A single-chain model for active gels I: active dumbbell model

Andrés Córdobaa, Jay D. Schieber*ab and Tsutomu Indeia
aDepartment of Chemical and Biological Engineering and Center for Molecular Study of Condensed Soft Matter Illinois Institute of Technology 3440 S. Dearborn St., Chicago, Illinois 60616, USA. E-mail: schieber@iit.edu
bDepartment of Physics Illinois Institute of Technology 3101 S. Dearborn St., Chicago, Illinois 60616, USA

Received 21st January 2014 , Accepted 26th March 2014

First published on 27th March 2014


Abstract

We have developed a single-chain theory to describe the dynamics of active gels. Active gels are networks of semiflexible polymer filaments driven by motor proteins that convert chemical energy from the hydrolysis of adenosine triphosphate to mechanical work and motion. In active gels molecular motors create active cross-links between the semiflexible filaments. We model the semiflexible filaments as bead-spring chains; the active interactions between filaments are accounted for using a mean-field approach, in which filaments have prescribed probabilities to undergo a transition from one motor attachment state into the other depending on the state of the probe filament. The level of description of the model includes the change in the end-to-end distance of the filaments, the attachment state of the filaments, and the motor-generated forces, as stochastic state variables which evolve according to a proposed differential Chapman–Kolmogorov equation. The motor-generated forces are drawn from a stationary distribution of motor stall forces that can be measured experimentally. The general formulation of the model allows accounting for physics that is not possible, or not practical, to include in available models that have been postulated on coarser levels of description. However, in this introductory manuscript, we make several assumptions to simplify the mathematics and to obtain analytical results, from which insight into the microscopic mechanisms underlying the dynamics of active gels can be gained. We treat the filaments as one-dimensional dumbbells, approximate the elasticity of the semiflexible filaments with a Hookean spring law, and assume that the transition rates are independent of the tension in the filaments. We show that even in this simplified form, the model can predict the buckling of individual filaments that is thought to be the underlying mechanism in the self-contraction of non-sarcomeric actin–myosin bundles [Lenz et al., Phys. Rev. Lett., 2012, 108, 238107]. The active dumbbell model can also explain the violation of the fluctuation–dissipation theorem observed in microrheology experiments on active gels [Mizuno et al., Science, 2007, 315, 370–373].


I. Introduction

Active gels are networks of semiflexible polymer filaments driven by motor proteins that can convert chemical energy from the hydrolysis of adenosine triphosphate (ATP) to mechanical work and motion. Active gels are ubiquitous in living tissue—the cell cytoskeleton is an active gel composed of many different types of filaments and motors. Active gels play a central role in driving cell division and cell migration,1,2 and have also been successfully prepared in vitro to study their mechanical and rheological properties.3–5 The semiflexible filaments that form active gels, such as actin and tubulin, are characterized by having a persistence length (length over which the tangent vectors to the contour of the filament remain correlated) that is much larger than the size of a monomer, and larger than the mesh size of the network, but typically smaller than the contour length of the filament. For instance, filamentous actin (F-actin) has persistence length of around 20 μm, while the mesh size of actin networks is estimated to be approximately 0.2 μm.6–8 This sets them apart from flexible networks where the persistence length of the polymeric chains is much smaller than the mesh size of the networks formed by those chains. An important mechanical characteristic of semiflexible networks is that they exhibit significant strain hardening for modest strains. A tension of a few pN can increase the modulus of a semiflexible network by a factor of 100.6,7

Molecular motors are proteins with a rigid, roughly cylindrical, backbone with clusters of binding heads on both ends that can attach to active sites along semiflexible filaments.9,10 In the absence of ATP they act as passive cross-links between the semiflexible filaments. In the presence of ATP molecular motors can “walk” along the filaments. The direction in which motors move is determined by the filament structural polarity.11,12 A molecular motor starts “walking” when an ATP molecule attaches to a binding head domain of the motor protein, which causes it to detach from the filament. Using the chemical energy from the hydrolysis of ATP the detached motor head moves towards the next attachment site along the filament contour and reattaches, in a process known as the Lymn–Taylor cycle.13 Each motor has at least two clusters of binding heads performing this same process. However, both of them do not necessarily detach at the same time. When the binding heads on one end of the motor are detached and moving towards the next attachment site, the heads on the other end can be attached to a different filament. This filament will feel an extensional or compressive force due to the motion of the motor. Therefore, in a network molecular motors can generate active, pair-wise interactions between filaments. The forces generated by the motors when they move along the filament are a function of the chemical potential difference between ATP and its hydrolysis products, which in itself is a function of the local concentration of ATP. In vivo molecular motors operate far from equilibrium. In typical active gels found in living cells the difference in chemical potential between ATP and its hydrolysis products is on the order of 10 kBT.12,14 Examples of molecular motors that have been extensively studied are myosin which moves along actin filaments; and kynesin and dynesin which move along microtubules.

Recent advances in experimental techniques have allowed the characterization of mechanical and rheological properties of active gels. These active polymeric networks have shown fundamental differences from their passive counterparts. The differences are not surprising given that these are materials in which molecular motors continuously convert chemical energy into mechanical work. Recent microrheology experiments3–5 on active gels have shown that the fluctuation–dissipation theorem (FDT) and the generalized Stokes–Einstein relation (GSER) are violated in active gels. The FDT is a central part of data analysis of passive microrheology experiments, where it is used to relate the position fluctuations of the probe bead to a frequency-dependent friction coefficient, from which, by using a generalized Stokes relation rheological properties can be extracted.15–17 The violation of FDT is observed as a frequency-dependent discrepancy between the material response function obtained from active and passive microrheology experiments. In the active experiments an external force is applied to the probe bead and the material response function is calculated from the bead position signal.17 In passive microrheology experiments, no external force is applied, and the material response function is calculated from the bead position autocorrelation function using the FDT.5 Other microrheology experiments in active gels3,5 also indicate that the activity of molecular motors can produce significant strain hardening of the active networks. This is observed in the non-Gaussian statistics of the probe bead position or as an overall progressive increase in the magnitude of the modulus of the gel upon addition of ATP.

From the theoretical perspective, in the recent past a considerable amount of work has been devoted to deriving simple generalizations or extensions of the FDT for out-of-equilibrium systems, such as active gels.18–20 In general, these extensions of FDT to out-of-equilibrium systems model the non-equilibrium forces as Brownian forces, but introduce an effective temperature,21–24 which is higher than the real temperature and is meant to account for the larger magnitude of the non-equilibrium fluctuations. This kind of approach can not explain the observations in the microrheology experiments in active gels5 since Brownian forces alone can not produce a frequency-dependent discrepancy between the material response obtained from the material's spontaneous stress fluctuations, and the material response obtained by applying a small external perturbation and observing the material response. Other works25,26 have modeled the attachment/detachment dynamics of motor forces as a stochastic jump process. This approach has been successful in describing some of the features observed in the microrheology experiments of active gels, such as diffusive behavior of tracer beads at frequencies where storage modulus of the gel has plateau behavior. However these models assume that motors can be described as force dipoles inside a continuum, which is an assumption that can easily break down for semiflexible networks, where the mesh size is smaller than the persistence length of the filaments. They also assume that the motors do not interact through the strain field in the network and neglect strain hardening. Given the level of description of such models, removal of these assumptions is difficult; therefore more microscopic models are required to elucidate the specific effect that these physical features have on the rheology of active gels.

Another important mechanical feature of active gels that has been extensively studied experimentally, is their capacity of self-contraction and self-organization.1,27–30 These mechanical features of active gels play a central role in cell division and motion. It has recently been shown that self-contraction in F-actin gels, is controlled by the buckling of individual filaments.27,28 This is known to be caused by the activity of myosin motors on F-actin filaments which support large tensions but buckle easily under piconewton compressive loads.1,8 Several works have used a continuum-mechanics level of description to model self-organization31,32 and rheology22,33,34 of active gels and fluids. More microscopic models have described active gels using a master equation for interacting polar rods; Aranson and Tsimring32 presented analytic and numerical results for rigid rods, later Head, Gompper, and Briels35 presented more detailed numerical simulations that account for filament semiflexibility. These models have been very successful in describing large scale phenomena such as the formation and dynamics of cytoskeletal patterns (e.g.: asters, vortices). However, the precise microscopic mechanisms underlying this process are still the subject of considerable experimental and theoretical investigation. Recently, a microscopic single-filament mean-field model to describe myosin-induced contraction of non-sarcomeric F-actin bundles was postulated by Lenz et al.27 In this work we use a similar description. However there are several issues in the level of description and mathematical formulation of Lenz et al.27 that we discuss and reformulate in this manuscript.

The main objective of this manuscript is to introduce a single-chain mean-field model for active gels. We present a level of description for active gels that has the minimum necessary components to predict mechanical and rheological features that have been observed in active gels. The general formulation of the model can account for many of the known microscopic characteristics of active gels. For example, motor concentration, the non-linear elasticity of semiflexible filaments, the dependence of motor attachment/detachment rates on the filament tension, and details about the motor-generated forces. However, in this introductory paper we make some assumptions that allow us to simplify the mathematics while still keeping most of the relevant physics in the model. Numerical simulations of more general formulations are postponed to a future manuscript. In Section II we give a detailed description of the model, and discuss the main assumptions, and parameters. In Section III we show that the active dumbbell model can predict the violation of the FDT observed in microrheology experiments of active gels. This is done by comparing the dynamic modulus of the gel obtained from the autocorrelation function of stress at the non-equilibrium steady state, with the modulus obtained from the stress response when a small perturbation is applied to the gel. In Section IV we illustrate the use of the active dumbbell model to make predictions about the underlying mechanism of self-contraction in active F-actin bundles.

II. Description of the active dumbbell model

In this section we introduce a single-chain mean-field model for an active bundle. Active networks are formed by semiflexible filaments and molecular motors that form active cross-links between filaments (Fig. 1A). In the presence of ATP, motors will go through detachment/attachment cycles13 in which they detach from one of the filaments that they are linking and step forward (in a direction determined by the filament's structural polarity) therefore exerting a force on the other filament to which they remained attached. Our model follows a single probe filament (illustrated in gray in Fig. 1A) and approximate its surroundings by an effective medium made of point-like motors that attach and detach from specific sites along the probe filament. The motors are assumed to form pair-wise interactions between filaments. When a motor is attached to the probe filament it is detached and steps forward on another filament in the mean-field and therefore pulls/pushes on the probe filament. The formulation of temporary network models provides a very useful mathematical and conceptual framework to model these dynamics.36–38 In Fig. 1B a probe filament is shown in more detail, each bead represents an active site along the filament where a motor can attach, and the springs represent the filament segments between this active sites. For the dumbbell version of the model (shown in Fig. 1C) there are only two active sites per filament (at the ends) where motors attach. The motors create the type of pair-wise active interactions between the dumbbells described above. To model these interactions we use a mean-field approach, in which filaments have certain probabilities to undergo a transition from one attachment/detachment state into another depending on the state of the particular filament. The integer variable s is used to label the attachment/detachment state of the dumbbell.
image file: c4ra02262a-f1.tif
Fig. 1 Sketch of the single-chain mean-field model for active gels. (A) Active bundle formed by polar filaments and motors (which can move towards the barbed end of the filaments). Motors attach and detach from the filaments. After detaching from a given filament a motor will step forward in that filament and will exert a force on the other filament where it is still attached. The gray filament indicates a probe filament whose dynamics are followed by the model. (B) The probe filament is represented by a bead-spring chain. Red beads represent attachment sites in the filament where a motor is attached, ζa is the friction coefficient of those beads. Blue beads represent sites in the filaments where no motor is attached ζd is the friction coefficient of those beads. [small script l]0 is the rest length of the filament before addition of ATP; r is the change in the end-to-end distance of the dumbbell due to motor activity. Fj is a motor-generated force acting on bead j. Motors generate a force on the filament only when attached. (C) Sketch of the attachment/detachment states of a dumbbell version of the model. The attachment/detachment states model the interaction of the probe filament with the mean-field.

We assume that before addition of ATP the distance between motors (acting as passive cross-links) is given by [small script l]0. This is the rest length of the filament segment and therefore there is no tension on the filaments before addition of ATP. After addition of ATP the motors (cross-links) become active and start detaching from and reattaching to the beads. τd is the average time a motor spends attached to a bead before detaching from it, whereas the average time a motor spends detached before reattaching is given by the model parameter τa. Active gels can have, besides motors, permanent passive cross-links; however in the model presented here, all cross-links are allowed to become active, we do not consider permanent cross-links in this manuscript. The force generated by a motor attached to bead j will be denoted Fj. Molecular motors can only move in one direction along the filament, determined by the filament's polarity. In this single-chain description we introduce this asymmetry by making all the forces Fj, that the motors exert on the beads of a given dumbbell, have the same sign (either positive or negative). Another force acting on the filament is the viscous drag from the surrounding solvent (which is mainly water for these biological networks). The frictional force from the surrounding solvent is characterized by a friction coefficient. Motors are expected to increase the friction coefficient of the filament when attached to an active site. Therefore this friction coefficient is allowed to take two different values: ζa when attached, and ζd < ζa if there is no motor attached to that bead. The change in the end-to-end length of the dumbbell due to the action of the motors is denoted r (see Fig. 1). We emphasize that the strands (filament segments between motors) have non-zero rest length [small script l]0 and the motors generate a change r in this length. The strands do not collapse to zero length, their end-to-end lengths fluctuate around [small script l]0 due to motor activity.

The general physical picture of our model, that is, the single-chain description and the representation of semiflexible filaments as bead-spring chains are the same as in a previous model proposed by Lenz et al.27 As discussed in Section I, this model has been used by its authors to study buckling in non-sarcomeric actin bundles. However the level of description (state variables of the model) and mathematical formulation of our model does not follow the work of Lenz et al.27 Instead we use a mathematical formulation similar to the one used in temporary network models of associating polymer chains.36–38

The following state variables (level of description) are used to construct the model of the active dumbbell Ω:{s, F1, F2, r}. Now let ψ(Ω) be the probability density describing the probability of finding an active dumbbell in state s with a change in its end-to end distance r due to motor forces F1 and F2 at time t. The time evolution for ψ(Ω) is given by the differential Chapman–Kolmogorov equation

 
image file: c4ra02262a-t1.tif(1)
where [small omega, Greek, circumflex]:{s, F1, F2} is a subspace of Ω and ε(t) is an externally applied strain in the direction of filament alignment. On writing eqn (1) we have assumed that the semiflexible segment can be described as a Fraenkel spring with spring constant kb. This assumption, although it is only expected to be valid for very small deformations (semiflexible filaments are known to strain harden under a tension of a few pN, which motors are known to generate6,7) allows us to proceed analytically with the solution of the model. For a semiflexible filament segment MacKintosh, Käs, and Janmey39 derived a relation between the persistence and rest lengths of the semiflexible filament and its linear elastic response constant image file: c4ra02262a-t2.tif, we use this relation throughout this paper. For F-actin filaments, the persistence length, [small script l]p is around 10 μm, [small script l]0 ∼ 1 μm, and therefore kb is on the order of 1 μN m−1.

The function A(s), in eqn (1) gives the connectivity of the beads as a function of the motor-attachment state, s, and is defined as,

 
image file: c4ra02262a-t3.tif(2)
where, as stated above, ζa is the friction coefficient of a bead when a motor is attached to it and ζd is the friction coefficient when there is no motor attached to it. For instance, several experimental measurements27,28 have shown that myosin II motors move along actin filaments at approximately 0.5–1 μm s−1, and they have an average stall force of 1 pN, therefore ζa is estimated to be around 1 μN·s m−1. ζa is expected to be larger than ζd since the motor attachment heads increase the cross-sectional area of the actin filament when they attach to it. Additionally to τa and τd it is convenient to label two additional time scales of the model τr,a = ζa/kb and τr,d = ζd/kb which are local relaxation times of the filament when a motor is attached to it and when there is no motor attached to it respectively. For myosin motors in actin gels τd is on the order of 200 ms while τa is usually between an order to two orders of magnitude smaller.25,27 Therefore in a typical active gel τa < τdτr,d < τr,a is expected.

The transition rate matrix image file: c4ra02262a-t55.tif in eqn (1) contains the transition rates between attachment/detachment states,

 
image file: c4ra02262a-t4.tif(3)
Where δ(...) is the Diract delta function, and for example, the second row of the first column gives the rate at which a dumbbell in state s = 0 jumps to state s = 1. The diagonal elements of the transition matrix are given by image file: c4ra02262a-t5.tif. The function p(F) is the probability distribution from which a motor force is drawn every time a motor attaches to a bead. Motor force distributions have been measured experimentally in actin–myosin bundles.27,28 The first moment of this distribution is approximately Fm = 1 pN, where Fm is usually referred to as the motor-stall force. The exact shape of these distributions can be incorporated into the model, however in this manuscript we will use simplifying assumptions about this distribution which will be discussed in Section III. Note that we have assumed that the motor force distribution is time independent, which is expected to be a good assumption as long as the kinetics of the ATP → work reaction in the motors are much faster than any of the time scales in the model. We have also assumed that the τa and τd are independent of time and of the tension on the filament. This might be unrealistic since it has been shown that the tension on the filaments can affect attachment and detachment rates of the motors.40 To summarize, we use a single-filament model of the active gel where binary active crosslinks with other filaments are accounted for by a mean-field of point-like motors that undergo transitions between attachment states determined by the phenomenological parameters τa, τd and p(F).

Note that in the model presented here p(F) is bead independent in contrast to the model by Lenz et al.27 In that model the transition rates do not depend on the motor force distribution but the motor force distribution depends on the position along the filament of the bead to which the motor is attached. Physically this assumption can be interpreted as the motors having spatial memory and being able to identify the particular position along the filament to which they are attaching. Presumably these assumptions are done to simplify the mathematics; however this leads to several issues in the solution and interpretation of the model predictions that have not been clearly resolved. We avoid such assumptions. Instead, F is kept as a stochastic state variable, p(F) is the same far all motors, and the transition rates depend on this motor force distribution. In other words, our model exists at a more-detailed level of description than the model of Lenz et al.27 In Section IV we will discuss the implications that these conceptual differences between the two models have in the explanation of the mechanisms underlying buckling and self-contraction in active gels.

In eqn (1) Brownian forces have not been included. To account for Brownian forces a term image file: c4ra02262a-t6.tif should be added inside the square bracket on the right hand side of eqn (1). In this manuscript we omit this term since it does not make any difference for the proof-of-concept calculations presented here. In other words, the version of the model presented in this manuscript represents an active limit, where the diffusive motion of the filaments due to thermal forces is negligible compared to the driven motion produced by motor activity. Since all terms of eqn (1) are linear, and motor and Brownian forces are in general uncorrelated,25,26 a solution that includes Brownian forces will simply be the superposition of the solution presented here plus the well known solution for a passive dumbbell.41 The passive part of the solution will not contribute to the active gel features we are investigating. For instance, it has been extensively shown that single-chain temporary network models that include Brownian forces, but not motor forces, satisfy the fluctuation-dissipation theorem.38,42

It is common for active gels to contain permanent passive cross-links such as biotin or α-actinin in actomyosin networks. In those cases myosin contracts F-actin into dense foci around the permanent cross-links. Once contracted, these aggregates can undergo further coalescence with each other and may form structures such as asters or vortices. The present formulation of our single-chain mean-field model does not yet seem appropriate for describing that type of phenomena. Other descriptions such as multi-chain models35 or coarser levels of description31,32 have been used to describe such observations. However other experiments28,43 have shown that local contraction can also occur in active gels containing motors alone. For instance sufficiently high density of myosin motors in the absence of passive cross-linkers has been shown to cause contractility in actomyosin bundles.28 In these conditions, the length scale over which contraction occurs within the network is proportional to the F-actin length, consistent with poor network connectivity by myosin motors. Through the addition of passive cross-linkers the length of contraction is increased to macroscopic length scales.43,44 The model presented here is expected to describe well the dynamics of active gels that do not contain permanent passive cross-links or sarcomeric organization.

In Sections IV and III we show that the model described above presents some of the main mechanical and rheological characteristics observed in active gels. The predictions here are expected to be only qualitative since several strong assumptions have to be made to obtain analytic solutions of the model: (i) aligned one-dimensional filaments, (ii) filaments with only two attachment sites at its ends (dumbbells), (iii) filaments modeled as Hookean springs, and (iv) attachment/detachment rates independent of filament tension. There is experimental evidence indicating that often some of these assumptions break down in active gels, but the calculations presented here illustrate the fundamental elements that a single-chain model needs to describe the main mechanical and rheological characteristics of active gels. Removal of these assumptions is straightforward with this approach.

III. Dynamic modulus of active gels

To test the validity of FDT in active gels, Mizuno et al.5 compared the complex compliance of actin–myosin networks measured with active and passive microrheology. In the active experiment they used optical tweezers to apply a small-amplitude oscillatory force to the bead. The complex compliance or the dynamic modulus of the medium can be calculated from the bead position signal.45 In a passive microrheology experiment no external force is applied to the bead or a static harmonic trap is used to hold the bead near its equilibrium position. In this case the autocorrelation of the bead position is used to estimate the dynamic modulus using the FDT and a generalized Stokes relation.15,16 Before addition of ATP the complex compliance of the actin–myosin network obtained with the passive and active techniques were identical. In gels activated with ATP a frequency-dependent discrepancy between the complex compliance obtained from the passive and active experiments appears at frequencies below 10 Hz. This discrepancy is due to a frequency-dependent increase in the magnitude of the bead fluctuations observed in the passive microrheology experiment, in the actin–myosin gels activated with ATP. These increased fluctuations have been linked to motor activity.25,26,46 Since the FDT does not account for motor activity, the dynamic modulus calculated from the fluctuations using the FDT does not agree with the modulus obtained from the active experiment.

In this section, we perform calculations with the active dumbbell model that test the validity of FDT in active gels in a way similar to the microrheology experiments of Mizuno et al.5 We begin by calculating the relaxation modulus of the active gel from the autocorrelation function of stress at the non-equilibrium steady-state. In other words, we apply in an active gel the famous Green–Kubo formula which relates the autocorrelation function of stress with the relaxation modulus of a material.47 In a second calculation we apply a small-amplitude oscillatory strain to the active gel and estimate the dynamic modulus from the observed stress. If the active dumbbell model satisfies the FDT the dynamic modulus obtained from those two calculations should be the same; if it does not, a frequency-dependent discrepancy should appear. Similar calculations are often performed for single-chain models of temporary networks, to demonstrate thermodynamic consistency (FDT compliance).38,42

A. Dynamic modulus of active gels from a Green–Kubo formula

For the derivation of the dynamic modulus we introduce the conditional probability ψst([small omega, Greek, circumflex], r; t|[small omega, Greek, circumflex]0, r0; 0) (where [small omega, Greek, circumflex]:{s, F1, F2}) that an active dumbbell at steady sate is in attachment state s with a change in its end-to end distance r due to motor forces F1 and F2 at time t when it had the initial conformation r0 and the initial attachment state s0 at t = 0. The time-evolution of ψst([small omega, Greek, circumflex], r; t|s0, r0; 0) obeys the same evolution equation as eqn (1) for an arbitrary initial state (r0, s0),
 
image file: c4ra02262a-t7.tif(4)
where the initial condition for the probability distribution function is given by ψst(ω, r; t = 0|[small omega, Greek, circumflex]0, r0; 0) = δ(rr0s,s0. The function A(s) was defined in eqn (2) and the transition rate matrix image file: c4ra02262a-t56.tif in eqn (3), both in Section II. Eqn (4) can be solved exactly using the same procedure outlined in Appendix A for eqn (1). If we make the motor force distribution a Dirac delta function centered around Fm, p(F) = δ(FFm), where Fm is a mean motor stall-force, all calculations can be performed analytically. At the end, the procedure reduces to solving a system of ordinary differential equations, eqn (A6a) and (A6b), with initial conditions 〈r{s,F}(t = 0) = 〈r{0,s} and 〈r2{s,F}(t = 0) = 〈r2{0,s}. The Green–Kubo formula requires the calculation of the autocorrelation function of stress at steady-state 〈σ(0)σ(t)〉st. Where σ, for this 1-dimensional system, is the normal stress in the direction of filament alignment. For a bead-spring single-chain model of a network, such as the one being considered, the macroscopic stress is related to the tension on the filaments by σ = −ncfr = nckbr2, where f = −kbr is the tension on the filament and nc is the number of filaments per unit volume.37,41 It is convenient to introduce the marginal probability density,
 
image file: c4ra02262a-t8.tif(5)

Using eqn (5) the autocorrelation function of stress at steady state can be calculated as,

 
image file: c4ra02262a-t9.tif(6)
where image file: c4ra02262a-t10.tif. The expressions for the conditional second moments of r for a given attachment state s that appear in eqn (6) are given in Appendix A, eqn (A9).

Using eqn (6) the non-equilibrium steady-state relaxation modulus of the active gel at the non-equilibrium steady-state can be expressed as,

 
image file: c4ra02262a-t11.tif(7)
where image file: c4ra02262a-t12.tif

The detailed procedure to obtain the full expression for the relaxation modulus is given in Appendix A. In the limit τr,aτr,d when τa/τd ≪ 1, simplified algebraic expressions can be obtained. Using eqn (6) and (7) the following simplified expression for the relaxation modulus is obtained,

 
image file: c4ra02262a-t13.tif(8)
where τr = ζa/kb = ζd/kb. A plot of this relaxation modulus for τa/τd = 0.1 and τr/τd = 4 is shown in Fig. 6A. A distinctive feature of this relaxation modulus, which makes it different from typical relaxation moduli of passive networks, is that it has a maximum located at image file: c4ra02262a-t14.tif with a magnitude given by image file: c4ra02262a-t15.tif. As we show in Section III C this feature is essential to the violation of FDT observed in active gels.

Also, by taking the one-sided Fourier transform of eqn (8), the non-equilibrium steady-state dynamic modulus of the active gel is obtained image file: c4ra02262a-t53.tif where image file: c4ra02262a-t54.tif is the one-sided Fourier transform,

 
image file: c4ra02262a-t16.tif(9a)
 
image file: c4ra02262a-t17.tif(9b)
Where GGK(ω) is the storage modulus and G′′GK(ω) is the loss modulus. eqn (9) are valid in the limit τr,aτr,d when τa/τd ≪ 1. A plot of this dynamic modulus for τa/τd = 0.1 and τr/τd = 4 is shown in Fig. 2B. Here, a maximum is also observed in GGK(ω) which corresponds to the maximum in GGK(t). Extrema do not occur in the relaxation or storage modulus of passive networks, and therefore this is a specific feature of active gels. The dynamic modulus of the active dumbbell gel also has features in common with the dynamic modulus of passive temporary networks, such as the high-frequency plateau in the storage modulus and the low frequency terminal zone in G′(ω) that goes as ω2.


image file: c4ra02262a-f2.tif
Fig. 2 (A) Non-equilibrium relaxation modulus for a gel formed by active dumbbells calculated from the autocorrelation function of stress at steady-state (Green–Kubo formula). (B) Non-equilibrium dynamic modulus for a gel formed by active dumbbells calculated from the autocorrelation function of stress at steady-state (Green–Kubo formula). The parameter values used in these figures are τr = ζa/kb = ζd/kb = 4τd and τa/τd = 0.1. It can be observed that a maximum appears in both G(t) and G′(ω). This maximum is characteristic of active gels and does not appear in passive temporary networks.

To further investigate how the shape of the dynamic modulus of the active network depends on the model parameters we performed calculations for different values of the ratios τa/τd and ζd/ζa. Fig. 3A shows how the shape of the dynamic modulus depends on the ratio τa/τd for a constant value of τr/τd. It can be observed that as the values of τa/τd increase, G′′(ω) and the maximum in G′(ω) decrease slightly in magnitude. In general, the shape of the dynamic modulus does not depend strongly on the ratio τa/τd. On the other hand, Fig. 3B shows how the position of the maxima in G′(ω) and G′′(ω) depends strongly on the ratio of the friction coefficients ζa/ζd. This can be understood by considering that when ζd is much smaller than ζa local relaxation of the filament upon motor detachment is faster than tension build-up when the motor is attached. Therefore stress relaxation in the gel occurs at shorter time scales (higher frequencies).


image file: c4ra02262a-f3.tif
Fig. 3 Non-equilibrium dynamic modulus for a gel formed by active dumbbells calculated from the autocorrelation function of stress at steady-state (Green–Kubo formula). (A) Dependence on the ratio τa/τd for a constant τr/τd = 10 ratio. (B) Dependence on the ratio ζd/ζa for a constant τa/τd = 0.1 ratio.

B. Dynamic modulus of active gels from a driven calculation

To show that the dumbbell model for active gels violates FDT in the way active gels do,5 we calculate the dynamic modulus observed in the stress response to an externally applied small-amplitude oscillatory strain. The calculation starts with eqn (1) with a small amplitude oscillatory strain applied to the bundle in the direction of filament alignment [small epsi, Greek, dot above](t) = ε0ω[thin space (1/6-em)]cos[thin space (1/6-em)]ωt. Where [small epsi, Greek, dot above] is the strain rate, ε0 is the strain amplitude and ω is the frequency. The detailed procedure to derive the dynamic modulus from this type of calculation and its general result are given in Appendix B. Moreover In the limit τr,aτr,d when τa/τd ≪ 1 the analytic expressions for G′ and G′′ can be simplified to
 
image file: c4ra02262a-t18.tif(10a)
 
image file: c4ra02262a-t19.tif(10b)

A plot of this dynamic modulus is shown in Fig. 4A. This dynamic modulus presents features similar to those obtained from the Green–Kubo formula. However there is a frequency-dependent discrepancy between the two moduli discussed in the next subsection.


image file: c4ra02262a-f4.tif
Fig. 4 A Comparison between the non-equilibrium dynamic modulus for an active gel calculated from, (i) the autocorrelation function of stress at steady-state (Green–Kubo formula, eqn (9)), (ii) the response to an external small-amplitude oscillatory strain, eqn (10). The parameter values used in these figures are τr = ζa/kb = ζd/kb = 2 and τa/τd = 0.005. (B) Fluctuation–dissipation relation for the active dumbbell model, violation of FDT vanishes in the limit τr/τd → ∞.

C. FDT in active gels

By comparing eqn (9) and (10) it can now be established if the active dumbbell model exhibits violation of FDT observed in microrheology experiments of active gels. Fig. 4A shows a comparison between the dynamic modulus of a typical actomyosin active gel obtained from the Green–Kubo formula, in Section III A, and the one obtained from the driven calculation, in Section III B. It can be observed that at high frequencies the two dynamic moduli are equal. In the limit τr,aτr,d when τa/τd ≪ 1, this high frequency plateau is given by image file: c4ra02262a-t20.tif. At lower frequencies G*GK begins to grow faster than G*D. Then both the storage and loss moduli reach maxima. The maximum in G*GK is larger than in G*D. At lower frequencies, below the maximum, GGK and GD decrease as ω2, while G′′GK and G′′D decrease as ω. The difference between G*GK and G*D persists as ω → 0.

Fig. 5A shows the imaginary part of the complex compliance J′′(ω) = 6π′′(ω) of the actomyosin gel studied by Mizuno et al.5 Where α(ω) is the complex compliance of the probe bead. In an active microrheology experiment the complex compliance is obtained from the relation rb(ω) = α(ω)ftrap. Where rb(ω) is the measured probe bead position and ftrap is a small amplitude oscillatory force, with frequency ω, applied with an optical trap. In the passive microrheology experiment, no external force is applied and the imaginary part of the complex compliance is obtained using the FDT, image file: c4ra02262a-t21.tif. Where image file: c4ra02262a-t22.tif is the autocorrelation function of the bead position. The complex compliance J*(ω) is related to the dynamic modulus by J*(ω)G*(ω) = 1. The symbols are experimental data obtained using passive and active microrheology, while the lines are fits used to transform J*(ω) to G*(ω). The data were taken at steady state after addition of ATP to the actomyosin network. In Fig. 5B the storage modulus obtained from the Green–Kubo calculation, eqn (9), and the one obtained from the driven calculation, eqn (10), are compared to the storage modulus of the actomyosin gel measured by active and passive microrheology by Mizuno et al.5 These experimental storage moduli are obtained from the fits to J′′(ω) shown in Fig. 5A using standard procedures.26,48,49


image file: c4ra02262a-f5.tif
Fig. 5 Predictions of the active dumbbell model of the dynamic modulus of active gels. (A) Imaginary part of the creep compliance, J′′(ω) = 6π′′(ω), of an actomyosin gel measured using passive and active microrheology. The symbols are experimental results by Mizuno et al.5 and the lines are fits used to convert J* to G*. (B) Comparison of the storage moduli, G′(ω), predicted by the active dumbbell model with the storage moduli determined from the microrheology experiments of Mizuno et al.5 The parameter values used in these figures are typical for actomyosin gels, τr = ζa/kb = ζd/kb = 2 and τa/τd = 0.005.

The frequency-dependent discrepancy between the response function of the active gel obtained from the driven and Green–Kubo calculations is in good qualitative agreement with the experimental observations of Mizuno et al.5 in microrheology experiments of actomyosin gels. For myosin motors τd ∼ 1–10 s, if we make frequency dimensional using this τd in Fig. 4A the discrepancy between GGK and GD appears around 10–50 Hz which agrees in order of magnitude with the results of Mizuno et al.5 It is important to emphasize that the parameters τa and τd used for the prediction shown in Fig. 5B were determined from other experiments.27,28 Only τr was fitted to the Mizuno et al.5 data. The model predicts correctly the characteristic maximum observed in the G′(ω) obtained from the passive microrheology technique. However, the active dumbbell model can not predict, simultaneously, the location of the maximum and the frequency where the discrepancy between the two moduli starts. This could be due to the very narrow relaxation spectrum of the dumbbell model in comparison with the real gel. Additional improvement in the prediction might be achieved by using a bead-spring chain with more beads, accounting for the polydispersity and non-linear elasticity of the actin filaments; these can increase the breadth of the relaxation spectrum of the gel. We note however that the predictions of the active dumbbell model are worst at lower frequencies, where the experimental observations indicates that G′(ω) goes to an elastic plateau, while the predictions continue to decay. However this discrepancy can be explained by the fact that the actomyosin gels prepared by Mizuno et al.5 also contain biotin crosslinks, while the active dumbbell model does not contain permanent passive crosslinks, but only active crosslinks (motors).

By taking the ratio of eqn (9) and (10) we can obtain the following relation for the active dumbbell model,

 
image file: c4ra02262a-t23.tif(11)
This relation is valid for τa/τd ≪ 1 and τr = ζa/kb = ζd/kb. The second term on the right hand side of eqn (11) gives the deviation from the FDT. In the limit τr/τd → ∞ this term vanishes and the FDT, GGK(ω) = GD(ω), is recovered. A plot of this relation, for different values of τr/τd, is shown in Fig. 4B. Note that the deviation from the equilibrium FDT in eqn (11) is additive and frequency-dependent and cannot, in general, be interpreted as an effective temperature. This result is in agreement with the conclusions of a recent and more general theoretical work by Ganguly and Chaudhuri20 in which an extension of the FDT for active systems was derived. Eqn (11) is one of the main conclusions of this paper, and is in contrast to other models for active gels in which motor activity is accounted for by means of an effective temperature.21–23 These latter works can not, in general, predict the correct frequency-dependent violation of the FDT due to the use of an effective temperature to model motor activity. Other works25,26 have used an attachment/detachment jump process to model motor dynamics. However those models were postulated on a continuum level of description, where motors are treated as force dipoles embedded within a continuum. In that picture each pole of the force dipole, representing the motor, is assumed to lie on a different filament. Those models have also been successful in describing some of the observations in the microrheology of active gels.

The microrheology experiments of Mizuno et al.5 can now be interpreted based on eqn (11). The frequency-dependent increase in the magnitude of the probe bead autocorrelation function in the passive microrheology of active gels is due to frequency-dependent stress fluctuations caused by motor activity. These fluctuations are therefore not related to the response function of the bead by the FDT. An additional, frequency-dependent term, as in eqn (11), is necessary to account for the magnitude and dynamics of motor activity. If such a modified FDT is used in the analysis of the passive microrheology data in active gels, it will account for frequency-dependent increase in the magnitude of the probe bead autocorrelation, yielding the same dynamic modulus obtained from the active technique. This conclusion agrees with previous works25,26 that have modeled the rheology of active gels on a continuum mechanics level of description but using a similar description of motor dynamics.

Other microrheology experiments in active gels3,5 have shown that motor-activity induces significant strain hardening of the semiflexible network. This can be observed in the statistics of bead displacements, which are Gaussian for non-active gels, but develop non-Gaussian tails when the gels are activated with ATP. The strain hardening is also observed as an overall increase in the magnitude of the dynamic modulus of the gel after addition of ATP. Since we have used a linear force law for the springs that represent the semiflexible filaments the current version of the model can not exhibit this feature. Non-linear spring force relations for semiflexible filaments6,50,51 could be used to model the strain-hardening in active gels. Accounting for these physics has the potential of further improving the predictions presented in Fig. 4B, which is explored in a future manuscript.

IV. Buckling and contraction in active bundles

Recent experiments1,27 in actomyosin bundles without sarcomeric organization have shown that self-contraction, upon addition of ATP, is related to the buckling of individual F-actin filaments that form the bundle. In the experiments, individual bundles were observed and F-actin buckling was found coincident with contraction. Prior to ATP addition, compact bundles with aligned F-actin are observed. Upon ATP addition, the frequency of buckles increases rapidly during contraction, and then diminishes once contraction stops. These F-actin buckles are dynamic, with their amplitude, curvature, and location changing over time. Motor-induced buckling of actin filaments has been shown to be a ubiquitous process in the self-organization of the cellular cytoskeleton.1

The main purpose of this section is to compare our model and results to the other single-chain mean-field description of active gels available in the literature.27 That model, as ours, treats the filaments as linear springs. However, as pointed in Section II, there are several differences in the level of description and mathematical formulation of the two models. Here we discuss in more detail how these differences reflect in a specific observable of the model. Lenz et al.27 specifically apply their model to study buckling and contraction in non-sarcomeric actomyosin bundles. Therefore using the model presented in Section II we calculate the fraction of buckled dumbbells, ϕB, as a function of time after addition of ATP in the absence of externally applied strain. Before addition of ATP the dumbbells have relaxed end-to-end length [small script l]0. Upon addition of ATP motor activity can change the end-to-end length of the dumbbells by an amount r. This change can cause compression (r < 0) or extension (r > 0) of the dumbbells. However F-actin filaments support large tensions but buckle easily under piconewton compressive loads.1,8 Therefore only dumbbells under compression (r < 0) buckle. A compressed dumbbell buckles when its tension f = −kbr reaches a buckling force threshold, FB. An estimation of this force threshold can be obtained by treating the filaments as thin elastic cylinders8,52 image file: c4ra02262a-t24.tif. For a typical F-actin filament, using the values presented in Section II, FB ∼ 0.1 pN.

To proceed with the calculation of ϕB we set ε(t) → 0 in eqn (1) and introduce the marginal probability function ϕ(s, r, t),

 
image file: c4ra02262a-t25.tif(12)
To perform the integrals in eqn (12) we make use of the algebraic expressions for the conditional moments given in Appendix A. Here as in Section III we assume p(F) = δ(FFm). Using this p(F) in eqn (A1) and then putting the result into eqn (12) we can now take the integrals with respect to F to obtain:
 
image file: c4ra02262a-t26.tif(13)
where δs,s is the Kronecker delta and the first and second conditional moments of r for a given s are given in Appendix A, eqn (A9). And J is a normalization constant.

On average, filaments in motor-attachment state s = 1 (see Fig. 1) undergo compression (〈rs=1 < 0); filaments in motor-attachment state s = 2 experience, on average, the same magnitude of extension (〈rs=1 = −〈rs=2). The end-to end distance of filaments in states s = 0 and s = 3 does not change, in average, due to motor activity (〈rs=3 = 0). To find the fraction of buckled filaments eqn (13) is integrated over all the compressed filaments whose tension exceeds FB and then summed over all the attachment states that is,

 
image file: c4ra02262a-t27.tif(14)

Eqn (14) gives the fraction of buckled filaments in the dumbbell as a function of time. Note that the first term on the right hand side of eqn (13) vanishes after performing the integral in eqn (14) since the end-to-end distance of the dumbbells on that attachment state does not change due to motor activity. In Fig. 6A the fraction of buckled filaments is shown as a function of time, for different values of the ratio ζd/ζa. To make the calculations and plots shown here we make time dimensionless by τd and choose as characteristic length scale Fm/kb. For these figure the ratio ζd/ζa is set to 0.1 and τr,a/τd is set to 10. The ratio between the buckling force of the filaments, FB, and the mean motor stall force Fm was set to 0.3, this value is inside the range observed experimentally in actomyosin networks.27,28 Upon addition of ATP at t = 0 there is a short lag time, after which the fraction of buckled filaments grows until reaching a steady state value that depends on the ratios ζa/ζd and τa/τd. The fraction of buckled filaments is never greater than 1/4, since this is the maximum fraction of filaments that can undergo compression due to motor activity. Since we have assumed p(F) = δ(FFm) the fraction of buckled filaments is expected to be smaller, for a given set of parameters, than in the real gel. A wider motor-force distribution will cause some filaments on attachment state s = 3 to also undergo compression. Qualitatively the shape of this curve coincides with the experimental observations of Lenz et al.27 Fig. 6B shows the steady state value of the fraction of buckled filaments as a function of the ratio τa/τd. It can be observed that for a given set of friction coefficients there is a critical values of the τa/τd at which buckling occurs. If, the detachment rate of the motors becomes comparable, or larger than the attachment rate then motor activity can not produce buckling of the semiflexible filaments. Here and through the rest of the paper it is useful to consider the limit τr,aτr,d when τa/τd ≪ 1, in this case the algebraic expressions simplify considerably. However, as mentioned in Section II, in real systems τr,a is larger than τr,d therefore the expressions for this limit are presented only to illustrate the general features of the model, and do not represent a specific physical system. For instance, in this limit the expression for 〈rs=1 at steady state simplifies to image file: c4ra02262a-t28.tif, and 〈r2(t = ∞)〉s={1,2,3} → 0. Therefore in this limit, the steady state does not depend on the friction coefficients and a filament buckles when an effective motor force that accounts for the attachment/detachment, image file: c4ra02262a-t29.tif, is larger than the buckling force of the filament FB.


image file: c4ra02262a-f6.tif
Fig. 6 (A) Fraction of buckled filaments in a bundle, ϕB, as a function of time for different values of the ratio ζd/ζa. τa/τd was set to 0.1. (B) Steady state fraction of buckled filaments as a function of τa/τd, for different values of the ratio ζd/ζa. For these plots FB/Fm = 0.3 and τr,a/τd was set to 10, where τr,aζa/kb.

Some previous attempts have been made to use microscopic models to describe buckling of filaments in active bundles,27,53 however, to our knowledge, this is the first time the dynamics of buckling formation have been calculated. This dumbbell version of our model is not sufficient to investigate the influence of motor concentration in the buckling dynamics. Lenz et al.27 have investigated the effect of motor concentration and found that there is an optimum concentration of motors for contraction to occur. High concentrations of motors suppress buckling while very sparse motors can not generate it. We will investigate the effect of motor concentration using our model in a future manuscript. There is however a fundamental difference in how the model presented in this paper and the model by Lenz et al.27 explain the mechanisms that cause buckling and self-contraction in active bundles. In the model presented here buckling arises because a fraction of the filaments in the bundle exist in a motor attachment state where contraction is favored (e.g.: state s = 1 for the dumbbell version). This is achieved by maintaining the motor forces as stochastic state variables instead of pre-averaging eqn (1) over them. The filaments on attachment states that undergo contraction buckle when the tension in them reaches FB. On the other hand in the model of Lenz et al.27 buckling arises from the dependence of p(F) on the position along the filament of the bead (active site) to which the motor attaches. In other words, in that model the underlying mechanism for buckling is postulated to be a spatial gradient of the motor stall forces.

V. Conclusions

We have introduced a single-chain mean-field model for active gels. The motors are assumed to create pair-wise active interactions between the filaments, which are taken into account by considering the probability for a motor to attach to (or detach from) the network active sites. We have postulated a minimum level of description that a single-chain mean-field model needs to reproduce the characteristic rheological and mechanical features of active gels. In the model presented here, the change in the end-to-end distance of the filaments, the attachment state of the filaments, and the motor-generated forces, are stochastic state variables that evolve according to a differential Chapman–Kolmogorov equation. The motor-generated forces are drawn from a stationary distribution of motor stall forces that can be measured experimentally.27,28 Most of the model parameters, which include attachment/detachment rates of motors, friction coefficients for filaments and motors, as well as parameters to characterized the elasticity of the semiflexible filaments, can be measured in independent experiments.

The general formulation of the model allows us to account for physics that are not available in models that have been postulated on coarser levels of description.25–27 These physics include motor concentration in the gel, non-linear elasticity of the semiflexible filaments and attachment/detachment rates dependent on the local filament tension. However, to simplify the mathematics and with the aim of obtaining analytical results we have made several assumptions in this introductory manuscript. We have assumed that motors attach only at the ends of the filament, this means we have represented the filaments as dumbbells. This is expected to be a reasonable approximation only in the limit of very low motor concentration. We have also assumed that the semiflexible filaments obey a linear elasticity law; again, this is expected to be a reasonable assumption only in the limit of very low motor concentration where strain hardening is negligible. Additionally we have assumed that the attachment/detachment transition rates are independent of the local filament tension. In a future manuscript we will present numerical results for a more general version of the model, where we examine the effect of relaxing some of the aforementioned assumptions.

We have shown that in its simplest form, the active dumbbell model, can qualitatively reproduce characteristic rheological and mechanical properties of active gels. For instance, the model can predict the buckling of individual filaments that is thought to be the underlying mechanism in the self-contraction of non-sarcomeric actin–myosin bundles.27,28 We have calculated the dynamics of buckling formation and found that they are mostly determined by the ratio of the friction coefficients of the filament to that of the motors. In general, buckling occurs if the ratio of detachment/attachment rates is below a critical value that is determined by the buckling force of the filaments, the mean motor stall force and the friction coefficients of filaments and motors. The active dumbbell model can also predict the violation of the fluctuation–dissipation theorem observed in microrheology experiments in active gels.5 We calculated and compared the dynamic modulus, of a network composed by active dumbbells, obtained from the autocorrelation function of stress at the non-equilibrium steady state and from the stress response after applying a small oscillatory strain. If the fluctuation–dissipation theorem were obeyed the two moduli would be equal.38,42,47 We find that for the active dumbbell model there is a frequency-dependent discrepancy between the moduli obtained from the two different methods. This discrepancy is in general determined by the attachment and detachment rates, as well as by the friction coefficients. When the ratio between the local relaxation time scale and the detachment time of the motors becomes very large, the FDT is recovered. In this limit, the dynamics of the system are completely dominated by the local relaxation of the filaments, and motor attachment/detachment dynamics have a negligible effect.

Appendix

A Analytic solution of the active dumbbell model

To solve eqn (1) analytically we begin by postulating a general form for the solution,
 
image file: c4ra02262a-t30.tif(A1)
where F = {F1, F2}. By putting eqn (A1) in eqn (1) it can be shown that this probability density function satisfies eqn (1) with [small psi, Greek, tilde](r, s; t|F), a conditional probability distribution for a given set of motor forces {F1, F2}, that evolves according to,
 
image file: c4ra02262a-t31.tif(A2)
where the function F(s) is given by,
 
image file: c4ra02262a-t32.tif(A3)

The transition rate matrix, image file: c4ra02262a-t52.tif, that appears in eqn (A2) whose elements give the transition rate at which a dumbbell that is in state s jumps to state s′ is defined as,

 
image file: c4ra02262a-t33.tif(A4)
with diagonal elements given by image file: c4ra02262a-t34.tif.

Now that the F dependence has been “factored-out” from [small psi, Greek, tilde](r, s; t|F), it is straightforward to obtain a solution from eqn (A2). Since the equation is linear in r, the solution is a Gaussian distribution,

 
image file: c4ra02262a-t35.tif(A5)
where J is the normalization factor and 〈r{s, F} ≔ ∫r[small psi, Greek, tilde](r, s; t|F)dr and 〈r2{s, F} ≔ ∫r2[small psi, Greek, tilde](r, s; t|F)dr are the first and second conditional moments of r for a given s and F. The evolution equations for these conditional moments can be derived from eqn (4) by multiplying both sides by r and r2, respectively. Integrating over r gives,
 
image file: c4ra02262a-t36.tif(A6a)
 
image file: c4ra02262a-t37.tif(A6b)

For a given set of initial conditions this system of coupled ordinary differential equations (eight equations, for the dumbbell version of the model) is solved for 〈r{s,F} and 〈r2{s,F}. These moments can, in general, be expressed as

 
r{s,F} = a1(s, t, X, 〈r{0,s})F1 + a2(s, t, X, 〈r{0,s})F2, (A7)
 
image file: c4ra02262a-t38.tif(A8)
where X:{kb, τd, τa, ζd, ζa} are model parameters; and 〈r{0,s} and 〈r2{0,s} are initial conditions for the first and second conditional moments of r for a given s, respectively. In general, the specific expressions for a1, a2, b1, b2 and b3 as a function t and s are calculated by solving eqn (A6a) and (A6b) after specifying values for kb, τd, τa, ζd, ζa, 〈r{0,s} and 〈r2{0,s}.

Once specific expressions for 〈r{s,F} and 〈r2{s,F} are obtained the probability density function of active dumbbell conformations can be obtained by putting back these moments into eqn (A5), and then inserting the result into eqn (A1).

A distribution of motor forces, p(F) must be specified before making further calculations with the model. For p(F) = δ(FFm) the conditional second moments of r for a given attachment state s that can be calculated analytically and are given by

 
image file: c4ra02262a-t39.tif(A9a)
 
image file: c4ra02262a-t40.tif(A9b)

And are therefore completely determined by the solution of eqn (A6a) and (A6b).

B Dynamic modulus from driven calculation

To calculate the dynamic modulus of the active gel observed in a driven experiment a small-amplitude oscillatory strain is applied to the bundle in the direction of filament alignment [small epsi, Greek, dot above](t) = ε0ω[thin space (1/6-em)]cos[thin space (1/6-em)]ωt. We need to calculate the macroscopic stress in the active gel that results from such a deformation. For a single-chain bead-spring model this is given by,
 
image file: c4ra02262a-t41.tif(B1)
where ϕ(s, r, t) is the marginal probability density defined in eqn (12). In the long-time limit the conditional moments 〈r{s,F} and 〈r2{s,F}, which are required to calculate ϕ(s, r, t) have the form:
 
r{s,F} = g1(s, ω)[thin space (1/6-em)]sin[thin space (1/6-em)]ωt + h1(s, ω)[thin space (1/6-em)]cos[thin space (1/6-em)]ωt, (B2a)
 
r2{s,F} = g2(s, ω)[thin space (1/6-em)]sin[thin space (1/6-em)]ωt + h2(s, ω)[thin space (1/6-em)]cos[thin space (1/6-em)]ωt. (B2b)

The coefficients g1 and g2 are associated with the material response that is in phase with the applied strain, while h1 and h2 are associated with the material response that is out of phase with the applied strain. By putting eqn (B2a) and (B2b) in eqn (A6a) and (A6b) a set of equations for g1, g2, h1 and h2 is obtained:

 
image file: c4ra02262a-t42.tif(B3a)
 
image file: c4ra02262a-t43.tif(B3b)
 
image file: c4ra02262a-t44.tif(B3c)
 
image file: c4ra02262a-t45.tif(B3d)
where 〈r{0,s} and 〈r2{0,s} are the non-equilibrium steady-state values for the first and second conditional moments, respectively; and the transition matrix image file: c4ra02262a-t51.tif was defined in eqn (A4). For the active dumbbell model this yields a system of 16 coupled linear equations. In general the solutions for g1 can be written as a function of F as:
 
g1 = a1,1(s, ω, X, 〈r{0,s})F1 + a1,2(s, ω, X, 〈r{0,s})F2 (B4)
where X:{kb, τd, τa, ζd, ζa} are model parameters. An equivalent equation can be written for h1. Note that the first subindex in the ai,j coefficients indicates whether the coefficient is used to express the solution for g or h; and the second subindex indicates if the coefficients multiply either F1 or F2. For g2 the solutions can be expressed as,
 
image file: c4ra02262a-t46.tif(B5)

And an equivalent definition for h2. As in Section III A we specify a motor force distribution p(Fj) = δ(FjFm) with this ϕ(s, r, t) can be written as in eqn (13) but with conditional moments of r for a given s are given by,

 
image file: c4ra02262a-t47.tif(B6a)
 
image file: c4ra02262a-t48.tif(B6b)

The dynamic modulus of the gel, observed in the driven or active experiment, can now be calculated by using eqn (B1) and (B6) and the relation σ = ε0[G′(ω)sin ωt + G′′(ω)[thin space (1/6-em)]cos[thin space (1/6-em)]ωt]. This gives in general,

 
image file: c4ra02262a-t49.tif(B7a)
 
image file: c4ra02262a-t50.tif(B7b)

Acknowledgements

We are grateful to the Army Research Office (grants W911NF-09-2-0071 and W911NF-09-1-0378) for financial support.

References

  1. M. S. e Silva, M. Depken, B. Stuhrmann, M. Korsten, F. C. MacKintosh and G. H. Koenderink, Active multistage coarsening of actin networks driven by myosin motors, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, 9408–9413 CrossRef PubMed.
  2. T. Surrey, F. Nédélec, S. Leibler and E. Karsenti, Physical properties determining self-organization of motors and microtubules, Science, 2001, 292, 1167–1171 CrossRef CAS PubMed.
  3. B. Stuhrmann, M. S. e Silva, M. Depken, F. C. MacKintosh and G. H. Koenderink, Nonequilibrium fluctuations of a remodeling in vitro cytoskeleton, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2012, 86, 020901 CrossRef.
  4. O. J. N. Bertrand, D. K. Fygenson and O. A. Saleh, Active, motor-driven mechanics in a DNA gel, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 17342–17347 CrossRef CAS PubMed.
  5. D. Mizuno, C. Tardin, C. F. Schmidt and F. C. MacKintosh, Nonequilibrium mechanics of active cytoskeletal networks, Science, 2007, 315, 370–373 CrossRef CAS PubMed.
  6. J. F. Marko and E. D. Siggia, Stretching DNA, Macromolecules, 1995, 28, 8759–8770 CrossRef CAS.
  7. F. Gittes and F. C. MacKintosh, Dynamic shear modulus of a semiflexible polymer network, Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top., 1998, 58, 1241–1244 CrossRef.
  8. C. Storm, J. J. Pastore, F. C. MacKintosh, T. C. Lubensky and P. A. Janmey, Nonlinear elasticity in biological gels, Nature, 2005, 435, 191–194 CrossRef CAS PubMed.
  9. A. F. Huxley, Muscular contraction, J. Physiol., 1974, 243, 1 CAS.
  10. T. L. Hill, E. Eisenberg, Y. D. Chen and R. J. Podolsky, Some self-consistent two-state sliding filament models of muscle contraction, Biophys. J., 1975, 15, 335–372 CrossRef CAS.
  11. A. Yamada and T. Wakabayashi, Movement of actin away from the center of reconstituted rabbit myosin filament is slower than in the opposite direction, Biophys. J., 1993, 64, 565–569 CrossRef CAS.
  12. F. Jülicher and J. Prost, Cooperative molecular motors, Phys. Rev. Lett., 1995, 75, 2618–2621 CrossRef.
  13. R. W. Lymn and E. W. Taylor, Mechanism of adenosine triphosphate hydrolysis by actomyosin, Biochemistry, 1971, 10, 4617–4624 CrossRef CAS.
  14. F. Jülicher, A. Ajdari and J. Prost, Modeling molecular motors, Rev. Mod. Phys., 1997, 69, 1269 CrossRef.
  15. T. Indei, J. D. Schieber and A. Córdoba, Competing effects of particle and medium inertia on particle diffusion in viscoelastic materials, and their ramifications for passive microrheology, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2012, 85, 041504 CrossRef.
  16. A. Córdoba, T. Indei and J. D. Schieber, The effects of compressibility, hydrodynamic interaction and inertia on two-point, passive microrheology of viscoelastic materials, Soft Matter, 2013, 9, 3521 RSC.
  17. J. D. Schieber, A. Córdoba and T. Indei, The analytic solution of Stokes for time-dependent creeping flow around a sphere: application to linear viscoelasticity as an ingredient for the Generalized Stokes–Einstein relation and microrheology analysis, J. Non-Newtonian Fluid Mech., 2013, 200, 3–8 CrossRef CAS PubMed.
  18. J. Prost, J.-F. Joanny and J. M. R. Parrondo, Generalized fluctuation–dissipation theorem for steady-state systems, Phys. Rev. Lett., 2009, 103, 090601 CrossRef CAS.
  19. P. Bohec, F. Gallet, C. Maes, S. Safaverdi, P. Visco and F. van Wijland, Probing active forces via a fluctuation–dissipation relation: Application to living cells, EPL, 2013, 102, 50005 CrossRef.
  20. C. Ganguly and D. Chaudhuri, Stochastic thermodynamics of active Brownian particles, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2013, 88, 032102 CrossRef.
  21. T. B. Liverpool, A. C. Maggs and A. Ajdari, Viscoelasticity of solutions of motile polymers, Phys. Rev. Lett., 2001, 86, 4171–4174 CrossRef CAS.
  22. T. B. Liverpool and M. C. Marchetti, Rheology of active filament solutions, Phys. Rev. Lett., 2006, 97, 268101 CrossRef CAS.
  23. T. B. Liverpool and M. C. Marchetti, Bridging the microscopic and the hydrodynamic in active filament solutions, EPL, 2007, 69, 846 CrossRef.
  24. J. Palacci, C. Cottin-Bizonne, C. Ybert and L. Bocquet, Sedimentation and effective temperature of active colloidal suspensions, Phys. Rev. Lett., 2010, 105, 088304 CrossRef.
  25. A. J. Levine and F. C. MacKintosh, The mechanics and fluctuation spectrum of active gels, J. Phys. Chem. B, 2009, 113, 3820–3830 CrossRef CAS PubMed.
  26. D. A. Head and D. Mizuno, Nonlocal fluctuation correlations in active gels, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2010, 81, 041910 CrossRef CAS.
  27. M. Lenz, T. Thoresen, M. L. Gardel and A. R. Dinner, Contractile units in disordered actomyosin bundles arise from F-actin buckling, Phys. Rev. Lett., 2012, 108, 238107 CrossRef.
  28. T. Thoresen, M. Lenz and M. L. Gardel, Reconstitution of contractile actomyosin bundles, Biophys. J., 2011, 100, 2698–2705 CrossRef CAS PubMed.
  29. K. Kruse and K. Sekimoto, Growth of fingerlike protrusions driven by molecular motors, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2002, 66, 031904 CrossRef CAS.
  30. K. Kruse, J. Joanny, F. Jülicher, J. Prost and K. Sekimoto, Asters, vortices, and rotating spirals in active gels of polar filaments, Phys. Rev. Lett., 2004, 92, 78101 CrossRef CAS.
  31. K. Kruse and F. Jülicher, Self-organization and mechanical properties of active filament bundles, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2003, 67, 051913 CrossRef.
  32. I. S. Aranson and L. S. Tsimring, Pattern formation of microtubules and motors: inelastic interaction of polar rods, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2005, 71, 050901 CrossRef.
  33. M. E. Cates, S. M. Fielding, D. Marenduzzo, E. Orlandini and J. M. Yeomans, Shearing active gels close to the isotropic–nematic transition, Phys. Rev. Lett., 2008, 101, 068102 CrossRef CAS.
  34. L. Giomi, T. B. Liverpool and M. C. Marchetti, Sheared active fluids: Thickening, thinning, and vanishing viscosity, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2010, 81, 051908 CrossRef.
  35. D. A. Head, G. Gompper and W. Briels, Microscopic basis for pattern formation and anomalous transport in two-dimensional active gels, Soft Matter, 2011, 7, 3116–3126 RSC.
  36. J. G. Hérnandez Cifre, T. H. Barenbrug, J. D. Schieber and B. H. A. A. van den Brule, Brownian dynamics simulation of reversible polymer networks under shear using a non-interacting dumbbell model, J. Non-Newtonian Fluid Mech., 2003, 113, 73–96 CrossRef.
  37. T. Indei and J. Takimoto, Linear viscoelastic properties of transient networks formed by associating polymers with multiple stickers, J. Chem. Phys., 2010, 133, 194902 CrossRef PubMed.
  38. T. Indei, J. D. Schieber and J.-i. Takimoto, Effects of fluctuations of cross-linking points on viscoelastic properties of associating polymer networks, Rheol. Acta, 2012, 51, 1021–1039 CrossRef CAS.
  39. F. C. MacKintosh, J. Käs and P. A. Janmey, Elasticity of semiflexible biopolymer networks, Phys. Rev. Lett., 1995, 75, 4425–4428 CrossRef CAS.
  40. K. I. Morozov and L. M. Pismen, Strain dependence of cytoskeleton elasticity, Soft Matter, 2012, 8, 9193–9199 RSC.
  41. R. B. Bird, C. F. Curtiss, R. C. Armstrong and O. Hassager, Dynamics of Polymeric Liquids, 2 Volume Set, Wiley-Interscience, 1996 Search PubMed.
  42. J. D. Schieber and K. Horio, Fluctuation in entanglement positions via elastic slip-links, J. Chem. Phys., 2010, 132, 074905 CrossRef PubMed.
  43. M. P. Murrell and M. L. Gardel, F-actin buckling coordinates contractility and severing in a biomimetic actomyosin cortex, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 20820–20825 CrossRef CAS PubMed.
  44. G. H. Koenderink, Z. Dogic, F. Nakamura, P. M. Bendix, F. C. MacKintosh, J. H. Hartwig, T. P. Stossel and D. A. Weitz, An active biopolymer network controlled by molecular motors, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 15192–15197 CrossRef CAS PubMed.
  45. M. Karim, S. C. Kohale, T. Indei, J. D. Schieber and R. Khare, Determination of viscoelastic properties by analysis of probe-particle motion in molecular simulations, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2012, 86, 051501 CrossRef.
  46. A. W. C. Lau, B. D. Hoffman, A. Davies, J. C. Crocker and T. C. Lubensky, Microrheology, stress fluctuations, and active behavior of living cells, Phys. Rev. Lett., 2003, 91, 198101 CrossRef CAS.
  47. D. Chandler, Introduction to modern statistical mechanics, Oxford University Press, 1987, vol. 1 Search PubMed.
  48. M. Baumgaertel and H. H. Winter, Determination of discrete relaxation and retardation time spectra from dynamic mechanical data, Rheol. Acta, 1989, 28, 511–519 CrossRef CAS.
  49. F. C. MacKintosh and A. J. Levine, Nonequilibrium mechanics and dynamics of motor-activated gels, Phys. Rev. Lett., 2008, 100, 18104 CrossRef CAS.
  50. M. Pasquali, V. Shankar and D. C. Morse, Viscoelasticity of dilute solutions of semiflexible polymers, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2001, 64, 020802 CrossRef CAS.
  51. P. T. Underhill and P. S. Doyle, Alternative spring force law for bead-spring chain models of the worm-like chain, J. Rheol., 2006, 50, 513 CrossRef CAS.
  52. F. Gittes, B. Mickey, J. Nettleton and J. Howard, Flexural rigidity of microtubules and actin filaments measured from thermal fluctuations in shape, J. Cell Biol., 1993, 120, 923–934 CrossRef CAS.
  53. M. Lenz, M. L. Gardel and A. R. Dinner, Requirements for contractility in disordered cytoskeletal bundles, New J. Phys., 2012, 14, 033037 CrossRef PubMed.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.