Motivated by the well-known fractal packing of chromatin, we study the Rouse-type dynamics of elastic fractal networks with embedded, stochastically driven, active force monopoles and force dipoles that are temporally correlated. We compute, analytically—using a general theoretical framework—and via Langevin dynamics simulations, the mean square displacement (MSD) of a network bead. Following a short-time superdiffusive behavior, force monopoles yield anomalous subdiffusion with an exponent identical to that of the thermal system. In contrast, force dipoles do not induce subdiffusion, and the early superdiffusive MSD crosses over to a relatively small, system-size-independent saturation value. In addition, we find that force dipoles may lead to “crawling” rotational motion of the whole network, reminiscent of that found for triangular micro-swimmers and consistent with general theories of the rotation of deformable bodies. Moreover, force dipoles lead to network collapse beyond a critical force strength, which persists with increasing system size, signifying a true first-order dynamical phase transition. We apply our results to the motion of chromosomal loci in bacteria and yeast cells’ chromatin, where anomalous sub-diffusion, MSD t ν with ν 0.4, was found in both normal and cells depleted of adenosine triphosphate (ATP), albeit with different apparent diffusion coefficients. We show that the combination of thermal, monopolar, and dipolar forces in chromatin is typically dominated by the active monopolar and thermal forces, explaining the observed normal cells vs the ATP-depleted cells behavior.

The slowed, subdiffusive motion of chromosomal loci within cells has recently become a focal point of research. This phenomenon has been particularly noted in experimental studies involving both normal cells, where active forces are prevalent, and cells depleted of adenosine triphosphate (ATP), where forces are purely thermal. Interestingly, in both cases, the observed subdiffusive exponents ( 0.4) remain consistent, a finding that current theories struggle to explain. Fractals, intricate self-similar patterns pervasive in nature, have drawn new interest due to their relevance to chromatin structure, which has been firmly established through experiments. Motivated by this fractal nature of chromatin, our theoretical investigation explores the dynamics of a fractal network influenced by active forces—both monopolar and dipolar. We aim to elucidate why chromosomal loci exhibit identical subdiffusive exponents in both normal and ATP-depleted cells.

Active matter is attracting increasing interest and spans diverse systems.1–6 Much effort has been recently devoted to the phase diagram of active systems, with emphasis on “motility-induced phase separation.” 3,7–11 Living systems, where the chemically stored energy is transformed to mechanical energy, can be regarded as a type of active matter, e.g., moving micro-organisms, in vivo intracellular dynamics, or in vitro active cytoskeleton dynamics.12–18 Earlier studies of such bio-systems aimed to obtain steady-state dynamics, in particular, its manifestation in the mean square displacement (MSD) of a probe monomer or particle.19–26 

In one such recent in vivo experiment, the motion of different chromosomal loci in bacteria and yeast cells was probed both in normal cells, in which active (“ac”) forces dominate, and in cells depleted of adenosine triphosphate (ATP), where fluctuations are governed by thermal (“th”) forces.25,26 It was found that in both cases, chromosomal loci perform anomalous sub-diffusion, MSD B t ν ( ν < 1), with identical exponents ν a c ν t h 0.4, yet different amplitudes B (from which an apparent diffusion coefficient is defined in Ref. 26 as B = 2 d D app, where d is the dimensionality); the ATP-depleted cells yield a lower amplitude. Telomeres in chromatin of mammalian cells also exhibit subdiffusion with a similar exponent, ν a c 0.4.23,24 Yet, the reason behind the equality of the two exponents reported in Ref. 26 is unclear.27 While passive (thermal) and active Rouse dynamics of linear chains predict identical subdiffusion exponents, their values are ν a c = ν t h = 1 / 2,28,29 larger than the observed values.

To explain these discrepancies, we take advantage of the fractal structure of chromatin (discussed next). Fractals, either disordered or deterministic, are characterized by a few broken dimensions:30–32 (i) the mass fractal (Hausdorff) dimension d f governing the scaling M ( r ) r d f of the mass M ( r ) enclosed in concentric spheres of radius r, (ii) the spectral dimension d s governing the scaling g ( ω ) ω d s 1 of the vibrational density of states (DOS) g ( ω ) with frequency ω,33,34 and (iii) the topological dimension d l, which governs the scaling M ( l ) l d l of the mass M ( l ) enclosed in concentric “spheres” of radius l in the topological (or “manifold”/“chemical”) space. One may also define, instead of d l, the chemical length (or the minimal path) dimension d min = d f / d l, which relates the real space distance r between two points on the fractal to the minimal path distance l between these points along the fractal network links, l r d min.32,35,36 To clarify the concept of topological space and d l, consider a linear chain—defining a 1D topological space–folded self-similarly in the 3D space without forming cross-links, thereby setting d l = 1 independent of d f (e.g., d f = 2 and 5 / 3 for Gaussian and self-avoiding chains, respectively), while cross-links may increase d l above 1. Similarly, a 2D sheet may be crumpled self-similarly in the 3D space without internal cross-links, corresponding to d l = 2 irrespective of d f. The three broken dimensions, d s, d l, and d f, obey the inequalities: 1 d s d l d f d, where d is the Euclidean embedding space dimension.

In recent years, significant insight has been gained in understanding chromatin structure during the interphase stage using various experimental methods.37–45 These experiments reveal a fractal-like density distribution and estimate the values for the (mass) fractal dimension d f (see Refs. 44 and 46 for summaries of values). In particular, Hi–C experiments,40 textural analysis,37 and neutron scattering38,39 yielded values of d f 2.8 2.9. The latter values of d f are nearly those predicted by the fractal-globule model,47–49 which builds on a linear polymer folded compactly in 3D space, thereby fractal dimensions d l = d s = 1 and d f = 3. Given the abundance of lamin A proteins making internal cross-links,24 we may expect that both d l and d s are somewhat larger than 1.

What are the anticipated effects of active forces acting on chromatin within living cells? These forces can arise from various sources.50–53 Here, we focus on two main categories: force monopoles and force dipoles. Force monopoles are external forces that cannot emerge from within the network, i.e., they are exerted by the cytoplasm and membrane (in prokaryotic cells) or by the cytoskeleton (in eukaryotic cells).52 On the other hand, force dipoles are internal forces predominantly found in the cytoskeleton, such as the actomyosin system. They might also be present in chromatin, potentially through various enzymes that actively interact with DNA.53 Higher-order multipoles, such as force quadrupoles, cannot be excluded completely when considering complicated enzyme or motor protein action but are likely rarer and weaker than the monopoles and dipoles.

Explanations have been proposed for the observed anomalous diffusion exponents in chromatin, ν a c ν t h 0.4; yet they do not account for the behaviors of both normal and ATP-depleted cells on the same footing. One approach, based on a linear chain topology, has been to associate ν t h < 1 / 2 in the presence of a viscoelastic solvent, whose frequency-dependent complex modulus behaves as a power-law, G ( ω ) ( i ω ) α, yielding ν t h = α / 2.25,54–56 However, a different exponent (also discussed briefly in  Appendix C and Sec. II A 1), ν a c = 3 α / 2 1, is predicted for an equivalent active system,55 contradicting the experimental results. Another explanation was based on Rouse dynamics (of a folded linear chain) yielding ν t h = 2 / ( 2 + d f ).57 Alternatively, it has been shown that this exponent is always ν t h = 1 d s / 2, irrespective of the fractal dimension;58–61 hence, ν t h = 1 / 2 is recovered for d s = 1. These two results for ν t h become identical for Gaussian fractal networks (i.e., in the absence of self-avoidance) for which d f = 2 d s / ( 2 d s ).62,63 This implies that d s = 6 / 5 corresponds to d f = 3, leading to ν t h = 0.4, consistent with the experimental results for ATP-depleted cells. From a different perspective, a single particle model, which uses the generalized Langevin equation with a delta-correlated active force that does not obey the fluctuation–dissipation theorem (FDT), yields ν a c = 2 ν t h 1 for ν t h 1 / 2 and ν a c = 0 for ν t h < 1 / 2 such that identical passive and active exponents are obtained only for normal diffusion, ν = 1.20,27,55,64 Active dynamics of critical percolation clusters—forming disordered fractals—of a 2D triangular network of springs have been studied previously; yet, the individual node MSD was not reported.65 Here, we aim to generalize such models to arbitrary fractals with potential applications to chromatin.

This paper is structured as follows: In Sec. II, we begin with an overview of the model, the general analytical theory that solves it (detailed in  Appendix A) that applies to any fractal, and the primary results for MSD. Next, in Sec. III, we present simulation results for the Sierpinski gasket, a well-known deterministic fractal, thereby validating our analytical results and discovering new findings for the case of force dipoles. In Sec. IV, we extend the present general theory of the active fractal network to chromatin relying on its fractal nature and show general agreement with experimental findings. We conclude in Sec. V by highlighting the role of active forces in active fractal dynamics and the significance of the nature of active forces in chromatin dynamics.

Here, we provide an outline of the analytical theory. The complete derivation is described in  Appendix A. Following the formalism developed for passive fractals,58–61 we consider a fractal network of beads connected by harmonic springs where, in addition to the white thermal noise, beads experience stochastic active forces distributed randomly but uniformly over the network. In the case of force monopoles, the force field F ( r , t ) at position r is
(1)
where f j ( t ) is an active force acting on bead j at position r j. For force dipoles,
(2)
where ϵ j is a randomly chosen vector from j to one of its nearest-neighbor beads ( | ϵ j | = b), and f j is taken parallel to ϵ j. The limit ϵ j 0 should be taken in the continuum limit of point force dipoles. A positive f j implies an inward (contractile) force dipole, and a negative f j implies an outward (extensile) one.1,66,67
The stochastic active forces are assumed to fluctuate independently in time between f 0 (probability p) and 0 (probability 1 p), following the random-telegraph process.68 The auto-correlation function of the forces, thus, follows:
(3)
where t is the lag time and τ is the force correlation time. We assume that the force variables f i’s, appearing in Eqs. (1)–(2), are not correlated in space. In addition, their directions are assumed isotropically distributed (see further discussion in  Appendix A).
To simplify the analytical treatment, we use the scalar elasticity Hamiltonian58,69 for the displacements u i = r i r 0 , i about the equilibrium positions { r 0 , i },
(4)
where ω o is the spring self-frequency, m is the bead mass, and i j stands for nearest-neighbor pairs connected by springs. The Langevin equations of motion in the overdamped limit, assuming local friction (i.e., a Rouse-type model), are
(5)
( i = 1 , , N), where j i stands for beads that are connected (as nearest-neighbors) to i, f i ( t ) is the active force acting on bead i, γ is the local friction coefficient, and ζ i is the thermal white noise obeying FDT,70  ζ i ( t ) ζ j ( t ) = 2 d k B T γ δ i , j δ ( t t ), where T is the temperature and k B is the Boltzmann constant. Hence, its variance is ζ i 2 = 2 d k B T γ, while that of the active noise—for t τ, where the active noise becomes effectively white—is f i 2 = 2 f 0 2 τ p ( 1 p ). This allows comparison between the relative strengths of thermal and active forces.
Using the normal modes of the fractal network Ψ α (for details, see  Appendix A), whose corresponding eigenfrequencies are ω α, and transforming Eq. (5) to the normal mode space, we obtain
(6)
Here, u α ( t ) is the amplitude of a normal mode Ψ α at time t and Γ α = m ω α 2 Λ α is the mode relaxation rate, where Λ α = 1 / γ is the mode mobility coefficient. ζ α ( t ) and F α ( t ) are, respectively, the mode-transformed thermal white noise and the active noise field; the latter takes the form: (i) For force monopoles,
(7)
and (ii) for force dipoles,
(8)
where δ f j ( t ) = f j ( t ) f 0 p f ^ j are the force fluctuations about their mean, and Ψ α ( r j ) Ψ α , j is the jth entry of normal mode α.
Next, we calculate the auto-correlation function of the active force field in mode space (Appendix A 1),
(9)
where W α = Q ω ξ: (i) Q = 1, ξ = 0, for force monopoles, and (ii) Q = 2 C 0 ω o 2 d s / d l, ξ = 2 d s / d l, for force dipoles. Here, ϕ is the (force density) fraction of force monopoles or dipoles in the system, and C 0 is a numerical constant related to the mean localization properties of the normal modes.30,33–36
Solving the Langevin equation (6) (see Appendix A 3 for details), expanding the MSD of an internal network bead, Δ r ( t ) 2 ( r i ( t ) r i ( 0 ) ) 2 = ( u i ( t ) u i ( 0 ) ) 2 , in terms of the normal modes, and performing disorder averaging, we obtain the MSD as the sum of a purely thermal contribution, Δ r ( t ) 2 th, and a purely active contribution, Δ r ( t ) 2 ac,
(10)
where
(11a)
(11b)
(where d is the Euclidean dimension).
Using the vibrational DOS g ( ω ) ω d s 1, we can approximate the sums in Eqs. (11a)(11b) to integrals over the frequency, where the lower and upper integration limits determine (respectively) the shortest, τ 0 = Γ ( ω 0 ) 1 = γ / m ω 0 2, and the longest, τ N Γ ( ω min ) 1 τ 0 N 2 / d s relaxation times of the system. As shown in Refs. 58–61, focusing on the time regime τ 0 t τ ( R g ), the thermal MSD exhibits subdiffusion,
(12)
with ν th = 1 d s 2, provided that d s < 2. The prefactor is B th d k B T γ d s / 2 1.
For a large system such that the frequency spectrum is dense, Δ r ( t ) 2 ac can also be evaluated analytically. Assuming short times such that t τ τ N, we find, for both force monopoles and force dipoles, superdiffusive, ballistic-type behavior,
(13)
where the amplitude follows the scaling: (i) for force monopoles, B acs ϕ f 0 2 τ d s / 2 provided that d s < 2, and (ii) for force dipoles, given that d s / 2 + d s / d l > 1, we have B acs ϕ f 0 2 τ 1. Otherwise, for rare fractal geometries where d s / 2 + d s / d l < 1, we get B acs ϕ f 0 2 τ d s / 2 d s / d l.

1. Force monopoles

For times much longer than the force correlation time such that τ t τ N, we find for force monopoles
(14)
where ν ac = 1 d s / 2, which is identical to the thermal subdiffusion exponent. The (purely) active amplitude is B ac ϕ f 0 2 τ p ( 1 p ) γ d s / 2 2 such that
(15)
Hence, if the active system combines thermal ( T > 0) and active forces (denoted by an “eff” subscript), e.g., as would be expected for normal cells, the MSD is Δ r ( t ) 2 = B eff t ν where B eff = B ac + B th, and the amplitude ratio is
(16)
In addition, for arrested center-of-mass (CM) translation motion, the MSD saturates at a value 2 r 2 ac where r 2 ac is the (active) static variance, Eq. (A51). The latter diverges with the lower integration limit ω min, and, thus, shows a Landau–Peierls-like instability r 2 ac N 2 / d s 1, similar to the passive case.71,72 Hence, for this particular case of force monopoles combined with thermal forces, it is possible to assign an effective temperature,
(17)
which can describe both the subdiffusion and saturation regimes of the MSD of an active system. Noteworthy, this applies exactly only for the second moment (i.e., MSD), and likely for higher moments, it is not appropriate since the active noise is Poissonianly distributed (i.e., the random telegraph process), while the thermal noise is Gaussianly distributed.73 An effective temperature is frequently used to characterize active or living systems,74–77 but clearly, it is not generally applicable (e.g., as shown next for force dipoles).

It is of interest to extend the subdiffusion exponents to the case of a viscoelastic solvent, whose frequency-dependent complex modulus behaves as a power-law, G ( ω ) ( i ω ) α, as discussed in Sec. I. This is presented in  Appendix C, where we obtain ν t h = α ( 1 d s 2 ) and ν a c = α ( 2 d s 2 ) 1, which reduce to the known results for a linear chain, d s = 1.25,54–56 Hence, for any fractal in viscoelastic solvent, the two exponents are equal only for α = 1, the viscous case. Interestingly, for α < ( 2 d s 2 ) 1, the active subdiffusion regime disappears.

2. Force dipoles

For force dipoles—since for typical fractals 1 d s / 2 d s / d l < 0an anomalous subdiffusion regime is absent, the MSD crosses over from the early ballistic motion t 2 to saturation at 2 r 2 ac. As now the latter diverges [see Eq. (A51)] with the upper integration limit ω 0, it becomes essentially independent of N. This implies that, surprisingly, for the combination of thermal and active forces, r 2 th may dominate the total static MSD for large enough systems due to the Landau–Peierls instability. Moreover, if force monopoles are also present, they are likely to dominate the evolution (as discussed in Sec. IV).

These analytical results are based on the scalar elasticity Hamiltonian [Eq. (4)], assuming a large system, N . To verify the sensitivity of our results to various assumptions, we performed Langevin dynamics (LD) simulations of a bead-spring Sierpinski gasket (Fig. 11), using different generations where the nth generation is denoted by S n, with n = 1 , 2 , . We solve Eq. (5) with H = 1 2 m ω 0 2 i j ( r i r j b r ^ i j ) 2 for the Hamiltonian of harmonic springs having an equilibrium distance b, where r ^ i j is the unit vector of the distance r i r j between beads i and j. For b = 0, it reduces to the scalar elasticity Hamiltonian discussed above. We follow the position of an internal bead (see Fig. 12) in a gasket and calculate its time-averaged MSD using the standard method for stationary noise. Throughout the simulation, we worked with dimensionless units, where distance is in units of b, force in units of m ω 0 2 b, and times are in units of τ 0 = γ / ( m ω 0 2 ), with k B T / m ω 0 2 b 2 = 1 (see details in  Appendix B).

Consider first the dynamics of a thermal (passive) network (previously studied58–61). We compare the MSD obtained from simulations, for both vanishing and non-vanishing b, with the numerical evaluation of Eq. (11a) for S 9, N = 9843. In Fig. 1(a), anomalous subdiffusion is observed spanning over three decades with exponents agreeing very well with the predicted theoretical value ν th = 1 d s / 2 = 0.317 for the Sierpinski gasket with d s = 1.365.33,78 The MSD amplitude of the analytical model, which is based on the scalar elasticity Hamiltonian, matches well with the LD simulation results for b = 0, while it differs from the MSD for b = 1. The short and long-time behavior represents regular diffusion, r 2 t, such that the short time is the independent bead diffusion, while the long represents the CM diffusion. For MSD of internal and peripheral beads, we find almost identical exponents but somewhat different amplitudes [see Fig. 1(b)], showing the faster motion of the peripheral bead than that of the internal bead.

FIG. 1.

Thermal network. Mean square displacement (MSD) of a bead in a passive Sierpinski gasket S 9 ( N = 9843): (a) MSD of an internal bead, estimated from LD simulations for b = 0 (square symbol) and b = 1 (triangle symbol) equilibrium bond length and analytical equation (11a) (circle symbol) with d = 2. (b) Comparison of the MSD of two different beads, internal and peripheral, for b = 1. Curves were fitted to a power law B th t ν th in the time range 5 < t < 50 000, yielding ν th 0.32 in all cases while having different amplitudes [specifically, ν th = 0.322 ± 0.001, 0.326 ± 0.001, and 0.327 ± 0.002 for (a) top to bottom symbols, and ν th = 0.326 ± 0.001 and 0.321 ± 0.001 for (b) top to bottom symbols].

FIG. 1.

Thermal network. Mean square displacement (MSD) of a bead in a passive Sierpinski gasket S 9 ( N = 9843): (a) MSD of an internal bead, estimated from LD simulations for b = 0 (square symbol) and b = 1 (triangle symbol) equilibrium bond length and analytical equation (11a) (circle symbol) with d = 2. (b) Comparison of the MSD of two different beads, internal and peripheral, for b = 1. Curves were fitted to a power law B th t ν th in the time range 5 < t < 50 000, yielding ν th 0.32 in all cases while having different amplitudes [specifically, ν th = 0.322 ± 0.001, 0.326 ± 0.001, and 0.327 ± 0.002 for (a) top to bottom symbols, and ν th = 0.326 ± 0.001 and 0.321 ± 0.001 for (b) top to bottom symbols].

Close modal

Turning to active fractal networks with force monopoles, to each bead, we assign an active force of random orientation. In Figs. 2(a) and 2(c) (drift velocity eliminated, see below), we depict the MSD of an internal bead for different values of active force parameters and arrested thermal motion ( T = 0). We observe a crossover at t τ, from short-time ballistic motion, MSD t 2, to subdiffusion at intermediate times. In Fig. 2(a), a crossover to ballistic motion appears at longer times, indicating the presence of a CM drift velocity. The latter emerges from the incomplete cancellation of the total force due to the system’s finite size. For N , i j f ^ i f ^ j = 0; however, in a finite system, this sum may have a residual (random) value, resulting in drift velocity v drift ϕ f 0 p γ N (Appendix B 1), which will nevertheless vanish if we average over many realizations of the force field.

FIG. 2.

Force monopoles. MSD of an internal bead of a Sierpinski gasket S 9 ( N = 9843) under stochastic force monopoles at each bead ( ϕ = 1) for various active force parameters with “on” probability p = 0.5, with arrested thermal motion ( T = 0) in (a)–(c). In (a), the MSD is calculated as is, while in (b)–(d), we subtract, using Eq. (B4) (see Appendix B 1), the drift motion of the CM, to suppress the ballistic motion at long times. (a) MSD at different τ for a given force amplitude f 0 = 1. The long-time t 2 behavior indicates the ballistic motion of the CM, associated with a drift velocity that emerges due to the finite size of the gasket. (b) MSD estimated from analytical equation (11b) and LD simulations for both vanishing and non-vanishing bond lengths for f 0 = 1 and τ = 1 with the CM drift subtracted, and the CM diffusion motion is observed at long times. (c) MSD of the active gasket for different force parameters. The subdiffusion exponent ν ac = 0.322 ± 0.003 is obtained from a power-law fit in the range [5–50 000]. MSD amplitudes, Eqs. (13)(14): B a c s = 0.0678 , 0.7132 , 0.0034 and B a c = 0.247 , 17.3 , 20.3 for ( τ , f 0 ) = ( 1 , 1 ) , ( 4 , 5 ), and ( 100 , 1 ), respectively. (d) MSD of the active gasket, including thermal fluctuation, T 0. The combination of active and thermal (passive) contributions to the MSD is compared with the purely thermal MSD. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

FIG. 2.

Force monopoles. MSD of an internal bead of a Sierpinski gasket S 9 ( N = 9843) under stochastic force monopoles at each bead ( ϕ = 1) for various active force parameters with “on” probability p = 0.5, with arrested thermal motion ( T = 0) in (a)–(c). In (a), the MSD is calculated as is, while in (b)–(d), we subtract, using Eq. (B4) (see Appendix B 1), the drift motion of the CM, to suppress the ballistic motion at long times. (a) MSD at different τ for a given force amplitude f 0 = 1. The long-time t 2 behavior indicates the ballistic motion of the CM, associated with a drift velocity that emerges due to the finite size of the gasket. (b) MSD estimated from analytical equation (11b) and LD simulations for both vanishing and non-vanishing bond lengths for f 0 = 1 and τ = 1 with the CM drift subtracted, and the CM diffusion motion is observed at long times. (c) MSD of the active gasket for different force parameters. The subdiffusion exponent ν ac = 0.322 ± 0.003 is obtained from a power-law fit in the range [5–50 000]. MSD amplitudes, Eqs. (13)(14): B a c s = 0.0678 , 0.7132 , 0.0034 and B a c = 0.247 , 17.3 , 20.3 for ( τ , f 0 ) = ( 1 , 1 ) , ( 4 , 5 ), and ( 100 , 1 ), respectively. (d) MSD of the active gasket, including thermal fluctuation, T 0. The combination of active and thermal (passive) contributions to the MSD is compared with the purely thermal MSD. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

Close modal

1. Subdiffusion

Calculating the MSD about the drift position using Eq. (B4) (Appendix B 1), we show in Fig. 2(c) that the intermediate-time subdiffusion regime becomes longer, MSD t ν ac with ν ac 1 d s / 2 0.32 as analytically predicted, and for t τ N, the long-time ballistic motion is replaced with t behavior associated with CM active diffusion. We can also observe the effects of the correlation time τ and force amplitude f 0 on the MSD ballistic and subdiffusion amplitudes confirming our prediction B a c s ϕ f 0 2 τ 0.682 ( t τ) and B a c ϕ f 0 2 τ ( t τ). Like the passive case, for b = 0, theory and simulations match exactly, see Fig. 2(b); for b = 1, the subdiffusion amplitude somewhat differs; yet, the exponent is unaltered.

2. Combined active and thermal motion

In Fig. 2(d), we show the combined effect of active and thermal forces for a few sets of parameters and compare them with the thermal MSD. When thermal forces are present, at short times t < τ 0, the thermal contribution dominates, resulting in diffusive behavior. At intermediate times, subdiffusion regimes with ν 1 d s / 2 0.32 can be observed for all cases; in one combination, having a long τ ( = 100 τ 0) and small f 0 ( = m ω 0 2 b), two consecutive such regimes appear, where the first is dominated by thermal motion and is followed by a second one, dominated by activity and having a much larger amplitude.

Next, we simulate force dipoles randomly assigned to a fraction ϕ = 20 % of total bonds (unless otherwise specified), where we either prohibit or do not prohibit dipoles having a common bead. In biological networks, the choice of distinct beads may arise due to the excluded volume interaction between motor proteins. During the simulations, we update each dipole orientation to remain parallel to r i j.

1. Absence of subdiffusion

We estimate the MSD for gasket S 9 (Fig. 3) for different force parameters. For times t τ, the MSD shows a ballistic behavior, t 2, which saturates to a constant value at t τ. Both the existence of the early ballistic motion and the absence of the intermediate subdiffusion regime for the Sirepinski gasket where 1 d s / 2 d s / d l = 0.5464 < 0 are in accord with the analytical predictions.

FIG. 3.

Force dipoles. MSD of an internal bead of gasket S 9 ( N = 9843) subjected to stochastic force dipoles in the absence of thermal motion. The force dipoles are randomly assigned to ϕ = 20 % of the network bonds. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

FIG. 3.

Force dipoles. MSD of an internal bead of gasket S 9 ( N = 9843) subjected to stochastic force dipoles in the absence of thermal motion. The force dipoles are randomly assigned to ϕ = 20 % of the network bonds. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

Close modal

2. Long-time rotational motion

In Fig. 3, at longer times, a ballistic-like rise in MSD for higher f 0 2 τ is observed, which requires further analysis. It is followed by oscillations that can be picked for smaller gaskets; see Fig. 4(a) for S 4. In the movies (snapshots displayed in Figs. 16–18 in  Appendix C, Multimedia available online) showing trajectories of a few gasket generations, we can observe persistent rotational motion of the network, either clockwise or anti-clockwise, that may explain the MSD oscillatory behavior. To confirm this, we define a mean square angular displacement (MSAD) associated with the pure rotational motion of an arbitrary bead i, located at a mean square distance R 2 from the CM, R 2 ( n ^ ( t ) n ^ ( 0 ) ) 2 , where n ^ ( t ) is the unit vector of R . In Fig. 4(a), we plot the MSAD together with the complete MSD; the inset (linear scale) emphasizes the motion at long times. The overlap of the two oscillatory curves is almost perfect. Importantly, even though force dipoles cannot generate a net torque, they can induce a mean rotational velocity (see discussion in Appendix B 2).

FIG. 4.

Force dipoles. MSD of an internal bead of gasket S 4 ( N = 42) subjected to stochastic force dipoles in the absence of thermal motion. (a) MSD Δ r 2 ( t ) , and mean square angular displacement (MSAD), R 2 ( n ^ ( t ) n ^ ( 0 ) ) 2 , of an internal bead with force dipoles randomly assigned to ϕ = 20 % of the network bonds. The inset shows both curves on a linear scale. (b) MSD of an internal bead of S 4 for ϕ = 20 % bonds (blue color) and ϕ = 100 % bonds (red color) associated with force dipoles. The force parameters are τ = 1.0, p = 0.5, and f 0 = 0.1. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

FIG. 4.

Force dipoles. MSD of an internal bead of gasket S 4 ( N = 42) subjected to stochastic force dipoles in the absence of thermal motion. (a) MSD Δ r 2 ( t ) , and mean square angular displacement (MSAD), R 2 ( n ^ ( t ) n ^ ( 0 ) ) 2 , of an internal bead with force dipoles randomly assigned to ϕ = 20 % of the network bonds. The inset shows both curves on a linear scale. (b) MSD of an internal bead of S 4 for ϕ = 20 % bonds (blue color) and ϕ = 100 % bonds (red color) associated with force dipoles. The force parameters are τ = 1.0, p = 0.5, and f 0 = 0.1. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

Close modal

To investigate the “persistent” rotational motion of the entire network, we compute the angular velocity Ω i of individual beads and subsequently determine the mean angular velocity (averaged over beads) Ω ¯ [Eq. (B15)]. The color map [Fig. 5(b)] displays the temporal fluctuation of angular velocity Ω i for individual beads. Similarly, fluctuation of Ω ¯ over time is shown in Fig. 5(c). However, the time average of mean angular velocity, Ω ¯ T W, for different time-averaging windows, T W [see Fig. 5(d)] is found to be non-vanishing and converges to a constant value for long values of T W, which depends on the (quenched) realization of the dipolar forces. Given the local fluctuation in Ω ¯, the non-zero value of Ω ¯ T W results in a “crawling” (non-coherent) rotation of a whole object, which appears as persistent rotational motion at long-time windows.

FIG. 5.

Force dipoles. Rotational motion of Sierpinski gasket S 4 subjected to stochastic force dipole with arrested thermal motion ( T = 0). (a) Schematic of S 4 ( N = 42) with indices from 1 to 42 assigned to different beads. (b) Color map showing the time-dependent angular velocity Ω i fluctuations of each bead. (c) Temporal mean angular velocity Ω ¯ [Eq. (B15)] and the mean radius-square-weighted angular velocity Ω | R | 2 ¯ [Eq. (B14)], averaged over beads at different times. (d) The time average of mean angular velocity, Ω ¯ T W, for different time-averaging windows, T W, converging to a non-vanishing constant value for long values of TW that depends on the distribution of the dipolar forces. (e) Ω | R | 2 ¯ for different simulation time step Δ t. The fluctuations seen in Ω | R | 2 ¯ improve with a smaller Δ t. (f) Ω ¯ T W for different simulation time step Δ t. The difference between the two curves is minor and effectively disappears when the time-averaging window gets long. The force parameters are τ = 1.0, p = 0.5, and f 0 = 0.5. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

FIG. 5.

Force dipoles. Rotational motion of Sierpinski gasket S 4 subjected to stochastic force dipole with arrested thermal motion ( T = 0). (a) Schematic of S 4 ( N = 42) with indices from 1 to 42 assigned to different beads. (b) Color map showing the time-dependent angular velocity Ω i fluctuations of each bead. (c) Temporal mean angular velocity Ω ¯ [Eq. (B15)] and the mean radius-square-weighted angular velocity Ω | R | 2 ¯ [Eq. (B14)], averaged over beads at different times. (d) The time average of mean angular velocity, Ω ¯ T W, for different time-averaging windows, T W, converging to a non-vanishing constant value for long values of TW that depends on the distribution of the dipolar forces. (e) Ω | R | 2 ¯ for different simulation time step Δ t. The fluctuations seen in Ω | R | 2 ¯ improve with a smaller Δ t. (f) Ω ¯ T W for different simulation time step Δ t. The difference between the two curves is minor and effectively disappears when the time-averaging window gets long. The force parameters are τ = 1.0, p = 0.5, and f 0 = 0.5. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

Close modal
FIG. 6.

Force dipoles: Radius of gyration R g of Sierpinski gasket S 4 (on a semi-log plot) at different f 0 for τ = τ 0 and p = 0.5: (a) contractile and (b) extensile force dipoles. Snapshots of the gasket at different f 0 are shown for both cases. R g and f 0 are in units of b and m ω 0 2 b, respectively.

FIG. 6.

Force dipoles: Radius of gyration R g of Sierpinski gasket S 4 (on a semi-log plot) at different f 0 for τ = τ 0 and p = 0.5: (a) contractile and (b) extensile force dipoles. Snapshots of the gasket at different f 0 are shown for both cases. R g and f 0 are in units of b and m ω 0 2 b, respectively.

Close modal
FIG. 7.

Force dipoles: Radius of gyration R g of a Sierpinski gasket (on a semi-log plot) at different f 0 for contractile force dipoles with p = 0.5: (a) for different generations, at given τ = τ 0, and (b) at different τ for gasket S 5. N = 42 , 123 , 366, and 1095 for S 4, S 5, S 6, and S 7, respectively. R g, f 0, and τ are in units of b, m ω 0 2 b, and τ 0, respectively.

FIG. 7.

Force dipoles: Radius of gyration R g of a Sierpinski gasket (on a semi-log plot) at different f 0 for contractile force dipoles with p = 0.5: (a) for different generations, at given τ = τ 0, and (b) at different τ for gasket S 5. N = 42 , 123 , 366, and 1095 for S 4, S 5, S 6, and S 7, respectively. R g, f 0, and τ are in units of b, m ω 0 2 b, and τ 0, respectively.

Close modal

Condition of net zero torque requires that mean radius-square-weighted angular velocity is zero, i.e., Ω | R | 2 ¯ = 0 [Eq. (B14)], which does not necessarily imply, Ω ¯ = 0. Here, R i is the position of the ith bead in the CM frame of reference. The time-dependent fluctuations of Ω | R | 2 ¯ presented in Fig. 5(c) are found to be two orders of magnitude smaller than that of Ω ¯, despite that for all beads, | R i | 2 > 1. The fluctuations seen in Ω | R | 2 ¯ reflect the minor simulation errors, which improve with a smaller simulation time step, Δ t. When the simulation time step is tenfold decreased, these fluctuations get significantly smaller, shown in Fig. 5(e). Nevertheless, in Fig. 5(f), we show that the difference in Ω ¯ T W between the two simulations (of different time steps) is minor and effectively vanishes when the time-averaging window gets long, demonstrating that the “crawling” rotation of the whole object is real.

We further argue that this rotational motion, whose direction and frequency randomly vary between realizations of the dipole spatial distribution [Fig. 13(a) in Appendix B 2], arises from a small residual anisotropy caused by the asymmetric distribution of forces associated with finite-size effects. As shown in Fig. 13(c), the onset of oscillations is delayed to longer times as N increases, reflecting a reduction in anisotropy. Furthermore, simulating a gasket where force dipoles are assigned to all bonds completely eliminates the long-term MSD oscillations, as shown in Fig. 4(b).

Interestingly, in triangular micro-swimmers that are subject to force dipoles and hydrodynamic interaction, rotational motion emerges too.79,80 In Fig. 15 (see  Appendix B), we show that even Rouse dynamics (i.e., without hydrodynamic interaction) of a triangular bead-spring, with anisotropy of force dipole strength (and random-telegraph fluctuations), is enough to produce rotational motion.

In fact, it has been suggested in general that, despite the vanishing total torque, a “crawling” rotational motion—associated with a given sequence of deformations—could be expected for a deformable body, once the rotational symmetry of (internal) forces, i.e., the force dipoles, is broken,81–84 similar to translational swimming in low Reynolds’ numbers. The broken symmetry due to a force dipole distribution and the system’s geometry can create localized deformations that can spread through the system in a coordinated manner. This sequence of deformations (shape changes) can cause the body to “crawl” or exhibit a slow, persistent, and directional (albeit non-torque-driven) rotational motion.

3. Network collapse

The above behavior for force dipoles does not apply above a critical force amplitude f 0 c m ω 0 2 b. For f 0 > f 0 c, the gasket collapses to a random shape maintaining dynamical fluctuations, but with rotational motion effectively arrested (see Fig. 14). The radius of gyration R g for contractile force dipoles against f 0 is shown in Fig. 6(a) for S 4 and in Fig. 7(a) for generations S 5 S 7. The transition persists and sharpens for larger (higher generation) gaskets, implying a true, activity-induced, first-order phase transition.

The collapse may be rationalized as a result of the dynamical “persistence length” 85,86—generalized to account for the spring resistance—rising above b, leading to f 0 c m ω 0 2 b / ( 1 e τ / τ 0 ) (see Appendix B 3), in agreement with our results for τ τ 0, Fig. 7(b). Contrary to contractile force dipoles [Fig. 6(a)], for extensile force dipoles R g rises again at larger forces, Fig. 6(b).

To connect with chromatin dynamics studies described in Ref. 26, we need to establish the parameter regime associated with chromatin. We hypothesize that the chemical distance between cross-links in chromatin is equal to the persistence length L p 250 nm,87,88 and that τ and f o are associated with cytoplasmic motor protein processivity times and forces. Cytoplasmic typical values are in the range f 0 = 1 10 p N, τ = 10 3 10 s.89–92 To obtain the dimensionless parameter range applicable to chromatin dynamics, we assume that the network segment between two beads behaves as a Gaussian 2D thermal spring with spring constant m ω 0 2 = 90 k B T L p 2 b 4 for semi-flexible polymers,93,94 and the bond length between two beads is b = L p. Taking the solvent viscosity (water) η = 1 × 10 3 Pa s, we estimate m ω 0 2 b = 90 k B T b = 1.5 pN and τ 0 = γ m ω 0 2 = 3 π η b 3 90 k B T = 4 × 10 4 s. Hence, for chromatin, the dimensionless values of force amplitude and force correlation time are in the range f 0 / ( m ω 0 2 b ) = 0.5–7 and τ / τ 0 = 1 10 4, respectively. These parameters yield MSDs of passive systems and purely active (i.e., T = 0) force monopole systems, exhibiting a subdiffusion regime with identical exponents, yet different amplitudes, as shown in Figs. 2(b)2(c).

To describe realistically untreated cells, we should consider the interplay of thermal forces (as white noise) with both active monopolar and dipolar forces. Moreover, with ν th = ν ac = 1 d s / 2 0.4 (the value observed in Ref. 26), we can predict d s 1.2, associated with a slightly cross-linked chain. This deviates from the original fractal globule model47 that ignores the presence of such cross-links and corresponds to the case where d s = 1. Indeed, Rouse dynamics of passive and force-monopole-active linear Gaussian chains do yield ν t h = ν a c = 1 / 2, consistent with our result with d s = 1.28,29 Our deduction of d s 1.2 does agree with experimental findings showing slight deviations from the fractal-globule model predictions for the static correlations and the fractal dimension.41 Furthermore, Bronstein et al. demonstrated the role as cross-linkers of lamin A proteins in eukaryotic cell chromatin.24 

In Fig. 8, we present the combination of thermal, monopolar, and dipolar forces (with comparison to the individual force source cases) for a fractal network ( N 10 4) with spectral dimension d s = 1.2. In Fig. 8(a), we estimate the MSD from the numerical integration of Eq. (A37), which does not include the CM motion, using ω min = ω 0 N 1 / d s for the lower limit and the spectral dimension d s = 1.2 as we suggest for chromatin. In Fig. 8(b), we add the CM motion to Eq. (A37) according to Eqs. (A33)(A34). We can still observe a significant subdiffusion regime described by an exponent ν 1 d s / 2 = 0.4, which emerges since the dipolar sources alone lead only to a moderate MSD saturation value [denoted as 2 r 2 ac in Eq. (A52)] allowing the combination of thermal and monopolar forces to take over. For comparison, in Fig. 9, we show the equivalent MSD results for a Sierpinski gasket ( N = 9843) from LD simulations. While the early evolution is similar to Fig. 8 (except for a different subdiffusive exponent), for force dipoles alone, we also observe the rise from the plateau at long times (associated with the slow rotation), which is not captured by the analytic theory. However, it is clear that the dipolar forces do not contribute to the subdiffusive motion induced by the active monopolar and thermal forces. Hence, we can assess that monopolar and thermal forces are the cause of the subdiffusive motion in chromatin.

FIG. 8.

Chromatin-like active network with force monopoles and dipoles both included: MSD of an arbitrary bead of a fractal network ( N = 9843) with spectral dimension d s = 1.2, subject to active forces (both dipolar and monopolar forces) and thermal forces, compared to motion with different force sources individually: (a) MSD without the CM motion, and (b) MSD with CM motion. The MSD is estimated from the numerical integration of Eq. (A37) with ω min = ω 0 N 1 / d s for the lower limit, using the spectral dimension d s = 1.2, that is predicted from our theoretical analysis ν = 1 d s / 2 considering the experimental value of subdiffusion exponent ν = 0.4 for chromatin in Ref. 26. For active forces, the force parameters are f 0 = 1, τ = 100, p = 0.5, and ϕ = 0.2 for both monopolar and dipolar forces such that the relative strength of active forces (variance) to thermal forces (variance) for the network is ϕ f 0 2 τ p ( 1 p ) / d = 2.5 (in a dimensionless unit). r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

FIG. 8.

Chromatin-like active network with force monopoles and dipoles both included: MSD of an arbitrary bead of a fractal network ( N = 9843) with spectral dimension d s = 1.2, subject to active forces (both dipolar and monopolar forces) and thermal forces, compared to motion with different force sources individually: (a) MSD without the CM motion, and (b) MSD with CM motion. The MSD is estimated from the numerical integration of Eq. (A37) with ω min = ω 0 N 1 / d s for the lower limit, using the spectral dimension d s = 1.2, that is predicted from our theoretical analysis ν = 1 d s / 2 considering the experimental value of subdiffusion exponent ν = 0.4 for chromatin in Ref. 26. For active forces, the force parameters are f 0 = 1, τ = 100, p = 0.5, and ϕ = 0.2 for both monopolar and dipolar forces such that the relative strength of active forces (variance) to thermal forces (variance) for the network is ϕ f 0 2 τ p ( 1 p ) / d = 2.5 (in a dimensionless unit). r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

Close modal
FIG. 9.

Active Sierpinski gasket with chromatin-like force monopoles and dipoles from LD simulations: MSD of an internal bead of Sierpinski gasket S 9 ( N = 9843) subjected to active forces (both dipolar and monopolar) and thermal forces, compared to motion with different force sources individually. Here, the MSD is estimated from LD simulations, and the force parameters are the same as in Fig. 8.

FIG. 9.

Active Sierpinski gasket with chromatin-like force monopoles and dipoles from LD simulations: MSD of an internal bead of Sierpinski gasket S 9 ( N = 9843) subjected to active forces (both dipolar and monopolar) and thermal forces, compared to motion with different force sources individually. Here, the MSD is estimated from LD simulations, and the force parameters are the same as in Fig. 8.

Close modal

To examine further the predicted MSD amplitudes (or “apparent diffusion coefficients”) in comparison with the experimental results,26 we consider in Fig. 10 the combination of thermal forces and active force monopoles in comparison with the purely thermal case (with dipolar forces ignored) for a chromatin-like fractal network where d s = 1.2. For small values of ϕ f 0 2 τ where ϕ f 0 2 τ p ( 1 p ) / ( d γ k B ) T, i.e., T eff T, the thermal contribution to the MSD dominates over the active contribution. For higher values of ϕ f 0 2 τ where T eff T, we observe the appearance of either a single subdiffusion regime dominated by active motion (for relatively small τ) or a thermally dominated subdiffusive regime crossing over to an actively dominated subdiffusive regime with the same exponent ν, albeit with a higher amplitude B eff. Importantly, this is quite similar to the observation in chromatin reported in Ref. 26 (see, e.g., Fig. 1 therein). The experimentally measured MSD amplitude ratios, B th / B eff, of ATP-depleted to normal cells26 (Table 1 therein) are theoretically reproduced depending on the combination ϕ f 0 2 τ. For example, for Saccharomyces cerevisiae, we need ϕ f 0 2 τ 100 to match the theoretical ratio with the experimental one, B th / B eff 0.12, and for E. coli, we require ϕ f 0 2 τ 13 to match the measurements.

FIG. 10.

Chromatin-like active network with force monopoles: MSD of an arbitrary bead of a fractal network ( N = 9843) with spectral dimension d s = 1.2, subjected to active monopolar forces and thermal forces together, compared to motion with thermal force alone for different force parameters: (a) MSD without CM motion and (b) MSD with CM motion. The MSD is estimated from the numerical integration of Eqs. (A37), the same as in Fig. 8. We note that the relative strength of active forces (variance) to thermal forces (variance) for the network is ϕ f 0 2 τ p ( 1 p ) / d (in a dimensionless unit). r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

FIG. 10.

Chromatin-like active network with force monopoles: MSD of an arbitrary bead of a fractal network ( N = 9843) with spectral dimension d s = 1.2, subjected to active monopolar forces and thermal forces together, compared to motion with thermal force alone for different force parameters: (a) MSD without CM motion and (b) MSD with CM motion. The MSD is estimated from the numerical integration of Eqs. (A37), the same as in Fig. 8. We note that the relative strength of active forces (variance) to thermal forces (variance) for the network is ϕ f 0 2 τ p ( 1 p ) / d (in a dimensionless unit). r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

Close modal

Some other recent MSD measurements of chromosomal locus in normal eukaryotic cells (i.e., no ATP depletion), either single-point MSD or two-point MSD,95–97 do show subdiffusion exponent, ν values closer to 1 / 2, and fractal-globule packing with d f = 3.95 Importantly, these results are in perfect accord with our general approach. Provided that d s = 1, as is the case for a fractal globule, the anomalous diffusion exponent ν = 1 d s / 2 will always be 1/2, regardless of the fractal dimension d f. It remains unknown whether ATP-depleted cells in such systems also exhibit the same exponent, a significant observation as we discussed. In case they would, for example, our general approach would enable us to infer that d s = 1 in these cells, as in the original fractal-globule model. To draw such a conclusion, it is imperative to have MSDs of both ATP-depleted and normal cells, where the values of ν are identical. This highlights how MSD measurements provide insights into the abundance of cross-links in chromatin.

Recent active, dipolar-force-based models for chromatin dynamics also appear to yield identical exponents, yet with ν t h ν a c 0.5.98,99 In Ref. 98, the active system is studied at finite temperatures, and the subdiffusion is found only for weak forces such that the system is thermally dominated.100 When the system becomes dominated by stronger forces,101 the MSD shows a crossover from super-diffusion at times shorter than the correlation time, to a constant value at time longer than the correlation time. Notably, a similar behavior is obtained in the present work only for dipolar forces.

In Ref. 99, the time window where anomalous subdiffusion was found numerically in the pure active case, 0.05 < t < 0.30,102 is less than one decade and is likely picking the crossover behavior from the t 2 to a constant value. This crossover from t 2 to a constant value is qualitatively identical to our findings for dipolar forces (e.g., Fig. 4). In contrast, our sub-diffusion regime for force-monopole ranges over a few decades (Fig. 2), demonstrating a true anomalous regime. Reference 99 claims to derive analytically ν = 1 / 2 for force dipoles (to support the numerical study) and ν = 3 / 2 for force monopoles, both for times shorter than the force correlation time. In contrast, our study yields ν = 2 for both force monopoles and dipoles in this time range.

To conclude this section, we comment again on the inability of the “power-law fluid” interpretation, G ( ω ) ( i ω ) α, to explain the observed equality of the active and thermal exponents on the same footing.25,26 Our extended exponents ν t h = α ( 1 d s 2 ) and ν a c = α ( 2 d s 2 ) 1 imply that ν a c ν t h for any d s, unless α = 1 (viscous fluid). Taking25  d s = 1 and α in the range 0.7–0.8 implies ν t h = 0.35 0.4 and ν a c = 0.05 0.2, which could be well discriminated experimentally. Furthermore, for d s = 1 and α < 2 / 3, one obtains ν a c = 0, i.e., the active subdiffusion is absent, and the MSD saturates to a constant value or crosses over to the CM diffusion.

We studied analytically arbitrary fractal networks that obey harmonic scalar elasticity and Rouse-type dynamics and are subject to active force monopoles and force dipoles. We also performed Langevin dynamics simulations of the Sierpinski gasket considering a nonzero equilibrium spring length for which the elasticity is not perfectly scalar. Our results show the absence of a true power-law subdiffusion regime when only force dipoles are at play within an ideal fractal network. However, when thermal forces, active force monopoles, or both are introduced into the system, their impact on MSD surpasses that of the dipoles, with the latter contributing only a constant saturation value at intermediate times. The resulting subdiffusion exponent is always ν = 1 d s / 2 irrespective of whether thermal forces, active force monopoles, or a combination of both predominate the dynamics. The LD simulations confirm the theory with high accuracy and show that the non-vanishing spring length does not alter this exponent. The value of the exponent ν depends solely on the spectral dimension d s and is independent of the fractal dimension d f of the self-similar structure. In particular, for a linear chain that folds in space without forming internal cross-links, we always have d s = 1 and ν = 1 / 2, regardless of whether d f = 2 for Gaussian chains, d f = 5 / 3 for self-avoiding chains, or d f = 3 for compact folding.

Therefore, we explain here the a priori surprising result of an identical subdiffusion exponent for both normal and ATP-depleted cells. Moreover, we showed that within the range of force strengths ( f 0), force correlation times ( τ), and force density ( ϕ), yielding different combinations of ϕ f 0 2 τ, we obtain amplitude ratios, B th / B eff, of ATP-depleted to normal cells that are similar to those experimentally observed.26 We conclude that the observed normal cell chromosomal-loci subdiffusion is caused by the combination of active force monopoles and thermal forces. However, we do not claim that dipole forces are absent; they simply do not contribute to this observable. We stress that our conclusions do not rely on the type of self-similarity for chromatin packing, in particular, whether it obeys the fractal-globule model,47,48 the loop-extrusion model,103 or a combination of the fractal-globule model with added cross-links.49,104

For force dipoles, we also discovered two other distinct behaviors that we consider fundamentally significant. First, we observed the collapse of the fractal to a compact object when the force amplitude f 0 exceeds a critical value, for which the dynamical persistence length, generalized to account for a harmonic potential, exceeds the equilibrium distance between beads. This is reminiscent of the previously studied “motility-induced phase transition” and other dynamical transitions in active systems.3,7–11,105 However, it should be noted that the structure in the collapsed state will be modified if an excluded volume interaction is included, presenting a low cutoff for the collapsed state size. We, therefore, did not delve into a detailed study of its structure and dynamics.

Second, below the critical force mentioned above, we observed a slow rotational, “crawling” motion of the whole fractal network, whose direction is random and chosen by the symmetry (isotropy) breaking associated with the random realization of the dipolar force locations. We have shown that this behavior is very similar to the one observed in triangular microswimmers for which the asymmetry is broken by choice.79,80 Even though the torque that each dipole exerts is vanishing, and so is the total torque, such a “crawling” rotational motion is possible, in general, for a deformable body81–84 but not for a rigid body. Thus, force dipoles, which are expected to exist in chromatin,53 should cause slow rotational motion of the chromosome (as in Fig. 4), provided that chromatin is free to rotate, e.g., when it is not anchored to the nuclear envelope in eukaryotic cells by lamin A (i.e., for lamin A depletion) or other linkers. Interestingly, a large-scale coherent motion has been experimentally observed in chromatin (possibly presenting local rotational motion of chromosomal territories) and backed up by simulations that include dipolar forces.106 Rotational motion has also been observed in cytoplasmic flow and proposed—conceptually in accord with our simulations—to be caused by actomyosin force dipoles.107,108

While our simulations were performed on a specific deterministic fractal, the Sierpinki gasket, our analytical theory, which agrees with the simulations, applies to any fractal, be it a deterministic or a disordered fractal, which makes our conclusions universal. We do not imply by any means that chromatin appears like the Sierpinski gasket; only that they are both fractal networks, which allows us to conclude about chromatin dynamics also from our simulations. Simulations of critical percolation clusters that form a disordered fractal below the percolation correlation length are currently underway to further check the general applicability of our analytical theory and the universality of our conclusions. Further work is also planned to strengthen the connection between the observed chromatin dynamics in normal and treated cells (e.g., ATP-depleted and lamin A-depleted cells) and our fractal model.24,26

S.S. acknowledges the BGU Kreitmann School postdoctoral fellowship. We thank Sam Safran, Itay Griniasty, Tom Witten, and Omer Granek for useful discussions and the BGU Avram and Stella Goldstein-Goren fund for support.

The authors have no conflicts to disclose.

Sadhana Singh: Formal analysis (equal); Methodology (equal); Software (lead); Visualization (lead); Writing – original draft (equal); Writing – review & editing (equal). Rony Granek: Conceptualization (lead); Formal analysis (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

For the scalar elasticity Hamiltonian given by Eq. (4), the eigenstates (i.e., normal modes), Ψ α, are solutions of the eigenvalue equation
(A1)
where { ω α } are the corresponding eigen-frequencies, Ψ α , i is the ith entry of normal mode α, and j i denotes beads connected by springs to bead i. On a fractal network, these eigenstates are known as “fractons,” the vibrational excitations of the fractal.30,33,34,36,58,78 The eigenstates Ψ α form an orthonormal set34 such that i Ψ α , i Ψ β , i = δ α , β and α Ψ α , i Ψ α , j = δ i , j. This allows to define a normal mode transform u α = i u i Ψ α , i and an inverse transform u i = α u α Ψ α , i.
On a fractal, the normal modes Ψ α are strongly localized in topological space, unlike the oscillatory behavior characteristic of uniform networks. A disorder-averaged eigenstate may be defined according to
(A2)
where dis denotes disorder averaging, i.e., averaging over all realizations of the fractal, keeping the bead i and j fixed, or averaging, within a given realization, overall different pair of beads i and j that have the same topological space distance | l i l j |. ( l i is the dimensionless coordinate of bead i in topological space.) Note that mode normalization implies Ψ α , i 2 dis = 1 / N.
It has been shown that, in topological space, Ψ ¯ ( ω α , l ) obeys the following scaling: form30,33,34
(A3)
where f ( y ) is the scaling function. For y 1, f ( y ) is exponentially decaying, and for y 1, f ( y ) 1 C 0 × y 2, where C 0 is a numerical constant.34–36 
To analyze the dynamics of an active fractal network, we follow the Langevin equations described in Refs. 58–61 for a passive fractal in the high damping limit, generalized for the active forces. In addition to the white thermal noise, beads experience stochastic active forces distributed randomly but uniformly over the network. The stochastic active forces are assumed to fluctuate independently of each other and to follow the random-telegraph (on–off) process,68 where the random variables take two values, 0 and f 0, describing the “off” and “on” states of the force. The auto-correlation function of the forces, thus, follows:
(A4)
where t is the lag time, p is the probability for an “on” state, and τ is the force correlation time.
We assume that the force directions are isotropically distributed. An isotropic distribution of f ^ i’s implies
(A5)
in an infinite network; however, in a finite system, this sum may have a residual (random) value. This sum will nevertheless vanish on a finite system if we perform in addition disorder averaging, in which the force field spans over all its possible realizations.

For monopoles, the directional distribution is isotropic by definition. In the case of dipoles, although they are taken to align with the bonds, their overall directional distribution remains (statistically) isotropic due to the random realization of their locations. For off-lattice disordered fractals, such as randomly branched polymers, this directional distribution is effectively continuous. For on-lattice disordered fractals, such as critical percolation clusters, or deterministic fractals, such as the Sierpinski gasket used for the simulations, the distribution is isotropic within the allowed discrete directions (three, in the case of Sierpinski). Note that the spatial and directional correlations between any two forces making a local dipole are already accounted for in Eq. (2).

For convenience, we consider the force fluctuations about their mean f 0 p f ^ i. Introducing
(A6)
we have
(A7)
It follows that { u i } will stand henceforth for the displacements in bead positions about the mean coordinates, as dictated by the mechanical balance between the mean forces and the network elasticity.
For a Rouse-type fractal network, the set of Langevin equations of motion [the same as Eq.(5)] of an active system are
(A8)
where γ is the local friction coefficient ( γ = 3 π η b in the case of Stokes drag, where η is the solvent viscosity and b is the bead diameter), ζ i is the thermal white noise obeying FDT,70 and f i is the active force acting on the ith bead.
Using the normal modes of the fractal network Ψ α, whose corresponding eigen-frequencies are ω α, and transforming Eq. (A8) to the normal mode space, we obtain
(A9)
Here, u α ( t ) is the amplitude of a normal mode α at time t, and Γ α = m ω α 2 Λ α is the mode relaxation rate, where Λ α is the mode mobility coefficient, Λ α = 1 / γ. ζ α ( t ) is the (mode-transformed) thermal white noise that obeys the fluctuation–dissipation theorem,
(A10)
and, the last term, F α ( t ) is the mode transformed active noise fluctuations (about their mean), which takes the form: (i) for force monopoles,
(A11)
and (ii) for force dipoles,
(A12)
where Ψ α ( r i ) Ψ α , i.

1. Force auto-correlation function in mode space

We first calculate the auto-correlation function of the active force field in mode space:

(i) for force monopoles,
(A13)
(ii) for dipoles,
(A14)
The averages in Eqs. (A13)(A14) are performed over both the stochastic evolution of the active forces and disorder:
(i) For force monopoles, we obtain, after disorder averaging, using Eq. (A2),
(A15)
and after ensemble (or time) averaging over the stochastic forces using Eq. (A7), and implementation of the normalization Ψ ¯ ( ω α , 0 ) = 1, we get
(A16)
where ϕ is the fraction of force monopoles. Note that Eq. (A16) is independent of the mode index α.
(ii) For force dipoles, we first use Eq. (A2),
(A17)
and after ensemble (or time) averaging over the stochastic forces, using Eq. (A7),
(A18)
where ϕ is now the fraction of force dipoles, leading to
(A19)
Note that transforming Eq. (A3) from topological space to real space, by using l ( r / b ) d f / d l, (where d f and d l are the fractal and topological dimensions, respectively), implies
(A20)
Using ϵ = b and the low-frequency expansion for the eigenstates, we have Ψ ¯ ( ω α , ϵ ) 1 C 0 ( ω α / ω o ) 2 d s / d l, leading to
(A21)
To summarize, for both types of active forces, we have
(A22)
where
(A23)
such that
(A24)
and
(A25)

2. Normal mode amplitude auto-correlation function

Solving the Langevin equation (A9), we obtain the auto-correlation function of the displacement at a steady state such that the absolute times t 1, t 2 are taken to infinity (or that active forces started to act at time ); therefore, all transient behaviors are completely relaxed. The time difference t = | t 2 t 1 | is finite. Hence, u α ( t ) u α ( 0 ) u α ( t 2 ) u α ( t 1 ) . We obtain
(A26)
and the mean square normal mode amplitude is
(A27)
In Eq. (A27), the first term is the thermal contribution and the second is the contribution of the stochastic force fluctuations. For τ , the latter reduces to
(A28)

3. Mean square displacement

Expanding the two-point correlation function60 in terms of the exact normal modes Ψ α ( r ) and performing disorder averaging, we find the required MSD as the sum of a purely thermal contribution, Δ r ( t ) 2 th, and a purely active contribution, Δ r ( t ) 2 ac,
(A29)
using the result of Eq. (A26) and (A27), we, thus, obtain
(A30)
where
(A31)
(A32)
In Eqs. (A31) and (A32), the term corresponding to ω α = 0 describes the translational motion and should be omitted from the sum. Nevertheless, the translational CM motion can be deduced analytically by taking the limit ω α 0. It follows that both thermal and active CM motions are diffusive; the thermal CM motion is
(A33)
and the active CM motion is
(A34)
Using the vibrational DOS g ( ω ) = n o ω d s 1, where n o = N d s / ω o d s is chosen such that 0 ω o d ω g ( ω ) = N, we can approximate the sum in Eqs. (A31) and (A32) to an integral over the frequency. Note that the lower and upper integration limits set the shortest τ 0 = Γ ( ω 0 ) 1 and longest relaxation times τ N = Γ ( ω min ) 1 where ω min ω o ( R g / b ) d f / d s ω o N 1 / d s of flexible network dynamics. As already described in Refs. 58–61, focusing on the time regime τ 0 t τ ( R g ), the thermal MSD exhibits subdiffusion,
(A35)
with ν th = 1 d s 2, provided that d s < 2, and the amplitude (prefactor)
(A36)
Here, C th is a numerical prefactor, and C th = Γ [ d s 2 ] / ( 2 d s ) in the Rouse model ( Γ [ x ] is the Gamma function).
An analytic evaluation of the other active origin terms is partly possible for a large system such that the frequency spectrum is dense. The integral approximation of Eq. (A32) is
(A37)
Focusing on the intermediate-time regime τ 0 t τ N and times shorter than the force correlation time such that τ 0 t τ τ N, we find superdiffusive, ballistic-type behavior Δ r ( t ) 2 ac t 2 for both force types, monopoles and dipoles. More precisely, we obtain
(A38)
where
(A39)
Setting the lower and upper limits of integration in Eq. (A39) to 0 and infinity, respectively, and performing the integration, we find, provided that 1 d s / 2 ξ / 2 > 0,
(A40)
where C acs is a numerical factor, C acs = 0 d ω ω d s + ξ 1 ω 2 + 1 = 1 2 B ( d s + ξ 2 , 1 d s + ξ 2 ), and B ( x , y ) is the Beta function.109 For force monopoles, this implies
(A41)
with C acs = 1 2 B ( d s 2 , 1 d s 2 ); hence, B acs ϕ f 0 2 τ d s / 2. In the case of force dipoles, we note that the condition 1 d s / 2 d s / d l > 0 is commonly not obeyed in familiar fractals; in the rare situation where it is obeyed, B acs ϕ f 0 2 τ d s / 2 d s / d l for force dipoles. Otherwise, for 1 d s / 2 d s / d l < 0, the integral in Eq. (A39) diverges with the upper limit ω 0, and we find
(A42)
such that B acs ϕ f 0 2 τ 1. Note that the influence of τ at times t τ does not contradict causality, as we describe the fluctuations at a steady state where the laboratory time [ t t o t, cf. Equation (B3)] has surpassed τ many times.
For times much longer than the force correlation time, t τ, Eq. (A37) can be approximated as
(A43)
Provided that 1 d s / 2 ξ / 2 > 0, setting the lower and upper limits of integration to 0 and infinity (respectively) and scaling of the integral leads to
(A44)
where ν ac = 1 d s / 2 ξ / 2, and
(A45)
Here, C ac is a numerical constant, C ac = 0 d ω ω d s + ξ 3 ( 1 e ω 2 ) = Γ ( d s + ξ 2 ) / ( 2 d s ξ ) where Γ ( x ) is a Gamma function.109 Thus, for force monopoles, ν ac = 1 d s / 2, which is identical to the thermal subdiffusion exponent, and
(A46)
hence, the ratio of the active to thermal MSD amplitudes is
(A47)
For force dipoles, since, as noted usually 1 d s / 2 d s / d l < 0, the integral in Eq. (A43) diverges with the upper bound. Therefore, an anomalous subdiffusion regime is absent, and the MSD crosses over rapidly to a constant value (see below).

4. Static mean square displacement

For t { τ N , τ }, the MSD saturates at
(A48)
in which the thermal and active contributions are also additive,
(A49)
As previously shown for a thermal (passive) fractal network with d s < 2, the thermal contribution exhibits the generalized Landau–Peierls instability r 2 th N 2 d s 1.71,72 The complete expression for static thermal MSD is
(A50)
For the active contribution to the static MSD, we use the second term in Eq. (A27) to give the mean square amplitude,
(A51)
The integral in Eq. (A51) may diverge with the lower limit ω min 0 ( N ), where ω min ω o N 1 / d s, giving rise to instability that resembles the Landau–Peierls instability. This divergence occurs if 1 d s / 2 ξ / 2 > 0, in which case
(A52)
Hence, for force monopoles ( ξ = 0) and with d s < 2, this yields an identical divergence with increasing system size similar to the thermal Landau–Peierls instability,
(A53)
i.e., r 2 ac N 2 d s 1. The active noise contribution will, thus, prevail, provided that the force magnitude and the correlation time are sufficiently large. For an active system with force monopoles, we may, thus, assign an effective temperature,
(A54)
that can faithfully describe both the anomalous subdiffusion regime and the static (saturation) MSD, 2 r 2 .

Turning to force dipoles since, as already mentioned in typical fractals, we mostly have 1 d s / 2 d s / d l < 0, Eq. (A52) does not apply, and r 2 ac—Eq. (A51)—diverges with the upper integration limit ( ω 0), becoming essentially independent of N. This implies that, surprisingly, r 2 th may dominate the total static MSD for large enough systems due to the Landau–Peierls instability.

Here, we apply the general analytical theory for arbitrary fractals, Eqs. (A31A32), and Langevin dynamics simulations, to the Sierpinski gasket (Fig. 11). We construct different generations of the Sierpinski gasket, in which the nodes (vertices) represent beads with identical masses m that are connected by identical harmonic springs with spring constant m ω 0 2 and equilibrium distance b, mimicking a bead-spring model of the fractal. We simulate the Sierpinski network dynamics under active force monopoles and dipoles, following Langevin dynamics simulations in the high damping limit. The set of Langevin equations of motion follow Eq. (5) with Hamiltonian H = 1 2 m ω 0 2 < i j > ( r i r j b r ^ i j ) 2, where r ^ i j is the unit vector of distance r i r j between beads i and j. For b = 0, this interaction potential reduces to the scalar elasticity Hamiltonian (4) studied analytically above and in the main text. The simulations, therefore, will compare this more realistic model predictions to those used in the analytical calculation.

FIG. 11.

The two-dimensional Sierpinski gasket at different generations S n, n = 1 , 2 , 3. At generation n = 1, Sierpinski gasket S 1 is an equilateral triangle with sides a. Generation n + 1, S n + 1, is constructed by placing three nth generation structures S n adjacent to each other such that vertices of three triangles S n are touching each other with three common nodes as shown in the figure. The same process is iterated to get the next higher generation of a Sierpinski gasket.

FIG. 11.

The two-dimensional Sierpinski gasket at different generations S n, n = 1 , 2 , 3. At generation n = 1, Sierpinski gasket S 1 is an equilateral triangle with sides a. Generation n + 1, S n + 1, is constructed by placing three nth generation structures S n adjacent to each other such that vertices of three triangles S n are touching each other with three common nodes as shown in the figure. The same process is iterated to get the next higher generation of a Sierpinski gasket.

Close modal
Without changing the notations (for simplicity), we transform to dimensionless variables as follows: r i r i / b, f i f i / ( m ω 0 2 b ) and ζ i ζ i / ( m ω 0 2 b ) and introduce a dimensionless time t t / τ 0 where τ 0 = γ / m ω 0 2 is the shortest relaxation time such that the reduced equation of motions are
(B1)
with a corresponding FD theorem,
(B2)
Here, d is the Euclidean embedding space dimension. In the simulations that follow, we have used k B T m ω 0 2 b 2 = 1. For an active force, we have taken various values of f 0, τ, and p = 0.5 throughout the simulations. The relative strength of the active force compared to the thermal force is expressed in dimensionless units as f 0 2 τ p ( 1 p ) d.
To solve the equation of motions numerically, we have used the three-value gear predictor–corrector algorithm.110 The integration step Δ t is 0.05 τ 0. To achieve equilibrium (before data acquisition), we first run the simulations for times much longer (two to three orders of magnitude longer) than the longest relaxation τ N. After reaching a steady state, the simulations were run sufficiently long to calculate the time-averaged MSD accurately. The time-averaged MSD of an arbitrarily chosen bead (the internal bead shown in Fig. 12) of the fractal network is calculated using the following familiar expression:
(B3)
where t is the lag time and t t o t is the time window of data acquisition, obeying t t o t t. In all figures presenting the analytical and simulation results, distances are in units of b, forces are in units of m ω o 2 b, and times are in units of τ 0 = γ / m ω o 2.
FIG. 12.

A two-dimensional Sierpinski gasket S 9 of generation n = 9. Here, the red mark shows the position of the internal and peripheral beads. For the MSD calculation of an arbitrary bead in numerical simulation, we have considered the position of the internal bead.

FIG. 12.

A two-dimensional Sierpinski gasket S 9 of generation n = 9. Here, the red mark shows the position of the internal and peripheral beads. For the MSD calculation of an arbitrary bead in numerical simulation, we have considered the position of the internal bead.

Close modal

1. Force monopoles

Force monopole directions are randomly picked from a uniform (isotropic) distribution. Thus, the total force on an infinite system ( N ), and hence the drift velocity, would vanish. For a finite network, the average total force on the fractal depends on the actual realization of the forces. If we average on many realizations, the mean total force would again vanish, with a standard deviation ϕ f 0 p N. Since the total friction is N γ, we estimate the typical magnitude of the drift velocity (whose direction is random) as v drift ϕ f 0 p γ N. The effect of this drift velocity can be seen in Fig. 2(a) at long times. To remove the effect of the drift velocity, the expression for MSD is modified as follows:
(B4)
where ( r ( t ) ) = 1 t tot t t i t tot t ( r ( t i + t ) r ( t i ) ) = v drift t is the mean displacement due to the drift velocity.

2. Force dipoles

Many active soft-matter systems can produce force dipoles. They emerge in various biological systems, such as “actomyosin,” i.e., an actin filament network in which myosin motors are dispersed.111 These internal forces must satisfy the conditions of zero net momentum transfer and zero angular momentum transfer such that the total force and torque vanish.1,66 This implies that for each force acting on a network bead i, there is an equal and opposite force acting on another connected bead j,
(B5)
(B6)
where F t o t , i = i H + f i is the total force acting on the ith bead. Thus, there is no net force on the system, and so there is no CM translational motion. Furthermore, during the simulations, each dipole orientation remains parallel to bond vector R i R j, where R i is the position vector in the CM frame of reference. Each force pair contributes zero torque to the system; therefore, the net torque on the system vanishes too, i.e.,
(B7)
(B8)
where T i = R i × F t o t , i (‘ ×” stands for the vector product) is the torque acting on the ith bead.
However, according to the equation of motion, each force is balanced by the friction force; hence,
(B9)
where v i is the velocity of the ith bead. Multiplying both sides of Eq. (B9) by R i, we can get the torque acting on the ith bead,
(B10)
Using v i = Ω i × R i where Ω i is the angular velocity, we obtain
(B11)
(B12)
where
(B13)
The condition of net zero torque on the network, therefore, implies
(B14)
which does not necessarily imply
(B15)
Hence, a “crawling” (non-coherent) rotation of the whole object with mean velocity Ω ¯ may exist under the action of active dipole forces; see Figs. 4(b)4(d). The effect of this rotational motion can be seen in the MSD as a crossover from a constant to oscillatory behavior, whose initial presentation is ballistic-like; see Fig. 13.
FIG. 13.

Active network with force dipoles: MSD of an internal bead of an active Sierpinski gasket S 4 (N=42): (a) MSD for three different realizations of the gasket, with f 0 = 0.1, p = 0.5, and τ = 1. (b) MSD for two different force amplitudes, with τ = 1 and p = 0.5. (c) MSD of three different generations of the gasket for f 0 = 0.1, p = 0.5, and τ = 1. In these figures, the crossover from constant to ballistic motion followed by oscillations is observed at long times, indicating a rotational motion with mean velocity 1 N Ω i, and the corresponding crossover time depends on this mean value. The different random realizations of force dipoles in (a), showing crossover at different times, indicate the random nature of the mean rotational velocity. Higher force amplitudes, as shown in (b), lead to an increase in the rotational frequency, whereas higher generations of gaskets shown in (c) depict a crossover at longer times, suggesting a decrease in the rotational frequency. (see also movie SM-1 to SM-3). r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

FIG. 13.

Active network with force dipoles: MSD of an internal bead of an active Sierpinski gasket S 4 (N=42): (a) MSD for three different realizations of the gasket, with f 0 = 0.1, p = 0.5, and τ = 1. (b) MSD for two different force amplitudes, with τ = 1 and p = 0.5. (c) MSD of three different generations of the gasket for f 0 = 0.1, p = 0.5, and τ = 1. In these figures, the crossover from constant to ballistic motion followed by oscillations is observed at long times, indicating a rotational motion with mean velocity 1 N Ω i, and the corresponding crossover time depends on this mean value. The different random realizations of force dipoles in (a), showing crossover at different times, indicate the random nature of the mean rotational velocity. Higher force amplitudes, as shown in (b), lead to an increase in the rotational frequency, whereas higher generations of gaskets shown in (c) depict a crossover at longer times, suggesting a decrease in the rotational frequency. (see also movie SM-1 to SM-3). r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

Close modal
FIG. 14.

Active network with force dipoles: Mean square angular displacement (MSAD), R 2 ( n ^ ( t ) n ^ ( 0 ) ) a c, of an internal bead of an active Sierpinski gasket S 4 (N=42) subjected to force dipoles for force amplitude f 0 below and above the critical force amplitude f 0 c with τ = 1 and p = 0.5. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

FIG. 14.

Active network with force dipoles: Mean square angular displacement (MSAD), R 2 ( n ^ ( t ) n ^ ( 0 ) ) a c, of an internal bead of an active Sierpinski gasket S 4 (N=42) subjected to force dipoles for force amplitude f 0 below and above the critical force amplitude f 0 c with τ = 1 and p = 0.5. r is in unit of b, f 0 in unit of m ω 0 2 b, and t and τ are in units of τ 0 = γ / ( m ω 0 2 ).

Close modal
FIG. 15.

Triangular micro-swimmer. (a) and (b) Schematic of a triangle, the smallest generation unit of a Sierpinski gasket, employed with force dipoles along each bond, with either clockwise decreasing or increasing strengths. (c) Time variation of the stochastic angular position, θ ( t ) [i.e., the angle between R ( 0 ) and R ( t )] of an arbitrary bead in the CM frame of reference with different order of forces (in the range of 0.1–0.5) along bonds for given τ = 1 and p = 0.5. Notice the continuous periodic change in θ (even though local fluctuations are present) from π to π and π to π, indicating anti-clockwise and clockwise rotations, respectively, with the direction of rotations depending on the order of asymmetric forces. Symmetric forces lead to random fluctuations of the angular position, indicating no explicit rotational motion of the triangle. This construct, consisting of a bead-spring triangle with force dipoles on each bond, is similar to the triangular micro-swimmer studied in Refs. 79 and 80 in the presence of hydrodynamic interactions. Notably, here, we find that the Rouse dynamics (i.e., local friction), together with anisotropy of force dipole strength and random-telegraph fluctuations, is enough to produce rotational motion.

FIG. 15.

Triangular micro-swimmer. (a) and (b) Schematic of a triangle, the smallest generation unit of a Sierpinski gasket, employed with force dipoles along each bond, with either clockwise decreasing or increasing strengths. (c) Time variation of the stochastic angular position, θ ( t ) [i.e., the angle between R ( 0 ) and R ( t )] of an arbitrary bead in the CM frame of reference with different order of forces (in the range of 0.1–0.5) along bonds for given τ = 1 and p = 0.5. Notice the continuous periodic change in θ (even though local fluctuations are present) from π to π and π to π, indicating anti-clockwise and clockwise rotations, respectively, with the direction of rotations depending on the order of asymmetric forces. Symmetric forces lead to random fluctuations of the angular position, indicating no explicit rotational motion of the triangle. This construct, consisting of a bead-spring triangle with force dipoles on each bond, is similar to the triangular micro-swimmer studied in Refs. 79 and 80 in the presence of hydrodynamic interactions. Notably, here, we find that the Rouse dynamics (i.e., local friction), together with anisotropy of force dipole strength and random-telegraph fluctuations, is enough to produce rotational motion.

Close modal

3. Network collapse: Dynamical persistence length in a harmonic potential

The notion of “dynamical persistence length” emerged in active systems, in analogy to the persistence length of polymers, to describe the motion under forces that exhibit temporal correlations.85 In the case of stochastic force that follows the random telegraph process, the correlation time τ can be thought of as “persistent time” for which the force remains in an “on” state with magnitude f 0 and the velocity is kept persistent. For a free particle moving under such active force, one can define a dynamical persistence length, x p, as the distance traveled in a time τ, thereby x p = f 0 τ / γ.85,86

In the case of a particle bound to a harmonic potential, we generalize the above notion by using the equation of motion,
(B16)
with x ( 0 ) = 0, whose solution is
(B17)
where τ 0 = γ / m ω 0 2. Using t = τ, we obtain the desired persistence length,
(B18)
which (obviously) coincides with the free-particle definition for τ τ 0. The critical force above which one expects a collapse of the network is obtained by equating x p = b, leading to
(B19)
This reduces to the familiar critical force f 0 c γ b / τ for τ τ 0 and converges to f 0 c m ω 0 2 b for τ τ 0.

While our focus in this paper has been on a viscous solvent, it is interesting to evaluate the asymptotic anomalous subdiffusion, in the case of force monopoles, to viscoelastic fluid. This was studied for a thermal system of a linear polymer and applied to chromatin modeled as a space-filling curve.25,55,56 For a “power-law solvent,” that is for a solvent whose complex modulus follows a power-law, G ( ω ) ( i ω ) α, it was shown that the thermal subdiffusion exponent is modified from ν t h = 1 / 2, the well-known Rouse subdiffusion exponent (that is associated in the present formalism with d s = 1) to ν t h = α / 2. Hence, if one assumes α = 0.8, one obtains ν t h = 0.4, which was experimentally observed. However, for an active linear polymer in a viscoelastic solvent, it was shown55 that ν a c = 3 α / 2 1 (omitting the contribution of the CM motion). Thus, for α = 0.8, one obtains ν a c = 0.2. Hence, a viscoelastic power-law solvent cannot explain the experimentally observed equality between the passive and active exponents, ν a c = ν t h = 0.4.

FIG. 16.

A snapshot of the video (fig15.mp4) showing the motion of Sierpinski gaskets for different force amplitudes f 0 of force dipoles. Multimedia available online.

FIG. 16.

A snapshot of the video (fig15.mp4) showing the motion of Sierpinski gaskets for different force amplitudes f 0 of force dipoles. Multimedia available online.

Close modal
FIG. 17.

A snapshot of the video (fig16.mp4) showing the motion of Sierpinski gaskets under different fraction ϕ of force dipoles. Multimedia available online.

FIG. 17.

A snapshot of the video (fig16.mp4) showing the motion of Sierpinski gaskets under different fraction ϕ of force dipoles. Multimedia available online.

Close modal
FIG. 18.

A snapshot of the video (fig17.mp4) showing the motion of different generations of gaskets under force dipoles. Multimedia available online.

FIG. 18.

A snapshot of the video (fig17.mp4) showing the motion of different generations of gaskets under force dipoles. Multimedia available online.

Close modal

Here, we generalize these results for an arbitrary fractal. We do not provide here a complete formal analytical derivation, as this is very lengthy and out of the scope of the present publication. Rather, we suggest a compelling argument for the above-stated result, which stems from similar works.25,54–56,112,113

Consider, first, the thermal MSD. For a viscous solvent, the bead friction coefficient γ is proportional to the solvent viscosity η, and from Eqs. (A35)(A36) (with ν th = 1 d s 2), we have
(C1)
Consider the MSD evaluated in the Fourier–Laplace plane (denoted by t i ω) for a viscoelastic fluid. Replacing η by the complex frequency-dependent viscosity η ( ω ), we have
(C2)
Using the identity G ( ω ) = i ω η ( ω ), we have
(C3)
and assuming G ( ω ) ( i ω ) α, we obtain
(C4)
Transforming back to the time domain, we find
(C5)
such that
(C6)
Next, consider the active network bead MSD in a viscoelastic solvent. Since for the viscous solvent case
(C7)
it follows that for the case of a viscoelastic solvent,
(C8)
leading to
(C9)
and with G ( ω ) ( i ω ) α, we obtain
(C10)
Transforming back to the time domain, we find
(C11)
i.e.,
(C12)
These exponents agree with the above-cited linear polymer case, d s = 1.

Thus, ν a c ν t h unless α = 1; i.e., equality is recovered only for the case of a viscous fluid studied in this paper. Hence, whatever the value of d s, a power-law embedding viscoelastic fluid cannot explain equality between the passive and active subdiffusion exponent.

1.
S.
Ramaswamy
, “
The mechanics and statistics of active matter
,”
Annu. Rev. Condens. Matter Phys.
1
,
323
345
(
2010
).
2.
M. C.
Marchetti
,
J. F.
Joanny
,
S.
Ramaswamy
,
T. B.
Liverpool
,
J.
Prost
,
M.
Rao
, and
R.
Aditi Simha
, “
Hydrodynamics of soft active matter
,”
Rev. Mod. Phys.
85
,
1143
1189
(
2013
).
3.
M. E.
Cates
and
J.
Tailleur
, “
Motility-induced phase separation
,”
Annu. Rev. Condens. Matter Phys.
6
,
219
244
(
2015
).
4.
G.
De Magistris
and
D.
Marenduzzo
, “
An introduction to the physics of active matter
,”
Phys. A
418
,
65
77
(
2015
).
5.
É.
Fodor
,
C.
Nardini
,
M. E.
Cates
,
J.
Tailleur
,
P.
Visco
, and
F.
Van Wijland
, “
How far from equilibrium is active matter
?”
Phys. Rev. Lett.
117
,
038103
(
2016
).
6.
É.
Fodor
and
M. C.
Marchetti
, “
The statistical physics of active matter: From self-catalytic colloids to living cells
,”
Phys. A
504
,
106
120
(
2018
).
7.
G.
Gonnella
,
D.
Marenduzzo
,
A.
Suma
, and
A.
Tiribocchi
, “
Motility-induced phase separation and coarsening in active matter
,”
C. R. Phys.
16
,
316
331
(
2015
).
8.
P.
Digregorio
,
D.
Levis
,
A.
Suma
,
L. F.
Cugliandolo
,
G.
Gonnella
, and
I.
Pagonabarraga
, “
Full phase diagram of active Brownian disks: From melting to motility-induced phase separation
,”
Phys. Rev. Lett.
121
,
098003
(
2018
).
9.
P.
Nie
,
J.
Chattoraj
,
A.
Piscitelli
,
P.
Doyle
,
R.
Ni
, and
M. P.
Ciamarra
, “
Stability phase diagram of active Brownian particles
,”
Phys. Rev. Res.
2
,
023010
(
2020
).
10.
C.
Maggi
,
M.
Paoluzzi
,
A.
Crisanti
,
E.
Zaccarelli
, and
N.
Gnan
, “
Universality class of the motility-induced critical point in large scale off-lattice simulations of active particles
,”
Soft Matter
17
,
3807
3812
(
2021
).
11.
M.
Paoluzzi
,
D.
Levis
, and
I.
Pagonabarraga
, “
From motility-induced phase-separation to glassiness in dense active matter
,”
Commun. Phys.
5
,
111
(
2022
).
12.
C.
Wilhelm
, “
Out-of-equilibrium microrheology inside living cells
,”
Phys. Rev. Lett.
101
,
028101
(
2008
).
13.
D.
Robert
,
T.-H.
Nguyen
,
F.
Gallet
, and
C.
Wilhelm
, “
In vivo determination of fluctuating forces during endosome trafficking using a combination of active and passive microrheology
,”
PloS One
5
,
e10046
(
2010
).
14.
T.
Toyota
,
D. A.
Head
,
C. F.
Schmidt
, and
D.
Mizuno
, “
Non-Gaussian athermal fluctuations in active gels
,”
Soft Matter
7
,
3234
3239
(
2011
).
15.
B.
Stuhrmann
,
M.
Soares e Silva
,
M.
Depken
,
F. C.
MacKintosh
, and
G. H.
Koenderink
, “
Nonequilibrium fluctuations of a remodeling in vitro cytoskeleton
,”
Phys. Rev. E
86
,
020901
(
2012
).
16.
M.
Guo
,
A. J.
Ehrlicher
,
M. H.
Jensen
,
M.
Renz
,
J. R.
Moore
,
R. D.
Goldman
,
J.
Lippincott-Schwartz
,
F. C.
Mackintosh
, and
D. A.
Weitz
, “
Probing the stochastic, motor-driven properties of the cytoplasm using force spectrum microscopy
,”
Cell
158
,
822
832
(
2014
).
17.
É.
Fodor
,
M.
Guo
,
N.
Gov
,
P.
Visco
,
D.
Weitz
, and
F.
Van Wijland
, “
Activity-driven fluctuations in living cells
,”
Europhys. Lett.
110
,
48005
(
2015
).
18.
G.
Livne
,
S.
Gat
,
S.
Armon
, and
A.
Bernheim-Groswasser
, “
Self-assembled active actomyosin gels spontaneously curve and wrinkle similar to biological cells and tissues
,”
Proc. Natl. Acad. Sci. U.S.A.
121
,
e2309125121
(
2024
).
19.
R.
Granek
and
S.
Pierrat
, “
Enhanced transverse diffusion in active biomembranes
,”
Phys. Rev. Lett.
83
,
872
875
(
1999
).
20.
A.
Caspi
,
R.
Granek
, and
M.
Elbaum
, “
Enhanced diffusion in active intracellular transport
,”
Phys. Rev. Lett.
85
,
5655
(
2000
).
21.
E.
Barkai
,
Y.
Garini
, and
R.
Metzler
, “
Strange kinetics of single molecules in living cells
,”
Phys. Today
65
(
8
),
29
35
(
2012
).
22.
A.
Ghosh
and
A. J.
Spakowitz
, “
Active and thermal fluctuations in multi-scale polymer structure and dynamics
,”
Soft Matter
18
,
6629
6637
(
2022
).
23.
I.
Bronstein
,
Y.
Israel
,
E.
Kepten
,
S.
Mai
,
Y.
Shav-Tal
,
E.
Barkai
, and
Y.
Garini
, “
Transient anomalous diffusion of telomeres in the nucleus of mammalian cells
,”
Phys. Rev. Lett.
103
,
018102
(
2009
).
24.
I.
Bronshtein
,
E.
Kepten
,
I.
Kanter
,
S.
Berezin
,
M.
Lindner
,
A. B.
Redwood
,
S.
Mai
,
S.
Gonzalo
,
R.
Foisner
,
Y.
Shav-Tal
et al., “
Loss of lamin A function increases chromatin dynamics in the nuclear interior
,”
Nat. Commun.
6
,
8044
(
2015
).
25.
S. C.
Weber
,
A. J.
Spakowitz
, and
J. A.
Theriot
, “
Bacterial chromosomal loci move subdiffusively through a viscoelastic cytoplasm
,”
Phys. Rev. Lett.
104
,
238102
(
2010
).
26.
S. C.
Weber
,
A. J.
Spakowitz
, and
J. A.
Theriot
, “
Nonthermal ATP-dependent fluctuations contribute to the in vivo motion of chromosomal loci
,”
Proc. Natl. Acad. Sci. U.S.A.
109
,
7338
7343
(
2012
).
27.
F. C.
MacKintosh
, “
Active diffusion: The erratic dance of chromosomal loci
,”
Proc. Natl. Acad. Sci. U.S.A.
109
,
7138
7139
(
2012
).
28.
D.
Osmanović
and
Y.
Rabin
, “
Dynamics of active rouse chains
,”
Soft Matter
13
,
963
968
(
2017
).
29.
R. G.
Winkler
and
G.
Gompper
, “
The physics of active polymers and filaments
,”
J. Chem. Phys.
153
,
040901
(
2020
).
30.
D.
Stauffer
and
A.
Aharony
,
Introduction to Percolation Theory
, 2nd ed. (
Taylor & Francis
,
London
,
1992
).
31.
A.
Bunde
,
S.
Havlin
, and
M.
Porto
, “
Are branched polymers in the universality class of percolation
?”
Phys. Rev. Lett.
74
,
2714
(
1995
).
32.
D.
Ben-Avraham
and
S.
Havlin
,
Diffusion and Reactions in Fractals and Disordered Systems
(
Cambridge University Press
,
2000
).
33.
S.
Alexander
and
R.
Orbach
, “
Density of states on fractals: «fractons»
,”
J. Phys. Lett.
43
,
625
631
(
1982
).
34.
S.
Alexander
, “
Vibrations of fractals and scattering of light from aerogels
,”
Phys. Rev. B
40
,
7953
7965
(
1989
).
35.
A.
Bunde
,
H. E.
Roman
,
S.
Russ
,
A.
Aharony
, and
A. B.
Harris
, “
Vibrational excitations in percolation: Localization and multifractality
,”
Phys. Rev. Lett.
69
,
3189
3192
(
1992
).
36.
J. W.
Kantelhardt
and
A.
Bunde
, “
Electrons and fractons on percolation structures at criticality: Sublocalization and superlocalization
,”
Phys. Rev. E
56
,
6693
(
1997
).
37.
A. J.
Einstein
,
H.-S.
Wu
,
M.
Sanchez
, and
J.
Gil
, “
Fractal characterization of chromatin appearance for diagnosis in breast cytology
,”
J. Pathol.
185
,
366
381
(
1998
).
38.
D.
Lebedev
,
M.
Filatov
,
A.
Kuklin
,
A. K.
Islamov
,
E.
Kentzinger
,
R.
Pantina
,
B.
Toperverg
, and
V.
Isaev-Ivanov
, “
Fractal nature of chromatin organization in interphase chicken erythrocyte nuclei: DNA structure exhibits biphasic fractal properties
,”
FEBS Lett.
579
,
1465
1468
(
2005
).
39.
D.
Lebedev
,
M.
Filatov
,
A.
Kuklin
,
A. K.
Islamov
,
J.
Stellbrink
,
R.
Pantina
,
Y. Y.
Denisov
,
B.
Toperverg
, and
V.
Isaev-Ivanov
, “
Structural hierarchy of chromatin in chicken erythrocyte nuclei based on small-angle neutron scattering: Fractal nature of the large-scale chromatin organization
,”
Crystallogr. Rep.
53
,
110
115
(
2008
).
40.
E.
Lieberman-Aiden
,
N. L.
Van Berkum
,
L.
Williams
,
M.
Imakaev
,
T.
Ragoczy
,
A.
Telling
,
I.
Amit
,
B. R.
Lajoie
,
P. J.
Sabo
,
M. O.
Dorschner
et al., “
Comprehensive mapping of long-range interactions reveals folding principles of the human genome
,”
Science
326
,
289
293
(
2009
).
41.
A.
Bancaud
,
S.
Huet
,
N.
Daigle
,
J.
Mozziconacci
,
J.
Beaudouin
, and
J.
Ellenberg
, “
Molecular crowding affects diffusion and binding of nuclear proteins in heterochromatin and reveals the fractal organization of chromatin
,”
EMBO J.
28
,
3785
3798
(
2009
).
42.
A.
Wax
,
C.
Yang
,
V.
Backman
,
K.
Badizadegan
,
C. W.
Boone
,
R. R.
Dasari
, and
M. S.
Feld
, “
Cellular organization and substructure measured using angle-resolved low-coherence interferometry
,”
Biophys. J.
82
,
2256
2264
(
2002
).
43.
K. J.
Chalut
,
J. H.
Ostrander
,
M. G.
Giacomelli
, and
A.
Wax
, “
Light scattering measurements of subcellular structure provide noninvasive early detection of chemotherapy-induced apoptosis
,”
Cancer Res.
69
,
1199
1204
(
2009
).
44.
A.
Bancaud
,
C.
Lavelle
,
S.
Huet
, and
J.
Ellenberg
, “
A fractal model for nuclear organization: Current evidence and biological implications
,”
Nucleic Acids Res.
40
,
8783
8792
(
2012
).
45.
J.
Yi
,
Y.
Stypula-Cyrus
,
C. S.
Blaha
,
H. K.
Roy
, and
V.
Backman
, “
Fractal characterization of chromatin decompaction in live cells
,”
Biophys. J.
109
,
2218
2226
(
2015
).
46.
E.
Iashina
and
S.
Grigoriev
, “
Large-scale structure of chromatin: A fractal globule or a logarithmic fractal
?”
J. Exp. Theor. Phys.
129
,
455
458
(
2019
).
47.
A. Y.
Grosberg
,
S. K.
Nechaev
, and
E. I.
Shakhnovich
, “
The role of topological constraints in the kinetics of collapse of macromolecules
,”
J. Phys.
49
,
2095
2100
(
1988
).
48.
A.
Grosberg
,
Y.
Rabin
,
S.
Havlin
, and
A.
Neer
, “
Crumpled globule model of the three-dimensional structure of DNA
,”
Europhys. Lett.
23
,
373
(
1993
).
49.
K. E.
Polovnikov
,
H. B.
Brandão
,
S.
Belan
,
B.
Slavov
,
M.
Imakaev
, and
L. A.
Mirny
, “
Crumpled polymer with loops recapitulates key features of chromosome organization
,”
Phys. Rev. X
13
,
041029
(
2023
).
50.
B.
Alberts
et al.,
Molecular Biology of the Cell
(4th ed.,
Garland Science
,
New York
,
2002
).
51.
F. C.
MacKintosh
and
A. J.
Levine
, “
Nonequilibrium mechanics and dynamics of motor-activated gels
,”
Phys. Rev. Lett.
100
,
018104
(
2008
).
52.
C.
Uhler
and
G.
Shivashankar
, “
Regulation of genome organization and gene expression by nuclear mechanotransduction
,”
Nat. Rev. Mol. Cell Biol.
18
,
717
727
(
2017
).
53.
R.
Bruinsma
,
A. Y.
Grosberg
,
Y.
Rabin
, and
A.
Zidovska
, “
Chromatin hydrodynamics
,”
Biophys. J.
106
,
1871
1881
(
2014
).
54.
H.
Vandebroek
and
C.
Vanderzande
, “
Transient behaviour of a polymer dragged through a viscoelastic medium
,”
J. Chem. Phys.
141
,
114910
(
2014
).
55.
H.
Vandebroek
and
C.
Vanderzande
, “
Dynamics of a polymer in an active and viscoelastic bath
,”
Phys. Rev. E
92
,
060601
(
2015
).
56.
K.
Polovnikov
,
M.
Gherardi
,
M.
Cosentino-Lagomarsino
, and
M.
Tamm
, “
Fractal folding and medium viscoelasticity contribute jointly to chromosome dynamics
,”
Phys. Rev. Lett.
120
,
088101
(
2018
).
57.
M. V.
Tamm
,
L. I.
Nazarov
,
A. A.
Gavrilov
, and
A. V.
Chertovich
, “
Anomalous diffusion in fractal globules
,”
Phys. Rev. Lett.
114
,
178102
(
2015
).
58.
R.
Granek
and
J.
Klafter
, “
Fractons in proteins: Can they lead to anomalously decaying time autocorrelations
?”
Phys. Rev. Lett.
95
,
098106
(
2005
).
59.
R.
Granek
, “
Proteins as fractals: Role of the hydrodynamic interaction
,”
Phys. Rev. E
83
,
020902
(
2011
).
60.
S.
Reuveni
,
J.
Klafter
, and
R.
Granek
, “
Dynamic structure factor of vibrating fractals: Proteins as a case study
,”
Phys. Rev. E
85
,
011906
(
2012
).
61.
S.
Reuveni
,
J.
Klafter
, and
R.
Granek
, “
Dynamic structure factor of vibrating fractals
,”
Phys. Rev. Lett.
108
,
068101
(
2012
).
62.
M. E.
Cates
, “
Brownian dynamics of self-similar macromolecules
,”
J. Phys. France
46
,
1059
1077
(
1985
).
63.
T.
Vilgis
, “
Flory theory of polymeric fractals-intersection, saturation and condensation
,”
Phys. A
153
,
341
354
(
1988
).
64.
A.
Caspi
,
R.
Granek
, and
M.
Elbaum
, “
Diffusion and directed motion in cellular transport
,”
Phys. Rev. E
66
,
011916
(
2002
).
65.
F. S.
Gnesotto
,
B. M.
Remlein
, and
C. P.
Broedersz
, “
Nonequilibrium dynamics of isostatic spring networks
,”
Phys. Rev. E
100
,
013002
(
2019
).
66.
R.
Aditi Simha
and
S.
Ramaswamy
, “
Hydrodynamic fluctuations and instabilities in ordered suspensions of self-propelled particles
,”
Phys. Rev. Lett.
89
,
058101
(
2002
).
67.
Y.
Hatwalne
,
S.
Ramaswamy
,
M.
Rao
, and
R. A.
Simha
,
Phys. Rev. Lett.
92
,
118101
(
2004
).
68.
C.
Gardiner
, Handbook of Stochastic Processes, and Its Applications to Physics, Chemistry and the Natural Sciences, Springer Series in Synergetics (Springer, New York, 1991).
69.
T.
Nakayama
,
K.
Yakubo
, and
R. L.
Orbach
, “
Dynamical properties of fractal networks: Scaling, numerical simulations, and physical realizations
,”
Rev. Mod. Phys.
66
,
381
(
1994
).
70.
R.
Kubo
, “
The fluctuation-dissipation theorem
,”
Rep. Prog. Phys.
29
,
255
(
1966
).
71.
R.
Burioni
,
D.
Cassi
,
M. P.
Fontana
, and
A.
Vulpiani
, “
Vibrational thermodynamic instability of recursive networks
,”
Europhys. Lett.
58
,
806
(
2002
).
72.
R.
Burioni
,
D.
Cassi
,
F.
Cecconi
, and
A.
Vulpiani
, “
Topological thermal instability and length of proteins
,”
Proteins
55
,
529
535
(
2004
).
73.
N. R.
Smith
and
O.
Farago
, “
Nonequilibrium steady state for harmonically confined active particles
,”
Phys. Rev. E
106
,
054118
(
2022
).
74.
D.
Loi
,
S.
Mossa
, and
L. F.
Cugliandolo
, “
Effective temperature of active matter
,”
Phys. Rev. E
77
,
051111
(
2008
).
75.
D.
Loi
,
S.
Mossa
, and
L. F.
Cugliandolo
, “
Effective temperature of active complex matter
,”
Soft Matter
7
,
3726
3729
(
2011
).
76.
L. F.
Cugliandolo
, “
The effective temperature
,”
J. Phys. A: Math. Theor.
44
,
483001
(
2011
).
77.
A.
Ghosh
and
N.
Gov
, “
Dynamics of active semiflexible polymers
,”
Biophys. J.
107
,
1065
1073
(
2014
).
78.
R.
Rammal
and
G.
Toulouse
, “
Random walks on fractal structures and percolation clusters
,”
J. Phys. Lett.
44
,
13
22
(
1983
).
79.
M. S.
Rizvi
,
A.
Farutin
, and
C.
Misbah
, “
Three-bead steering microswimmers
,”
Phys. Rev. E
97
,
023102
(
2018
).
80.
M. S.
Rizvi
,
P.
Peyla
,
A.
Farutin
, and
C.
Misbah
, “
Deformable microswimmer in an external force field
,”
Phys. Rev. Fluids
5
,
033101
(
2020
).
81.
A.
Shapere
and
F.
Wilczek
, “
Self-propulsion at low Reynolds number
,”
Phys. Rev. Lett.
58
,
2051
(
1987
).
82.
A.
Shapere
and
F.
Wilczek
, “
Gauge kinematics of deformable bodies
,”
Am. J. Phys.
57
,
514
518
(
1989
).
83.
A.
Shapere
and
F.
Wilczek
, “
Geometry of self-propulsion at low Reynolds number
,”
J. Fluid Mech.
198
,
557
585
(
1989
).
84.
E.
Putterman
and
O.
Raz
, “
The square cat
,”
Am. J. Phys.
76
,
1040
1044
(
2008
).
85.
C.
Bechinger
,
R.
Di Leonardo
,
H.
Löwen
,
C.
Reichhardt
,
G.
Volpe
, and
G.
Volpe
, “
Active particles in complex and crowded environments
,”
Rev. Mod. Phys.
88
,
045006
(
2016
).
86.
X.
Cao
,
B.
Zhang
, and
N.
Zhao
, “
Crowding-activity coupling effect on conformational change of a semi-flexible polymer
,”
Polymers
11
,
6
(
2019
).
87.
G.
Wedemann
and
J.
Langowski
, “
Computer simulation of the 30-nanometer chromatin fiber
,”
Biophys. J.
82
,
2847
2859
(
2002
).
88.
K.
Bystricky
,
P.
Heun
,
L.
Gehlen
,
J.
Langowski
, and
S. M.
Gasser
, “
Long-range compaction and flexibility of interphase chromatin in budding yeast analyzed by high-resolution imaging techniques
,”
Proc. Natl. Acad. Sci. U.S.A.
101
,
16495
16500
(
2004
).
89.
Y.
Cui
and
C.
Bustamante
, “
Pulling a single chromatin fiber reveals the forces that maintain its higher-order structure
,”
Proc. Natl. Acad. Sci. U.S.A.
97
,
127
132
(
2000
).
90.
C. P.
Johnstone
,
N. B.
Wang
,
S. A.
Sevier
, and
K. E.
Galloway
, “
Understanding and engineering chromatin as a dynamical system across length and timescales
,”
Cell Syst.
11
,
424
448
(
2020
).
91.
B.
Fierz
and
M. G.
Poirier
, “
Biophysics of chromatin dynamics
,”
Annu. Rev. Biophys.
48
,
321
345
(
2019
).
92.
C. Y.
Zhou
,
S. L.
Johnson
,
N. I.
Gamarra
, and
G. J.
Narlikar
, “
Mechanisms of atp-dependent chromatin remodeling motors
,”
Annu. Rev. Biophys.
45
,
153
181
(
2016
).
93.
F.
MacKintosh
,
J.
Käs
, and
P.
Janmey
, “
Elasticity of semiflexible biopolymer networks
,”
Phys. Rev. Lett.
75
,
4425
(
1995
).
94.
C. P.
Broedersz
and
F. C.
MacKintosh
, “
Modeling semiflexible polymer networks
,”
Rev. Mod. Phys.
86
,
995
1036
(
2014
).
95.
D. B.
Brückner
,
H.
Chen
,
L.
Barinov
,
B.
Zoller
, and
T.
Gregor
, “
Stochastic motion and transcriptional dynamics of pairs of distal DNA loci on a compacted chromosome
,”
Science
380
,
1357
1362
(
2023
).
96.
P.
Mach
,
P. I.
Kos
,
Y.
Zhan
,
J.
Cramard
,
S.
Gaudin
,
J.
Tünnermann
,
E.
Marchi
,
J.
Eglinger
,
J.
Zuin
,
M.
Kryzhanovska
et al., “
Cohesin and CTCF control the dynamics of chromosome folding
,”
Nat. Genet.
54
,
1907
1918
(
2022
).
97.
M.
Gabriele
,
H. B.
Brandão
,
S.
Grosse-Holz
,
A.
Jha
,
G. M.
Dailey
,
C.
Cattoglio
,
T.-H. S.
Hsieh
,
L.
Mirny
,
C.
Zechner
, and
A. S.
Hansen
, “
Dynamics of CTCF-and cohesin-mediated chromatin looping revealed by live-cell imaging
,”
Science
376
,
496
501
(
2022
).
98.
S.
Brahmachari
,
T.
Markovich
,
F. C.
MacKintosh
, and
J. N.
Onuchic
, “Temporally correlated active forces drive chromosome structure and dynamics,”
PRX Life
2
, 033003 (
2024
).
99.
S.
Put
,
T.
Sakaue
, and
C.
Vanderzande
, “
Active dynamics and spatially coherent motion in chromosomes subject to enzymatic force dipoles
,”
Phys. Rev. E
99
,
032421
(
2019
).
100.
See Fig. 2(f) in Ref. 98, light-blue curve, and long times.
101.
See Fig. 2(f) in Ref. 98, other colors.
102.
See Fig. 3, left panel, in Ref. 99.
103.
S.
Brahmachari
and
J. F.
Marko
, “
Chromosome disentanglement driven via optimal compaction of loop-extruded brush structures
,”
Proc. Natl. Acad. Sci. U.S.A.
116
,
24956
24965
(
2019
).
104.
S.
Subramanian
and
S. M.
Murray
, “
Subdiffusive movement of chromosomal loci in bacteria explained by DNA bridging
,”
Phys. Rev. Res.
5
,
023034
(
2023
).
105.
T.
Nemoto
,
E.
Fodor
,
M. E.
Cates
,
R. L.
Jack
, and
J.
Tailleur
, “
Optimizing active work: Dynamical phase transitions, collective motion, and jamming
,”
Phys. Rev. E
99
,
022605
(
2019
).
106.
D.
Saintillan
,
M. J.
Shelley
, and
A.
Zidovska
, “
Extensile motor activity drives coherent motions in a model of interphase chromatin
,”
Proc. Natl. Acad. Sci. U.S.A.
115
,
11442
11447
(
2018
).
107.
F. G.
Woodhouse
and
R. E.
Goldstein
, “
Spontaneous circulation of confined active suspensions
,”
Phys. Rev. Lett.
109
,
168105
(
2012
).
108.
K.
Suzuki
,
M.
Miyazaki
,
J.
Takagi
,
T.
Itabashi
, and
S.
Ishiwata
, “
Spatial confinement of active microtubule networks induces large-scale rotational cytoplasmic flow
,”
Proc. Natl. Acad. Sci. U.S.A.
114
,
2922
2927
(
2017
).
109.
I. S.
Gradshteyn
, and
I. M.
Ryzhik
,
Table of Integrals, Series, and Products
(
Academic Press
,
2014
).
110.
M.
Allen
and
D.
Tildesley
,
Computer Simulation of Liquids
(
Oxford University Press
,
1987
).
111.
K.
Drescher
,
J.
Dunkel
,
L. H.
Cisneros
,
S.
Ganguly
, and
R. E.
Goldstein
, “
Fluid dynamics and noise in bacterial cell–cell and cell–surface scattering
,”
Proc. Natl. Acad. Sci. U.S.A.
108
,
10940
10945
(
2011
).
112.
R.
Granek
, “
Membrane surrounded by viscoelastic continuous media: Anomalous diffusion and linear response to force
,”
Soft Matter
7
,
5281
5289
(
2011
).
113.
K.
Nishi
,
F. C.
MacKintosh
, and
C. F.
Schmidt
, “
Multiscale microrheology using fluctuating filaments as stealth probes
,”
Phys. Rev. Lett.
127
,
158001
(
2021
).