active entanglement-tracking microrheology directly ...home.sandiego.edu/~randerson/active_t.pdf ·...

6
Active Entanglement-Tracking Microrheology Directly Couples Macromolecular Deformations to Nonlinear Microscale Force Response of Entangled Actin Tobias T. Falzone and Rae M. Robertson-Anderson* Department of Physics, University of San Diego, San Diego, California 92110, United States * S Supporting Information ABSTRACT: We track the deformation of discrete entangled actin segments while simultaneously measuring the resistive force the deformed laments exert in response to an optically driven microsphere. We precisely map the network deformation eld to show that local microscale stresses can induce lament deformations that propagate beyond mesoscopic length scales (60 μm, >3 persistence lengths l p ). We show that the lament persistence length controls the critical length scale at which distinct entanglement deformations become driven by collective network mechanics. Mesoscale propagation beyond l p is coupled with nonlinear local stresses arising from steric entanglements mimicking cross-links. E ntangled networks of the archetypical semiexible polymer actin, which display complex viscoelastic stress responses, are responsible for most mechanically driven cellular processes and exhibit powerful tunable properties for novel biomimetic materials. 16 The complexity of actin networks, compared to exible polymers, stems from the large persistence length of actin laments (l p 17 μm) relative to their size (520 μm) and the length between entanglements (l e 1 μm). Despite widespread interest in actin network mechanics, 710 the lament dynamics that give rise to stress response, including the propagation of lament deformations surrounding an induced strain, remain controversial. 11,12 Recently, established microrheology techniques can characterize molecular-level stress response by using passively diusing or actively driven microscale probes (beads) to sense network dynamics. However, two-point microrheology, which measures correla- tions between two diusing beads in the network, is limited to the near-equilibrium regime, 13,14 and microscope rheometers that image embedded beads during bulk strain 15,16 cannot determine localized stress propagation. Both techniques also indirectly infer network deformation from foreign probes that can underestimate or disrupt network mechanics. Previous studies probing actin network mechanics have largely focused on cross-linked networks and linear regimes where laments are only weakly perturbed. 6,1721 However, steric entanglement dynamics, especially those far from equilibrium, remain relatively unexplored. 3,22 Cross-linking entangled laments has been suggested to inhibit nonane (nonuniform) network deformation typically exhibited by entangled networks by suppressing lament-bending modes. 15,16,23 Resulting ane (uniform) response, driven solely by lament-stretching modes, is predicted to give rise to nonlinear stress response not apparent in entangled net- works. 15,24 However, we recently revealed microscale stress response for entangled actin that exhibited key nonlinear properties such as stress-stiening and power-law relaxation when strain rates γ̇ exceeded the disentanglement rate (i.e., rate at which individual deformed entanglement lengths relax, τ ent 1 ). 22, 25 Relaxation dynamics indicated dilation and subsequent healing of the classical reptation model entangle- ment tube surrounding each lament. 26,27 Here, we demonstrate a new widely applicable technique that tracks submicron segments of entangled actin laments while simultaneously measuring the stress induced in tracked laments by a bead driven through the network (Figure 1). We characterize the network deformation eld and couple induced microscale stress to network deformations over 3 orders of length scales: from below the network entanglement length (l e 0.42 μm) 7,25 to >3× the largest length scale of the network (the persistence length l p ). We show that the persistence length dictates the distance from the applied strain R at which noncontinuum deformations give way to collective network mechanics. We demonstrate that for strain rates above the disentanglement rate (γ̇ τ ent 1 3 s 1 ), reduced nonane deformations similar to those of cross-linked networks lead to nonlinear stresses that propagate beyond mesoscale distances. For all measurements, a trapped bead (radius r = 2.25 μm) is pulled 10 μm through the network at speeds v = 1, 5, and 10 Received: September 15, 2015 Accepted: October 14, 2015 Letter pubs.acs.org/macroletters © XXXX American Chemical Society 1194 DOI: 10.1021/acsmacrolett.5b00673 ACS Macro Lett. 2015, 4, 11941199

Upload: others

Post on 15-Jun-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Active Entanglement-Tracking Microrheology Directly ...home.sandiego.edu/~randerson/Active_T.pdf · Active Entanglement-Tracking Microrheology Directly Couples Macromolecular Deformations

Active Entanglement-Tracking Microrheology Directly CouplesMacromolecular Deformations to Nonlinear Microscale ForceResponse of Entangled ActinTobias T. Falzone and Rae M. Robertson-Anderson*

Department of Physics, University of San Diego, San Diego, California 92110, United States

*S Supporting Information

ABSTRACT: We track the deformation of discrete entangledactin segments while simultaneously measuring the resistiveforce the deformed filaments exert in response to an opticallydriven microsphere. We precisely map the network deformationfield to show that local microscale stresses can induce filamentdeformations that propagate beyond mesoscopic length scales(60 μm, >3 persistence lengths lp). We show that the filamentpersistence length controls the critical length scale at whichdistinct entanglement deformations become driven bycollective network mechanics. Mesoscale propagation beyondlp is coupled with nonlinear local stresses arising from stericentanglements mimicking cross-links.

Entangled networks of the archetypical semiflexible polymeractin, which display complex viscoelastic stress responses,

are responsible for most mechanically driven cellular processesand exhibit powerful tunable properties for novel biomimeticmaterials.1−6 The complexity of actin networks, compared toflexible polymers, stems from the large persistence length ofactin filaments (lp ≈ 17 μm) relative to their size (∼5−20 μm)and the length between entanglements (le ∼ 1 μm). Despitewidespread interest in actin network mechanics,7−10 thefilament dynamics that give rise to stress response, includingthe propagation of filament deformations surrounding aninduced strain, remain controversial.11,12 Recently, establishedmicrorheology techniques can characterize molecular-levelstress response by using passively diffusing or actively drivenmicroscale probes (beads) to sense network dynamics.However, two-point microrheology, which measures correla-tions between two diffusing beads in the network, is limited tothe near-equilibrium regime,13,14 and microscope rheometersthat image embedded beads during bulk strain15,16 cannotdetermine localized stress propagation. Both techniques alsoindirectly infer network deformation from foreign probes thatcan underestimate or disrupt network mechanics.Previous studies probing actin network mechanics have

largely focused on cross-linked networks and linear regimeswhere filaments are only weakly perturbed.6,17−21 However,steric entanglement dynamics, especially those far fromequilibrium, remain relatively unexplored.3,22 Cross-linkingentangled filaments has been suggested to inhibit nonaffine(nonuniform) network deformation typically exhibited byentangled networks by suppressing filament-bendingmodes.15,16,23 Resulting affine (uniform) response, driven solelyby filament-stretching modes, is predicted to give rise to

nonlinear stress response not apparent in entangled net-works.15,24 However, we recently revealed microscale stressresponse for entangled actin that exhibited key nonlinearproperties such as stress-stiffening and power-law relaxationwhen strain rates γ exceeded the disentanglement rate (i.e., rateat which individual deformed entanglement lengths relax,τent

−1).22,25 Relaxation dynamics indicated dilation andsubsequent healing of the classical reptation model entangle-ment tube surrounding each filament.26,27

Here, we demonstrate a new widely applicable technique thattracks submicron segments of entangled actin filaments whilesimultaneously measuring the stress induced in trackedfilaments by a bead driven through the network (Figure 1).We characterize the network deformation field and couple

induced microscale stress to network deformations over 3orders of length scales: from below the network entanglementlength (le ≈ 0.42 μm)7,25 to >3× the largest length scale of thenetwork (the persistence length lp). We show that thepersistence length dictates the distance from the applied strainR at which noncontinuum deformations give way to collectivenetwork mechanics. We demonstrate that for strain rates abovethe disentanglement rate (γ ≈ τent

−1 ≈ 3 s−1), reducednonaffine deformations similar to those of cross-linkednetworks lead to nonlinear stresses that propagate beyondmesoscale distances.For all measurements, a trapped bead (radius r = 2.25 μm) is

pulled 10 μm through the network at speeds v = 1, 5, and 10

Received: September 15, 2015Accepted: October 14, 2015

Letter

pubs.acs.org/macroletters

© XXXX American Chemical Society 1194 DOI: 10.1021/acsmacrolett.5b00673ACS Macro Lett. 2015, 4, 1194−1199

Page 2: Active Entanglement-Tracking Microrheology Directly ...home.sandiego.edu/~randerson/Active_T.pdf · Active Entanglement-Tracking Microrheology Directly Couples Macromolecular Deformations

μm/s, which we express as strain rates via γ = 3v/√2r andnormalize by γC ≅ 3 s−1.28 By analogy to constant ratemacrorheology shear strain measurements, in which affinedeformation is in the direction of applied strain and nonaffine isperpendicular to strain,16 we refer to filament displacementsparallel to the strain path (x) as affine and perpendiculardisplacements (y) as nonaffine. We determine the dependenceof tracked segment displacements on R by calculating thecorresponding ensemble average within an annulus of radius R(Figure 1c). For affine motion, we quantify the ensemble-averaged maximum displacement xmax induced by the strain,maximum recovery distance xrec following strain (Figure 1d),and fractional recovery ( f rec = xrec/xmax; Figures 2 and 3). Whilexmax quantifies how much the network is deformed by thestrain, xrec quantifies how much it retracts back to startingconfiguration. Fractional recovery of 0 indicates completedissipation or fluidization, while f rec = 1 implies completeelasticity or reversibility.We find that maximum filament displacement xmax and

fractional recovery f rec (Figure 2) display spatial crossoversfrom noncontinuum filament dynamics to that of self-averagingnetwork mechanics at R ≈ lp. The slowest applied strain (0.3γC)displaces segments near the bead path up to ∼4× more thanstrains above γC, indicating more fluid-like (less constrained)deformations, which are exponentially suppressed to xmax/e at∼lp (Figure 2a). Conversely, more constrained deformations(smaller xmax) for γ > γ C, indicative of more rigidentanglements, propagate well beyond lp, with xmax linearlydecaying 2-fold from R ≈ lp to 2lp. Previous studies show tracerbeads in cross-linked networks subject to bulk shear straindisplay similar linear dependence of affine network deformationon distance from the applied strain,15 suggesting (i)entanglement dynamics beyond lp are dictated by bulkmechanics, and (ii) entanglements can mimic cross-links for

rates above the disentanglement rate (∼γC). Correspondingforce measurements for γ > γC show nonlinear stress-stiffeningsimilar to that of cross-linked networks (Figure SI1b).22 Thus,stiff entanglements near the bead path, able to propagatedeformations over several lp, give rise to nonlinear forceresponse (i.e., stress-stiffening), as opposed to γ < γC linearstress response resulting from fluid-like filament deformationsunable to propagate beyond ∼lp.This rate-dependent crossover to stiffening and entangle-

ment rigidity (strain-induced cross-linking) is also apparent inthe fractional deformation recovery f rec for R < lp (Figure 2b)showing nearly complete elastic recovery ( f rec ≥ 0.7) for γ > γCversus effectively irreversible dissipation ( f rec < 0.3) for γ < γC.Beyond ∼lp, fractional recovery for all γ no longer varies with R,indicating self-averaging network mechanics. For γ < γC, thedeformation field displays increased elasticity compared to R <lp deformations ( f rec increases to 0.5); while for γ > γC this self-averaging regime displays relatively increased dissipation ( f recdecreases to ∼0.6). Magnetic tweezer experiments on cross-linked microtubules have found a similar coexistence ofdissipation and recovery which authors interpret as due to afraction of cross-links rupturing (allowing for flow) while othersare maintained (providing elasticity).29 Similarly, we canunderstand the partial dissipation at R > lp as due to ∼50−70% of entanglements acting as effective cross-links, while∼30−50% are free to reorient on the time scale of the strain.While our slowest strain is slower than the disentanglement rateτent

−1 ≈ 3 s−1, allowing for substantial fluidization/reorientationof entanglements near the bead path, it is >10× faster than therate at which entire deformed entanglement tubes (eachformed by many entanglements) relax (i.e., the disengagementrate τD

−1 ≈ 0.07 s−1).22,25 Because bulk network elasticity canonly be completely released at rates below τD

−1 (∼γC/50), we

Figure 1. Active entanglement-tracking microrheology. (a) Cartoon of tracking labeled segments along an actin filament. (b) Microscope image offilament with interspersed ∼0.45 μm labeled segments that are centroid-tracked during experiments. (c) Tracking segments at varying distances Rfrom strain path. Orange circles represent bead position before/after strain and yellow line represents strain path. Dotted lines outline annulipositioned every 4.5 μm from strain path (R = 6.75, 11.25, ..., 29.25 μm). All tracks within each annulus are used to determine R dependence ofsegment deformation. All tracks i for R = 22.5 μm at γ = 0.3γC are shown. (d) Tracked segment displacements (xi(t),yi(t)) and induced force aremeasured during the following: equilibrium [no trap movement]; strain [trapped bead pulled 10 μm through network at constant rates γ = 0.3, 1.7,3.3γC]; recovery [no trap movement]. Average x-displacement ⟨xi(t)⟩ of the tracked filaments shown in (c) (blue triangles), force exerted on bead(green circles), and trap position (red line, axis not shown) are shown.

ACS Macro Letters Letter

DOI: 10.1021/acsmacrolett.5b00673ACS Macro Lett. 2015, 4, 1194−1199

1195

Page 3: Active Entanglement-Tracking Microrheology Directly ...home.sandiego.edu/~randerson/Active_T.pdf · Active Entanglement-Tracking Microrheology Directly Couples Macromolecular Deformations

find nonzero f rec at length scales beyond lp for all measuredstrain rates.This spatially dependent recovery is also consistent with our

measured force relaxations that indicate nonclassical entangle-ment tube dilation (Figure SI1).22 Strain-induced dilationshould be strongest near the strain where entanglements areeasily forced apart (dilating entanglement tubes), but as thestress propagates along entanglements, the impact of thedisturbance on any one entanglement segment is reduced(limiting dilation) and all segments display collective networkbehavior. Tube dilation leads to faster relaxation dynamics, sofor γ < γC, entanglements closest to the strain path can almostcompletely relax on the time scale of strain (dissipation, smallf rec). However, during relaxation, these entanglements redis-tribute induced stress to neighboring entanglements, furtherfrom the strain site, in less dilated tubes. These more confinedentanglements cannot relax as quickly, resulting in enhancedmemory (increased f rec). For γ > γC increased dissipationbeyond lp indicates a shift from a microscale stress-stiffeningregime evidenced by virtually complete recovery to a bulkregime that displays enhanced dissipation and, thus, stress-softening. This spatial dependence on nonlinear stress responsecan explain why bulk rheology measurements of entangled actinonly exhibit stress-softening dynamics (with stress-stiffeningreserved for cross-linked networks).15,30,31

We note that while overall behavior of the two rates above γCare quite similar compared to 0.3γC, the intermediate strain(1.7γC) exhibits more dissipative features than for 3.3γC.Specifically, following the 3.3γC strain, filaments close to theapplied strain completely recover ( f rec ≈ 1) dissipating to f rec ≈0.7 in the continuum phase; while at 1.7γC, filaments onlypartially recover near the strain ( f rec ≈ 0.7) reducing to f rec ≈0.5 in the continuum phase. This enhanced dissipation likelyarises from the closeness of 1.7γC to the crossover rate, allowingfor partial entanglement relaxation on the time scale of thestrain. In other words, some of the entanglements are mobile,analogous to ruptured cross-links, while others are fixed,providing the partial recovery.29

Filament recovery distance xrec (Figure 3a) corroborates thatthe persistence length mediates bulk mechanics. Namely, fordistances beyond lp, xrec is independent of strain rate andexponentially decays over a length scale R ≈ 2lp (xrec ∼ e−R/2.4lp).Simulations of cross-linked networks subject to local forceperturbations have also predicted exponential decay of networkdeformations beyond a critical distance from the perturbation,

Figure 2. Induced entanglement deformation and subsequent recoverydisplay spatial and rate-dependent dynamic crossovers. (a) Maximumaffine displacement xmax and (b) fractional recovery ( f rec = xrec/xmax) vsR for all strain rates γ, with dashed lines at spatial crossover length ∼lp.

Figure 3. Network deformation recovery displays distance-dependenttube dilation. (a) Affine recovery distance xrec vs R for all strain rates γ.Beyond R ≈ lp (dashed lines), filament recovery exponentially decaysas xrec ∼ exp(−R/2.4lp). (b) Recovery rates vs R, measured via fittingeach ⟨xi(t)⟩ during recovery (Figures 1d and SI2) to a singleexponential. Displayed recovery rates for each R, which areindependent of γ, are averaged over all three γ and normalized bythe classically predicted disengagement rate of undilated entanglementtubes (τD

−1 ≈ 0.07 s−1).22

ACS Macro Letters Letter

DOI: 10.1021/acsmacrolett.5b00673ACS Macro Lett. 2015, 4, 1194−1199

1196

Page 4: Active Entanglement-Tracking Microrheology Directly ...home.sandiego.edu/~randerson/Active_T.pdf · Active Entanglement-Tracking Microrheology Directly Couples Macromolecular Deformations

which authors interpret as due to self-averaging behavior.32 Thecorresponding length scale λ identified for cross-linkednetworks is mediated by the distance between cross-links lCand the bending rigidity of the filaments, scaling as λ ∼(lClp)

1/2.23,30 Thus, if we remove fixed cross-links from thesystem, then the most obvious length scale controllingcontinuum behavior is indeed the persistence length.Decreasing recovery distance with increasing R is consistent

with entanglement tubes becoming less dilated further from thestrain site; that is, filaments are more confined, inhibitingretraction to starting positions. Distance-dependent tubedilation is further displayed by measured recovery rates ofdeformed filaments (Figure 3b). All filament deformationsduring recovery decay exponentially in time (Figures 1d andSI2) with γ-independent relaxation rates that decrease withincreasing R. Filaments closest to the bead path decay at a rateof ∼0.3 s−1, corresponding to ∼5τD−1,22 indicating thatentanglement tubes closest to the strain are dilated ∼2.3×the equilibrium diameter dT (dT

−2 ∼ τD). At the furthest pointfrom strain, relaxation slows to ∼1.4τD−1, corresponding to∼1.1dT, showing that tubes far from the strain maintain aclassically predicted size.Force measurements show that tube dilation leads to

exponential force relaxation for γ < γC (linear response) versustwo-phase power-law relaxation for γ > γC (nonlinear response;Figure SI1).22 While for γ < γC, exponential force decay rates of5τD

−1 corroborate the corresponding recovery data near thestrain (Figure 3b), the distinct power-law relaxation for γ > γC isunique to the force response. Two-phase force relaxation ispredicted to arise from initial fast tube contraction, followed byslow reptation, coupled with further contraction.22,27 Thisnonlinear force response is coupled with filament deformationsthat propagate over several lp with minimal dissipation. Thus,we can understand power-law force relaxation as a second ordereffect of distant entanglements needing to relax before highlystressed entanglements close to the strain can completelyrecover. Put differently, nonlinear force response arises fromsubstantial collective mechanics beyond lp contributingsignificantly to force exerted at the microscale strain site.Network deformation is largely dominated by affine (parallel

to strain path) entanglement displacements, which have beenthe focus of the discussion above. However, rate-dependentnonaffine network deformation is also seen in Figures 4 andSI3, which map maximum segment displacements induced bythe strain (x max + ymax) averaged over (2.5 μm)2 regions of thenetwork, colorized by the directionality of ymax. By analogy toprevious studies characterizing nonaffine response to uniformshear strains,15,16,33 we quantify nonaffinity by a normalizednonaffinity parameter N(R) = ymax

2/(xmax2 + ymax

2) (Figure 4d).We find that, for γ < γC, N is independent of R, with an averagevalue that is greater in magnitude than that for γ > γCdeformations. This R-independence and increased magnitudehave been previously reported for minimally cross-linked actinnetworks subject to macroscopic shear strain, due to bendingmodes dominating the response.15,16 As cross-linking increases,bending modes are suppressed and N is expected to scale withR as N ∼ Ra, where 0 < a < 2 depends on the level of cross-linking. This bending/nonaffinity suppression is the source ofstress-stiffening and nonlinearity driven principally by stretch-ing modes.23,30 In accord with this model, we find that, for γ >γC, N ∼ R0.1, signifying that strain rates above γC indeed inducecross-linking of entanglements. We further quantify the γ-dependence of nonaffinity by determining the R-averaged

nonaffine deformations ⟨ymax2(R)⟩ for all strain rates (Figure

4e). We find that ⟨ymax2(R)⟩ decreases with increasing γ,

demonstrating that at higher strain rates entangled networkstransition to elastically driven dynamics dominated by

Figure 4. Displacement maps show γ-dependent nonaffine filamentdeformation indicative of strain-induced cross-linking of entangle-ments. (a−c) Arrows represent maximum filament displacements (x max+ ymax) averaged over (2.5 μm)2 regions, colorized by ymax direction[positive (blue), negative (red)]. (d) Nonaffinity parameter N vs R forall γ (Figure 2 color scheme). Dashed line shows power-law scaling N∼ R0.1. (e) R-averaged nonaffine deformations ⟨ymax

2(R)⟩ vs γ 2 showthat nonaffine deformations decrease with increasing γ.

ACS Macro Letters Letter

DOI: 10.1021/acsmacrolett.5b00673ACS Macro Lett. 2015, 4, 1194−1199

1197

Page 5: Active Entanglement-Tracking Microrheology Directly ...home.sandiego.edu/~randerson/Active_T.pdf · Active Entanglement-Tracking Microrheology Directly Couples Macromolecular Deformations

stretching modes, similar to that of cross-linked networks.Substantial nonaffine response of cross-linked networks is onlypredicted for length scales below λ ∼ (lClp)

1/2.23,30−32 Strain-induced cross-linking of all entanglements would imply lC ≈ leand λ ≈ 4 μm, while our measurement range is ∼2λ < R < 15λ.Thus, we expect our nonaffinity parameter to scale similarly tobulk measurements as described by the scaling above. Further,while λ is predicted to also mediate the crossover to continuummechanics for cross-linked networks, our results indicate thatthe length scales controlling nonaffinity and noncontinuummechanics are two distinct parameters for entangled networks.As the microscale strain is not as simple as uniform bulk

shear strain, we expect to see a footprint of the bead path andgeometry in the deformation field. Previous simulations ofconstant force microsphere perturbations, similar to magnetictweezer measurements, predict that entangled filaments pile upin front of the bead and vacate the wake behind it, leading to anosmotically controlled restoring force on the bead.34−36 Thisosmotic pressure difference leads to steric repulsion of filamentsaway from the leading edge of the bead and filament attractiontoward the trailing edge, which is exactly what we see in Figures4a−c and SI3. The trailing-edge wake is most apparent forhigher strain rates due to inability of entanglement segments torearrange or bend during strain, further corroborating stress-stiffening behavior (Figure SI1b). The repulsion/attraction ismost evident for 0.3γC, as entanglements can easily flow andredistribute in response to the disturbance.In conclusion, we demonstrate an innovative technique able

to directly track individual entanglement segments overmesoscopic length scales while measuring microscale forces inresponse to local strains. We show that for entangledsemiflexible polymers, the persistence length mediates thecrossover from hierarchical deformations to continuum net-work mechanics; while the disentanglement rate mediates theonset of nonlinearity. The mesoscale spatial crossover at lp islikely unique to semiflexible filaments (lp > le) in whichindividual entanglement segments along a filament aremechanically connected via the persistence length, preventingindependent relaxation and continuum mechanics at lengthscales below lp. Strain-induced cross-linking of entanglementsfor γ > τent

−1 leads to discrete elastic deformations near thestrain path (R < lp) coupled with long-range stress propagation;with bulk network deformations beyond lp contributingsignificantly to the force exerted at the strain site. Rigidentanglements < lp from the strain give rise to signaturenonlinear dynamics, including stress-stiffening and suppressednonaffinity that are quenched over a length scale lp asentanglements soften and become increasingly mobile.

■ EXPERIMENTAL METHODSDiscrete Filament Labeling. Multisegmented fluorescent actin

filaments were generated via simultaneous annealing and polymer-ization of labeled and unlabeled actin, forming filaments withinterspersed fluorescent-labeled sections ∼0.45 μm in length (Figure1b).Sample Preparation. Discretely labeled actin filaments and BSA-

treated 4.5 μm microspheres are embedded in 0.5 mg/mL entangledactin networks.Data Collection. An Olympus IX70 microscope imaged the

network and served as the optical trap base.37 We precisely move thetrapped bead through the network using a piezoelectric mirror whilecapturing videos of the network (2.5 fps) and measuring the forceexerted on a trapped bead (20 kHz).22

Data Processing/Analysis. We use Crocker and Weeks’ particletracking algorithms to obtain positions of labeled segments and linkpositions in time.38 From the resulting tracks, we quantify thedisplacement of each segment i in both the parallel/affine (xi(t)) andperpendicular/nonaffine (yi(t)) directions. We determine the depend-ence of segment displacement on R by selecting all tracks within a 4.5μm wide annulus of radius R centered on the bead path (Figure 1c)and calculating the corresponding ensemble average (⟨xi(t)⟩, ⟨yi(t)⟩).Error bars are determined by bootstrapping over 100 repeatingsubsets.39

All methods are further described in the Supporting Information.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acsmacro-lett.5b00673.

Expanded experimental methods, Figures SI1, SI2, andSI3 (figure content described in text) (PDF).

■ AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected] authors declare no competing financial interest.

■ ACKNOWLEDGMENTSThis research was funded by an NSF CAREER Award, GrantNo. 1255446.

■ REFERENCES(1) Wang, B.; Guan, J.; Anthony, S. M.; Bae, S. C.; Schweizer, K. S.;Granick, S. Phys. Rev. Lett. 2010, 104 (11), 118301.(2) Chhabra, E. S.; Higgs, H. N. Nat. Cell Biol. 2007, 9 (10), 1110−1121.(3) Stricker, J.; Falzone, T.; Gardel, M. L. J. Biomech. 2010, 43 (1),9−14.(4) Bausch, A.; Kroy, K. Nat. Phys. 2006, 2 (4), 231−238.(5) Buehler, M. J.; Yung, Y. C. Nat. Mater. 2009, 8 (3), 175−188.(6) Kas, J.; Strey, H.; Tang, J.; Finger, D.; Ezzell, R.; Sackmann, E.;Janmey, P. Biophys. J. 1996, 70 (2), 609−625.(7) Gardel, M.; Valentine, M.; Crocker, J. C.; Bausch, A.; Weitz, D.Phys. Rev. Lett. 2003, 91 (15), 158302.(8) Chen, D. T.; Wen, Q.; Janmey, P. A.; Crocker, J. C.; Yodh, A. G.Annu. Rev. Condens. Matter Phys. 2010, 1, 301.(9) Wen, Q.; Basu, A.; Janmey, P. A.; Yodh, A. G. Soft Matter 2012, 8(31), 8039−8049.(10) Granick, S.; Bae, S. C.; Wang, B.; Kumar, S.; Guan, J.; Yu, C.;Chen, K.; Kuo, J. J. Polym. Sci., Part B: Polym. Phys. 2010, 48, 2542.(11) Larson, R. G. J. Polym. Sci., Part B: Polym. Phys. 2007, 45 (24),3240−3248.(12) Likhtman, A. E. J. Non-Newtonian Fluid Mech. 2009, 157 (3),158−161.(13) Squires, T. M.; Mason, T. G. Annu. Rev. Fluid Mech. 2010, 42(1), 413.(14) Starrs, L.; Bartlett, P. Faraday Discuss. 2003, 123, 323−334.(15) Liu, J.; Koenderink, G.; Kasza, K.; MacKintosh, F.; Weitz, D.Phys. Rev. Lett. 2007, 98 (19), 198304.(16) Basu, A.; Wen, Q.; Mao, X.; Lubensky, T.; Janmey, P. A.; Yodh,A. Macromolecules 2011, 44 (6), 1671−1679.(17) Koenderink, G.; Atakhorrami, M.; MacKintosh, F.; Schmidt, C.Phys. Rev. Lett. 2006, 96 (13), 138307.(18) Hinsch, H.; Frey, E. arXiv preprint arXiv:0907.1875 2009.(19) Semmrich, C.; Larsen, R. J.; Bausch, A. R. Soft Matter 2008, 4(8), 1675−1680.

ACS Macro Letters Letter

DOI: 10.1021/acsmacrolett.5b00673ACS Macro Lett. 2015, 4, 1194−1199

1198

Page 6: Active Entanglement-Tracking Microrheology Directly ...home.sandiego.edu/~randerson/Active_T.pdf · Active Entanglement-Tracking Microrheology Directly Couples Macromolecular Deformations

(20) Chaudhuri, O.; Parekh, S. H.; Fletcher, D. A. Nature 2007, 445(7125), 295−298.(21) Gardel, M.; Shin, J.; MacKintosh, F.; Mahadevan, L.;Matsudaira, P.; Weitz, D. Science 2004, 304 (5675), 1301−1305.(22) Falzone, T. T.; Blair, S.; Robertson-Anderson, R. M. Soft Matter2015, 11, 4418.(23) Levine, A. J.; Head, D.; MacKintosh, F. J. Phys.: Condens. Matter2004, 16 (22), S2079.(24) Wilhelm, J.; Frey, E. Phys. Rev. Lett. 2003, 91 (10), 108103.(25) Isambert, H.; Maggs, A. Macromolecules 1996, 29 (3), 1036−1040.(26) Sussman, D. M.; Schweizer, K. S. Macromolecules 2012, 45 (7),3270−3284.(27) Sussman, D. M.; Schweizer, K. S. Macromolecules 2013, 46 (14),5684−5693.(28) Squires, T. M. Langmuir 2008, 24 (4), 1147−1159.(29) Valentine, M. T.; Perlman, Z. E.; Gardel, M. L.; Shin, J. H.;Matsudaira, P.; Mitchison, T. J.; Weitz, D. A. Biophys. J. 2004, 86 (6),4004−4014.(30) Head, D. A.; Levine, A. J.; MacKintosh, F. Phys. Rev. Lett. 2003,91 (10), 108102.(31) Head, D. A.; MacKintosh, F.; Levine, A. J. Phys. Rev. E: Stat.Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2003, 68 (2), 025101.(32) Head, D.; Levine, A.; MacKintosh, F. Phys. Rev. E 2005, 72 (6),061914.(33) DiDonna, B.; Lubensky, T. Phys. Rev. E 2005, 72 (6), 066619.(34) Ter-Oganessian, N.; Pink, D.; Boulbitch, A. Phys. Rev. E 2005,72 (4), 041511.(35) Ter-Oganessian, N.; Quinn, B.; Pink, D. A.; Boulbitch, A. Phys.Rev. E 2005, 72 (4), 041510.(36) Uhde, J.; Feneberg, W.; Ter-Oganessian, N.; Sackmann, E.;Boulbitch, A. Phys. Rev. Lett. 2005, 94 (19), 198102.(37) Chapman, C. D.; Lee, K.; Henze, D.; Smith, D. E.; Robertson-Anderson, R. M. Macromolecules 2014, 47 (3), 1181−1186.(38) Crocker, J. C.; Weeks, E. R. Particle tracking using IDL. http://www.physics.emory.edu/faculty/weeks//idl/.(39) Efron, B. Biometrika 1981, 68 (3), 589−599.

ACS Macro Letters Letter

DOI: 10.1021/acsmacrolett.5b00673ACS Macro Lett. 2015, 4, 1194−1199

1199