Disentangling Entanglements in Biopolymer Solutions

Philipp Lang and Erwin Frey∗ 1Arnold Sommerfeld Center for Theoretical Physics and Center for NanoScience, Department of Physics, Ludwig-Maximilians-Universität München, Theresienstrasse 37, 80333 München, Germany

Reptation theory has been highly successful in explaining the unusual material properties of entangled solutions. It reduces the complex many-body dynamics to a single-polymer description where each polymer is envisaged to be confined to a tube through which it moves in a snake-like fashion. For flexible , reptation theory has been amply confirmed by both experiments and simulations. In contrast, for semiflexible polymers experimental and numerical tests are either limited to the onset of reptation, or were performed for tracer polymers in a fixed, static matrix. Here we report Brownian dynamics simulations of entangled solutions of semiflexible polymers, which show that curvilinear motion along a tube (reptation) is no longer the dominant mode of dynamics. Instead, we find that polymers disentangle due to correlated constraint release which leads to equilibration of internal bending modes before polymers diffuse the full tube length. The physical mechanism underlying terminal stress relaxation is rotational diffusion mediated by disentanglement rather than curvilinear motion along a tube.

Dense solutions of polymers are viscoelastic: While agarose networks seem to support Odijk’s scaling result[3] 2 they respond like a fluid to low-frequency stresses, they τr ∼ `pL . However, these experimental results do not act like a cross-linked elastic network at high frequen- settle the actual controversy, as the polymer diffuses in a cies. These intriguing material properties are attributed fixed, static matrix and not in an entangled polymer solu- to the extended structure of polymers, which makes topo- tion. Experiments on entangled solutions are sparse and logical interactions particularly important: Polymers can mostly investigate the dynamics in the plateau regime, effortlessly slide past each other but are not allowed to where polymers experience tube confinement but do not cross each other’s path. These ‘entanglements’ mutu- yet show curvilinear motion along the tube[16–18]. Un- ally restrict the accessible configuration space, turning fortunately, these studies provide no explicit informa- the dynamics of polymer solutions into a difficult many- tion on the dependence of the terminal relaxation time body problem. Efforts to incorporate these features into on polymer length and persistence length[19]. Further- a single-polymer mean-field model led to the famous tube more, whether or not reptation, i.e. curvilinear motion model[1–3]. In this model, the dynamic topological con- along some ‘primitive path’, is the actual mechanism straints on the motion of a given polymer are represented of stress relaxation has never been tested experimen- as a static, confining tube[1]. By this means, the single- tally or by means of computer simulations. While there polymer dynamics in an entangled solution is reduced are experiments reporting the direct observation of fila- to the curvilinear of its center-of-mass ment dynamics within a tube and sliding motion along along the long axis of the tube[2], termed ‘reptation’. For the tube[20, 21], following the polymer dynamics over flexible polymers, reptation theory is well established and longer time scales poses a formidable experimental chal- also fairly predictive[4–7], though there are still many in- lenge and has not yet been realized. Similarly, previous teresting open questions[8]. Brownian dynamics simulations span the regime up to Reptation theory has also been employed to elucidate intermediate time scales where the tube forms, but do the viscoelastic properties of entangled solutions of semi- not extend deeply enough into the terminal relaxation flexible polymers, which play an important role in deter- regime[13]. mining the material properties of biopolymer solutions[9]. While scaling theories for the tube width[10, 11] have Here we present a large-scale Brownian dynamics sim- been convincingly verified, both experimentally[12] and ulation study and a complementary scaling approach for numerically[13], the predictions of reptation theory for the dynamics of entangled solutions of semiflexible poly- the long-time dynamics and the ensuing (terminal) stress mers. We consider a monodisperse solution of worm-like arXiv:1802.03702v1 [cond-mat.soft] 11 Feb 2018 relaxation remain controversial for several reasons. First, chains with length L and persistence length `p that in- different scaling theories lead to conflicting results for the teract by volume exclusion only. The polymer density is given by the number ν of polymers per unit volume, dependence of the terminal relaxation time τr on poly- q mer length , persistence length , and mesh size : 3 L `p ξ with the mesh size defined as ξ := νL ; for an illustra- 7 4 2 Results range from[3] τr ∼ L /ξ through[10] τr ∼ `pL tion refer to Fig.1. We are mainly interested in the to[14] ∼ 2/3 3. Recent experimental studies[15] τr (`p/ξ) L dynamics of semiflexible polymers (`p & L) at densities of the Brownian motion of carbon nanotubes in porous where the polymer length is much larger than the mesh size, L  ξ, but lies below the threshold of the isotropic- to-nematic transition. For example, in a typical solution of actin filaments of length L ≈ 10 µm at a number den- ∗ [email protected] 3 sity ν ≈ 1 µm− the average mesh size is ξ ≈ 0.5 µm, and 2

a) b) 10 2 t1/2

2 3 10 3/4 /L t 2 R 4 10 p/L =0.44 p/L =1

5 p/L =5 10

5 4 3 2 1 0 1 10 10 10 10 10 10 10 t/ ℓp d FIG. 2. Dynamics of theFIG. mean-square 2. end-to-end dis- tance. a, δR2(t) of a single filament in an entangled poly- 휉 mer solution, with time rescaled in units of the diffusion time 3 τd = ζL /(6kB T ) where ζ is the friction coefficient and T the temperature. Simulations were performed for filaments with the indicated persistence lengths in a solution with mesh size FIG. 1. Illustration of an entangled polymer solution. ξ/L = 0.086. In each case, there are three distinct dynamic Typical configuration of a tracer filament (red) with persis- regimes: free relaxation following a t3/4 power law, an in- tence length ` /L = 0.44 in an entangled solution with mesh p termediate regime where the tube reorganises and topologi- size ξ/L = 0.1. For better visibility of the surrounding chains cal constraints are relaxed, and a final regime where bending (grey) only those in direct proximity to the tracer filament modes are fully relaxed at their respective equilibrium value are shown. The dynamics of an entangled solution on differ- δR2 (t) = L4/(45`2). b, Equilibration time of internal modes, ent time scales is illustrated in the Supplementary Movie 1. eq p τeq/τd, for a tracer filament of length L (with ξ/L = 0.086) in a solution of polymers with length Ln as a function of Ln/L. To a very good approximation we find a linear dependence of τ on L , which excludes a standard constraint-release mech- therefore much smaller than both and ≈ m[22]. eq n L `p 17 µ anism based on curvilinear motion of the surrounding chains. The polymer dynamics is governed by a Langevin equa- tion,

∂U theoretical predictions[14, 23–26]. It is followed by ζ∂ rk = − +η , (1) t i k an intermediate regime where δR2(t) increases more ∂ri slowly, indicating that the internal bending modes are k constrained by the surrounding11 polymers. We refer to where ri (t) is the position vector of bead i ∈ {1,...,N} on polymer k. The potential U accounts for the bend- the corresponding time scale marking the crossover from ing energy of each filament as well as the mutual steric free to constrained relaxation as the ‘entanglement time’ interaction between the filaments, ζ denotes the friction τe. Finally, we observe at some ‘equilibration time’ τeq coefficient, and η is Gaussian white noise with an am- that the MSD end-to-end distance begins to saturate. plitude determined by the Stokes-Einstein relation; for As the measured saturation value is identical to the 2 1 4 2 further details see Methods. Our Brownian dynamics thermal equilibrium value[22, 27] δReq = 45 L /`p , the simulations employ a standard bead-spring algorithm, as bending modes are now fully relaxed and thus no longer explained in Methods. constrained by the surrounding polymers. In other words, the internal bending modes are disentangled.

RESULTS Initial entanglement is described by Odijk- Semenov scaling. According to the scaling theory for To learn how the polymers first entangle and then semiflexible polymers by Odijk and Semenov[10, 11], disentangle, i.e. how topological constraints emerge topological constraints imposed by neighbouring poly- and are released, we studied the dynamics of individual mers can effectively be described by a tube of diameter 6/5 1/5 (tracer) polymers in an entangled polymer solution; d ∼ ξ `p− , which leads to a restriction of the bending see Supplementary Movie 1. We began our analysis fluctuations of the confined polymer for lengths larger 2 1/3 by measuring the time evolution of the mean-squared than the ‘entanglement length’ Le ∼ (d `p) (for an displacement (MSD) of the end-to-end distance (Meth- illustration see Supplementary Movie 2). This implies ods), δR2(t), thus focusing on the relaxation of the that the dynamics of the end-to-end distance will begin internal modes and disregarding any global translation to deviate from free relaxation at the corresponding 4 or rotation of the polymer. We find three distinct ‘entanglement time’ τe ∼ Le/`p. All these scaling re- regimes (Fig.2): Initially, the MSD end-to-end distance sults for the physics of initial entanglement are in full increases as a power law, δR2(t) ∼ t3/4; this agrees with accordance with our Brownian dynamics simulations previous experimental observations for freely relaxing (Supplementary Note 2): To avoid ambiguities, we semiflexible polymers[22] as well as corresponding chose to define the tube diameter d as that value of 3 the transverse fluctuations g1, (t) (see Methods) where ⊥ it starts to deviate from the t3/4-behaviour of free polymers (Supplementary Fig. 7). This determines both the tube diameter d (Supplementary Fig. 8) and the 4 entanglement time τe, and with τe ∼ Le/`p also the entanglement length Le (Supplementary Fig. 9), with all numerical results fully supporting Odijk-Semenov scaling.

Bending fluctuations equilibrate (disentan- gle) faster than the diffusion time. The caging effect of the tube on the bending fluctuations is only transient, as the polymers surrounding the tracer polymer themselves move through bending fluctuations, centre-of-mass and rotational diffusion. Hence, at some equilibration time τeq the tube will effectively begin to ‘dissolve’ (Supplementary Movie 2, and Supplementary Fig. 10). Surprisingly, we find that this happens well FIG. 3. Spatial correlations of transverse fluctua- before the polymer diffuses the full length of the tube, tion. The covariance cov(∆t, ∆x) of transverse fluctua- 3 k τeq < τd, where the ‘diffusion time’ τd = ζL /(6kBT ) tions r⊥,i of beads at position i on neighbouring chains k is defined as the time a rod takes to diffuse its own and l during a time interval ∆t, as a function of the dis- k l length L. A comprehensive analysis of our Brownian tance ∆x = ri (t + ∆t) − ri(t + ∆t) , and normalised with re- dynamics simulations (Supplementary Note 3) shows spect to cov(∆t, 0): cor(∆t, ∆x) := cov(∆t, ∆x)/cov(∆t, 0). Simulations were performed for an entangled solution with that the equilibration time τeq is independent of the persistence length (Supplementary Fig. 11) and given by ξ/L = 0.086 and `p/L = 1, and at different times t. As ex- τ = 2.7×10 4 L5/ξ2 (Supplementary Fig. 12). pected, for very small times t . τe (red curve), there is no eq − correlation. In contrast, for times in the intermediate regime, τe . t . τeq (here t = 0.02τd), there are clear spatial correla- Constraint release is due to correlated motion tions spanning a distance of about three times the mesh size of filaments and their neighbouring filaments. ξ. These correlations vanish for times larger than the equili- What then drives the release of the N ∼ L/ξ topolog- bration time, e.g. for times equal to the diffusion time, t = τd. ical constraints and thereby determines× the equilibra- tion time τeq? As one expects that each of these con- straints is released independently, the equilibration time 2 should scale as τeq ∼ N τ , where τ denotes the re- scaling laws but rather to affirm that correlated motion laxation time of a single× topological× constraint.× At first of the tracer polymer and its neighbouring polymers is sight our simulations seem to suggest that the constraint responsible for constraint release. release time equals the diffusion time, τ = τd (Supple- To further substantiate these many-body effects, we mentary Fig. 13), but this is clearly inconsistent× with measured how the spatial correlations between the poly- τeq < τd. To disentangle these puzzling results, we ex- mer’s fluctuations perpendicular to its end-to-end dis- ploited a unique strength of Brownian dynamics simula- tance evolve over time (Fig.3); for a definition of the tions, namely that they allow one to adjust the length correlation function see Eq. (11)–(13) in Methods. Our and friction coefficient of each individual polymer in the numerical data show clear spatial correlations in precisely solution (Supplementary Note 3). First, we investigated that time window in which tube reorganisation takes the dynamics of a tracer polymer of length L in an en- place. Moreover, these correlations span several mesh tangled solution of polymers with a shorter length Ln, sizes, strongly suggesting that many-body effects are re- 2 2 and found τeq ∼ N LnL (Fig.2b, and Supplementary sponsible for tube reorganisation and the cocommittant Fig. 14). This scaling× behaviour can not be explained by relaxation of internal bending modes. We have also per- a classical constraint release mechanism[5, 28] based on formed simulations for dilute solutions with ξ/L = 0.51 sliding motion of the shorter filaments only, as the latter and found no correlations in the transverse displacements cr 2 3 would predict τeq ∼ N Ln. Instead, our numerical results of neighbouring chains. suggest that the interplay× between the dynamics of the Taken together, we conclude that constraint release tracer polymer and its surrounding polymers drives con- (disentanglement) of internal modes is driven by the straint release. To test this hypothesis we varied the rela- correlated motion of a given polymer and its surrounding tive value of the friction coefficient of the tracer polymer, polymers, and leads to an equilibration of internal bend- ζ, and the surrounding polymers, ζn, (Supplementary ing modes on a time scale τeq smaller than the diffusion 2 3 α 3 β Fig. 15). We find τeq ∼ N (ζL ) (ζnLn) with α+β = 1 time τd. As a consequence, for times larger than τeq and the numerical data× consistent with α = 2/3 and the tube is dissolved insofar as the bending modes β = 1/3. These results are not meant to indicate strict are equilibrated. However, this does not suffice for 4

101 4 2 τr ∼ `pL /ξ (Fig.4b, and Supplementary Fig. 17), which a) 2 2 b) const 100 δR /L is clearly distinct from Odijk’s[10] result ∼ 2 . 2 3 τOdijk `pL δe 10− 1 R 4 10− ) Moreover, all our simulation data collapse on a univer-

2 ξ/L sal scaling curve τr = τDoi τˆ(`p/`∗) (Fig.4b), confirming 10− x p )( ξ/L = 0.224 d ∝ that there is a crossover from Doi’s rigid rod scaling 3 4 10− 10− /τ ξ/L = 0.172

r regime[29] to a qualitatively different disentanglement

4 τ ξ/L = 0.142 10− ( ξ/L = 0.100 regime by reducing the persistence length below some 5 3 2 10− ξ/L = 0.086 threshold value, `p < `∗ := L /ξ . This constitutes an in- 5 p 10− 4 2 0 3 2 1 0 1 10− 10− 10 10− 10− 10− 10 10 termediate asymptotic scaling regime extending over at 2 3 t/τd x = (`p/L)(ξ/L) least two decades (10− < x < 1) in the scaling variable 2 3 x = `p/`p∗ = `pξ /L . It is precisely this regime which FIG. 4. Terminal relaxation. a, Time evolution of the is most relevant for entangled solutions of cytoskeletal mean-square end-to-end distance δR2(t) (triangles), and the biopolymers[19] as well as carbon nanotubes[15]. mean-square changes in the orientation δe2 (t) (circles) of R We can explain the observed terminal relaxation a tracer polymer for entangled solutions with ξ/L = 0.15, building on our results for the equilibration of internal `p/L = 5 (red), and ξ/L = 0.086, `p/L = 0.44 (yellow). Note that terminal relaxation sets in long after relaxation of inter- modes by the following physical picture: After an initial nal bending modes. b, Scaling plot of the terminal relaxation phase of confinement to an Odijk-Semenov tube, the 4 ∗ time τr/τDoi with τDoi = τd(L/ξ ) as a function of `p/`p with correlated motion of each polymer and its surrounding ∗ 3 2 `p = L /ξ , as determined from a fit of the asymptotic relax- polymers leads to a reorganisation of the topological 2 ation of δeR(t) to the functional form 2−2 exp(−t/τr). The constraints, and thereby to an equilibration of internal 5 2 numerical data collapse to a master curve with two distinct bending modes, within a time τeq ∼ L /ξ . As a conse- ∗ regimes: For small values of x = `p/`p, the terminal relax- quence, there is a mean-square angular rotation of the ation time τr increases linearly with x, whereas for x 1 it 2 & polymer due to bending fluctuations[9, 27], hδθ i ∼ L/`p, becomes independent of x, indicating that one has reached and the ensuing rotational diffusion constant scales the rigid rod limit where Doi’s scaling results are valid[29]. 2 2 4 as Dr ∼ hδθ i/τeq ∼ ξ /(L `p). This is identical to the scaling of the terminal relaxation time as we find it 4 2 in our simulations, τr ∼ `pL /ξ (Fig.4b). Hence, mechanical stresses to relax in a solution of semiflexible terminal relaxation in solutions of entangled semiflexible polymers, as the latter requires that correlations in the polymers is not primarily due to diffusive motion along polymer’s orientation also must vanish. the tube’s backbone (reptation), but rather due to disentanglement of internal bending modes mediated by Terminal relaxation. To study this terminal re- the release of topological constraints and the ensuing rotational diffusion of the polymer’s orientation, induced laxation regime we measured the auto-correlation func- 2 2 2 by bending modes (hδθ i ∼ L/`p). The constraint release tion for the polymer orientation, δeR(t) := h[e(t)−e(0)] i, where e(t) = R(t)/R(t) denotes the instantaneous, nor- for the internal modes is facilitated by the correlated malised end-to-end vector of the polymer. Our sim- dynamics of each polymer and its surrounding polymers. ulations show that terminal relaxation happens much later than the relaxation of the internal modes (Fig.4a). Comparison of entangled solutions with frozen To determine the terminal relaxation time τr we per- environments. In order to better understand the ori- 2 gin of these differences and the role of many-body ef- formed a least-squares fit of δeR(t) close to saturation to 2−2 exp(−t/τr) (Supplementary Fig. 16), as expected fects in the dynamics of entangled solutions, we also for rotational diffusion[14]. For stiff chains, where the studied the dynamics of a reference system where a maximal mean-square amplitude of the bending modes tracer polymer moves in a frozen environment, similar 2 3 2 δr ∼ L /`p is less than the squared mesh size ξ , we re- as in recent experiments studying the Brownian motion ⊥ 4 cover Doi’s[29] result, τDoi ∼ τd(L/ξ) (Fig.4b). Hence, of carbon nanotubes in a fixed agarose network[15]. In in this asymptotic limit, terminal relaxation is facili- good quantitative agreement between our simulations in tated by Doi’s ‘reptation – tube rotation’ mechanism[29]: the frozen environment and the experimental results[15], after diffusing its own length L within a tube of di- the rotational relaxation time exhibits a crossover from ameter d ∼ ξ2/L in a time of the order of the diffu- Doi scaling[3] to Odijk scaling[10] with increasing poly- sion time τd, the rod becomes confined to a new tube, mer length L (Fig.5): For rather stiff polymers with 2 1/3 which is tilted relative to the previous one by an an- L/(ξ `p) . 2, the relaxation time agrees with Doi’s 7 4 gle δθ ∼ d/L. The validity of Doi’s effective reduction to prediction, τr ∼ L /ξ , while for semiflexible chains with 2 1/3 2 a single-chain problem in the stiff limit has also been L/(ξ `p) & 2, one finds Odijk’s prediction, τr ∼ `pL . shown recently in computer simulations of needle-like Moreover, our Brownian dynamic simulations in a frozen rigid rods[30, 31]. In contrast, for δr2 > ξ2, where tube environment are also consistent with previous numerical confinement of bending modes is significant,⊥ another re- results of rigid rods in a random array of fixed obstacles laxation mechanism sets in, which allows terminal relax- where τr is reported to follow Doi’s prediction[30, 32–34]. ation to take place orders of magnitude faster. We find Comparing the dynamics in a fixed environment with 5

101 experiment simulation ) 100 /L p ` )( r /τ

d 10 1

τ − (

2 10− 0 1 2 3 4 5 6 7 2 1/3 1/3 L/(`pξ ) = (`p∗/`p)

FIG. 5. Scaling of the rotational relaxation time for a tracer polymer in an array of frozen polymer chains. The numerical results for the rotational diffusion con- −1 stant Dr ∼ τr in systems of different density and flexibility, rescaled as (τd/τr)(`p/L), are plotted as a function of rescaled 2 1/3 ∗ 1/3 polymer length L/(`pξ ) = (`p/`p) . As in the experi- mental system, where carbon nanotubes diffuse in a porous agarose network (squares, data extracted from Ref.[15]), our simulations (solid circles) show that with increasing filament length L there is a crossover from Doi’s scaling result for rigid 7 4 rods, τDoi ∼ L /ξ , (solid line) to Odijk’s scaling result for 2 semiflexible polymers, τOdijk ∼ τDoi `p/L ∼ `pL . To fit our definition, which uses the saturation time τr instead of the diffusivity, the experimental data has been rescaled by the necessary factor of 0.5. Our simulations agree quantitatively well with the experimental data[15].

the dynamics in an entangled solution, we find that all FIG. 6. Comparison between dynamics in a fixed ar- key observables show qualitatively different behaviour ray and an entangled solution. Key quantities character- ising the dynamics of a tracer polymer in a fixed array (green (Fig.6): The relaxation of the internal modes and the squares) or an entangled solution (yellow triangles) for a sys- orientational correlations is retarded in a frozen environ- tem with ξ/L = 0.086 and `p/L = 3. To emulate a fixed array ment, and the equilibration time τ˜eq, as well as the ter- we used an equilibrated configuration of the entangled solu- minal relaxation time τ˜r, now agree well with the time tion and froze the configuration of all but one tracer polymer, it takes to diffuse the tube length, τ˜eq = τd, and Odijk’s which is reminiscent of the experimental situation in Ref.[15]. 2 scaling result, τOdijk ∼ `pL , respectively. While for an In a fixed array, relaxation of a tracer polymer’s orientation 2 entangled solution the correlation function for the trans- (δeR) is much slower and follows the scaling predictions of verse displacement (g1, ) of the center of mass exhibits a Odijk and Doi. Moreover, the MSD of the polymer’s fluctu- ⊥ slanted plateau, it shows a clear flat plateau in a frozen ations transverse to the end-to-end vector (g1,⊥) exhibits a environment, indicating that the dynamics of the tracer clearly visible plateau in a fixed array, but continues to relax polymer is confined to a persistent tube in the time win- in an entangled solution. At asymptotically large times, the transverse MSD and likewise the longitudinal MSD (g1,k) is dow between the entanglement time τe and the tube dif- slowed down in a fixed array. Finally, the relaxation of inter- fusion time τd. These predictions for the various key nal bending modes (δR2) shows a qualitatively different be- observables may be tested using microrheology experi- haviour: While there is an intermediate regime where δR2(t) ments. Taken together, the comparison between the dy- appears to saturate, which reflects caging in a tube, this is not namics of entangled solutions and fixed arrays reaffirm the case for an entangled solution. There internal modes con- that tube reorganisation, and the ensuing relaxation of tinuously relax, indicating ongoing tube reorganisation with topological constraints, must be based on the combined concomitant relaxation of topological constraints. effect of the dynamics of a given polymer and its sur- rounding polymers. 6

DISCUSSION In all our simulations we chose units such that kBT = 1, and the friction per unit length is ζ = 1. Also, we used The results presented here challenge our current un- the contour length of the polymers as our basic unit of derstanding of the dynamics of entangled polymer solu- length. For later reference, an actin filament with a con- tions. While both asymptotic regimes, flexible polymers tour length of L = 10 µm and a diameter of 5 nm in a and rigid rods, are well described within the framework solution with a of η = 0.1 Pa s at a temperature 4 of classical reptation theory, our current data show that of 20◦C has a disengagement time of τd ≈ 1.5 × 10 s. semiflexible polymer dynamics clearly exhibits strong The position of bead i on polymer chain k at time t k many-body correlation effects. These correlation effects is denoted by ri (t) with 1 ≤ i ≤ N and 1 ≤ k ≤ M. Each lead to a fast equilibration (’disentanglement’) of the in- bead i is connected to its neighbours by FENE springs[43] ternal bending modes significantly before a polymer had with the interaction potential time to diffuse the full length of its confining tube. As "  k 2# k X ai − a0 a consequence, terminal relaxation is facilitated by the UFENE = −B ln 1 − , (2) amax correlated release of topological constraints and the en- 1 a0 + amax the namic correlations found in the glassy behaviour of dense k colloidal systems[35]. However, it remains a challenge to FENE potential is UFENE = ∞. We chose the spring con- identify the essence of the many-body physics responsi- stant B ≈ 1000 kBT . With this choice the deviation in ble for tube renewal leading to the relaxation of bending bond length from the equilibrium value was always be- modes. low 0.05 a0 in our simulations. The bending stiffness of To test the proposed disentanglement mechanism un- the polymers is described by a standard worm-like chain derlying terminal relaxation experimentally one actually model[44, 45] with the bending energy given by does not need to measure the terminal relaxation time N 2 `pkBT X− explicitly. It suffices to measure the equilibration time of U k = 1 − tk · tk  , (3) WLC a i i+1 the internal modes and validate the predicted scaling of 0 i=1 the equilibration time. This should be feasible well within where tk = (rk − rk)/|rk − rk| is a normalised bond the time window accessible to experiments on entangled i i+1 i i+1 i vector (tangent vector). For the mutual (steric) inter- solutions of cytoskeletal filaments, carbon nanotubes, or action between the beads we used the Weeks-Chandler- short DNA filaments. Anderson (WCA) potential[46], which for bead i on chain k reads for rij ≤ σ ACKNOWLEDGMENTS  !12 !6  i,k X σ σ U = A  − 2 + 1 , (4) WCA rkl rkl This project was supported by the German Excellence l,j ij ij Initiative via the program “NanoSystems Initiative Mu- where the sum extends over all other beads in the sim- nich” (NIM). ulation box, i.e. over all polymers l and beads j except kl k l bead i of chain k, and rij = |ri − rj| denotes the dis- METHODS tance between a pair of beads (i, j) on chains k and l. i,k For distances rij > σ, the potential vanishes: UWCA = 0. We chose for the range of the interaction. This Brownian dynamics simulation. We implemented σ = 0.9 a0 choice prevents chain crossings and at the same time spu- a basic bead-spring algorithm[36, 37] to simulate the rious oscillations of neighbouring beads in a chain due Brownian dynamics of polymer chains in an entan- to their WCA interaction. In order to avoid crossing gled solution, using standard interactions as discussed of chains or the overlapping of different beads we used previously[4, 34, 38, 39], and explained in detail be- a strong potential by setting the parameter A = 20 k T low. Our simulations comprise M polymer chains with B (see Supplementary Note 1). Finally, we implemented a 164 ≤ M ≤ 1106 in a cubic simulation volume with edge cell-linked list algorithm to evaluate the occurring colli- length X = 1.35 L, and periodic boundary conditions. sions which presorts the beads according to their posi- Each polymer is represented by a linear sequence of N tions before testing for polymer collisions. This led to a beads connected by springs. For low densities we used significant decrease in the runtime of our simulations. N = 45, while for densities above ξ/L ≤ 0.1 we used a The Langevin equation for the entangled polymer so- finer discretisation with N = 60; this ensures that the lution reads polymers are thin enough such that the density of the entangled solution is sufficiently below the threshold to ∂rk(t) ∂ U i,k ({rk}) ζ i = − total i + ηk(t) , (5) the nematic phase[40–42]. k i ∂t ∂ri 7 where ζ denotes the friction coefficient of a single bead, the center monomer parallel and perpendicular to the k k k k ηi (t) is Gaussian white noise with mean zero and co- orientation of the end-to-end vector e = R /R , respec- k l variances given by hηi (t) ηj(t0)i = 6 ζ δijδkl kBT δ(t−t0). tively: The total potential acting on bead i of polymer k reads: h  i2 U i,k i,k U k U k . k k k total := UWCA + FENE + WLC g1, (t) := rN/2(t) − rN/2(0) · e (0) , (8) We used uniformly distributed random numbers gen- k  2 erated by a maximally equidistributed combined Taus- h k k i g1, (t) := rN/2(t) − rN/2(0) − g1, (t) . (9) worthe generator[47] for the noise. These kinds of ran- ⊥ k dom number generators have been shown to amount to the same behaviour in the dynamics of polymers on time Finally, a quantity which allows to measure the terminal scales significantly above one time step as Gaussian white relaxation of stresses in the solution is given by the mean- noise within the statistical errors while being significantly square changes of the direction of the end-to-end vector faster[13, 37, 38]. To calculate the time evolution, the 2  k k 2 Langevin equation, Eq. (5), is integrated via a semi- δeR(t) := h e (t)−e (0) i . (10) implicit Euler algorithm[48] with a time step of (in our 4 5 Definition of the covariance for transverse dis- units) 1 × 10− for systems with N = 45 and 2 × 10− for N = 60, respectively. placements. To quantify the many-body effects in en- We performed extensive tests to ensure the reliabil- tangled polymer solutions, we were looking for correla- ity of our Brownian dynamics simulations (Supplemen- tions in the dynamics of neighbouring polymers. Since tary Note 1). We find good agreement of both the the tube is mainly constraining the fluctuations of a poly- tangent-tangent correlations (Supplementary Fig. 1) and mer transverse to its end-to-end vector, and the tube it- the mean-square end-to-end distance (Supplementary self is due to the presence of neighbouring chains, we Fig. 2) for freely relaxing polymers with known analytical investigated correlations in these transverse fluctuations. results[14, 23–26, 44]. Moreover, we have tested that our We considered the magnitude of the displacement of bead simulation algorithm does not show spurious chain cross- i on polymer k during a time interval ∆t, perpendicular k ings (Supplementary Fig. 3). Finally, we have tested for to polymer’s end-to-end vector e (t + ∆t) finite size effects (Supplementary Fig. 4), and that our k k  k k  r ,i := P (t + ∆t) · ri (t) − ri (t + ∆t) , (11) results for the mean-square displacement of the center ⊥ ⊥ monomer (Supplementary Fig. 5) and the mean-square where changes of the end-to-end-vector are largely independent of the finite filament thickness (Supplementary Fig. 6). Pk (t) = 1 − ek(t) ⊗ ek(t) (12) Quantities of interest. To characterise the dynamics ⊥ of individual chains within the entangled polymer solu- is a projection operator onto the end-to-end vector. Sim- tion we studied the following quantities of interest where ilar to the work of Doliwa and Heuer[49] on the cage ef- the averages h...i indicate averages over all M polymers fect in colloidal systems, we asked for correlations in the in the simulation box, and over three independent real- k transverse displacements r ,i of neighbouring polymers, isations. We chose to characterise the dynamics of the and define their covariance⊥ as internal bending modes of a tracer polymer in terms of k l k l its tangent-tangent correlation function, cov(∆t, ∆x) := r ,ir ,i − hr ,iihr ,ii . (13) ⊥ ⊥ ⊥ ⊥

Cij := hti · tji , (6) The average in Eq. (13) is taken for a given chain k with all other chains k =6 l at a given time ∆t and a dis- k l and the mean-square displacements (MSD) of the end- tance ∆x = ri (t + ∆t) − ri(t + ∆t) , and we have also to-end distance Rk = |Rk|, performed a moving time window average over a win- dow of size 15 τd discretised in subintervals of size ∆t. 2  k k 2 δR (t) := h R (t) − R (0) i . (7) Moreover, for simulations with ∆t < 0.2τd and ∆t = τd, we averaged over 4 or 12 independent realisations, re- In order to characterise the center-of-mass motion of a spectively. For specificity, we used the beads at position tracer filament we used the mean-square displacement of i = N/5 and the equivalent beads at i = 4N/5.

[1] Edwards, S. F. The statistical mechanics of polymerized polymer systems. Part 1. Brownian motion in the equilib- material. Proc. Phys. Soc. 92, 9 (1967). rium state. J. Chem. Soc. Faraday Trans. 74, 1789–1801 [2] de Gennes, P.-G. Reptation of a polymer chain in the (1978). presence of fixed obstacles. J. Chem. Phys. 55, 572 [4] Kremer, K. & Grest, G. S. Dynamics of entangled lin- (1971). ear polymer melts: A molecular-dynamics simulation. J. [3] Doi, M. & Edwards, S. F. Dynamics of concentrated Chem. Phys. 92, 5057 (1990). 8

[5] McLeish, T. Tube theory of entangled polymer dynamics. [28] Viovy, J. L., Rubinstein, M. & Colby, R. H. Constraint Adv. Phys. 51, 1379–1527 (2002). release in polymer melts: Tube reorganization versus [6] Everaers, R. et al. Rheology and microscopic topology tube dilation. Macromol. 24, 3587–3596 (1991). of entangled polymeric liquids. Science 303, 823–826 [29] Doi, M. & Edwards, S. F. Dynamics of rod-like macro- (2004). molecules in concentrated solution. Part 1. J. Chem. Soc. [7] Hou, J.-X., Svaneborg, C., Everaers, R. & Grest, G. S. Faraday Trans. 74, 560–570 (1978). Stress relaxation in entangled polymer melts. Phys. Rev. [30] Höfling, F., Munk, T., Frey, E. & Franosch, T. Entangled Lett. 105, 068301 (2010). dynamics of a stiff polymer. Phys. Rev. E 77, 060904 [8] Rubinstein, M. – The ugly duckling (2008). story: Will polymer physics ever become a part of [31] Leitmann, S., Höfling, F. & Franosch, T. Tube concept “proper” physics? J. Polym. Sci. B Polym. Phys. 48, for entangled stiff fibers predicts their dynamics in space 2548 (2010). and time. Phys. Rev. Lett. 117, 097801 (2016). [9] Broedersz, C. P. & MacKintosh, F. C. Modeling semiflex- [32] Munk, T., Höfling, F., Frey, E. & Franosch, T. Effective ible polymer networks. Rev. Mod. Phys. 86, 995 (2014). perrin theory for the anisotropic diffusion of a strongly [10] Odijk, T. The statistics and dynamics of confined or en- hindered rod. Europhys. Lett. 85, 30003 (2009). tangled stiff polymers. Macromol. 16, 1340–1344 (1983). [33] Höfling, F., Frey, E. & Franosch, T. Enhanced diffusion [11] Semenov, A. N. Dynamics of concentrated solutions of of a needle in a planar array of point obstacles. Phys. rigid-chain polymers. Part 1. Brownian motion of per- Rev. Lett. 101, 120605 (2008). sistent macromol. in isotropic solution. J. Chem. Soc. [34] Nam, G., Johner, A. & Lee, N.-K. Reptation of a semi- Faraday Trans. 2 82, 317–329 (1986). flexible polymer through porous media. J. Chem. Phys. [12] Romanowska, M. et al. Direct observation of the tube 133, 044908 (2010). model in F-actin solutions: Tube dimensions and curva- [35] Berthier, L. & Biroli, G. Theoretical perspective on the tures. Europhys. Lett. 86, 26003 (2009). glass transition and amorphous materials. Rev. Mod. [13] Ramanathan, S. & Morse, D. C. Simulations of dynamics Phys. 83, 587 (2011). and viscoelasticity in highly entangled solutions of semi- [36] Fixman, M. Simulation of polymer dynamics: 1 General flexible rods. Phys. Rev. E 76, 010501 (2007). theory. J. Chem. Phys. 69, 1527–1537 (1978). [14] Granek, R. From semi-flexible polymers to membranes: [37] Grassia, P. S., Hinch, E. J. & Nitsche, L. C. Computer- anomalous diffusion and reptation. J. Phys. II 7, 1761– simulations of brownian-motion of complex systems. J. 1788 (1997). Fluid Mech. 282, 373–403 (1995). [15] Fakhri, N., MacKintosh, F. C., Lounis, B., Cognet, L. [38] Kremer, K. & Grest, G. S. Simulations for structural and & Pasquali, M. Brownian motion of stiff filaments in a dynamic properties of dense polymer systems. J. Chem. crowded environment. Science 330, 1804–1807 (2010). Soc. Faraday Trans. 88, 1707–1717 (1992). [16] Liu, J. et al. Microrheology probes length scale depen- [39] Paul, W., Binder, K., Heermann, D. W. & Kremer, K. dent rheology. Phys. Rev. Lett. 96, 118104 (2006). Dynamics of polymer solutions and melts. reptation pre- [17] Semmrich, C. et al. Glass transition and rheological re- dictions and scaling of relaxation times. J. Chem. Phys. dundancy in F-actin solutions. Proc. Natl. Acad. Sci. U 95, 7726–7740 (1991). SA 104, 20199–20203 (2007). [40] Dijkstra, M. & Frenkel, D. Simulation study of the [18] Perkins, T. T., Quake, S. R., Smith, D. E. & Chu, S. isotropic-to-nematic transitions of semiflexible polymers. Relaxation of a single DNA molecule observed by optical Phys. Rev. E 51, 5891 (1995). microscopy. Science 264, 822–826 (1994). [41] Khokhlov, A. R. & Semenov, A. N. Liquid-crystalline [19] Hinner, B., Tempel, M., Sackmann, E., Kroy, K. & Frey, ordering in the solution of long persistent chains. Physica E. Entanglement, elasticity, and viscous relaxation of A 108, 546–556 (1981). actin solutions. Phys. Rev. Lett. 81, 2614 (1998). [42] Onsager, L. The effects of shape on the interaction of [20] Käs, J., Strey, H. & Sackmann, E. Direct imaging of colloidal particles. Ann. N.Y. Acad. Sci. 51, 627–659 reptation for semiflexible actin filaments. Nature 368, (1949). 226–229 (1994). [43] Grest, G. S. & Kremer, K. Molecular dynamics simula- [21] Perkins, T. T., Smith, D. E. & Chu, S. Direct observation tion for polymers in the presence of a heat bath. Phys. of tube-like motion of a single polymer chain. Science Rev. A 33, 3628 (1986). 264, 819–822 (1994). [44] Kratky, O. & Porod, G. Röntgenuntersuchung gelöster [22] Le Goff, L., Hallatschek, O., Frey, E. & Amblard, F. Fadenmoleküle. Recl. Trav. Chim. 68, 1106–1122 (1949). Tracer studies on F-actin fluctuations. Phys. Rev. Lett. [45] Saitô, N., Takahashi, K. & Yunoki, Y. Statistical me- 89, 258101 (2002). chanical theory of stiff chains. J. Phys. Soc. Jpn. 22, [23] Farge, E. & Maggs, A. C. Dynamic scattering from semi- 219–226 (1967). flexible polymers. Macromol. 26, 5041–5044 (1993). [46] Weeks, J. D., Chandler, D. & Andersen, H. C. Role of [24] Frey, E. & Nelson, D. R. Dynamics of flat membranes repulsive forces in determining the equilibrium structure and flickering in red blood cells. J. Phys. I (France) 1, of simple liquids. J. Chem. Phys. 54, 5237 (1971). 1715–1757 (1991). [47] Galassi, M. et al. GSL–GNU Scientific Library: Refer- [25] Hallatschek, O., Frey, E. & Kroy, K. Propagation and ence manual (2011). relaxation of tension in stiff polymers. Phys. Rev. Lett. [48] Giordano, N. J. & Nakanishi, H. Computational Physics 94, 77804 (2005). (Addison-Wesley, New York, 2005). [26] Kroy, K. & Frey, E. Dynamic scattering from solutions [49] Doliwa, B. & Heuer, A. Cage effect, local anisotropies, of semiflexible polymers. Phys. Rev. E 55, 3092 (1997). and dynamic heterogeneities at the glass transition: A [27] Wilhelm, J. & Frey, E. Radial distribution function of computer study of hard spheres. Phys. Rev. Lett. 80, semiflexible polymers. Phys. Rev. Lett. 77, 2581 (1996). 4915 (1998).