simulation (extrapolation) of the diagnostic process (without doing any real observation) could be helpful as a heuristic strategy in the study of diagnosis problems. A diagnosis problem usually starts with a (probabilistic) model of (well-defined) sign and disease variables which describes the (statistical) dependencies of the variables; such an "effective" model of the signs and diseases may come from a microscopic model of the system (human body or biological cell) with emergent macroscopic behaviors that are inter- preted as diseases. Various modeling frameworks have been developed and used in medical diagnostics: (i) probabilistic models and belief networks, (ii) neural networks and machine learning methods, and (iii) a complex network approach to the problem. Bayesian belief networks provide a probabilistic framework to study the sign-disease de- pendencies [6–10]. The belief networks are represented by tables of conditional probabilities that show how the state of a variable in an acyclic directed graph depends on the state of the parent variables. The above information along with a few simplifying assumptions then are used to infer the marginal sign and disease probabilities for a given set of findings [7]. An- other approach is to use an artificial neural network to represent the complex relationships of the sign/disease variables [11–13]. The model parameters here are obtained in a learning process using the machine learning techniques [14]. Finally, in a network approach to the problem, one constructs a (weighted) symptom-disease network with connections relating the signs to the diseases. This network along with other complementary data, e.g., gene- disease, RNA-disease, protein-disease, metabolite-disease and disease-disease networks, are then utilized as information resources by a diagnostic algorithm [15–20]. 2 In addition to the model, an efficient inference algorithm is needed to estimate the marginal sign and disease probabilities [21, 22]. When the number of initially observed signs is too small to make a diagnosis, we need to suggest a number of new medical tests to know the value of several other relevant signs. For this aim, we need an appropriate objective function and optimization algorithm to choose the more informative signs, which can lead us to the right diagnosis with a smaller number of observations. In Refs. [23, 24] we proposed probabilistic models of signs and diseases which can systematically take into ac- count the effects of different types of sign-disease, disease-disease, and sign-sign interactions; the models are indeed graphical models of the sign and disease variables with a number of interaction factors, each one connecting a small subset of disease variables to the associated sign variables [22]. We introduced approximate inference and optimization algorithms to deal with such probabilistic models, and studied the effects of the model structure and the objective function on the performances of the diagnostic algorithms. The models we consider are natural generalizations of the simpler probabilistic models studied in previous works [7, 8, 10], which usually assume that only one disease is behind the findings (exclusive diseases assumption), or, the diseases act independently on the signs (causal independence assumption). Moreover, for computational simplicity, it is usually assumed that there is no disease-disease and sign-sign interactions. In Ref. [23], we showed that such interactions can significantly improve the accuracy of diagnosis without resorting to the exclusive diseases or the causal independence assumption. In this paper, we elaborate more on the nature and behavior of the macroscopic states of the probabilistic models we introduced in the above studies. We employ the mean-field approximation to study the possible changes in the (macroscopic) state of the system as the number of observed signs increases, and to estimate the sign and disease marginal probabil- ities [25]. For the objective function, we choose a function that favors the more polarizing tests, which result in disease probabilities that are closer to zero or one [23]. This could be useful especially when the gap between the most probable diseases and the other diseases is small. Moreover, this objective function is easier to compute than a maximum-likelihood function that is typically taken in these problems. Starting from an initial set of observed signs, we use an approximate (Monte Carlo) optimization algorithm to find a sequence of candidate observations that maximizes the above objective function [24]. However, instead of the true value of the "observed" sign at each step, we assume the outcome is given by 3 the most probable value of the sign obtained from the model by the approximate inference algorithm. We show that this strategy is able to improve the quality of diagnosis compared to the case that is merely based on the direct implications of the initial findings. Interest- ingly, the improvement is observed for nontrivial cases when the system undergoes a phase transition to an ordered phase; i.e., where the effect of observed signs can propagate in the system to influence the state of the other signs (for a similar phenomenon see [26]). II. MAIN DEFINITIONS The microscopic state of the system (patient or cell) is identified by the sign values S = {Si : i = 1, . . . , NS}, where for simplicity we work with binary sings Si = ±1. The probability of being in state S is given by P (S). The probability P (S) (the model) is parametrized by a set of couplings K(t), which in general depend on real time t. The conditional probability of the unobserved signs depends on the subset of the observed signs O(t) = {j = 1 . . . , NO(t)} with the values denoted by So(t) = {Sj : j ∈ O(t)}. The macroscopic states (or phenotypes, or diseases) of the system (for large number of signs NS → ∞) can be identified by the Gibbs states of P (S) [27]. We label these macroscopic states with D, for diseases, with D = 0 representing the healthy state. The average of an unobserved sign in state D is denoted by hSiiD. A pure Gibbs state is characterized by the clustering property; i.e., hSiSjiD − hSiiDhSjiD goes exponentially to zero by the distance of sign nodes i and j in the interaction graph of the sign variables induced by P (S). A mixed state is composed of more than one pure states. In this way, the state of a disease pattern D or a cluster of similar disease patterns are represented by the statistical properties of the sign variables in the associated pure or mixed Gibbs states. We start by asking several interesting questions: • What is P (S) and how does it (or the couplings K) change with time? Here we need a dynamical model of the system to study the stochastic evolution of the sign variables. In the following, we shall assume some reasonable structures for the model and leave this problem for future studies. Instead of going from the model P (S) to the macroscopic states, we start from the diseases D and obtain the model from the joint probability of the sign and disease variables P (S; D) = P (SD)P0(D). Then, the model is obtained by summing over the disease variables P (S) = PD P (SD)P0(D). 4 The conditional probability P (SD) can be a decreasing function of the distance of S and a reference sign configuration S(D). Here the Si(D) represent the most probable symptoms of disease D. These models could be useful (in the absence of the realistic models) as benchmarks in the study of a diagnosis problem. • Do we see a significant change of behavior with time in P (S)? For example, from weak sign correlations to a regime of strong correlations. Typically, we encounter strong correlations close to a phase transition from one macroscopic state to another state. As we will see, even simple (but plausible) models of signs and diseases can exhibit both continuous and discontinuous phase transitions as the strength of the sign and disease interactions are varied. In particular, the phase transition can be induced by increasing the number of observed signs for given strength of the interactions. • How does the structure of P (S) affect the diagnosis? Here we need an efficient and approximate inference algorithm to compute the sign and disease probabilities. It is easy to obtain very good estimations of these marginal probabilities as long as there is only one macroscopic (pure) state, or there are a number of symmetry-related states. Otherwise, the above algorithms will not converge or will need a very large computation time to provide a fair sampling of the probability distribution. We will see how the convergence and quality of an approximate inference algorithm which is based on the mean-field approximation affect the diagnostic performance. Consider a set of ND binary variables D = {Da = 0, 1 : a = 1, . . . , ND}, where Da = 0, 1 shows the absence or presence of disease a. Each disease is assigned a positive weight Wa, to take into account the significance of diseases. In the following we assume the Wa are uniformly distributed in (0, 1). The joint probability distribution of the sign and disease variables (i.e., the model) is identified by P (S; D) = P (SD)P0(D). Here P0(D) is the prior probability distribution of diseases, which could depend on the patient's characteristics such as gender and age and disease properties such as duration of a disease, mortality rate and transmission rate among others. In the following, we shall assume the prior probability is factorized as P0(D) ∝ exp(Pa K 0 expected number of present diseases. aDa). The parameters K 0 a can also be used to control the Let Ptrue(SD) to be the true (or empirically estimated) probability distribution of the sign variables given the disease hypothesis D. In practice, we may have access only to a 5 small subset of marginal probabilities of the above probability distribution. For instance, suppose we are given the sign probabilities Ptrue(Sinodisease), Ptrue(Si, SjonlyDa), and Ptrue(Si, SjonlyDa, Db) conditioned on the presence of no diseases, the presence of only one disease, and the presence of only two diseases, respectively. Using the maximum entropy principle, for the conditional probability distribution of the signs we take P (SD) = 1 Z(D) φ0(S) ×Ya φa(SDa) ×Ya<b φab(SDa, Db), (1) where the partition function Z(D) is obtained from normalization. The disease interaction factors (φ0, φa, φab) can in general be parametrized by the cou- plings of all the possible multi-sign interactions. As customary in maximum entropy model- ing, assuming an exponential family, the parameters sufficient to describe the above marginal probabilities are involved in the one-sign terms (K 0 i Si), the one-disease-one-sign interactions (K a i DaSi), the one-disease-two-sign interactions (K a teractions (K ab i DaDbSi), and finally the two-disease-two-sign interactions (K ab ijDaSiSj), the two-disease-one-sign in- ij DaDbSiSj). More precisely, the disease interaction factors are given by φ0(S) = ePi K 0 i Si, φa(SDa) = eDa[Pi K a i Si+Pi<j K a ijSiSj], φab(SDa, Db) = eDaDb[Pi K ab i Si+Pi<j K ab ij SiSj ]. (2) (3) (4) Figure 1 shows the interaction graph of the sign and disease variables related by the above interaction factors. We use Ma, ka and Mab, kab for the number and connectivity of one- disease and two-disease interaction factors, respectively. In principle, the information provided by the marginal probabilities of the true (or empir- ical) probability distribution is sufficient to determine the model parameters [28–31]. Note that φ0(S) is responsible for the probability of observing S in the absence of any diseases, where the most probable value is Si = −1. It is reasonable to assume that in the healthy case each sign takes the positive value with a small probability independently of the other sign values. The joint probability distribution of the sign and disease variables can be rewritten as, P (S; D) ∝ exp(−H(S; D)), where the energy function reads as follows H(S; D) = −Xi hi(D)Si −Xi<j Jij(D)SiSj + ln Z(D) − ln P0(D). (5) 6 Here, the partition function and the new couplings are: ePi hi(D)Si+Pi<j Jij(D)SiSj , i +Xa K a i Da +Xa<b K ab i DaDb, Z(D) =XS hi(D) = K 0 Jij(D) =Xa K a ijDa +Xa<b K ab ij DaDb. (6) (7) (8) From the above model, we can extract simpler models depending one the maximum number of disease and sign variables that are involved in the interactions; for instance, we could have the D1S1 (one-disease-one-sign), D1S2 (one-disease-two-sign), D2S1 (two-disease-one-sign), and D2S2 (two-disease-two-sign) models. In the following, we consider only the D1S1 and D2S1 models, where we can exactly compute the partition function Z(D) = Qi (2 cosh hi(D)). For these models, we can also exactly compute the model parameters given the true marginal probabilities, K 0 i = K a i = K ab i = 1 2 1 2 1 2 Ptrue(Si = −1nodisease)(cid:19) , ln(cid:18) Ptrue(Si = +1nodisease) Ptrue(Si = −1onlyDa)(cid:19) − K 0 ln(cid:18) Ptrue(Si = +1onlyDa) Ptrue(Si = −1onlyDa, Db)(cid:19) − K 0 ln(cid:18) Ptrue(Si = +1onlyDa, Db) i , i − K a i − K b i . (9) (10) (11) For a given subset O of observed signs with values So, the disease probabilities are obtained from P (Da = 1So) = 1 Z(So)XD Dae−H(DSo), (12) where H(DSo) = − log L(DSo) is the log-likelihood function H(DSo) = −Xa K 0 aDa −Xi∈O So i hi(D) +Xi∈O ln (2 cosh hi(D)) , (13) and Z(So) = PD exp(−H(DSo)) is another normalization constant. As before, the prior probability distribution is P0(D) ∝ exp(Pa K 0 aDa). It is easy to show that the marginal probability of an unobserved sign is given by: P (Si = 1So) = 1 Z(So)XD (cid:18) 1 + tanh hi(D) 2 (cid:19) e−H(DSo). (14) The approximate equations for the D1S2 and D2S2 models can be found in [23]. 7 III. THE HOMOGENEOUS FULLY-CONNECTED MODELS The direct problem of inferring the marginal sign/disease probabilities from the above models can be solved exactly as long as the model parameters do not depend on the sign or disease labels. The thermodynamic limit here is defined by the limit ND, NS, NO → ∞ such that γ = ND/NS and no = NO/NS remain finite. To provide some order of magnitude, it is useful to mention that in Internist (a probabilistic model for internal diseases [7]) the number of diseases is about 500 and the number of associated signs is around 4000. In addition, we need to scale the model parameters as K a a = κ0 D, and K 0 i /ND, K ab i = κab i = κa i = κ0 i /N 2 i , K a K 0 ij/(NSND), K ab ij = κa D), a; the scaling ensures that the energy ij /(NSN 2 ij = κab function is extensive (proportional to NS). The sign and disease probabilities are obtained by minimizing the following free energy with respect to x = P (D = 1) and yu = P (S = 1) (for an unobserved sign), f (x, y) = −γS(x) − (1 − no)S( 1 + yu 2 ) + S( 1 + z(x) 2 ) − h(x)(y − z(x)) − 1 2 J(x)(y2 − z2(x)) − γκ0 ax. (15) The value of the observed signs enters in y = noyo + (1 − no)yu with yo = (Pi∈O So and yu = (Pi /∈O Si)/(NS − NO). Here z is the solution to z = tanh(h(x) + J(x)z), and S(x) = −x log x − (1 − x) log(1 − x) is the Gibbs-Shannon entropy function. Moreover, the i )/NO, effective field h(x) = κ0 ij x2 (see Appendix A for the derivations). Each local or global minimum of the free energy can be considered i x2 and the coupling J(x) = κa i + κa 2κab 2κab ijx + 1 i x + 1 as a macroscopic state of the system. Figure 2 shows how the sign and disease probabilities change with the number of observations, when all the observed signs have a positive value (see also Fig. 7 in Appendix A). As the figures show, a new macroscopic state can appear continuously or discontinuously depending on the value of the model parameters. IV. THE INHOMOGENEOUS MODELS: MEAN-FIELD APPROXIMATION In this section, we find an estimation of the sign and disease probabilities for arbitrary couplings K. To this end, we write Da = hDai + δDa and Si = hSii + δSi where the δDa = Da − hDai and δSi = Si − hSii are small deviations from the mean values. The mean- field (MF) approximation here is obtained by neglecting the second order deviations in a 8 Taylor expansion around the mean values [25]. In the following, we shall restrict ourselves to the D1S1 and D2S1 models, where the normalization function Z(D) can be computed exactly; for the D1S2 and D2S2 models we need also to compute this function within the MF approximation (see Appendix B). In this way, the MF approximation for the sign and disease probabilities are obtained by solving the self-consistency equations xa = exp(ha(x))/(1 + exp(ha(x))), with P (Da = 1) = xa and P (Si = 1) = (1 + tanh hi(x))/2. Here, the effective fields experienced by the sign and disease variables are given by K a i xa +Xa<b K ab i xaxb, ha(x) = K 0 [So i − tanh(hi(x))](K a hi(x) = K 0 i +Xa a +Xi∈O K ab i xb). (16) (17) i +Xb6=a The equations are solved by iteration starting from random initial values for the xa. The time complexity of this algorithm is of order NON 3 D in a fully-connected model. The fixed points of these equations are considered as the macroscopic states of the system (Gibbs states). As long as there is only one macroscopic state, the iteration algorithm converges easily to the single fixed point of the equations. Non-convergence of the iteration algorithm is a signature of the presence of more than one fixed point. To check the performances of the algorithms, we shall assume that the true model is given by an exponential probability distribution Ptrue(SD) ∝ exp(−βH(S; S(D))). Here S(D) gives the most probable symptoms of disease pattern D, and H(S; S(D)) is the Hamming distance (number of different elements) of the two sign configurations. Moreover, β is a positive parameter that controls the structure of the true model around the symptoms S(D); the diseases are more clearly distinguished for larger values of β. We assume that each element Si(D) (for i = 1 . . . , NS) takes the positive and negative values with equal probability, except for the healthy case (D = 0) where all the elements are negative. Given the true model, we use the true marginal probabilities to construct e.g. the D2S1 model. Suppose that we are given a subset O of NO observed signs with values So. A simple diagnostic procedure works by computing the posterior disease probabilities conditioned on the observations P (DaSo). Then, the most probable diseases or those that have a probability greater than a threshold value, are reported as the diagnosed diseases; in the following, we shall assume that the most probable diseases, within a small window of size 9 δPD = 0.01, are the present ones. Figure 3 displays the accuracy of such a diagnosis with the D1S1 and D2S1 models for a small number of sign and disease variables. The figure also shows the probability gap between the most probable disease(s) and the other diseases. A patient with disease pattern D and NO initial observed signs from the most probable symptoms S(D) is presented to the model for diagnosis; a disease pattern is chosen with a probability proportional to the weights Wa of the present diseases in D. From [23] we know that the D1S1 and D2S1 models work well so long as the number of present diseases in D, denoted by D, is less than or equal to two; that is why we choose patients with a small number of diseases. As the figure shows, we obtain more accurate predictions as the parameter β increases. The situation is different when we have to resort to an approximate inference algorithm. We see in Fig. 4 that the MF approximation does not provide accurate estimations of the sign and disease marginal probabilities for large β, where the algorithm does not converge. Here the best performances are observed for intermediate values of β. V. DIAGNOSIS BY SIMULATION OF THE DIAGNOSTIC PROCESS It may happen that the information provided by the initial number of observations are not enough to reach a reliable diagnosis, especially in the early stages of the diseases. Thus, we need a good strategy to choose the most informative signs for the next observations. Here the goal could be to reach the right diagnosis with a minimum number of the medical tests [23, 24]. Thus, for the objective function we propose an increasing function of the polarizations (deviations from the neutral value) in the posterior disease probabilities. The optimal choice then is provided by the most polarizing observation conditioned on the value of the previous observations. In contrary to the maximum likelihood function which is computationally hard to compute, the above objective function can easily be computed given the posterior disease probabilities. And, one can easily incorporate the importances of the diseases (the Wa) into the objective function, to assign more weight to polarization of the more important diseases. In a sequential diagnostic process of length T , we do the medical tests one by one and at each step we obtain the true value of the observed sign (this is called Diags-I in [24]). To obtain an optimal sequence of medical tests, one has to simulate in advance a diagnostic process of T observations without doing any real observation (this is called Diags-II in 10 [24]). This simulation of the observation process, or extrapolation from the initial set of observations, is proposed here as another heuristic approach to diagnosis to fully exploit the statistical dependencies of the variables provided by the model in addition to the initial medical tests. To this end, we use the mean-field approximation to compute the posterior disease probabilities. Then, the (Monte Carlo) optimization algorithm of Ref. [24] is used to maximize an objective functional of T observations, T Xt=1 1 Pa Wa Xa Wa(cid:12)(cid:12)(cid:12)(cid:12) 1 ! . 2(cid:12)(cid:12)(cid:12)(cid:12) E[O(T )] = P (Da = 1) − (18) Here O(T ) = {j1, . . . , jT } is the sequence of observations. One can also add the cost or relevance of the observed signs to this objective function [24]. The optimization algorithm starts from a random sequence of T observations, uses the marginal sign probabilities to generate a sequence of new observations, and accepts the suggested sequence if the objective function increases. Note that the above problem is indeed a stochastic optimization problem, where the objective function depends on the stochastic outcomes of the observations [32, 33]. To simplify the computation, we assume that the observed sign j at each time step takes the most probable value identified by the marginal probability P (SjSo) conditioned on the value of the previous observed signs. Figure 5 shows how the above objective function and the optimization algorithm perform. The figure displays the changes the first right and wrong diagnosis times compared to a random sequence of T observations [24]; the first right diagnosis time TR is the first time (number of observations) the probability of having a right disease becomes larger than a threshold value, here Pth = 0.9. Similarly we define the first wrong diagnosis time TW . In Fig. 6, we compare the accuracy of the diagnosis with the D2S1 model before and after extrapolation for T = NO/2 steps. Here the sign/disease marginal probabilities are computed by the MF approximation. Similar comparisons are shown also in Figs. 3 and 4. VI. DISCUSSION In summary, depending on the model and the strength of the model parameters, new macroscopic states can appear as the number of observed signs increases. This could be helpful because the disease probabilities are usually more informative within such states. On 11 the other hand, this affects the algorithm convergence and consequently the quality of the sign and disease probabilities which are computed by the approximate inference algorithm. More advanced and accurate algorithms can of course improve the quality of inference, but at the expense of more computational time [23]. We showed that simulation of the diagnostic process provides a useful strategy for diag- nosis when a naive approach that is based on the direct implications of the observed signs is not very helpful. In other words, this strategy works in the ordered phase of the system where the values of the observed signs significantly affect the probability distribution of the unobserved signs; the classical example is a ferromagnetic spin system in the ordered (low temperature) phase where the values of the boundary spins determine the physical (Gibbs) state of the system. In this way, we can define a critical number of initial observations which are needed to enter such an ordered state, for systems that display a phase transition. Here, for the sake of efficiency, we assumed that each "observed" sign in the simulation takes the most probable value predicted by the model. Moreover, we used a very naive optimization algorithm to find the optimal sequence of observations. A more accurate study should consider the stochastic nature of the simulated observations, and employ a more sophisticated optimization algorithm, e.g., simulated annealing. Finally, it would be inter- esting to have a microscopic (or phenomenological) model of patient (or an ensemble of patients) to study the time evolution of the sign probability distribution, and the emergent macroscopic (disease) states. VII. PERSPECTIVES An accurate medical diagnosis from a limited number of findings (e.g. at the early stages of diseases) should exploit all the statistical information on the sign/disease dependencies observed in the clinical and laboratory data. Such interdependencies are emerging due to the advancements in omics technologies and progress in population studies and aging research (e.g. identification of co-occurrence of age-related diseases). We note that the existing datasets lack the necessary probabilistic information needed for our approach, as such new data need to be generated. These studies will be the subject of our future works and in the current article, we are primarily addressing the mathematics and statistical physics communities. 12 Let us recall briefly the kind of statistical data we need to construct the models studied in this paper. First, note that these models have been obtained from an expansion around the healthy state where the number of involved diseases is small (D = 1, 2) [23]. On the other hand, given a disease hypothesis D, it is usually assumed that the sign variables are uncor- related in a zero-order approximation of the signs statistics [7]. Here, the necessary data are encoded in the conditional probabilities P (SionlyDa) (in D1S1 model) or P (SionlyDa, Db) (in D2S1 model). Obviously, we expect to have two-sign correlations, or higher-order sign correlations, even in presence of only a single disease. But taking into account these corre- lations considerably increases the computational complexity of the problem. Additionally, it is in practice very difficult to obtain statistically good clinical data which capture the higher-order correlations. Nevertheless, in the end, it is the collection of available empirical data that determines the structure of the model. The method can in principle be applied to any diagnostic problem to infer the macro- scopic state (phenotype) of the system from a limited number of evidences. This could be, for instance, the problem of assigning a state to a biological cell or a complex electronic device. In particular, assignment of state to a cell is a major challenge in immunology and cancer biology and it has complicated developing therapies for cancer and autoimmunity. We envision that our approach will be generically applied to a wide range of problems in medicine, science and technology. Appendix A: The homogeneous fully-connected models As long as the model parameters do not change with the sign or disease labels, we can write all the quantities in terms of the collective variables x = (Pa Da)/ND and y = (Pi Si)/NS. Then for large number of signs (NS → ∞), we get 1 + z(x) ln Z(D) ≈ S( ) + h(x)z(x) + J(x)z2(x), (A1) 2 1 2 1 NS 1 NS H(S; D) ≈ −h(x)(y − z(x)) − 1 2 J(x)(y2 − z2(x)) + S( 1 + z(x) 2 ) − γκ0 ax, (A2) where γ = ND/NS, and h(x) = κ0 J(x) = κa 1 2 ij x2. κab i + κa i x + 1 2 ijx + κab i x2, 13 (A3) (A4) Here we take the scaling K 0 K a i , K 0 i = κ0 1 ND 1 i = K a ij = κa ij, K ab ij = NSND a = κ0 a, i , K ab κa i = 1 N 2 D κab i , 1 NSN 2 D κab ij . Moreover, z is the solution to z = tanh(h(x) + J(x)z). which minimizes the following free energy f (z) = −S( 1 + z 2 ) − h(x)z − 1 2 J(x)z2, Here, for brevity, we defined the Shanon entropy function (A5) (A6) (A7) (A8) (A9) S(p) = −p ln p − (1 − p) ln(1 − p). (A10) To take into account the value of the observed signs, we write y = noyo + (1 − no)yu with yo = (Pi∈O So i )/NO, yu = (Pi /∈O Si)/(NS − NO), and no = NO/NS. In this way, the grand partition function is given by Z(So) ≃Z 1 0 dxZ 1 −1 dyue−NS f (x,y), (A11) At the end, the self-consistency equations for the x and yu variables in the thermodynamic limit (NS → ∞), are obtained by minimizing the following free energy f (x, y) = −γS(x) − (1 − no)S( 1 + yu 2 ) − h(x)(y − z(x)) − 1 2 J(x)(y2 − z2(x)) + S( 1 + z(x) 2 ) − γκ0 ax. (A12) Figure 7 shows how the above free energy behaves when the model parameters in the D2S1 model are varied. Appendix B: The inhomegeuous models: mean-field approximation For the D1S1 and D2S1 models we can compute some quantities exactly, therefore, we present the mean-field approximation for these models separately. 14 1. In the absence of the sign-sign interactions Here we write Da = xa + δDa where xa = hDai, and δDa = Da − xa is a small deviation from the mean value. Then, the local field experienced by sign i is hi(D) = hi(x) +Xa K a i δDa +Xa<b where K ab i (δDaxb + xaδDb + δDaδDb) = hi(x) + δhi, (B1) hi(x) = K 0 i +Xa K a i xa +Xa<b K ab i xaxb. (B2) The Hamiltonian can be written as H(DSo) = −Xa K 0 aDa −Xi∈O So i [hi(x) + δhi] +Xi∈O ln (2 cosh[hi(x) + δhi]) , (B3) Expanding the last term up to the first order deviations δDa, we get H(DSo) ≈ H0 −Xa ha(x)Da, (B4) with ha(x) = K 0 a +Xi∈O [So i − tanh(hi(x))](K a i +Xb6=a K ab i xb). (B5) Then, the average values xa are obtained by the following self-consistency equations: xa = eha(x) 1 + eha(x) , The equations are solved by iteration starting from random initial values for the xa. 2. In the presence of the sign-sign interactions In general, the partition function in the MF approximation reads Z(D) ∝Yi 2 cosh[hi(D) +Xj6=i Jij(D)zj(D)]! , where the zi are solutions to zi(D) = tanh[hi(D) +Xj6=i Jij(D)zj(D)]. 15 (B6) (B7) (B8) Define δDa = Da − hDai and δSi = Si − hSii. For brevity, we take xa = hDai and yi = hSii. Note that yi = So i is fixed for i ∈ O (the subset of observed signs). Then, to first order in the δDa and δSi, we have H(S; D) ≈ H0 −Xi /∈O [hi(x) +Xj6=i Jij(x)yj]Si −Xa ha(x, y, z)Da, (B9) where ha(x, y, z) = K 0 Ba ij(x)yiyj Ba a +Xi i (x)yi +Xi<j tanh[hi(x) +Xj6=i −Xi Here, the new introduced local fields are Jij(x)zj] × Ba i (x) +Xj6=i [Ba ij(x)zj + Jij(x)χa j ]! (B10) Ba i (x) = K a Ba ij(x) = K a K ab i xb, K ab ij xb, i +Xb6=a ij +Xb6=a and the susceptibility χa i is given by χa i = ∂zi ∂xa = (1 + tanh2[hi(x) +Xj6=i Jij(x)zj]) In summary, the mean-field equations read as follows, × Ba i (x) +Xj6=i [Ba ij(x)zj + Jij(x)χa j ]! . (B11) (B12) (B13) (B14) (B15) (B16) (B17) xa = eha(x,y,z) 1 + eha(x,y,z) , yi = So i ∈ O i , yi = tanh[hi(x) +Xj6=i zi = tanh[hi(x) +Xj6=i Jij(x)yj], i /∈ O Jij(x)zj]. We solve the equations by iteration starting from random initial values for the xa, yi(i ∈ O), zi, and the χa i . [1] Ledley, R.S., Lusted, L.B. (1959) Reasoning foundations of medical diagnosis; symbolic logic, probability, and value theory aid our understanding of how physicians reason. Science 130:9-21. 16 [2] Miller, R.A. and Geissbuhler, A. (1999). Clinical Diagnostic Decision Support Systems-An Overview, Springer New York, Page 3-34, ISBN: 978-1-4757-3903-9. [3] Papadakis, M., McPhee, S.J., Rabow, M.W. (2016). Current Medical Diagnosis and Treat- ment. 55 edition, LANGE CURRENT Series, 1920 pages, ISBN: 0071845097. [4] Hartmann, Alexander K., and Heiko Rieger, eds. New optimization algorithms in physics. John Wiley & Sons, 2006. [5] Mezard, M., and Montanari, A. (2009). Information, physics, and computation. Oxford Uni- versity Press. [6] Spielgelharter, D. J. (1987). Probabilistic Expert Systems in Medicine. Statistical Science, 2, 3-44. [7] Shwe, M. A., Middleton, B., Heckerman, D. E., Henrion, M., Horvitz, E. J., Lehmann, H. P., and Cooper, G. F. (1991). Probabilistic diagnosis using a reformulation of the INTERNIST- 1/QMR knowledge base. Methods of information in Medicine, 30(4), 241-255. [8] Heckerman, D. E., and Shortliffe, E. H. (1992). From certainty factors to belief networks. Artificial Intelligence in Medicine, 4(1), 35-52. [9] Miller, R.A. (1994). Medical diagnostic decision support systemspast, present, and future: a threaded bibliography and commentary. J Am Med Inform Assoc 1:827. [10] Nikovski, D. (2000). Constructing Bayesian networks for medical diagnosis from incomplete and partially correct statistics. Knowledge and Data Engineering, IEEE Transactions on, 12(4), 509-516. [11] Baxt, William G. "Use of an artificial neural network for data analysis in clinical decision- making: the diagnosis of acute coronary occlusion." Neural computation 2.4 (1990): 480-489. [12] Penedo, Manuel G., et al. "Computer-aided diagnosis: a neural-network-based approach to lung nodule detection." IEEE Transactions on Medical Imaging 17.6 (1998): 872-880. [13] Khan, Javed, et al. "Classification and diagnostic prediction of cancers using gene expression profiling and artificial neural networks." Nature medicine 7.6 (2001): 673. [14] Murphy, K. P. (2012). Machine learning: a probabilistic perspective. MIT press. [15] Goh, Kwang-Il, et al. "The human disease network." Proceedings of the National Academy of Sciences 104.21 (2007): 8685-8690. [16] Barabsi, Albert-Lszl, Natali Gulbahce, and Joseph Loscalzo. "Network medicine: a network- based approach to human disease." Nature reviews. Genetics 12.1 (2011): 56. 17 [17] Gustafsson, Mika, et al. "Modules, networks and systems medicine for understanding disease and aiding diagnosis." Genome medicine 6.10 (2014): 82. [18] Sun, Kai, et al. "Predicting disease associations via biological network analysis." BMC bioin- formatics 15.1 (2014): 304. [19] Liu, Wei, et al. "Integrative analysis of human protein, function and disease networks." Sci- entific reports 5 (2015). [20] Suratanee, A., and Plaimas, K. (2015). DDA: A Novel Network-Based Scoring Method to Identify DiseaseDisease Associations. Bioinform Biol Insights. 9: 175186. [21] Cooper, G. F. (1990). The computational complexity of probabilistic inference using Bayesian belief networks. Artificial intelligence, 42(2), 393-405. [22] Jordan, M. I. (2004). Graphical models. Statistical Science, 140-155. [23] Abolfazl Ramezanpour and Alireza Mashaghi. Toward First Principle Medical Diagnostics: On the Importance of Disease-Disease and Sign-Sign Interactions. Frontiers in Physics 5: 32 (2017). [24] Abolfazl Ramezanpour and Alireza Mashaghi. Uncovering a hidden disease pattern by simu- lating the clinical diagnostic process. Scientific Reports 8(1):2436 (2018). [25] Opper, Manfred, and David Saad, eds. Advanced mean field methods: Theory and practice. MIT press, 2001. [26] Mzard, Marc, and Andrea Montanari. "Reconstruction on trees and spin glass transition." Journal of statistical physics 124.6 (2006): 1317-1350. [27] Krzakaa, Florent, et al. "Gibbs states and the set of solutions of random constraint satisfaction problems." Proceedings of the National Academy of Sciences 104.25 (2007): 10318-10323. [28] Kappen, H. J., and Rodriguez, F. B. (1998). Efficient learning in Boltzmann machines using linear response theory. Neural Computation, 10(5), 1137-1156. [29] Tanaka, T. (1998). Mean-field theory of Boltzmann machine learning. Physical Review E, 58(2), 2302. [30] Ricci-Tersenghi, F. (2012). The Bethe approximation for solving the inverse Ising problem: a comparison with other inference methods. Journal of Statistical Mechanics: Theory and Experiment, 2012(08), P08015. [31] Nguyen, H. C., Zecchina, R., Berg, J. Inverse statistical problems: from the inverse Ising problem to data science. Advances in Physics, 66 (3), 197-261 (2017). 18 P(SD) P (D) 0 b a P(S0) j i α Prior Disease Interaction Sign Leak FIG. 1. The interaction graph of disease variables (left circles) and sign variables (right circles) related by Ma one-disease and Mab two-disease interaction factors (middle squares) in addition to interactions induced by the leak probability (right square) and the prior probability of diseases (left square). In general, an interaction factor α = a, ab is connected to kα signs and lα diseases [23]. [32] Birge, J. R., and Louveaux, F. (2011). Introduction to stochastic programming. Springer Science & Business Media. [33] Altarelli, F., Braunstein, A., Ramezanpour, A., and Zecchina, R. (2011). Stochastic matching problem. Physical review letters, 106(19), 190601. 19 ) 1 = D P ( ) 1 = D P ( 1 0.8 0.6 0.4 0.2 0 0 0.1 0.2 (b) κ a=κ ab=0,κ i ij ij a=2: κi κi κi ab=0 ab=2 ab=3 1 0.8 0.6 0.4 0.2 ) 1 = S P ( 0.5 0 0 0.1 0.2 0.3 0.4 0.5 a=2: (a) κ a=κ ab=0,κ i ij ij κi ab=0 κi ab=2 κi ab=3 0.4 0.3 NO/NS NO/NS 1 0.8 0.6 0.4 0.2 0 0 a=2: ab=1,κ (c) κ a=κ i ij ij κi ab=0 κi ab=2 κi ab=3 0.1 0.2 0.3 0.4 0.5 NO/NS ) 1 = S P ( 1 0.8 0.6 0.4 0.2 0 0 a=2: ab=1,κ (d) κ a=κ i ij ij κi ab=0 κi ab=2 κi ab=3 0.1 0.2 0.3 0.4 0.5 NO/NS FIG. 2. The sign and disease probabilities in the homogenuous fully-connected models vs the fraction of observed signs NO/NS . We assume that all the observed signs are positive. The prior disease probabilities and the leak sign probabilities are P0(Da = 1) = P (Si = +10) = 0.001. Panels (a),(b) show the probability of observing a positive sign P (S = 1) and the probability of having a disease P (D = 1) for the D1S1 and D2S1 models (K a ij = K ab ij = 0). Panels (c),(d) display the above probabilities for the D1S2 and D2S2 models (K a ij = K ab ij = 1). 20 ) N T + P T ( y c a r u c c a 2 1.8 1.6 1.4 1.2 1 (a) D1S1,D=1: β=0.02 β=0.05 β=0.10 β=0.02,T=NO/2 0 2 4 6 8 10 12 14 16 18 20 ) N T + P T ( y c a r u c c a 2 1.8 1.6 1.4 1.2 1 (b) D2S1,D=1: β=0.05 β=0.1 β=0.5 β=0.05,T=NO/2 0 2 4 6 8 10 12 14 16 18 20 ) N T + P T ( y c a r u c c a 2 1.8 1.6 1.4 1.2 1 (c) D2S1,D=2: β=0.1 β=1.0 β=4.0 β=1.0,T=NO/2 0 2 4 6 8 10 12 14 16 18 20 NO NO NO 0.15 0.1 0.05 0 D1S1,D=1,β=0.02: ∆p12 ∆p23 (d) 0 2 4 6 8 10 12 14 16 18 20 0.2 0.15 0.1 0.05 0 D2S1,D=1,β=0.05: ∆p12 ∆p23 (e) 0 2 4 6 8 10 12 14 16 18 20 1 0.8 0.6 0.4 0.2 0 (f) D2S1,D=2,β=1.0: ∆p12 ∆p23 0 2 4 6 8 10 12 14 16 18 20 NO NO NO FIG. 3. Comparing the accuracy (true positive plus true negative) of the diagnosis (panels a,b,c) and the relative gaps (panels d,e,f) in the disease probabiltiies (sorted by magnitude) ∆p12 = (P1 − P2)/P1 and ∆p23 = (P2 − P3)/P2 for the D1S1 and D2S1 models using an exhastive inference algorithm. We consider the cases in which only one or two diseases are present (D = 1, 2). The prior disease probabilities are chosen such that NDP0(Da = 1) = D. The model parameters are obtained from the exponential true model for different values of β. The filled circles in the top panels show the results after a simulation process of T = NO/2 steps, where NO is the initial number of the observed signs with known true values. For the model structure we take a fully- connected graph of ND = 5, NS = 20 variables with Ma = 5, Mab = 10 interaction factors, and connectivities ka = kab = 20. The data are results of 2000 independent realizations of the problem. 21 y t i l i b a b o r p e c n e g r e v n o c ) N T + P T ( y c a r u c c a 1 0.8 0.6 0.4 0.2 0 2 1.8 1.6 1.4 1.2 1 (a) D1S1,D=1:β=0.1 β=0.5 β=1.0 0 10 20 30 40 50 60 70 80 90 100 NO (d) D1S1,D=1:β=0.1 β=0.5 β=1.0 β=0.1,T=NO/2 y t i l i b a b o r p e c n e g r e v n o c ) N T + P T ( y c a r u c c a 1 0.8 0.6 0.4 0.2 0 2 1.8 1.6 1.4 1.2 1 (b) D2S1,D=1:β=0.1 β=0.5 β=1.0 0 10 20 30 40 50 60 70 80 90 100 NO (e) D2S1,D=1:β=0.1 β=0.5 β=1.0 β=1.0,T=NO/2 (c) D2S1,D=2:β=0.1 β=1.0 β=2.0 0 10 20 30 40 50 60 70 80 90 100 NO (f) D2S1,D=2:β=0.1 β=1.0 β=2.0 β=1.0,T=NO/2 y t i l i b a b o r p e c n e g r e v n o c ) N T + P T ( y c a r u c c a 1 0.8 0.6 0.4 0.2 0 2 1.8 1.6 1.4 1.2 1 0 10 20 30 40 50 0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100 NO NO NO FIG. 4. Comparing the convergence probability (panels a,b,c) and accuracy (true positive plus true negative) of the diagnosis with and without extrapolation (panels d,e,f) for the D1S1 and D2S1 models using the MF approximation. We consider the cases in which only one or two diseases are present (D = 1, 2). The prior disease probabilities are chosen such that NDP0(Da = 1) = D. The model parameters are obtained from the exponential true model. The filled circles show the results after a simulation process of T = NO/2 steps, where NO is the initial number of the observed signs with known true values. For the model structure we take a random graph of ND = 50, NS = 500 variables with Ma = 50, Mab = 500 interaction factors, and connectivities ka = kab = 400. The data are results from at least 100 independent realizations of the problem. 22 0.1 0 -0.1 4 0.1 0 -0.1 4 (a) D1S1,D=1,β=0.1: P(∆TR <0) P(∆TW <0) (b) D2S1,D=1,β=1: P(∆TR <0) P(∆TW <0) 0.6 0.5 0.4 0.3 0.2 0.1 (c) D2S1,D=2,β=1: P(∆TR <0) P(∆TW <0) 0.4 0.3 0.2 0.1 6 8 10 NO 12 14 16 0 10 20 30 40 50 60 0 10 20 30 NO 50 60 70 40 NO 1 0 -1 -2 -3 -4 -5 12 14 16 -6 10 (d) D1S1,D=1,β=0.1: ∆TR ∆TW 6 8 10 NO (e) D2S1,D=1,β=1: ∆TR ∆TW 20 30 40 50 60 NO 2 1 0 -1 -2 -3 -4 -5 10 (f) D2S1,D=2,β=1: ∆TR ∆TW 20 30 40 NO 50 60 70 FIG. 5. The improvment in the statistics of the first right and wrong diagnosis times (TR, TW ) after maximizing the objective function E[O(T )] with the D1S1 and D2S1 models using the MF approximation. The algorithm starts from a random sequence of T = NO/2 observations, where NO is the initial number of the observed signs with known true values. Panels (a,b,c) show the probabilities P (∆TR,W < 0) of decreasing the corresponding times by the algorithm. Panels (d,e,f) show the average values ∆TR;W of the changes in the corresponding times by the algorithm. We consider the cases in which only one or two diseases are present (D = 1, 2). The prior disease probabilities are chosen such that NDP0(Da = 1) = D. The model parameters are obtained from the exponential true model. For the model structure we take a random graph of ND = 50, NS = 500 variables with Ma = 50, Mab = 500 interaction factors, and connectivities ka = kab = 400. The data are results of at lesat 100 independent realizations of the problem. 23 ) P T ( e v i t i s o p e u r t ) N F ( e v i t a g e n e s l a f 0.6 0.5 0.4 0.3 0.2 0.1 0 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 ) P F ( e v i t i s o p e s l a f ) N T + P T ( y c a r u c c a 0.03 0.02 0.01 0 1.6 1.5 1.4 1.3 1.2 1.1 1 (a) D2S1,D=2,β=1: T=0 T=NO/2 80 60 40 100 20 NO (c) D2S1,D=2,β=1: T=0 T=NO/2 20 40 60 80 100 NO (b) D2S1,D=2,β=1: T=0 T=NO/2 20 40 60 80 100 NO (d) D2S1,D=2,β=1: T=0 T=NO/2 80 60 40 100 20 NO FIG. 6. Comparing the accuracy of the diagnosis with and without extrapolation for the D2S1 model using the MF approximation. We consider the cases in which only two diseases are present (D = 2). The prior disease probabilities are chosen such that NDP0(Da = 1) = D. The model parameters are obtained from the exponential true model. The filled circles show the results after a simulation process of T = NO/2 steps, where NO is the initial number of the observed signs with known true values. For the model structure we take a random graph of ND = 50, NS = 500 variables with Ma = 50, Mab = 500 interaction factors, and connectivities ka = kab = 400. The data are results of at lesat 100 independent realizations of the problem. 24 9 8 κi a=2 κi a=3 κi a=4 ) x ( f 7 6 5 0 9 8 κi a=2 κi a=3 κi a=4 ) y ( f 7 6 5 0.001 (a) no=0.1, κ ab=1 i 8.5 8 κi ab=4 κi ab=5 κi ab=6 (b) no=0.1, κ a=1 i 0.2 0.4 0.6 0.8 1 x (c) no=0.1, κ ab=1 i ) x ( f 7.5 7 6.5 0 0.2 0.4 0.6 0.8 1 x (d) no=0.1, κ a=1 i 8.5 8 κi ab=4 κi ab=5 κi ab=6 ) y ( f 7.5 7 6.5 0.001 0.01 0.1 1 (1+y)/2 0.01 0.1 1 (1+y)/2 FIG. 7. The free energy landscape as a function of the sign and disease probabilities in the homogenuous fully-connected D2S1 model. We assume that all the observed signs are positive. Here P (D = 1) = x, P (S = +1) = (1 + y)/2, and no = NO/NS. The prior disease probabilities and the leak sign probabilities are P0(Da = 1) = P (Si = +10) = 0.001. 25
1109.4140
5
1109
2017-05-07T22:49:11
Neuron as a reward-modulated combinatorial switch and a model of learning behavior
[ "physics.bio-ph", "q-bio.NC" ]
This paper proposes a neuronal circuitry layout and synaptic plasticity principles that allow the (pyramidal) neuron to act as a "combinatorial switch". Namely, the neuron learns to be more prone to generate spikes given those combinations of firing input neurons for which a previous spiking of the neuron had been followed by a positive global reward signal. The reward signal may be mediated by certain modulatory hormones or neurotransmitters, e.g., the dopamine. More generally, a trial-and-error learning paradigm is suggested in which a global reward signal triggers long-term enhancement or weakening of a neuron's spiking response to the preceding neuronal input firing pattern. Thus, rewards provide a feedback pathway that informs neurons whether their spiking was beneficial or detrimental for a particular input combination. The neuron's ability to discern specific combinations of firing input neurons is achieved through a random or predetermined spatial distribution of input synapses on dendrites that creates synaptic clusters that represent various permutations of input neurons. The corresponding dendritic segments, or the enclosed individual spines, are capable of being particularly excited, due to local sigmoidal thresholding involving voltage-gated channel conductances, if the segment's excitatory and absence of inhibitory inputs are temporally coincident. Such nonlinear excitation corresponds to a particular firing combination of input neurons, and it is posited that the excitation strength encodes the combinatorial memory and is regulated by long-term plasticity mechanisms. It is also suggested that the spine calcium influx that may result from the spatiotemporal synaptic input coincidence may cause the spine head actin filaments to undergo mechanical (muscle-like) contraction, with the ensuing cytoskeletal deformation transmitted to the axon initial segment where it may...
physics.bio-ph
physics
Neuron as a reward-modulated combinatorial switch and a model of learning behavior Marat M. Rvachev∗ April 27, 2013 Abstract This paper proposes a neuronal circuitry layout and synaptic plasticity principles that allow the (pyramidal) neuron to act as a "combinatorial switch". Namely, the neuron learns to be more prone to generate spikes given those combinations of firing input neurons for which a previous spiking of the neuron had been followed by a positive global reward signal. The reward signal may be mediated by certain modulatory hormones or neurotransmit- ters, e.g., the dopamine. More generally, a trial-and-error learning paradigm is suggested in which a global reward signal triggers long-term enhancement or weakening of a neuron's spiking response to the preceding neuronal input firing pattern. Thus, rewards provide a feedback pathway that informs neurons whether their spiking was beneficial or detrimental for a particular input combination. The neuron's ability to discern specific combinations of firing input neurons is achieved through a random or predetermined spatial distribution of input synapses on dendrites that creates synaptic clusters that represent various permuta- tions of input neurons. The corresponding dendritic segments, or the enclosed individual spines, are capable of being particularly excited, due to local sigmoidal thresholding in- volving voltage-gated channel conductances, if the segment's excitatory and absence of inhibitory inputs are temporally coincident. Such nonlinear excitation corresponds to a particular firing combination of input neurons, and it is posited that the excitation strength encodes the combinatorial memory and is regulated by long-term plasticity mechanisms. It is also suggested that the spine calcium influx that may result from the spatiotemporal synaptic input coincidence may cause the spine head actin filaments to undergo mechani- cal (muscle-like) contraction, with the ensuing cytoskeletal deformation transmitted to the axon initial segment where it may modulate the global neuron firing threshold. The tasks of pattern classification and generalization are discussed within the presented framework. 1 Introduction The field of reinforcement learning (RL) solves the problem of sequential decision making by an agent receiving delayed numerical rewards [1]. The field can be viewed as originating from two major threads: the idea of learning by trial and error that started in the psychology of animal learning (e.g., [2]), and the problem of optimal control and its solution using value functions and dynamic programming [3]. An important branch of the RL theory is the temporal difference ∗Email: [email protected] 1 (TD) class models for the phasic activity of midbrain dopamine neurons [4, 5, 6]. The dopamine activity is believed to encode a reward prediction error (RPE) signal that guides learning in the frontal cortex and the basal ganglia [7, 8, 9, 10]. Most scholars active in dopamine studies believe that the dopamine signal adjusts synaptic strengths in a quantitative manner until the subject's estimate of the value of current and future events is accurately encoded in the frontal cortex and basal ganglia [11]. This paper considers the problem of instantaneous decision making by an agent receiving immediate rewards within an RL-type framework. A trial-and-error learning paradigm is suggested in which the reward signal modulates memory in (cortical) neurons that act as combinatorial switches. The reward signal may come from an "elementary" reward generator such as that reflecting pain or satisfaction of hunger; it may also involve an RPE-type or "critic"-type [1] signal mediated by dopamine and/or other agents that could convey positive as well as negative reward components as was first suggested in [12]. The first contributing thread to the presented model, as in the classical RL theory, is the idea of learning by trial and error and reinforcement of favorable outcomes. The idea, as expressed in Edward Thorndike's "Law of Effect" [2], is: "Of several responses made to the same situation those which are accompanied or closely followed by satisfaction to the animal will, other things being equal, be more firmly connected with the situation, so that, when it recurs, they will be more likely to recur; those which are accompanied or closely followed by discomfort to the animal will, other things being equal, have their connections to the situation weakened, so that, when it recurs, they will be less likely to occur. The greater the satisfaction or discomfort, the greater the strengthening or weakening of the bond." This idea is widely regarded as a basic principle underlying much behavior [13, 14, 15, 16]. The second contributing thread is a novel idea that, given proper neuronal circuitry layout, pyramidal neurons can process information by switching the neuron output based on active input neuron combinations. This idea builds on the Two-Layer Neural Network (TLNN) model for the pyramidal neuron [17]. Additional computational advantages that could make the idea possible may be provided by mechanical force generated at the dendritic spines and stretch- activation of Na+ channels at the axon initial segment. An interesting feature of the presented framework is its ability to distil reusable abstract concepts about the environment, making learning with the low-dimensional feedback signal, the reward, efficient. 1.1 Problem formulation The following organism-level learning problem is posed. For simplicity, the neuronal activity states are considered to be binary: "firing" or "not firing". Given an arbitrary combination X of firing neurons in a (perhaps sensory) input layer L1, activate a corresponding "optimal" combination Y ∗(X) of firing neurons in a (perhaps motor) output layer L2 (Fig. 1(a)). The optimal combination Y ∗(X) is defined as one that produces the motor behavior that results in a positive global reward signal R in the organism. As such, Y ∗(X) can be an arbitrary combination of L2 neurons from a combinatorics perspective. The reward signal R, in biolog- ical terms, may be mediated by certain modulatory neurotransmitters or hormones that are diffusely delivered to generally trainable neurons. It is assumed that in biological systems R can be activated by evolutionarily hardwired circuits, such as when hunger is satisfied, as well as by higher mental processes, e.g., due to the organisms' subjective evaluation of the motor 2 Input x1 x2 x3 x4 xm-1 xm Pattern X in L1 Input L1:X={xi} Enhance X's excitation of Y Weaken X's excitation of Y "Action" or "guessing" mechanism R y1 y2 y3 yn-1 yn L2:Y*(X)={yk} Output (a) Pattern Y in L2 Output Yes Global reward signal positive? No (b) Figure 1: The organism-level learning problem and an outline of the suggested solution. (a) Formulation of the problem. Neurons xi, i = 1, . . . , m in layer L1 connect to neurons yk, k = 1, . . . , n in layer L2. A pattern of excitations X = {xi}, if responded to by a pattern of excitations Y = {yk}, elicits a positive or negative reward R resulting from the interaction of the generated motor behavior with the environment. The problem is: given an arbitrary X, excite Y ∗(X) that would lead to positive R. (b) Outline of the suggested solution. Learning proceeds by trial and error. Excitation of pattern X excites a pattern Y (X), possibly with the help of an "action" mechanism (e.g., depolarization to all L2 neurons until a certain level of the aggregate L2 output activity is achieved, as discussed in Sec. 2.3). A "guessing" mechanism introduces variations in the excited patterns Y . X's excitation of those Y that lead to positive (negative) R is enhanced (weakened). behavior as being satisfactory given the sensory inputs. It is suggested that the learning process proceeds in a trial-and-error fashion. Given a firing combination X variations are introduced in the firing combination Y with the X's excitation of those Y that lead to positive R being enhanced while X's excitation of those Y that lead to negative R being weakened (Fig. 1(b)). Details of this suggested process are discussed in more detail in Sec. 5. First, a more elementary learning task is considered: given an arbitrary firing combination X long-term strengthen excitation of an L2 neuron yk, specifically by X, if the subsequent reward R is positive. Conversely, long-term weaken excitation of yk, specifically by X, if R is negative (Fig. 2). 2 Solution to the single-neuron combinatorial switching prob- lem 2.1 Local dendritic integration as the basis for combinatorial memory The following mechanism is posited as the solution and is illustrated in Figs. 3 and 4. L1 neurons connect to the yk dendrites at random or predetermined locations, forming spa- tially localized (and possibly overlapping) "synapse neighborhoods" Nj that contain various permutations of input neurons. Sufficient depolarization of the dendritic and/or spine inte- 3 Input x1 x2 x3 x4 xm-1 xm L1:X={xi} R L2 yk Output Figure 2: The single-neuron learning problem. L1 neurons xi connect to an L2 neuron yk. Long-term enhance (weaken) yk excitation by those combinations X for which the following yk excitation resulted in a positive (negative) R. The enhancement and weakening of excitation may involve long-term potentiation (LTP) and long-term depression (LTD) processes that are influenced by both the combinatorics of the problem and the reward R, as suggested in Sec. 2.2.1. rior within the jth neighborhood, caused by the temporal coincidence of the neighborhood's excitatory and absence of inhibitory inputs, causes Nj excitation. The Nj excitation drives local input-output function Fj that has a "combinatorial memory" input-output component Cj that possesses the following properties: 1) Cj expression is long-term enhanced (weakened) if the neighborhood Nj is excited, this is closely followed by a back-propagating action poten- tial (BPAP) at Nj, and the immediately following R is positive (negative), and 2) compared to other drivers of neuron stimulation, Cj can substantially contribute to the yk excitation. Note that the input-output function Cj is driven by Nj excitation that itself is caused by the spatiotemporal coincidence of inputs. This confers Cj combinatorial specificity. As an example, assuming, as we do throughout this paper, that all inputs {xi} are 1 or 0, i.e., active or inactive, and also that all synaptic weights wji from xi to Nj are +1, -1 or 0 (corresponding to excitatory, inhibitory synapses and the absence of synaptic contact, respectively), a simple Cj can be written as where γj is the weight on Nj, n∗ Cj(X) = γjH(nj − n∗ j = Pwji>0 wji > 0 is the number of Nj excitatory synapses, nj = Pwji>0 wjixi is the number of active Nj excitatory synapses (n∗ j ≥ nj ≥ 0), nj = −Pwji<0 wjixi ≥ 0 is the number of active Nj inhibitory synapses and H(n) is the step j − nj), function: (1) (2) H(n) = (cid:26) 1 0 if n ≥ 0, if n < 0. In Eq. 1 the argument to H() is less than 0 unless 1) n∗ j = nj, i.e., all excitatory synapses in Nj are active and 2) nj = 0, i.e., all inhibitory synapses in Nj are inactive (recall that n∗ j ≥ nj ≥ 0 and nj ≥ 0). Weights γj are increased (decreased) if Nj is excited, this j > 0, n∗ 4 (a) x1 x2 x3 x4 xm-2 xm-1 xm L1:X={xi} yk Axon ΣαjFj N1: {x1, x2, x3, x4} (b) nj=Σwjixi N2: N3: N4: N5: N6: {x2, {x1, x4, x2, xm-1} x3, xm-2} {x2, x4, xm-2} {x1, x2, xm-1} {x1, x2, x4} Nl-2: Nl-1: Nl: {x3, x4, xm-2, xm-1, xm} {x4, xm-1, xm} {xm-1, xm, xm} Dendrite n1 n2 n3 N1 N2 N3 nl Nl Fj(nj) g αj yk Figure 3: Solution to the single-neuron learning problem shown in Fig. 2. (a) Inputs from xi form spatially localized clusters, or "neighborhoods", Nj, j = 1, . . . , l, on the yk dendrites. In the figure, a set of xi below an Nj denotes neurons projecting into the cluster. Activation of excitatory synapses simultaneous with a lack of activation of inhibitory synapses in cluster Nj produces output Fj that has a "combinatorial memory" component Cj. Fj generated at all yk neighborhoods superimpose. Cj expression is regulated as suggested in Fig. 4. (b) An equivalent neural network diagram for (a). The synaptic weight from xi to Nj is wji. Neighborhood weights αj are shown in filled circles. Both figures (a) and (b) are essentially a reinterpretation of Fig. 1 in [17]. is closely followed by a BPAP at Nj and the immediately following R is positive (negative). Note that in this formulation weights γj are independent from wji and learning may proceed with changing γj and unchanged wji. It is easy to see that the existence of the input-output functions Cj can in principle solve the single-neuron learning problem posed in Fig. 2, assuming that yk firings cause BPAPs that propagate to all Nj (we do assume this here and below). Indeed, insofar as the features of X are represented in the corresponding set of excited neighborhoods NX = {Nj}, the corresponding set CX = {Cj} will be strengthened and will enhance yk excitation when X is presented, if an earlier X presentation was followed by yk firing and the subsequent R was positive. This immediately follows from the Cj training rules (Fig. 4). The combinatorial specificity of Cj should ensure that the enhanced yk excitation will be specific to the pattern X and similar patterns (see Sections 3 and 6 for numerical simulations). Conversely, Cj training 5 Pattern X in L1 Enhance Cj expression Weaken Cj expression Do not change Cj expression Corresponding yk's set ΝX={Nj} nonlinearly excited, its functions Cj expressed "Action" or "guessing" mechanism Neuron yk fires Is BPAP observed at site Nj∈ΝX? Yes Global reward signal R positive? No No Yes Figure 4: Suggested learning rules for the combinatorial memory Cj. The excitation of a pattern X in L1 leads to the excitation of the corresponding set of neighborhoods NX = {Nj} in yk. The input-output function Cj is long-term enhanced (weakened) if the neighborhood Nj is excited, i.e., Nj ∈ NX, this is closely followed by a BPAP at Nj and the following R is positive (negative). rules will cause a weakened yk excitation by those X for which a previous yk excitation was followed by negative R. In Sec. 2.2 we digress into looking for a plausible physical (electrical or mechanoelectrical) Cj realization in biological neurons, and starting in Sec. 2.3 we model networks of neurons possessing Cj. 2.2 Possible physical realizations of the combinatorial memory 2.2.1 Electrical mechanism The local input coincidence detection that confers Cj the combinatorial nature can be related to the local sigmoidal thresholding of postsynaptic potentials (PSPs) that results from nonlinear activation of voltage-dependent NMDA, Ca2+ and Na2+ currents [20, 21, 22, 17]. Using a detailed compartmental model of a hippocampal CA1 pyramidal neuron, the neuron input- output function was shown to behave as a two-layer "neural network" with the output given by [17] l y = g Xj=1  αjs(nj)  , 6 (3) Subunit input-output functions Neuron activation function V m , t u p t u o t i n u b u S 4.5 4 3.5 3 2.5 2 1.5 1 0.5 0 0 + (n) s s(n) s- (n) s lin(n ) = 0.3 3 4 2 n 2 1 8 Net number of excitatory synapses (n) 3 6 7 4 5 z H , t u p t u o n o r u e N 70 60 50 40 30 20 10 0 9 g(x) 0 10 20 30 40 50 60 70 80 Induced local field, mV Figure 5: The subunit input-output function from [17]: s(n) = 1/(1 + exp((3.6 − n)/0.2)) + 0.3n + 0.0114n2. Also shown are estimated subunit input-output functions after a uniform 25% increase and decrease in efficacy of all subunit synapses: s+(n) = s(1.25n) and s−(n) = s(0.75n). The function slin(n) = 0.3342n is the linear fit to s(n) below n = 3. (b) The neuron global activation function from [17]: g(x) = 0.96x/(1 + 1509exp(−0.26x)). where nj is the net number of excitatory synapses driving the jth dendritic "subunit", s(n) is the subunit input-output function, αj is the weight on the jth subunit, l is the number of subunits in the cell, and g(x) is the global output nonlinearity (Fig. 5). In the study the subunits were assumed to correspond physically to long, thin unbranched terminal dendrites of the apical and basal tree. The strength of each synapse was scaled to yield an equal (5 mV) peak EPSP locally at each synapse for the input intensities simulated. We estimate the effect of a uniform 25% increase and decrease in efficacy of all subunit synapses on the subunit input-output function in Fig. 5(a), assuming that the scaling in synaptic strength is equivalent to the scaling in the number of active synapses. Lets suppose the subunit's "combinatorial memory" input-output component is given by C(n) = s(n) − slin(n), where slin(n) = 0.3342n is the linear fit to s(n) below the threshold (Fig. 5(a)). Indeed, as shown in Fig. 6, C(n) is increased (decreased) and its "activation threshold" is lowered (raised) following a uniform increase (decrease) in the synapse efficacy, making C(n) a good candidate for a trainable detector of local input coincidences if the linear component slin(n) can be subtracted from s(n) in the signal analysis. In fact, it will be shown in Sec. 3.2 that the lowered C(n) activation threshold when the synaptic efficacy is increased is not likely to improve the neuron performance in the task of generalization. However, in 7 Candidate for combinatorial memory function in neuron V m , n o i t c n u f t t u p u o - t u p n I 3 2.5 2 1.5 1 0.5 0 -0.5 0 + (n)-s lin(n) li n ( n ) s - s ( n ) s s-(n)-slin(n) 5 6 2 1 8 Net number of activated synapses (n) 3 4 7 9 Figure 6: Estimated effect of a uniform 25% increase and decrease in efficacy of all subunit synapses on the function C(n) = s(n) − slin(n) (black lines). Plotted in red is a similar hypothetical combinatorial memory function C ∗(n) = γH(n − ntr) with a trainable weight γ and activation threshold ntr. order for C(n) to act as the combinatorial memory defined in Sec. 2.1, the individual synapse plasticity has to be influenced by both the combinatorics of the problem and the reward R. For example, C(n) will conform to the combinatorial memory plasticity rule if the synaptic plasticity behaves as: LTP (LTD) is induced in a synapse if the synapse and its subunit are excited, this is immediately followed by a BPAP at the subunit and the immediately following R is positive (negative). 2.2.2 Mechanical (muscle-like) mechanism An interesting Cj realization is possible if the free calcium entering a spine during a spatiotem- poral synaptic input coincidence event elicits spine actin filament contraction, e.g., through calcium-activated actin interaction with myosin or another actin-binding protein. The ensuing cytoskeletal and cytoplasmic stresses, the magnitudes of which could encode the combinato- rial memory, could be transmitted along the dendritic shaft to the yk's axon initial segment (Fig. 7). At the initial segment these stresses, superimposed with those generated at other dendritic sites, could regulate the global yk excitation threshold via stretch-modulating Na+ voltage-gated ion channels (Nav) [23, 24]. The use of the mechanical force would provide the second dimension to the neuron's computational machinery, disentangling the spatiotemporal 8 Glutamate receptors Actin cytoskeleton Dendrite Microtubules Actin cross-linking proteins Myosin Cytoskeletal stresses induced by actin-myosin contractility Corresponding cytoplasmic pressure gradients Figure 7: The influx of Ca2+ into a spine could elicit (muscle-like) contraction in the spine actin cytoskeleton. Induced cytoskeletal stresses and cytoplasmic pressure gradients are shown as green and blue arrows, respectively. These mechanical forces could propagate along the dendrite, as shown. At the axon initial segment, the ensuing stresses could modulate the global neuron firing threshold. coincidence detection mechanism, which would be electrical and based on local nonlinear volt- age summation, from the Cj readout mechanism, which would be mechanical. The spine head volume and the associated quantity of actin filaments would then reflect the Cj magnitude, rather similarly to how the muscle cell volume and strength reflect the memory of previous exercise. The advantage of having additional mechanical memory can be seen from the following simulation. Using simulation results from [17], we modeled a pyramidal neuron with 37 dendrites (the "subunits", or "branches") connecting to the apical trunk (Fig. 8). As in [17], the neuron inputs were excitatory only. The dendrite input-output functions s(n) = 1/(1 + exp((3.6 − n)/0.2)) + 0.3n + 0.0114n2 and the global output nonlinearity g(x) = 0.96x/(1 + 1509exp(−0.26x)) that in [17] were combined using Eq. 3, were modified with two versions of "mechanical memory", C ′ j H(nj − 4), as (4) jH(nj − 4) and C ′′ j = 1, . . . , 37, Sj(nj) = s(nj) + C ′ j(nj), j(nj) = γ ′ j (nj) = γ ′′ 37 y = g  x0 + Xj=1(cid:2)αjSj(nj) + C ′′ j (nj)(cid:3)  , (5) where nj is the net number of active inputs in subunit j and αj, γ ′ j and x0 are constants. The function C ′ j(nj) represented an added output invoked in the subunit j by a local mechan- ical memory mechanism (e.g., the local actin cytoskeleton contractility stretch-modulating gating of the membrane ion channels). The function C ′′ j (nj) represented modulation of the yk firing threshold by a global mechanical memory mechanism. Both mechanisms were activated when nj was at least four. The task posed for the neuron was to detect coincident activation of at least 4 inputs on any of the branches, signaling it by firing with the frequency at least 40 Hz. j, γ ′′ 9 n1 n2 n3 N1 N2 N3 n37 N37 s(nj)+Cj'(nj) nj=Σwjixi αj, Cj''(nj) g yk Figure 8: Equivalent neural network for simulation of a pyramidal neuron with hypothetical mechanical memory. Compared to the TLNN model [17] expressed by Eq. 3, each subunit has an additional "local" mechanical memory output C ′ j to the global nonlinearity input. j as well as it contributes C ′′ The coincidence threshold was set at 4 because the electrical subunit functions s(n) display a sharp increase at n = 4 (Fig. 5(a)) which should help the neuron detect such patterns. For the purely electrical neuron (γ ′ j = 0) the global output g(x) threshold was allowed to vary via the parameter x0 to improve the detection performance. j = γ ′′ The total of 7000 input patterns were generated, 3500 with four active synapses on a branch (the coincidence pattern), and 3500 with at most three active synapses on a branch (the no- coincidence pattern). To sample a wide range of input intensities, patterns were selected as follows. First, the total number of active synapses Ne was chosen at 30, 35, 40, 45, 50, 55 or 60. Five no-coincidence patterns were generated by randomly distributing Ne active inputs on the 37 branches, restricting the number of active inputs to at most three per branch. One coincidence pattern was generated for each c from 1 to 5, by assigning four active inputs to c random branches; the remaining (Ne − 4c) active inputs were randomly distributed to the remaining (37 − c) branches, restricting the number of active inputs per branch to at most three. The above procedure for generating 10 patterns was repeated 100 times for each Ne, yielding 7000 patterns overall (7 x 10 x 100). To simplify, all αi were assumed to be equal and were scaled in a pilot run with γ ′ j = 0 so that a reasonable sub-40-Hz output was observed for the 7000 patterns, giving α = αi = 1.7 (Fig. 9(a)). Note that in the more realistic neuron model [17], the ratio of the average couplings αi for the "first-order" branches (those connected directly to the apical trunk) to that for the higher-order branches was from 1.41 to 1.73 depending on technique used (see Fig. 4 in [17]). However, we do not expect that using a distribution of values for αi would change our conclusions below. j = γ ′′ Fig. 9(b) illustrates the performance in classifying the 7000 patterns for the three neuron models. The fraction of correct classifications for the purely electrical neuron (γ ′ j = 0) as a function of the global output threshold x0 was at best 64%. It can be seen that such poor performance stems mainly from the relatively large variance in the overall input intensity Ne: although each subunit "tries" to detect the coincidence, background signals from other subunits hinder the detection at the neuron level. The addition of either type of mechanical memory improved the performance dramatically, with more than 90% correct classification for γ ′ j ≥ 6 mV and γ ′′ j ≥ 11 mV. j = γ ′′ 10 z H , t u p t u o n o r u e N 80 70 60 50 40 30 20 10 0 Calibration of the coupling parameter α Classification performance for various neuron models The average and the range of the neuron output n o i t l a c i f i s s a c t c e r r o c t n e c r e P 100 90 80 70 60 50 γ'=γ''=0, vary x0 x0=γ''=0, vary γ' x0=γ'=0, vary γ'' 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 α 0 5 15 10 x0, γ' or γ'' (mV) 20 25 30 Figure 9: (a) The average and the range of output for the electrical neuron (γ ′ j = 0) for the 7000 input patterns, as a function of α = αi. The error bars show the minimum and maximum neuron output over all the input patterns. (b) Classification performance for the 7000 patterns for the three neuron models: purely electrical (red circles), with local mechanical memory (green squares) and with global mechanical memory (blue triangles). In all cases α = αi = 1.7. The "mechanical memory" weights for the 37 subunits were assumed equal: γ ′ = γ ′ j , j = 1, . . . , 37. No parameter fine tuning was required to obtain the results shown other than setting α to 1.7. For α equal to 1 (2.5) the best classification performance was 63%, 100%, 100% (64%, 86%, 86%) for the electrical and the two mechanical neurons, respectively. j, γ ′′ = γ ′′ j = γ ′′ For the neuron with the global mechanical memory and a large γ ′′ j ≥ 11 mV), the stretch activation at the axon initial segment effectively acts in a digital manner, providing a feedback signal that a structural modification has occurred somewhere in the dendritic tree, much like a BPAP is thought to tell the neuron the axon has fired. Several morphological observations favor such memory model for pyramidal neurons. Unlike many other neuron types, the pyramidal neurons have rather straight dendrites that tend to branch at small angles, which should facilitate transmission of the cytoskeletal and cytoplasmic stresses along the dendrite length. The dendritic microtubules are linear, quite rigid and invade the spines, where they likely link to actin cytoskeleton [25, 26], which should also facilitate the transmission. j (e.g., γ ′′ A 1-ms cytoplasmic pressure pulse propagates along a 1-µm-diameter unmyelinated axon In the "high viscosity" regime with the velocity 1.1 m/s and decay length 0.18 mm [24]. 11 applicable to such pulses, these quantities scale as √ωd and q d ω , respectively, where ω is the central frequency of the wave packet and d is the axon diameter [24] (see also [29]). Assuming that unmyelinated axons and dendrites have similar mechanical properties, the propagation parameters for cytoplasmic pressure pulses should be similar for dendrites. In particular, a 10-ms pressure pulse in a 1-µm-diameter dendrite should travel with 0.35 m/s velocity and 0.57 mm decay length; for a 100-ms pulse these should change to 0.11 m/s and 1.8 mm, respectively. These values are consistent with the idea that mechanical forces can be transferred through the lengths of the pyramidal neuron dendrites and that the forces can be produced and transmitted sufficiently rapidly so as to be associated with the spike initiating event. Given these observations, it is suggested that the mechanical mechanism for the combinato- rial memory may have evolved relatively recently, culminating in the creation of the pyramidal neuron in higher animals, which allowed the neuron to more specifically respond to the com- binatorial aspects of inputs. Note that the mechanical mechanism suggested here confers a functional role to the spine head volume and high spine actin content, roles of which are still enigmatic [28, 26, 30]. 2.3 Neuronal circuitry layout Fig. 10(a) shows a suggested neuronal circuitry layout for the learning process and memory readout. Lets assume that in an untrained system presented with an X in L1 the postsynaptic integration does not suffice to excite an L2 neuron yk. In a learning trial, yk is activated by additional depolarization created by increased excitation or reduced inhibition from one or more "guess" (G) or "action" (A) neurons that connect to yk in dominant positions, such as near the axon initial segment. Alternatively, the "guess" or "action" neurons could connect to yk at the apical tuft, where they could generate the Ca2+ dendritic spikes propagating towards the soma and driving initiation of the action potentials [28]. The general learning scheme with a global reinforcement signal R broadcast by a critic to all neurons and the neurons receiving "empiric" synapses driven by random spike trains from an external experimenter was first suggested in [31]. Output neurons could be structurally connected to inputs in a similar, although not nec- essarily identical, manner, such as when closely spaced L2 neurons sprawl basal dendrites in a plane, into which L1 axons diffusely and randomly project (Fig. 10(a)). This connectivity would be conducive to increasing the learning power of the system, as each L2 neuron would roughly be equal in its ability to learn how to react to an arbitrary combination X. In an untrained system, given an X in L1, all L2 neurons should then be similarly close to the ac- tivation threshold. Following learning, the L2 neurons trained to react positively to X should be closer to the activation threshold than others. The actual memory readout could proceed using the "action" neurons that deliver similar rising levels of depolarization to all L2 neurons, e.g., via somatic or apical tuft connections, until a certain criterion such as a predefined level of the aggregate L2 output activity is met. 12 Input from "action" or "guess" neurons Input from "action" or "guess" neurons Input from "action" or "guess" neurons Input from L1 Output layer L2 Ca2+ spikes originating in apical tufts may drive trial variations in L2 and Li firing patterns Input from L1 yk Intermediate layer Li yk Output layer L2 L2 output (a) L2 output (b) Figure 10: Suggested neuronal architecture for the learning process and memory readout. (a) Learning with one input layer, L1 (red), and one output layer, L2 (black). Input from L1 diffusely projects into L2 dendrites that are arranged in a plane. The "guess" and "action" inputs (green) connect at the soma or the apical tufts. (b) Learning with an added intermediate layer Li (blue). Layer L1 diffusely projects into both L2 and Li, while Li diffusely projects into L2. The "guess" and "action" neurons connect at the apical tufts or the soma (not shown). First, Li neurons learn to fire for the important to the organism combinations X in L1. The reduced dimensionality signals are then used in further L2 learning. In the neocortex, L2 could correspond to the layer V pyramidal neurons and Li to the layer II/III pyramidal neurons. 3 Pattern classification and generalization We first consider several standard benchmark problems for binary input neural networks and then consider the problem of sparse input classification and generalization. 3.1 Standard benchmark problems Fig. 11 shows the equivalent neural network for a trivial solution to the problem of classification of binary patterns randomly assigned to two classes within the presented framework. m input neurons form l = 2m synaptic clusters of size m synapses each (one synapse per input neuron in each cluster) on the yk dendrites, with each cluster j representing one of the 2m unique patterns X = {xi} using the set of weights wji from xi to Nj that take on values of 1 or -1 for an active or inactive xi, respectively, in the pattern X being represented. The cluster 13 Input x1 Hidden +1 +1 +1 +1 N1 x2 -1 +1 N2 γ2 γl Σ yk γ1 Output xm -1 -1 -1 N2m "Trial" input Figure 11: Equivalent neural network for a single-neuron solution to the problem of classifica- tion of binary patterns X = {xi} randomly assigned to two classes. The hidden computation layer contains units Nj, j = 1, . . . , 2m that correspond to synaptic clusters on the yk den- drites. Each unit Nj represents one of the 2m unique patterns X using its m input weights wji ∈ {−1, 1}. The unit input-output function is given by γjH(nj − n∗ j − nj) with the notation defined in Eq. 1. Here we assume n∗ j ≥ 0. Output of yk is the simple sum of its inputs. j can be equal to 0, i.e., n∗ input-output functions are given by Eq. 1 and are summed to yield the neuron output y = l Xj=1 γjH(nj − n∗ j − nj). (6) It can be seen that for every input pattern X only one neighborhood j is "excited", i.e., its H(nj − n∗ j − nj) is equal to 1, leading to y = γj. Initially, all γj are set to 0. A weight γj is set to 1 if its neighborhood Nj is excited, this is followed by a BPAP at Nj, i.e., yk fires, and the following global reward signal R is positive: Set γj = 1 if H(nj − n∗ j − nj) = 1, yk fires and the following R > 0. (7) Given a random mapping {Xi} → X(0), X(1) training proceeds as follows. A pattern Xi ∈ X(1) is excited, this is followed by a "trial" yk firing and the delivery of positive R. The weight γj whose Nj was excited by Xi is set to 1, in accordance with the learning rule in Eq. 7. The procedure is repeated for every Xi ∈ X(1), resulting in yk being trained to respond with output 0 to all Xi ∈ X(0) and with 1 to all Xi ∈ X(1) thus solving the classification problem. Note that this solution requires 2m(m + 1) synaptic contacts and at most one training pass or "epoch". A special case of the above classification problem is the N -bit party problem in which the neuron is required to output 1 if the number of 1s in its N inputs is odd, and 0 otherwise. Within the presented framework this is solved similarly, resulting in γj equal to 1 for those Xi that have odd number of active inputs, and to 0 otherwise. As in the classification problem, the solution requires 2m(m + 1) synaptic contacts for m inputs. Table 1 compares the speed and efficiency of solving 7- to 10-bit parity problems in the presented model to several other neural network models. The presented model is much more efficient in training time, but much less efficient in the number of synapses used compared to the back-propagation [32] and 14 Parity problem RMCS Number of links (epochs) 7-bit 8-bit 9-bit 10-bit 1024(1) 2304(1) 5120(1) 11264(1) Back-propagation[32] Number of links (epochs) 127(781) 161(1953) N/A N/A EPNet[33] Number of links (epochs) 34.7(177417) 55(249625) N/A N/A FCEA[34] Number of links (epochs) 64(1052) 81(3650) 100(6704) 121(9896) Table 1: Comparison of speed and efficiency of solving 7- to 10-bit parity problems in the presented RMCS (Reward-Modulated Combinatorial Switch) model to other models. In the RMCS parity problems are solved exactly in one pass. For m inputs the number of links used is 2m(m + 1). Back-propagation model uses the N -2N -1 configuration, fully connected from inputs to hiddens and from hiddens to output [32]. EPNet and FCEA are evolutionary algorithms that combine architectural evolution and weights optimization [33, 34]. "N/A" denotes "Not available". evolutionary [33, 34] models listed. The presented model uses only one, albeit much more intricate, neuron, compared to many more in the other models. Fig. 12 shows one of possible solutions to the XOR problem (the 2-bit parity problem) within the presented framework. As in the above solution to the binary pattern classification problem, the cluster excitation thresholds are set to the number of their excitatory synapses n∗ j . However, here a cluster may receive more than one synapse from the same input neuron. 3.2 Sparse coding problems To study how the presented framework performs in the arbitrary X → Y mapping task posed in Sec. 1.1 in a sparse coding regime the following simulation setup was created. Here for clarity we modify the notation. Ni input neurons form synaptic clusters on dendrites of No output neurons, Nc clusters in total per output neuron, each cluster having exactly nc excitatory and no inhibitory synapses (the equivalent neural network diagram is similar to the one depicted in Fig. 14(a) for a problem considered later). Np different randomly generated binary input patterns, each having exactly Ne active inputs, are randomly assigned to the No outputs, Np/No patterns per output. The task is to train the system to classify the Np patterns into the assigned No output classes, i.e., make it fire the correct output neuron given presented patterns. The total number of synapses per output neuron, ncNc, was restricted to N (max) . s Training proceeded as follows. The Np patterns were presented sequentially t times each, with Nnoise additional random inputs activated at each presentation. Each presentation was followed by a "trial firing" of the correct output neuron. Positive reward R was delivered. Cluster weights γj, initially set at 0, were incremented by 1 for all the synaptic clusters that were excited on the neuron that fired. For this purpose a synaptic cluster was defined as excited if at least nlearn (nlearn ≤ nc) its synapses were excited. In the testing phase, the same Np patterns were again presented sequentially, each pattern only once, with additional Nnoise random inputs activated at each presentation. The output neuron with the largest Γ = P γj, where the sum is over the clusters that had at least nrecall (nrecall ≤ nc) active inputs, was the one that fired. In cases when several neurons had the 15 (a) x1 x2 L1 (b) Input x1 x2 yk Axon +1 +1 +1 +1 C1+C2 -1 -1 N1: {x1, x1, x2} N2: {x1, x2, x2} Dendrite C1=H(2x1-2-x2), C2=H(2x2-2-x1) Hidden +2 -1 -1 -2 N1 +2 -2 N2 Output Σ+1 +1 yk Figure 12: (a) A solution to the XOR problem within the presented framework. The yk output is the sum of the cluster input-output functions, C1 = H(2x1−2−x2) and C2 = H(2x2−2−x1). For every binary input pattern X at most one cluster is active. (b) The equivalent neural network diagram for (a). same largest Γ one of them was randomly selected to fire. Table 2 shows the results of the experiment for Ni = 30, No = 10, Np = 1000, N (max) = 40000, t = 1, Nnoise = 0 for different values of nc and Ne. Here we used nlearn = nrecall = nc, i.e., all nc cluster inputs were required to be excited for a cluster to learn and to output what it learned. t = 1 and Nnoise = 0 means that there was no randomness introduced during learning and testing, i.e., the task was to simply memorize presented patterns. In one cluster generation method (Ndupl = 1), each synapse was randomly generated from the Ni inputs with equal probability 1/Ni. In the second method (Ndupl = 0), after clusters were generated as described above the clusters that had more than one contacting synapse from the same input neuron were eliminated. s From the results in Table 2 it is rather clear that the best performance in the memorization task for pattern sizes Ne ≤ 10 was for the setup with the cluster size nc of 3 or 4. This can be understood by noting that for Ni = 30 and with the limitation of N (max) = 40000, the maximum cluster size nc that would allow the exact representation of all possible patterns , is nc = 3. Table 3 shows of size Ne = nc, i.e., maximum nc such that the maximum number of inputs Ni for which all patterns of size Ne = nc can be exactly represented by individual neuron clusters using at most N (max) (Ni−nc)!nc! ≤ N (max) ncNi! s = 40000 synapses. s s Another interesting conclusion from the results in Table 2 is that the elimination of the clusters that had more than one input from the same neuron (i.e., using Ndupl = 0) improves the memorization performance for nc = 3 and 4 dramatically as the clusters representing lower-dimensional pattern features are eliminated. Table 4 shows the results of experiments in which random noise was added to the patterns during learning and testing. The same base pattern was presented up to three times during learning. In addition, a cluster was allowed to increase its weight and/or contribute to the neuron output when fewer than all its nc synapses were active (nlearn < nc and nrecall < nc, respectively). The conclusion drawn from these experiments is that using nlearn and nrecall 16 nc Ne 1 100 2 29 55 100 29 100 Ndupl 1 1 1 1 1 1 0 0 0 0 0 0 1 2 3 4 5 6 1 2 3 4 5 6 3 19 40 67 19 62 100 4 20 42 78 34 20 57 100 38 5 21 42 77 39 33 21 52 99 79 13 6 20 39 75 49 39 31 20 47 98 94 34 10 7 21 37 74 53 48 40 21 44 96 98 73 17 8 22 38 67 61 59 45 22 41 89 99 88 34 10 19 34 59 63 73 58 19 36 76 97 98 86 15 22 28 39 50 60 22 29 41 59 83 Table 2: Percentage of patterns classified correctly into No = 10 classes, for Np = 1000 different random binary patterns generated on Ni = 30 inputs, for various values of the pattern size Ne and the synaptic cluster size nc, for the model of Sec. 3.2. Clusters were (Ndupl = 1) or were not (Ndupl = 0) allowed to have more than one input from the same neuron. For Ne = 1, 2 the number of generated patterns was 30 and 30·29/2=435, respectively. The limit of N (max) = 40000 synapses per output neuron. For nc = 1, 2, 3 and Ndupl = 1, all 30nc clusters resulting from all possible input permutations were simulated for each output neuron. For nc = 2, 3 and Ndupl = 0, the 30nc clusters were simulated and those with more than one input from the same neuron eliminated. No random noise added to the patterns. All nc cluster inputs must be excited for the cluster to learn and to output what it learned (nc = nlearn = nrecall). Each measurement used a single simulated set of neurons and their interconnections, and the same initial random number seed. s equal to nc is optimal in most cases. The only situation when using nlearn and nrecall both lower than nc helped the classification was when the typical presented pattern was not represented in a cluster but was represented in a subcluster. However, in almost all cases using nlearn = nrecall = nc − 1 was more optimal than using nlearn = nc with nrecall = nc − 1. That is, the lowering of the cluster excitation threshold after the cluster was trained (as was the case for the function C(n) in Fig. 6) was typically less optimal than keeping the cluster excitation threshold at a constant, lower or higher, level. Another conclusion from Table 4 is that the system exhibited rather high "generalization" performance. For example, the 1000 patterns were classified correctly in 80% of cases for the cluster size 4 and the pattern size 6 when the patterns were presented 3 times during learning and one additional random input was activated during both learning and testing. As can be seen from the above simulations, having synaptic clusters that are more specific to the combinatorics of patterns X is desirable for better pattern memorization but this may require a large number of clusters. Lets assume that nlearn = nrecall = nc. The following question is posed: for Ni input neurons connected to an output neuron yk, Ne the (constant) number of excited neurons in a random binary input pattern X, nc the synaptic cluster size and random Ni to yk connectivity, how many clusters Nc are needed to fully represent an 17 nc Ni 1 40000 2 200 3 44 4 23 5 17 6 15 7 14 Table 3: The maximum number of inputs Ni for which all binary patterns of size Ne = nc may be exactly represented by individual clusters of size nc using at most 40000 synapses. That is, maximum Ni such that Ni!nc (Ni−nc)!nc! ≤ 40000. arbitrary X (without necessarily exactly representing X combinatorics)? The probability that a synapse receives active input from X is Ne/Ni. Therefore, approximately, the probability that a cluster is excited, i.e., all its inputs are excited, is ( Ne )nc . One needs at least Ne/nc Ni clusters to represent X, leading to Nc ≈ Ne )nc . Note that, as expected, for nc = 1, i.e., nonclustered synapses, this leads to Ns = ncNc = Ni synapses needed to represent an arbitrary pattern X. For Ne = nc, i.e., exact encoding of X combinatorics in clusters, one needs roughly ncNc = Ne( Ni Ne )Ne synapses. Using the above formula for Nc and assuming, for the sake of argument, Ni = 100 and Ne = 10, yields Ns = ncNc = 10nc+1. Comparing this to 50000, the number of contacting synapses for a pyramidal neuron [28], suggests the cluster size of not more than 3 to 4. Note that for Ns = 50000, Ni = 100, Ne = 10 and nc = 3, roughly Ns( Ne )nc = 50 synapses are in Ni the excited clusters for the typical firing pattern. ( Ni Ne nc On the other hand, as discussed in Sec. 6, most of the learned activity of higher organisms could be considered a form of combinatorial switching if the combinatorial switching idea is taken to its logical extreme. Then, the language, as an artificial human construct designed for ease of communication, should reflect the switching dynamics that the involved neurons are "comfortable" operating on. There are about 40-50 sounds in a typical language and 4-5 sounds in a typical word, which may suggest, roughly, 4-5 as the typical pattern size (Ne) and 40-50 as the number of input neurons (Ni). Here for simplicity the issue of sound ordering within words is ignored. Using the formulas above with Ne = 5 and Ni = 50, the approximate number of contacting synapses per neuron needed to represent an arbitrary word is 5 · 10nc . Again, comparing this to the experimentally observed 50000 synapses per neuron suggests the cluster size of not more than 4. One could also hypothesize that Ne (Ni) can be related to the number of syllables comprising a typical word (the total number of syllables), the number of words in a typical sentence (the total number of words), and the number of the elements in the writing of a typical letter (the total number of the elements). 4 Training of intermediate layers To further reduce the dimensionality of inputs through encoding of frequently occurring and significant to the organism input patterns an intermediate layer Li could be trained using the neuronal architecture suggested in Fig. 10(b). The axons from L1 randomly project into the basal dendrites of both the intermediate layer Li and the output layer L2 while the Li axons randomly project into the L2 basal dendrites. The apical tufts of both Li and L2 receive the guessing and action driving signals emanating from distal brain areas. Note that the sprawling planar arrangement of Li and L2 basal dendrites and random synapse connectivity 18 nc t Nnoise nlearn nrecall Ne 3 3 3 3 3 3 3 3 3 3 4 4 4 4 4 4 4 4 4 1 1 1 3 3 3 3 3 3 1 1 1 3 3 3 3 3 3 0 0 0 1 1 1 2 2 2 0 0 0 1 1 1 2 2 2 3 2 3 3 2 3 3 2 3 4 3 4 4 3 4 4 3 4 3 2 2 3 2 2 3 2 2 4 3 3 4 3 3 4 3 3 100 55 26 63 27 27 33 23 23 10 99 10 13 44 37 17 26 26 4 100 47 23 79 27 22 42 19 19 38 96 28 40 53 42 34 29 32 5 99 41 24 80 26 22 47 22 21 79 87 30 68 44 36 46 29 32 6 98 33 21 76 26 20 46 22 21 94 76 34 80 41 36 56 28 29 7 96 31 21 70 24 20 43 22 20 98 63 33 84 36 34 58 26 29 8 89 31 21 64 26 21 42 22 20 99 51 27 87 31 28 57 23 24 10 76 27 17 50 21 16 36 20 16 97 39 26 73 28 24 48 23 22 Table 4: Percentage of patterns classified correctly into No = 10 classes, for Np = 1000 different random binary patterns generated on Ni = 30 inputs, for various values of the pattern size Ne and the synaptic cluster size nc, for the model of Sec. 3.2. During learning, each of the 1000 patterns was presented t times with Nnoise additional random inputs activated. During learning, a cluster weight γj was increased by 1 if the cluster had at least nlearn active synapses and its neuron fired. During testing, each of the 1000 patterns was presented once with Nnoise additional random inputs activated. A cluster weight γj contributed to the neuron output if the cluster had at least nrecall active inputs. N (max) = 40000, Ndupl = 0. For nc = 3, all 30nc clusters resulting from all possible input permutations were simulated after which those with more than one input from the same neuron eliminated. Each measurement used a single simulated set of neurons and their interconnections, and the same initial random number seed. s should increase the learning power of the system, as discussed in Sec. 2.3. First, L2 neurons are trained to respond to certain L1 combinations. As a side effect, some Li neurons learn to become more responsive to the frequently occurring L1 patterns that are followed by positive R. This learning could occur without the BPAP signals in Li neurons if their plasticity can be induced only by the local neighborhood excitation and the subsequent receipt of positive R. Then, the trained Li neurons would become more excited when the learned L1 patterns are presented and could themselves be fired by the guessing or action mechanisms, thus making the reduced dimensionality inputs available to L2 neurons for further learning. 19 5 Solution to the multi-neuron combinatorial switching prob- lem As suggested in Sec. 1, the organism-level learning problem of finding an optimal Y ∗(X) (Fig. 1(a)) is solved using a trial-and-error search as variations are introduced in the firing combination Y . As discussed in Sec. 2.3, the signals driving the trial variations could come in the form of the Ca2+ dendritic spikes originating in the apical tufts. The apical tufts of pyramidal neurons in cortical layers II/III and V are known to receive input from distal cortical areas and nonspecific thalamic projections, with the inputs generally having different origins than those that form synapses with more proximal apical or basal dendrites [28]. It is evident that a proper allocation of behaviors to various L2 neurons or groups of neurons can increase learning efficiency. For example, assume that L2 has n trainable binary- state neurons. Random search for an optimal combination Y ∗(X) for a certain X, assuming for simplicity that only a single Y ∗(X) exists, would consume ∼ 2n trials. This compares to only n trials if one neuron can be trained at a time in any order, or roughly n(n + 1)/2 trials if one neuron can be trained at a time, but in a particular order that also has to be found by trial and error. The latter training strategies would be possible if L2 neurons drove complementary motor behaviors, such as movements of legs and arms, or rough movements of a leg and finer movements of the leg. The optimal for learning layout of L2 and Li neurons should certainly be subject to major evolutionary pressures. We have so far considered independent learning for each yk neuron. However, the excitation of patterns in L2 could be coordinated, e.g., if an Li neuron drove excitation of several L2 neurons. In mammals, it is evident that the "combinatorially trainable" layer L2 and Li neurons are likely primarily located in the neocortex where they can store complex behaviors. It is suggested that the hippocampus, situated at the edges of the neocortex and indirectly projected into by it, is a major site for generation of basic cognitive and higher global reward signals, or what we suggest may be experienced as "feelings" or "emotions" in humans and higher animals, based on the hippocampus's observation of the neocortical and other brain activity, including the more primary positive and negative rewards generated in other brain regions (Fig. 13). The mechanism of generation of the global reward signal, which probably is an RPE-type signal [7, 8, 9, 10], is not the main subject of this paper. However, we note that the mechanism may employ the combinatorial switching principles discussed here to classify synaptic input patterns, although the ability to evaluate the temporal relationships between input signals and gauge the magnitude of the primary rewards would also be needed. 6 Discussion Many actions of humans and higher animals seem to fit into the following paradigm: given a combination of sensory inputs, generate an appropriate for the combination action that can be altered through learning. It would be an elegant solution of nature if individual neurons, with some help of auxiliary neuronal circuitry, in fact exhibited this basic behavior -- at the single-neuron level expressed as the combinatorial switching of the neuron's output. Indeed, pyramidal neuron connectivity suggests just that: barring necessity for system redundancy, why would a neuron's axon make multiple seemingly randomly distributed connections with 20 Input Pattern X in L1 Enhance X's excitation of Y Weaken X's excitation of Y Output Neocortex Pattern Y in L2 Generate basic reward signals Yes Basic rewards positive? No "Action" or "guessing" mechanism Hippocampus Compare evolution of (cortical) input patterns and the subsequent basic rewards to that recorded previously; generate positive reward signal if observe more positive aggregate basic reward given similar inputs No Higher rewards positive? Yes Figure 13: The suggested process of learning in mammals. The diagram is an elaboration of the process illustrated in Fig. 1(b). The hippocampus plays a role of an "observing body" or "critic" [1] that generates global reward signals. The "basic", or "primary", rewards shown may also be partly generated in the hippocampus. The framework suggests that the higher reward signals may have structurally evolved as an extension of more elementary signals such as pain or hunger. another neuron's dendrites, unless there was a combinatorial aspect that is used? On the other hand, it is widely accepted that higher organisms try to learn to respond to the environment's inputs to achieve positive and avoid negative feelings and emotions [2, 13, 14, 15, 16]; and that following these subjective learning goals is ultimately connected to the achievement of the organisms' survival and evolutionary objectives. The idea that emotions play a critical role in learning can be demonstrated with the fol- lowing example. Consider a toddler learning how to kick a ball to hit a real or imaginary target (creating an implicit, or procedural, memory) by repeatedly kicking the ball and ob- serving its trajectory. What is the mechanism that causes the motor activity associated with more successful trials to be memorized better than that associated with less successful trials, thus allowing the technique improvement? One could suggest that the child consciously and voluntarily, using some mental picture of the process, chooses to remember the movements associated with more successful trials. This would likely require a corresponding cognitive mechanism implemented at the neural level. However, this paper suggests that the positive emotions that accompany the child's realization that an attempt was successful already provide a convenient mechanism for relaying the signal of long-term memorization of the preceding spiking response to the neurons responsible for the more advantageous behavior. Indeed, the reason that emotional responses in humans and higher animals are delivered to a large num- ber (or all, via hormones) of trainable neurons [9] may be that the exact site of the neurons being trained, given the complexities of the sensory-motor signal flows, is not easily locatable 21 from the perspective of the emotion generating systems, which themselves may be scattered throughout the nervous system. An interesting question is: why would the paradigm of combinatorial switching, in which the ability to classify input patterns into output patterns can be considered a multi-neuron im- plementation, be successful in our world? The answer appears to be that, from a fundamental perspective, the world around us is indeed usefully classifiable, which is in large part driven by the repeating motives in the terrestrial environment and the life organization into similarly behaving species as well as the similarities across the species. (On an even deeper level, these regularities may be viewed as stemming from the invariance of the physical laws in space and time.) A wolf that has learned how to catch a rabbit is more likely to catch another rabbit, as well as another alike animal, in a similar terrestrial environment. The key to efficient learning with a low-dimensional feedback signal (the reward, or the "emotional response") may be the ability to distil reusable concepts in relatively few learning trials. As an illustration of these ideas consider the following simple learning model. An untrained and hungry test subject has 12 sensory neurons connecting to 3 motor neurons (Fig. 14(a)). All the neurons operate in an "on" or "off" regime. The subject is seated at a table on which apples (rounded symmetrical shape, stem on top, smooth surface) or stones (rounded symmetrical shape, no stem on top, rough surface) are placed one at a time (Fig. 14(b)). The apples and stones can be of 1 of 3 sizes (small, medium or large) and 1 of 3 colors (red, yellow or green). Each of the 3 motor neurons drives an action: eating the object on the table, pushing it off the table, or doing nothing, in which case the object is removed from the table following a delay. Each of the sensory neurons fires if its assigned object feature is present: rounded shape, symmetrical shape, stem on top, no stem on top, smooth surface, rough surface, red, yellow or green color, small, medium or large size (the total of 12 features, one feature per sensory neuron). The sensory neurons connect to the motor neuron dendrites at random locations, forming Nc clusters on each motor neuron, each cluster having nc excitatory synapses. A cluster is defined as being excited if all nc its synapses are excited. Each cluster is initially assigned a weight of 0. A neuron fires in a "learned" excitation if at least M its clusters with weights of at least 1 are excited. A weight of 0.25 is added to a cluster for eating an apple, and 0.1 for pushing an object off the table, if 1) all the cluster's synapses are excited, 2) this is immediately followed by a trial firing of the cluster's neuron and 3) this is immediately followed by a positive reward. A cluster's weight is reset to 0 if 1) all the cluster's synapses are excited, 2) this is immediately followed by a trial or nontrial (learned) firing of the cluster's neuron and 3) this is immediately followed by a negative reward (Fig. 14(d)). Positive reward is generated for eating an apple or pushing an object off the table. Negative reward is generated for eating a stone, doing nothing, or doing more than one action simultaneously (i.e., at least two motor neurons fire). After an object is placed on the table, the subject attempts to execute a memorized action. If there is no memorized action (i.e., less than M clusters with the weight of at least 1 are excited on each of the motor neurons) a random motor neuron fires in a trial firing. A computer program RMCLS (Reward-Modulated Combination Learning System) imple- mented the above learning algorithm. To complicate the problem for the subject and to test its deductive reasoning, no green or large apples and no small or red stones were presented during learning, while green large apples and small red stones were presented during testing. 22 (a) Input x1 x2 x12 (1) N2 (1) NNc (2) N1 (2) N2 (2) NNc (3) N1 (3) N2 (3) NNc Guessing input Guessing input y2 Push off Guessing input y3 Do nothing Dendritic clusters (1) N1 R Output (b) Object features: Rounded shape, x1 X Symmetrical shape, x2 Stem on top, x3 No stem on top, x4 Smooth surface, x5 Rough surface, x6 y1 Eat Apple Stone X X X X X X X (c) Additional features of eight learning (L) and four testing (T) objects: Apple Small, x7 Medium, x8 Large, x9 Green, x 12 Yellow, x 11 Red, x 10 Stone L L L L T T Small, x7 Medium, x8 Large, x9 Green, x 12 Yellow, x 11 Red, x 10 T L T L L L (d) Learning rewards and "correct" (C) test actions: Do nothing, y3 Push off, y2 Apple Stone Eat, y1 +0.25, C -1 +0.1 +0.1, C -1 -1 Figure 14: (a) Equivalent neural network diagram for the problem simulated in Sec. 6. The 12 binary inputs form Nc clusters of size nc synapses on the dendrites of each of the 3 output neurons. Each cluster outputs 1 if its weight is at least 1 and its inputs are coincident. Weights are modified during learning. An output neuron yk fires if the sum of its inputs is at least M or if it receives the "guessing" input. (b) Definition of inputs x1, . . . , x6 for the 2 object types: "X" denotes an input of "1", the input is "0" otherwise. (c) Definition of inputs x7, . . . , x12 for the 8 learning and 4 testing objects. The inputs are "0" or "1" depending on the "L"/"T" position in the matrices. (d) Rewards elicited by the firing of each of the 3 output neurons for the 2 object types. The reward of "-1" resets a cluster weight to 0 if the cluster was excited and its neuron fired. If more than 1 output fires simultaneously the reward of "-1" is generated. "C" denotes the correct output neuron for the purpose of testing. 23 RMCLS learning curves for nc=4, Nc=10000 100 10 1 % , s t s e t r u o f e h t f o y n a n o r e w s n a t c e r r o c n I M=70, random trials M=70, round-robin trials M=1, round-robin trials % , s t s e t r u o f e h t f o y n a n o r e w s n a t c e r r o c n I RMCLS learning curves for Nc=4*12nc, round-robin trials 100 10 1 nc=1, M=6 nc=2, M=34 nc=3, M=155 nc=4, M=620 20 40 60 80 100 120 140 160 20 40 60 80 100 120 140 160 Total number of objects presented Total number of objects presented Figure 15: (a) Learning curves for the RMCLS model for nc = 4, Nc = 10000, with the random or round-robin trial firing of the output neurons. The minimum number of excited clusters with the weight at least 1 required to fire a neuron, M , was either 70 or 1. Each curve is the average over 1000 statistically independent simulations. (b) Learning curves for the RMCLS model for nc = 1, 2, 3, 4 and M = 6, 34, 155, 620, respectively. Nc = 4 · 12nc and round-robin output neuron trials. The values of M were roughly optimal for the learning performance given the values of the other parameters. Each curve is the average over 1000 statistically independent simulations. Specifically, the subject was presented with a random sequence of 8 objects: small red apple, small yellow apple, medium red apple, medium yellow apple, medium yellow stone, medium green stone, large yellow stone, large green stone (Fig. 14(c)). After each presentation it was recorded whether the subject would have had correct responses (i.e., eating apples and pushing off stones), if tested, to the 4 test objects: large green apple, large red apple, small red stone, medium yellow stone (Fig. 14(c)). In some cases the system was not able to learn responses to all 4 test objects even after a large number of trials. For nc = 4, Nc = 10000 and M = 70, the subjects learned to pass all the tests correctly, after a large number of trials, in 95.5% of cases (Fig. 15(a) shows the corresponding learning curve). In the other 4.5% of cases the subjects typically learned the response of pushing both apples and stones off the table. This usually occurred when "pushing off" was, at random, tried many more times than "eating" when an apple was presented; therefore, the subjects had learned to "push off" apples before having tried to "eat" many of them. To make the trial neuron firings more regular the algorithm was modified to select the firing neurons sequentially 24 in a round-robin. Then, the subjects learned to pass the tests correctly in 98.3% of cases (Fig. 15(a) shows the learning curve). For nc = 4, Nc = 10000 and M = 1 (with the round-robin motor neuron trials), the subjects learned the 4 correct responses in only 15.3% of cases. The most common reason for failing a test was due to motor neurons being excited by rarely occurring clusters that represented low-dimensional object features. For example, a cluster with 2 inputs coming from the "rounded shape" sensory neuron and 2 inputs from the "red" sensory neuron would cause all rounded red objects to be classified as edible if the training object sequence happened to have many red apples. Note that out of the 124 = 20736 clusters representing all possible ordered permutations of 4 out of 12 inputs, 1, 14, 36 and 24 clusters encode the excitation of 1, 2, 3 or 4 particular input neurons, respectively. Therefore, requiring a minimum number of excited clusters to fire a neuron assigned lower importance to one- and two-dimensional object features relative to three- and four-dimensional features. Next, for each nc from 1 to 4 (and the round-robin motor neuron trials) the optimal for learning M was searched for, using a large Nc, Nc = 4 · 12nc , so that all possible input combinations were likely to occur in the clusters. For nc = 1 the test performance was best when M was equal to 6, with the 4 correct test responses generated in only 34.8% of cases after a large number of trials; for nc = 2, >90% correct responses were obtained for M from 33 to 35 (which represented 5.7-6.1% of all clusters); for nc = 3, >95% correct responses were for M from 115 to 197 (1.6-2.9% of clusters); and for nc = 4, >95% correct responses were for M from 339 to 904 (0.41-1.09% of clusters). All these measurements were made using 500 statistically independent simulations for each value of nc and M . Clearly, the systems with combinatorial memory (nc > 1) performed much better than those without. It is interesting that the range of M/Nc when the test success rate was greater than 95% was the highest for nc = 3. As expected, for low Nc the test performance deteriorated. For example, for nc = 4, Nc = 1000 and M = 7 the correct responses to the 4 tests were learned in 87.8% of cases. Although the RMCLS algorithm is simple, it does suggest that learning in the reward- modulated combinatorial switching framework can be rather efficient, via deduction of reusable abstract concepts. In order to deduce the reusable abstract concepts the system needs to learn in situations that display both these concepts and variability in other features. The system deduces the reusable concepts by accumulating weights for the synapse clusters that represent the concept features. Note that the resulting behavior can be described as "deductive reasoning" and will probably appear intelligent to an external observer. It is evident that in biological neuronal systems the analogues of parameters nc, M and Nc are likely to evolve to suit a particular neuron's operating environment. The presented framework does not involve value functions and is more alike to policy space search RL algorithms (e.g., evolutionary algorithms [35]) than to value-function-type RL algorithms [1]. This introduces limitations compared to many well-known value-function RL implementations. Also, it is conceptually possible to apply standard neural network RL algorithms such as the policy gradient method within the presented framework. However, learning through direct interaction with the environment seems more appropriate given that the system tends to have a very large number of weights that are expected to provide a "built-in machinery" for memorization and generalization, and it would probably be difficult to perturb that many weights in a controlled manner while searching in the policy space. 25 In summary, it is suggested that pyramidal neurons can process information by switch- ing the neuron output based on active input neuron combinations. A trial-and-error learning paradigm is presented in which an (RPE-type) reward signal that itself may adjust over time modulates the combinatorial memory that stores learned behaviors. An experimental verifica- tion of the proposed mechanisms, including the putative mechanical or muscle-like contribu- tions that can provide computational advantages to the single-neuron combinatorial switching, is needed. Acknowledgments The author gratefully thanks Prof. Michael A. Rvachov for prompting exploration of this subject and for many useful discussions. The author also thanks anonymous reviewers for their thorough and helpful reviews. References [1] R. S. Sutton, A. G. Barto, Reinforcement Learning: An Introduction (MIT Press, Cam- bridge, MA, 1998). [2] E. L. Thorndike, Animal Intelligence (Hafner, Darien, CT, 1911), p. 244. [3] R. E. Bellman, Dynamic Programming (Princeton University Press, Princeton, 1957). [4] P. R. Montague, P. Dayan, C. Person, T. J. Sejnowski, Bee foraging in uncertain environ- ments using predictive hebbian learning, Nature 377 (1995) 725 -- 728. [5] P. R. Montague, P. Dayan, T. J. Sejnowski, A framework for mesencephalic dopamine systems based on predictive Hebbian learning, J Neurosci 16 (1996) 1936 -- 1947. [6] W. Schultz, P. Dayan, P. R. Montague, A neural substrate of prediction and reward, Science 275 (1997) 1593 -- 1599. [7] R. R. Bush, F. Mosteller, A mathematical model for simple learning, Psychol Rev 58 (1951) 313 -- 323. [8] R. R. Bush, F. Mosteller, A model for stimulus generalization and discrimination, Psychol Rev 58 (1951) 413 -- 423. [9] W. Schultz, Predictive reward signal of dopamine neurons, J Neurophysiol 80 (1998) 1 -- 27. [10] W. Schultz, Behavioral theories and the neurophysiology of reward, Annu Rev Psychol 57 (2006) 87 -- 115. [11] P. W. Glimcher, Understanding dopamine and reinforcement learning: the dopamine reward prediction error hypothesis, Proc Natl Acad Sci USA 108 (Suppl 3) (2011) 15647 -- 15654. 26 [12] N. D. Daw, S. Kakade, P. Dayan, Opponent interactions between serotonin and dopamine, Neural Netw 15 (2002) 603 -- 616. [13] E. R. Hilgard, G. H. Bower, Theories of Learning, 4th edn. (Prentice Hall, Englewood Cliffs, NJ, 1975). [14] D. C. Dennett, Brainstorms: Philosophical Essays on Mind and Psychology (MIT Press, Cambridge, MA, 1981). [15] D. T. Campbell, Blind variation and selective survival as a general strategy in knowledge processes, in Self-Organizing Systems, eds. M. C. Yovits and S. Cameron (Pergamon, New York, 1960), pp. 205 -- 231. [16] G. Cziko, Without Miracles: Universal Selection Theory and the Second Darwinian Rev- olution (MIT Press, Cambridge, MA, 1995). [17] P. Poirazi, T. Brannon, B. Mel, Pyramidal neuron as two-layer neural network, Neuron 37 (6) (2003) 989 -- 999. [18] J. Tanaka, Y. Horiike, M. Matsuzaki, T. Miyazaki, G. Ellis-Davies, H. Kasai, Protein syn- thesis and neurotrophin-dependent structural plasticity of single dendritic spines, Science 319 (5870) (2008) 1683 -- 1687. [19] H. Kasai, M. Fukuda, S. Watanabe, A. Hayashi-Takagi, J. Noguchi, Structural dynamics of dendritic spines in memory and cognition, Trends Neurosci 33 (3) (2010) 121 -- 129. [20] A. Polsky, B. Mel, J. Schiller, Computational subunits in thin dendrites of pyramidal cells, Nat Neurosci 7 (6) (2004) 621 -- 627. [21] R. A. Silver, Neuronal arithmetic, Nat Rev Neurosci 11 (2010) 474 -- 489. [22] M. Hausser, B. Mel, Dendrites: bug or feature?, Curr Opin Neurobiol 13 (3) (2003) 372 -- 383. [23] M. M. Rvachev, Alternative model of propagation of spikes along neurons, (2003) arXiv:physics/0301063v2 . [24] M. M. Rvachev, On axoplasmic pressure waves and their possible role in nerve impulse propagation, Biophys Rev Lett 5 (2) (2010) 73 -- 88. [25] F. Korobova, T. Svitkina, Molecular architecture of synaptic actin cytoskeleton in hip- pocampal neurons reveals a mechanism of dendritic spine morphogenesis, Mol Biol Cell 21 (1) (2010) 165 -- 176. [26] P. Hotulainen, C. Hoogenraad, Actin in dendritic spines: connecting dynamics to function, J Cell Biol 189 (4) (2010) 619 -- 629. [27] M. London, M. Hausser, Dendritic computation, Annu Rev Neurosci 28 (2005) 503-532. [28] N. Spruston, Pyramidal neurons: dendritic structure and synaptic integration, Nat Rev Neurosci 9 (2008) 206 -- 221. 27 [29] P. Ciarletta, M. B. Amarb, Peristaltic patterns for swelling and shrinking of soft cylindrical gels, Soft Matter, 8 (2012) 1760 -- 1763. [30] H. Kasai, T. Hayama, M. Ishikawa, S. Watanabe, S. Yagishita, J. Noguchi, Learning rules and persistence of dendritic spines, Eur J Neurosci 32 (2) (2010) 241 -- 249. [31] I. R. Fiete, H. S. Seung, Gradient learning in spiking neural networks by dynamic per- turbation of conductances, Phys Rev Lett, 97(4) (2006) 048104. [32] G. Tesauro, R. Janssens, Scaling relationships in back-propagation learning: dependence on predicate order, Complex Systems, 2 (1988) 39 -- 44. [33] X. Yao, Y. Liu, A new evolutionary system for evolving artificial neural networks, IEEE Trans Neural Netw, 8 (3) (1997) 694 -- 713. [34] J. M. Yang, C. Y. Kao, A robust evolutionary algorithm for training neural networks, Neural Comput Appl, 10 (2001) 214 -- 230. [35] D. E. Moriarty, A. C. Schultz, J. J. Grefenstette, Evolutionary algorithms for reinforce- ment learning, J Artif Intell Res, 11 (1999) 241 -- 276. 28
1102.4213
1
1102
2011-02-21T13:11:04
Pattern formation and coexistence domains for a nonlocal population dynamics
[ "physics.bio-ph" ]
In this communication we propose a most general equation to study pattern formation for one-species population and their limit domains in systems of length L. To accomplish this we include non-locality in the growth and competition terms where the integral kernels are now depend on characteristic length parameters alpha and beta. Therefore, we derived a parameter space (alpha,beta) where it is possible to analyze a coexistence curve alpha*=alpha*(\beta) which delimits domains for the existence (or not) of pattern formation in population dynamics systems. We show that this curve has an analogy with coexistence curve in classical thermodynamics and critical phenomena physics. We have successfully compared this model with experimental data for diffusion of Escherichia coli populations.
physics.bio-ph
physics
Pattern formation and coexistence domains for a nonlocal population dynamics Jefferson A.R. da Cunha3,4, André L.A. Penna2,4, and Fernando A. Oliveira1,4∗ Instituto de Física - Universidade de Brasilia, Brazil1 FGA-Universidade de Brasília, Brazil2 Instituto de Física Universidade Federal de Goiás, Goiânia, Brazil3 International Center for Condensed Matter Physics CP 04455, 70919-970 Brasilia DF, Brazil,4 (Dated: August 14, 2018) Abstract In this communication we propose a most general equation to study pattern formation for one- species population and their limit domains in systems of length L. To accomplish this we include non-locality in the growth and competition terms where the integral kernels are now depend on characteristic length parameters α and β. Therefore, we derived a parameter space (α, β) where it is possible to analyze a coexistence curve α∗ = α∗(β) which delimits domains for the existence (or not) of pattern formation in population dynamics systems. We show that this curve has an analogy with coexistence curve in classical thermodynamics and critical phenomena physics. We have successfully compared this model with experimental data for diffusion of Escherichia coli populations. PACS numbers: 89.75.Kd, 89.75.Fb, 05.65.+b 1 1 0 2 b e F 1 2 ] h p - o i b . s c i s y h p [ 1 v 3 1 2 4 . 2 0 1 1 : v i X r a ∗ [email protected] 1 2 Introduction - In recent years the phenomenon of pattern formation has been intensively studied to describe the spatial distribution of species in population dynamics. This amazing behavior of populations, observed under certain conditions, can be modeled by nonlinear equations of reaction and diffusion-type [1 -- 7]. Such mathematical models provide a rich structure to include a variety of intra- or inter-specific interactions among species [8, 9], as well as permitting to describe many forms of effects of dispersal with or without memory effects [10 -- 12]. On the other hand, the overwhelming majority of the studies have shown that Fickian- type diffusion [13 -- 15] is unable to describe spreading of species in population equations formulated through reaction-diffusion models. Moreover, this is not the only possible criti- cism that one can find at the ordinary nonlinear reaction-diffusion approach. In population dynamics context, one believes that there is no real justification for assuming that interac- tion among species are, in fact, local. There are many models in which such an assumptions become clearly unwarrantable, as for example, in competition of one-species in a habitat where the system is rapidly equilibrated or in typical biological interactions where the indi- viduals intercommunicate via chemical means. Of course, these are typical nonlocal effects which should not be overlooked. Other forms of nonlocal growth and interaction effects have been also observed when we deal with wide variety of biological fields, such as epidemic spread in network [16 -- 18], embryological development, and bacterial growth [19], where the density of individuals involved are not small and the analysis of local or short-range diffusive flux is not sufficient accurate to understand the dynamical aspects of these phenomena. In these cases we need to include the contribution of long-range effects, and then to analyze the domains limits for the existence of patterns in these systems. The present communication is an attempt to build a most general equation to study effects of pattern formation in one-species population dynamics. Our starting point is to write an equation which includes nonlocal growth and interaction terms involving long-range effects in the system. This equation can be written as ∂u(x, t) ∂t = aZΩ gα(x − x′)u(x′, t)dx′ − bu(x, t)ZΩ fβ(x − x′)u(x′, t)dx′ , (1) where u(x, t) describes the population density with growth a and competition b terms. Then we denote by gα(x − x′) the correlation growth function, which weights the growth of a population in the domain Ω for a specific growth length parameter α. We call fβ(x − x′) the 3 correlation competition function which weights the interaction among the constituents of the population for a competition length parameter β in the domain Ω. In the intuit to modeling population dynamics, we have assumed that the kernels are symmetric functions, such that fβ → 0 and gα → 0 as x − x′ → ∞. An of advantages of Eq. (1) is that it provides an useful concept to describe a great variety of long-range diffusive effects in physical systems, by permitting we can absorb all higher derivatives (diffusion and dissipation terms) in only integral term. Moreover this model is then parametrized mainly by α and β quantities, and this allows us to suppose on the existence of values (α, β) for which there are pattern formation. Indeed the relation between these length domains offers a simplest model to deal with quantitative estimates of experimental data related to the growth dynamics of bacteria, for example. In this case, specific domains which show the existence (or not) of patter formation, which incorporates long-range effects as well, is important for the physical description of spreading of individuals with nonlocal dynamics. Starting from Eq. (1) we can derive important connections with classical population models by carrying out appropriated limits. For example, if fβ(x−x′) = gα(x−x′) = δ(x−x′) we get the logistic equation [15] ∂u(x, t) ∂t = au(x, t) − bu(x, t)2. (2) One can also consider gα(x − x′) with a finite range, such that we can expand the growth term as aZΩ gα(x − x′)u(x′, t)dx′ = ∞ Xm=0 ay2m (2m)! ∂2m ∂x2m u(x, t) , where y = x − x′ and the k-moments are yk = Z ykgα(y)dy . (3) (4) Using the above procedure with fβ(y) = δ(y) and retaining the first two terms in Eq. (3) we get the ordinary Fisher equation ∂u(x, t) ∂t = D ∂2u(x, t) ∂x2 + au(x, t) − bu2(x, t). (5) Here we show a very important point: the first gain with the nonlocal growth term is the possibility to connect the growth rate a with the diffusion constant D = ay2 2 . (6) 4 Note that this equation shows that a species with a large growth rate has "more need" for diffuse behavior, i.e. a large growth rate creates a large pressure proportional to the concentration gradient D ∂ ∂x u(x, t) which increases the diffusion. It shows that the diffusion is intrinsically related to the existence of the nonlocal growth term gα, i.e. it is necessary to exist a second moment y2 6= 0 in according to Eq. (4). The higher order terms (m > 1) in expansion of Eq. (3) yield the dispersive terms. It is interesting to note that if gα is not even the first derivative yields the convective term v ∂ ∂xu(x, t), where we can obtain the convective velocity as v = ya. On the other hand, if we have an asymmetric gα it corresponds to a convective drift [21]. If we keep the expansion up to second order and fβ(y) in the second integral, we get nonlocal Fisher equation ∂u(x, t) ∂t = au(x, t) + D ∂2u(x, t) ∂x2 − bu(x, t)ZΩ f (x − x′)u(x, t)dx , (7) which has been widely used by many authors to discuss pattern [9]. Note that is Eq. (1) is most general, incorporating all the previous equations. Perturbative analysis - In the study of pattern formation through a model of popu- lation dynamics it is usual to calculate a quantity known as the growth rate of pattern γ [9, 10] which leads to the pattern formation in the system. Therefore, we first shall start with the perturbative analysis through the function u(x, t) = a b + ǫ exp(cid:16)ikx + φ(k)t(cid:17), (8) where a/b is the homogeneous steady state solution, constant in space and time. The term ǫ exp(ikx) exp(φt) is a perturbation to the steady state that will grow or die out, depending on the values of the wavenumbers k. Substituting Eq. (8) into Eq. (1) and retaining only first order perturbative terms, we find a dispersion relation between the complex pattern growth rate φ and the wavenumber k, given by φ(k) = a(cid:16)Fc{gα(y)} − Fc{fβ(y)}(cid:17) + ia(cid:16)Fs{gα(y)} − Fs{fβ(y)}(cid:17) , (9) where Fc{·} and Fs{·} are, respectively, the Fourier cosine and sine transform of the influence function gα(y) and fβ(y) (assumed to be even). Therefore, only the real part of the complex growth rate φ(k) = γ(k)+iξ(k), where γ(k) = a(cid:16)RΩ gα(y)cos(ky)dy−RΩ fβ(y)cos(ky)dy−1(cid:17), i.e. the growth rate of pattern γ(k), will be important to determine whether the perturbation with wavenumber k will die out or will generate a pattern, for negative or positive values, 5 respectively. Now let us consider the simple case of the square interaction influence function given by fβ(y) = 1 2β(cid:2)Θ(β − y)Θ(β + y)(cid:3) , (10) where Θ refers to the Heaviside function and β is the cut-off range (0 < β < L, where L is the size of the system). If we consider similar relation for gα with cut-off 0 < α < L, from the condition (10), γ(k) is given by γ(k) = a(cid:16) sin(kα) kα − sin(kβ) kβ − 1(cid:17). (11) Therefore we can study self-organization of the equation (1) considering that the system depends physically on the domain of the functions gα and fβ. In this case, pattern formation appears when wave numbers k, in growth rate of pattern, obey the condition γ(k) > 0. Note that for α = 0, γ(k) is larger. This will happen for lower diffusive systems D ≈ 0. The Eq. (11) is plotted in Fig. 1 for different values of the growth length α = (0.01, 0.03, 0.09, 0.40) with competition length parameter β = 0.45 fixed. In this figure, we verified that when α < β we may have γ(k) > 0. If α → β, the function γ(k) becomes negative. This behavior of γ(k) is very important to determine if we have a large or a negligible amplitude of pattern. We show in Fig. 1 that pattern formation appears for values α < β with γ(k) > 0. This behavior is also verified later with numerical results, see Fig. 4, and discussed through experimental values. ) k ( γ 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8 -1 -1.2 β=0.45 a=1.0 α=0.001 α=0.03 α=0.09 α=0.40 0 5 10 15 20 k 25 30 35 40 FIG. 1: The real part of growth exponent γ(k) as a function of k plotted for different values of the correlation length of growth α with a length interaction of individuals β fixed. The pattern formation appear for those values of k for which γ is positive. Numerical method - To solve Eq. (1) numerically, we applied the Operator Splitting Method (OSM) [20]. By this method, the operator of the differential equation is split into several parts, which act additively on u(x, t). If we write Eq. (1) as 6 where T is the total operator, then ∂u(x, t) ∂t = T u(x, t), T u(x, t) = Tgrowu(x, t) + Tintu(x, t), with Tgrowu(x, t) = aZΩ gα(x − x′)u(x′, t)dx′ Tintu(x, t) = −bu(x, t)ZΩ fβ(x − x′)u(x′, t)dx′. (12) (13) (14) (15) In the latter equations Tgrow and Tint are nonlocal growth and nonlocal interaction operators, respectively. In our numerical calculations, we have used periodic boundary conditions u(x = 0, t) = u(x = L, t) with spatial period L. For each part of the operator, we apply a known difference scheme for updating the function u(x, t) from step j to step j + 1. In Fig. 2 we show the evolution of u(x, t). We start with a distribution of individuals u(x, 0) = 1 Γ exp(cid:20)− (x − x0)2 2σ2 (cid:21) , (16) where Γ = p π 2 σherf(cid:16) x0√2σ(cid:17) + erf(cid:16) L−x0√2σ (cid:17)i, and we see the evolution to a state which exhibits pattern. We use σ = 0.3, x0 = 0.5 and L = 1.0. The spacial and time increments are δx = 1 × 10−3 and δt = 1 × 10−2. In all simulations we use a = b = 1.0. Several numerical experiments show the final state independent of the initial conditions [21]. These simulations are fundamental to show the pattern formation that appears after a long time for bacterial growth [8 -- 11, 22]. Similar simulations are used to compose Fig. 3 and Fig. 4. In Fig. 3, we show the evolution after 60000 time steps where the density has reached its final form. Each curve is similar to the simulations described in Fig. 2. The first column is for β = 0.07 and for up to down α = (0.009, 0.012, 0.019). The second column has β = 0.11 and for up to down α = (0.012, 0.020, 0.030). For each β the lower curve represents the hight value of α, for which there is no more pattern formation, i.e. for α ≥ α∗, where we get u(x, t) = a/b. 7 u(x,t) u(x,t) 10 8 6 4 2 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 x 0 10 20 30 t 40 50 60 70 FIG. 2: The typical pattern formation on density u(x, t) as a function of x and t in arbitrary units. The growth rate and interaction rate are a = b = 1.0. The competition length parameter β = 0.15 and the growth length parameter α = 0.009. In Fig. 4, we show the region in the space (α, β) where pattern can exist. For each β the points represent the α∗ above which there is not more pattern, such as described in Fig. 3. The shadow area is limited by the coexistence curve α∗(β) = P (β)(βc − β)µ, (17) which is the best fit of the points. P (β) is a polynomial with no roots in the region 0 < β < βc. For β → βc we get from Eq. (10) βc = 1/2, i.e the function fβ(y) weights equally the all the space 0 < x < L, and consequently we have no pattern formation for β ≥ βc. Finally, from the data we have estimated the value µ = 0.53 ± 0.06 at the vicinity of βc for the exponent in Eq. (17). Experimental data - Starting from Eq. (10) we can compute y2 = α/3. By inserting this result into Eq. (6) we get α = r 6D a . (18) Now using the experimental values for a = (2.23 ± 0.2) × 10−4s−1 and D = (2.2 ± 0.2) × 10−5cm2s−1 obtained by N. Perry [22], for systems with Escherichia coli populations, we can estimate the value of α, given by α = (7.70 ± 0.09)mm. Then to form pattern in a finite system of length L, β must be inside the shadow area of Fig. 4. In fact, the value of α as obtained in Eq.(18) is determined by the coefficient of diffusion D of system and it establishes 5 4 β=0.07 α=0.009 ) x ( u 3 2 1 0 5 4 0 0.2 0.4 0.6 0.8 1 x β=0.07 α=0.012 ) x ( u 3 2 1 0 5 4 0 0.2 0.4 0.6 0.8 1 x β=0.07 α=0.019 ) x ( u 3 2 1 0 0 0.2 0.4 0.6 0.8 1 x 8 6 5 4 ) x ( u 3 2 1 0 6 5 4 ) x ( u 3 2 1 0 6 5 4 ) x ( u 3 2 1 0 β=0.11 α=0.012 0 0.2 0.4 0.6 0.8 1 x β=0.11 α=0.020 0 0.2 0.4 0.6 0.8 1 x β=0.11 α=0.030 0 0.2 0.4 0.6 0.8 1 x FIG. 3: Snapshots of the stationary state u(x) for some values of competition length parameter β and growth length parameter α with correlation competition function and correlation growth function Eq. (10). In this snapshots we consider the growth rate a and competition rate b as 1.0. For a fixed β as α increases the pattern disappears. 0.1 0.08 0.06 0.04 0.02 α No Pattern Pattern 0 0 0.1 0.2 0.3 0.4 0.5 β FIG. 4: The phase diagram of critical correlation growth length αc as a function of critical competition length parameter βc. The Pattern and Non Pattern region indicate the separation of large-amplitude patterns and negligible-amplitude patterns for Eq.(16). a lower referential limit for the presence (or not) of pattern formation. For β ≪ L, we are in 9 the linear part of Eq. (17) and we get α = 0.3β, therefore we have pattern for β > 25.4mm. In this case, only experimental values of width of influence function β > 25.4mm permit patterns, which are in concordance with our theoretical and numerical results. Moreover our formulation allow us to analyze pattern formation as an interplay between two length parameters α and β. It is important to note that the fact that 0 < α < β is not just a curiosity of the theory, it is one of its major result. Without a finite value of α, there will be no diffusion, which is fundamental for any species, and so reproduction and propagation are associated. Consequently one should expect a non null α. On the other hand if α is too large, bonds are tight, and they may face difficult to meet and consequently to reproduce. This phenomena described here for pattern formation in bacterial colony can be observed in large animals with migrate habits, such as deer and woolf, they travel with the only family, we call this phenomena the faithful sailor travel. Conclusion- The presence of memory, non-locality in time, have been used to explain ergodicity violation in particle diffusion [23, 24]. Since pattern formation implies in ergodicity breaking, one could expect that a nonlocal space kernel would yield that. Consequently, we proposed here a new formulation for population dynamics, which includes a growth and a competitive nonlocal terms. The presence of two kernels gα(x) and fβ(x) demand the existence of a growth length parameter α and of a competition parameter β. Particular values for the kernels yield most of the previews formulations of population dynamics. We obtain a domain region 0 < α < β where patterns may arise, a coexistence curve similar to those in phase transition, and a direct connection between the diffusion constant D the growth rate a and the mean square deviation y2 = R fα(y)y2dy which is a function of α. More results can be obtained from this formulation, however there are some restrictions, we need more detailed dynamical experiments in growth, in such way that we can propose more elaborated kernels. Those present here, Eq (4), gives us a rough idea of the dynamics. More accurate gα(x) and fβ(x) will permit us to get a better description and a generalization for two dimensions. [1] M.C. Cross and P.C. Hohenberg, Rev. Mod. Phys. 65, 851 (1993). [2] P.C. Fife, J. Chem. Phys. 64, 554 (1976). 10 [3] A.M. Zhabotinsky, M. Dolnik, and I. R. Epstein, J. Chem. Phys. 103, 10306 (1995). [4] B. Legawiec and A. L. Kawczynski, J. Phys. Chem. A 101, 8063 (1997). [5] V.K. Vanag, A.M. Zhabotinsky, and I.R. Epstein, Phys. Rev. Lett. 86, 552 (2000). [6] P. De Kepper, E. Dulos, J. Boissonade, A. De Wit, G. Dewel, and P. Borckmans, J. Stat. Phys. 101, 495 (2000). [7] A.L. Kawczynski and B. Legawiec, Phys. Rev. Lett. 63, 021405 (2001). [8] A.M. Delprato, A. Samadani, A. Kudrolli and L. S. Tsimring, Phys. Rev. Lett. 87, 158102 (2001). [9] M.A. Fuentes, M. N. Kuperman, and V. M. Kenkre, Phys. Rev. Lett. 91, 158104 (2003). [10] V. M. Kenkre, Physica A 342, 242 (2004). [11] M.A. Fuentes, M.N. Kuperman, and V.M. Kenkre, J. Phys. Chem. B 108, 10505 (2004). [12] A.S. Mikhailov and K. Showalter, Phys. Rep. 425, 79 (2006). [13] M. G. Clerc, E. Tirapegui, and M. Trejo, Phys. Rev. Lett. 97, 176102 (2006). [14] D. Bolster, D.A. Benson, T. Le Borgne, and M. Dentz, Phys. Rev. E 82, 021119 (2010). [15] J. D. Murray, Mathematical Biology, 2nd ed (Springer, New York, 1993). [16] E. Kenah and J.M. Robins, Phys. Rev. E 76, 036113 (2007). [17] M. Barthelemy, A. Barrat, R. Pastor-Satorras, and A. Vespignani, Phys. Rev. Lett. 92, 178701 (2004). [18] M. E. J. Newman, Phys. Rev. E 66, 016128 (2002). [19] J. Mueller and W. van Saarloos, Phys. Rev. E 65, 061111 (2002). [20] H.P. William, A.S. Teukolsky, T.V. Vetterling and P.B. Flannery, Numerical Recipes in C, 2nd ed (Cambridge University Press, New York, 1992). [21] J.A.R. da Cunha, A.L.A. Penna, M.H. Vainstein, R. Morgado, and F.A. Oliveira, Phys. Lett. A, 373, 661 (2009). [22] N. Perry, J. R. Soc. Interface 2, 379 (2005). [23] L.C. Lapas, R. Morgado, M.H. Vainstein, J.M. Rubi, and F. A. Oliveira, Phys. Rev. Lett. 101, 230602 (2008). [24] S. Burov, R. Metzler, E. Barkai, PNAS, 107, 13228 (2010).