arxiv: v1 [cond-mat.soft] 6 Dec 2017

Similar documents
arxiv: v2 [cond-mat.soft] 13 Feb 2015

Inhomogeneous elastic response of amorphous solids

This false color image is taken from Dan Howell's experiments. This is a 2D experiment in which a collection of disks undergoes steady shearing.

Workshop on Sphere Packing and Amorphous Materials July 2011

arxiv: v1 [cond-mat.soft] 29 Sep 2016

Jamming and the Anticrystal

Relevance of jamming to the mechanical properties of solids Sidney Nagel University of Chicago Capri; September 12, 2014

arxiv:cond-mat/ v2 [cond-mat.soft] 28 Mar 2007

Thermal fluctuations, mechanical response, and hyperuniformity in jammed solids

Indiana University, January T. Witten, University of Chicago

Active Stresses and Self-organization in Cytoskeletal Networks

Xiaoming Mao. Department of Physics and Astronomy, University of Pennsylvania. Collaborators: Tom Lubensky, Ning Xu, Anton Souslov, Andrea Liu

arxiv: v1 [cond-mat.soft] 7 Nov 2016

Xiaoming Mao Physics, University of Michigan, Ann Arbor. IGERT Summer Institute 2017 Brandeis

arxiv: v2 [cond-mat.soft] 27 Jan 2011

Modeling athermal sub-isostatic fiber networks

Normal Modes of Soft-Sphere Packings: from High to Physical Dimensions

arxiv:cond-mat/ v2 [cond-mat.soft] 18 Dec 2006

Inverse Design (and a lightweight introduction to the Finite Element Method) Stelian Coros

Structural Signatures of Mobility in Jammed and Glassy Systems

Granular materials (Assemblies of particles with dissipation )

2.1 Strain energy functions for incompressible materials

Continuum Mechanics and the Finite Element Method

Chapter 2. Rubber Elasticity:

Fig. 1. Circular fiber and interphase between the fiber and the matrix.

Normal modes in model jammed systems in three dimensions

Linearized Theory: Sound Waves

A MASTER EQUATION FOR FORCE DISTRIBUTIONS IN POLYDISPERSE FRICTIONAL PARTICLES

Mechanics PhD Preliminary Spring 2017

Elasticity of biological gels

3.22 Mechanical Properties of Materials Spring 2008

Distinct regimes of elastic response and deformation modes of cross-linked cytoskeletal and semiflexible polymer networks

Exploiting pattern transformation to tune phononic band gaps in a two-dimensional granular crystal

The response of an idealized granular material to an incremental shear strain

Constitutive Equations

Fundamentals of Linear Elasticity

Jamming. Or, what do sand, foam, emulsions, colloids, glasses etc. have in common? Or, transitions to rigidity in disordered media

Lecture 7, Foams, 3.054

Game Physics. Game and Media Technology Master Program - Utrecht University. Dr. Nicolas Pronost

arxiv: v2 [cond-mat.soft] 7 Nov 2014

The bulk modulus of covalent random networks

Structural signatures of the unjamming transition at zero temperature

The effect of plasticity in crumpling of thin sheets: Supplementary Information

Local Anisotropy In Globally Isotropic Granular Packings. Kamran Karimi Craig E Maloney

MODELING OF CONCRETE MATERIALS AND STRUCTURES. Kaspar Willam. Isotropic Elastic Models: Invariant vs Principal Formulations

Force, relative-displacement, and work networks in granular materials subjected to quasistatic deformation

Uncoupling shear and uniaxial elastic moduli of semiflexible biopolymer networks:

Classification of Prostate Cancer Grades and T-Stages based on Tissue Elasticity Using Medical Image Analysis. Supplementary Document

Measurements of the yield stress in frictionless granular systems

Biomechanics. Soft Tissue Biomechanics

HIGHLY ADAPTABLE RUBBER ISOLATION SYSTEMS

Two-dimensional ternary locally resonant phononic crystals with a comblike coating

collisions inelastic, energy is dissipated and not conserved

Soft Bodies. Good approximation for hard ones. approximation breaks when objects break, or deform. Generalization: soft (deformable) bodies

arxiv:cond-mat/ v1 [cond-mat.soft] 29 May 2002

SEMM Mechanics PhD Preliminary Exam Spring Consider a two-dimensional rigid motion, whose displacement field is given by

Rubber Elasticity (Indented text follows Strobl, other follows Doi)

Wave Propagation Through Soft Tissue Matter

A phenomenological model for shear-thickening in wormlike micelle solutions

An Atomistic-based Cohesive Zone Model for Quasi-continua

Advantages of a Finite Extensible Nonlinear Elastic Potential in Lattice Boltzmann Simulations

MECHANICS OF MATERIALS

THE PHYSICS OF FOAM. Boulder School for Condensed Matter and Materials Physics. July 1-26, 2002: Physics of Soft Condensed Matter. 1.

PHYS 101 Lecture 34 - Physical properties of matter 34-1

Testing Elastomers and Plastics for Marc Material Models

Chapter 2: Elasticity

Abvanced Lab Course. Dynamical-Mechanical Analysis (DMA) of Polymers

Ultralight Cellular Composite Materials with Architected Geometrical Structure. Maryam Tabatabaei and Satya N. Atluri

Module-4. Mechanical Properties of Metals

Nonlinear stability of steady flow of Giesekus viscoelastic fluid

arxiv: v1 [physics.comp-ph] 23 Aug 2017

A FAILURE CRITERION FOR POLYMERS AND SOFT BIOLOGICAL MATERIALS

PEAT SEISMOLOGY Lecture 2: Continuum mechanics

1. Background. is usually significantly lower than it is in uniaxial tension

Testing and Analysis

Critical scaling near the yielding transition in granular media

1. A pure shear deformation is shown. The volume is unchanged. What is the strain tensor.

Cover Page. The handle holds various files of this Leiden University dissertation

Pre-yield non-affine fluctuations and a hidden critical point in strained crystals

Failure process of carbon fiber composites. *Alexander Tesar 1)

16.21 Techniques of Structural Analysis and Design Spring 2003 Unit #5 - Constitutive Equations

BIOEN LECTURE 18: VISCOELASTIC MODELS

Understand basic stress-strain response of engineering materials.

Nonlinear analysis in ADINA Structures

Introduction to Continuous Systems. Continuous Systems. Strings, Torsional Rods and Beams.

Quasistatic behavior and force transmission in packing of irregular polyhedral particles

Mechanical Properties of Polymer Rubber Materials Based on a New Constitutive Model

Conservation of mass. Continuum Mechanics. Conservation of Momentum. Cauchy s Fundamental Postulate. # f body

Statistical Mechanics of Jamming

PIANO SOUNDBOARD UNDER PRESTRESS: A NUMERICAL APPROACH

Soft Matter PAPER. Dynamic Article Links C < Cite this: DOI: /c2sm07058h 2 Tianxiang Su and Prashant K.

JournalName. Stress anisotropy in shear-jammed packings of frictionless. 1 Introduction

Surface stress and relaxation in metals

The Finite Element Method II

Mechanics of Granular Matter

Frictional Jamming & MRJ/RLP crossover

ALGORITHM FOR NON-PROPORTIONAL LOADING IN SEQUENTIALLY LINEAR ANALYSIS

Lecture 15 Strain and stress in beams

arxiv: v1 [cond-mat.stat-mech] 6 Mar 2008

Emergent SO(3) Symmetry of the Frictionless Shear Jamming Transition

Transcription:

Origins of the Poynting effect in sheared elastic networks arxiv:72.278v [cond-mat.soft] 6 Dec 27 Karsten Baumgarten and Brian P. Tighe Delft University of Technology, Process & Energy Laboratory, Leeghwaterstraat 39, 2628 CB Delft, The Netherlands (Dated: December 7, 27) Poynting observed that some elastic solids dilate when sheared, while others contract. We determine the origins of this nonlinear effect in random spring networks, a minimal model for, e.g., biopolymer gels and solid foams. We show that the sign of the Poynting effect is determined by the microscopic Grüneisen parameter, which quantifies how stretching shifts a network s vibrational mode frequencies. Spring networks contract (a negative Poynting effect) because their frequencies shift upwards. We find that the amplitude of the Poynting effect is sensitive to the shear and bulk moduli, which can be tuned with a network s connectivity and its preparation protocol. Finally, we demonstrate that normal stresses cause the material to stiffen at finite strain. PACS numbers: Shear stresses can induce dilation, contraction, or normal stresses in a material even when the material is isotropic. While such mechanical coupling is often associated with granular matter [, 2], Kelvin predicted and Poynting demonstrated similar phenomena in elastic materials [3, 4]. Typical materials such as rubber dilate (volumetric strain ɛ > ) or develop compressive stresses (pressure change p > ). Yet in exceptional cases such as semiflexible polymers in the cytoskeleton and extracellular matrix, shear induces contraction or tension this is the negative Poynting effect [5]. Despite its long history, models of the Poynting effect are largely phenomenological or contain strong approximations [5 ]. Fig. illustrates the Poynting effect during pure shear at fixed volume (ɛ = ) of an isotropic random spring network in 2D. In addition to a shear stress q proportional to the shear strain γ, there is also an inherently nonlinear pressure change that grows initially as p γ 2. The quadratic scaling is due to isotropy, which requires p (or ɛ, if p is held fixed) to be an even function of γ. One contribution to the Poynting effect comes from asymmetry (hence nonlinearity) in the springs force-extension curve. If one assumes that nodes displace affinely, a neg- FIG. : (a) Pure shear strain γ applied to an unstressed spring network. Each node s area is proportional to its contribution to the pressure p; circles (squares) are tensile (compressive). (b) The initial growth of the shear stress q is linear in γ, while p is quadratic and negative (tensile). ative (positive) effect occurs when stretching produces a larger (smaller) force than compression by the same amount [5, 7]. This explanation is incomplete, however, as negative normal stresses also occur in simulations of fiber networks with linear stretching and bending that allow non-affine fluctuations [2, 3]. Here we bridge the gap between micromechanics and the Poynting effect, making no assumptions regarding non-affinity, nonlinear force laws, or bending. We focus on the initial growth of p and ɛ through the coefficients χ ɛ = [( 2 ) ] p γ 2 ɛ and χ p = [ ( 2 ) ɛ γ 2 p ]. () χ ɛ and χ p are evaluated in the initial condition ( ). Their subscript distinguishes strain control (fixed ɛ = ) from stress control (fixed p = p p = ). We derive exact expressions for the coefficients in hyperelastic solids, a common class of materials including rubbers, foams, and tissue. The expressions relate χ ɛ and χ p to a network s vibrational modes and the microscopic Grüneisen parameter Γ n [4], which quantifies how volumetric strain shifts the frequency ω n of the n th mode, [ ( ) ] ω n Γ n =. (2) ω n ɛ γ We validate our predictions numerically in random networks of linear springs (Fig. ), which are widely studied as minimal models of, e.g., polymer networks, foams, and glasses [5 23]. We show that the sign of the Poynting effect in spring networks is negative and set by the Grüneisen parameter, which can be motivated theoretically. We focus on marginally rigid spring networks close to the isostatic state (mean coordination z = z c + z, with z c 4 in 2D), and study scaling with z. Spring networks. For concreteness, we first illustrate the Poynting effect in random spring networks. We consider networks of N = 24 harmonic springs in a periodic unit cell with initial side lengths L = L 2 = L.

2 3 2 2 3.85 2 z G K χ ɛ Kχ p Eq. 6. 2 z FIG. 2: The shear modulus G, bulk modulus K, and Poynting coefficients χ ɛ and χ p as a function of excess coordination z for (a) packing derived and (b) randomly cut spring networks. Networks are prepared in two ways. Packing derived () networks are prepared by generating bidisperse packings of soft repulsive disks close to the jamming transition [24 26]. Each contact between disks is then replaced by a spring with stiffness k and a rest length l ij equal to its initial length l ij, so p = [4, 23, 27 29]. To prepare randomly cut () networks, we start from a network with mean coordination z 6 and randomly remove springs, with a bias towards highly connected nodes [7, 9, 2]. All numerical results are presented in dimensionless units by setting k and the average disk size in the initial packing to unity. and networks are indistinguishable by eye, but their elastic moduli are qualitatively different (Fig. 2, open symbols). In networks, the shear modulus G z µ vanishes continuously with the distance to the isostatic point with µ =, while the bulk modulus K jumps discontinuously to zero at z c [3]. In contrast, in networks both G and K vanish continuously with z µ, with µ. [7, 9, 2]. We consider deformations combining pure shear strain and volumetric expansion, such that lattice vectors of the unit cell are transformed by the deformation gradient [ ] + γ F = ( + ɛ) ( + γ). (3) The corresponding Cauchy stress tensor is [ ] p q σ =. (4) p + q When networks are sheared using strain control, ɛ is held fixed at zero while γ is increased incrementally. At each step the elastic energy ij V ij = (k/2) ij (l ij l ij )2 is minimized with respect to the node positions using FIRE [3]. The resulting p and q are determined from σ αβ = /(L L 2 ) ij f ij l ij (ˆn ij,β ˆn ij,β ), where f ij = V ij / l ij and ˆn ij is a unit vector pointing from node i to j. For stress controlled simulations, γ and ɛ are also allowed to vary while the energy is minimized subject to p = and a prescribed q [32]. Because finite-sized systems are never perfectly isotropic, plots of p or ɛ versus γ contain a linear contribution with a prefactor that vanishes as N [33]. To estimate the Poynting coefficients, we symmetrize p and ɛ by shearing both forward (γ > ) and backward (γ < ). Fig. 2 presents our first main result, the Poynting coefficients for and networks over a range of z. In all cases the Poynting effect is negative. There is an apparent equality between χ ɛ and Kχ p (motivated below), albeit with fluctuations at the lowest z. There is a notable difference in how the and Poynting coefficients scale with with z. In networks χ ɛ and Kχ p diverge, with an empirical fit to / z λ giving λ.85. In contrast, in networks χ ɛ and Kχ p are flat (λ = ). Hence the Poynting coefficients depend on both preparation and shearing protocols, and in three out of four cases they diverge at the isostatic point. Microscopic theory. We now develop exact expressions for the Poynting coefficients, beginning with the relation between χ ɛ and χ p. In a hyperelastic material, the pressure p = (/2)χ ɛ γ 2 due to shearing at fixed ɛ must be equal to the pressure from a two-step process: first shearing to γ at constant p, followed by a volumetric strain ɛ = (/2)χ p γ 2 that reverses the volume change induced in the first leg. The second step changes pressure by p = Kɛ = (/2)Kχ p γ 2, and therefore χ ɛ = Kχ p. We next relate χ ɛ to the shear modulus G(ɛ) = (/2)[( q/ γ) ɛ ] γ= after a volumetric strain. The total differential of the strain energy density is dw = S : de, where E = (F T F )/2 is the Green-Lagrange strain. The second Piola-Kirchoff stress S is related to the more experimentally-relevant Cauchy stress via σ = FSF T /J, where J = det F. Hence [ dw = 2( + ɛ) 2 p dɛ + ɛ + q dγ ]. (5) + γ Using the Maxwell relation of Eq. (5), one finds χ ɛ = 2G () 4G(), (6) where the prime indicates differentiation with respect to ɛ. Earlier work neglected the difference between the various stress and strain measures in nonlinear elasticity, but still arrived at the same result [6, 8]. Numerical evaluation of Eq. (6) is in good agreement with direct measurements of χ ɛ and Kχ p, as shown in Fig. 2. We now relate χ ɛ to discrete degrees of freedom. Network elasticity is encoded in the the extended Hessian H = 2 U/ q 2, where the the 2N +-component vector q contains the node positions and shear strain γ [34]. The

3 4 3 2 4 3 2 z 4. 4.2 4.4 4.8 z 4. 4.2 4.4 4.8 4.5 4.29 4.52 4.5 4.3 4.5 2 3 ω/g Γω 2 /K Γω 2 /K 2 2 ω FIG. 3: The product versus eigenfrequency ω in and networks at varying coordination z. D is the density of states and Λ 2 is a measure of modes coupling to shear. FIG. 4: The Grüneisen parameter Γ for and RD networks, scaled by the prediction of Eq. () and plotted for ω > ω. Symbols match the legends in Fig. 3 shear modulus can be written as a sum over the nonrigid body eigenmodes of H, /G = (v/n) n Λ2 n/ωn, 2 where v = JL 2 /N, ωn 2 is the squared eigenfrequency of the n th eigenvector, and Λ n /N is its component along the strain coordinate [34]. Letting D(ω), Λ 2 (ω) and Γ(ω) denote the density of states and averages of Λ 2 n and the Grüneisen parameter Γ n in the interval [ω, ω + dω), and replacing sums with integrals, we find and, from Eq. (6), χ ɛ = 2vG 2 G = v dω, (7) ω2 [ Γ ω 2 2 ln ] Λ2 D Λ 2 D dω. (8) ln ω Eq. (8) is a central result: it explicitly relates the Poynting effect to vibrational modes. Note that the sign of χ ɛ is controlled by Γ and the logarithmic derivative of Λ 2 D. Application to networks. We now evaluate Eq. (8) in the context of spring networks, focusing on the scaling of χ ɛ with z. Close to the isostatic state, both and networks display an anomalous abundance of soft modes that dominate the response to forcing [35, 36]. The modes appear above a characteristic frequency ω, and for scaling analysis the density of states is well approximated by a window function between ω and ω O(k /2 ) [35, 36]. Following Ref. [34], we assume that all soft modes couple similarly to shear, so Λ 2 const. Hence Eqs. (7) and (8) give ω G and χ ɛ G 2 ω ω Γ dω. (9) ω2 The sign and form of Γ can be rationalized with scaling arguments. Perturbing a network along mode n carries an energetic cost U ωn, 2 so Γ ω 2 ( U/ ɛ). U can be expanded in powers of the relative normal and transverse motions, u ij and u ij respectively, between connected nodes. The well-known result is U = (/2) ij [k(u ij )2 (f ij /l ij )(u ij )2 ], where the force f ij and length l ij are evaluated prior to the perturbation [37]. In a network that has previously undergone a small volumetric strain ɛ from its unstressed state, the typical force will be proportional to the pressure p = Kɛ, and so U/ ɛ K(u ) 2 N. The typical transverse motion of a mode can be estimated from the trial modes of Ref. [36], which give u /N /2 (much larger than the normal motion u ωu when ω is small). Therefore and, by Eq. (9), Γ K/ω 2, () χ ɛ K/G. () This remarkably simple expression for χ ɛ correctly predicts the sign of the Poynting effect and captures all of the phenomenology in Fig. 2. It relates the qualitatively different behavior of χ ɛ in and networks to the differences in their shear and bulk moduli, predicting λ = and λ = µ =. On a qualitative level, it explains that the Poynting effect in spring networks is negative because tension is stabilizing. Finally, the strength of the Poynting effect grows near isostaticity because tension couples to transverse motions, which dominate soft modes and cause strong non-affine fluctuations [29, 36]. The above scaling arguments rely on two essential approximations, namely that const and Γ K/ω 2 above ω G. We now validate them by direct numerical evaluation. In Fig. 3, is plotted as a function of ω/g for both and networks. As expected, in

4 q/ (Gγ) 2p/ ( χ ɛ γ 2) 3 γχ ɛ / GK 3 γχ ɛ / GK FIG. 5: Master curves for shear stress q and pressure p of and networks sheared to finite strain γ at fixed ɛ =. both cases there is a broad plateau above ω. In Fig. 4 we plot the ratio of Γ to K/ω 2 ; Γ is estimated from a linear fit of ω n versus ɛ after a series of small volumetric strain steps. In networks the ratio approaches a positive constant as z, indicating that Eq. () becomes increasingly accurate as the isostatic point is approached. At finite z there is a slow upturn with increasing ω. We attribute this to a subdominant correction to scaling, consistent with the observation that a power law fit to χ ɛ and Kχ p in networks gives a somewhat smaller value of λ than. The same ratio has a more complex form in networks, including a sign change for the lowest z, but it also approaches a low frequency plateau in the isostatic limit. Finite strain. The Poynting coefficients quantify the leading order dependence of p and ɛ on γ. We now show normal stress effects lead to strain stiffening when a network is sheared to finite strain at fixed volume. There has been no prior study of networks at finite strain, while studies of networks did not report normal stresses. shear stresses were shown to stiffen beyond some vanishing strain scale γ [7] (unlike sphere packings, which soften [38, 39]). The secant modulus q/γ in networks satisfies q/(gγ) = Q(γ/γ ), with Q for x and Q x θ with θ > for x [7]. It is natural to make a similar ansatz for the pressure, 2p χ ɛ γ 2 = P(γ/γ ), (2) where P for x and P x φ for x. The scaling functions Q and P are plotted in Fig. 5. In Ref. [7] it was argued that γ G, in agreement with our data for networks but not networks. However, we find that we can collapse data from both network types using { GK γ z λ +µ /2 for networks χ ɛ z λ+µ for networks. (3) In order for shear stress and pressure to remain finite when z, we must have θ = µ /(λ + µ /2).74 and φ = λ /(λ +µ /2).63 (using λ =.85), as well as θ = µ /(λ +µ ). and φ = λ /(λ + µ ). These are all in good agreement with numerics (see Fig. 5). The strain γ can be motivated by assuming that normal stress stiffens the material. The secant modulus can then be expanded for small strains as q 2Gγ + [ q/ p] p(γ) + χ2 ɛγ 2 2Gγ GK. (4) In the last step we have used Eq. (6) and dropped numerical prefactors. Balancing terms gives Eq. (3); it represents the extrapolated scale where leading order scaling ceases to dominate. The data collapse in Fig. 5 provides a posteori confirmation of our assumption that normal stresses induce stiffening. Conclusion We have derived exact expressions for the Poynting coefficients in hyperelastic materials, and validated them numerically in packing derived and randomly cut spring networks. Spring networks display a negative Poynting effect, whose origin can be traced to the stabilizing influence of tension on a network s vibrational modes. Microscopically, the amplitude of the effect is controlled by the coupling between tension and relative transverse motions, which explains the correlation between normal stress and non-affinity [3], and results macroscopically in a coefficient χ ɛ that scales with the ratio K/G. Eq. (8) is applicable in any 2D hyperelastic material hence our results can lend insight to the Poynting effect in other elastic networks, including fiber networks (e.g. [2, 3, 4 43]). The scaling arguments for and Γ presented here are specific to spring networks; they must be modeled or evaluated anew for each material. While our calculations and numerics are all in 2D, extension to 3D is straightforward, and we do not expect the underlying physics to change. We have shown that Poynting coefficients and stiffening behavior are highly sensitive to the linear elastic moduli G and K. Recent work has demonstrated how to prepare spring networks using a biased cutting protocol to target essentially any positive value of K/G [44, 45]. Our results suggest that the same techniques could allow to select for desirable nonlinear mechanical properties. Acknowledgments. We acknowledge financial support from the Netherlands Organization for Scientific Research (NWO) and the use of supercomputer facilities provided by NWO Physical Sciences. [] O. Reynolds, in Proc. Brit. Assoc. (885), p. 896. [2] J. Ren, J. A. Dijksman, and R. P. Behringer, Phys. Rev. Lett., 832 (23). [3] W. Thomson, in Encyclopaedia Britannica (Adam and Charles Black, Edinburgh, 878).

5 [4] J. Poynting, Proceedings of the Royal Society of London 82, 546 (99). [5] P. A. Janmey, M. E. McCormick, S. Rammensee, J. L. Leight, P. C. Georges, and F. C. MacKintosh, Nature Materials 6, 48 (27). [6] D. Weaire and S. Hutzler, Philosoph. Mag. 83, 2747 (23). [7] A. R. Cioroianu and C. Storm, Phys. Rev. E 88, 526 (23). [8] B. P. Tighe, Granular Matter 6, 23 (24). [9] F. Meng and E. M. Terentjev, Soft Matter 2, 6749 (26). [] H. C. de Cagny, B. E. Vos, M. Vahabi, N. A. Kurniawan, M. Doi, G. H. Koenderink, F. C. MacKintosh, and D. Bonn, Phys. Rev. Lett. 7, 2782 (26). [] C. Horgan and J. Murphy, Soft Matter 3, 496 (27). [2] C. Heussinger, B. Schaefer, and E. Frey, Phys. Rev. E 76, 396 (27). [3] E. Conti and F. C. MacKintosh, Phys. Rev. Lett. 2, 882 (29). [4] N. Xu, V. Vitelli, A. J. Liu, and S. R. Nagel, EPL 9, 56 (2). [5] S. Feng, M. Thorpe, and E. Garboczi, Physical Review B 3, 276 (985). [6] D. J. Jacobs and M. F. Thorpe, Phys. Rev. E 53, 3682 (996). [7] M. Wyart, H. Liang, A. Kabla, and L. Mahadevan, Phys. Rev. Lett., 255 (28). [8] B. P. Tighe and J. E. S. Socolar, Phys. Rev. E 77, 333 (28). [9] W. G. Ellenbroek, Z. Zeravcic, W. van Saarloos, and M. van Hecke, EPL 87, 344 (29). [2] B. P. Tighe, Phys. Rev. Lett. 9, 6833 (22). [2] M. Sheinman, C. P. Broedersz, and F. C. MacKintosh, Phys. Rev. E 85, 28 (22). [22] E. Lerner, E. DeGiuli, G. Düring, and M. Wyart, Soft Matter, 585 (24). [23] C. Buss, C. Heussinger, and O. Hallatschek, Soft Matter 2, 7682 (26). [24] C. S. O Hern, L. E. Silbert, A. J. Liu, and S. R. Nagel, Phys. Rev. E 7, 4332 (24). [25] M. van Hecke, J. Phys. Cond. Matt. 22, 33 (2). [26] D. J. Koeze, D. Vågberg, B. B. Tjoa, and B. P. Tighe, EPL 3, 54 (26). [27] H. Mizuno, L. E. Silbert, and M. Sperl, Phys. Rev. Lett. 6, 6832 (26). [28] K. Baumgarten, D. Vågberg, and B. P. Tighe, Phys. Rev. Lett. 8, 98 (27). [29] K. Baumgarten and B. P. Tighe, Soft Matter 3, 936 (27). [3] C. S. O Hern, L. E. Silbert, A. J. Liu, and S. R. Nagel, Phys. Rev. E 68, 36 (23). [3] E. Bitzek, P. Koskinen, F. Gähler, M. Moseler, and P. Gumbsch, Phys. Rev. Lett. 97, 72 (26). [32] S. Dagois-Bohy, B. P. Tighe, J. Simon, S. Henkes, and M. van Hecke, Phys. Rev. Lett. 9, 9573 (22). [33] C. P. Goodrich, S. Dagois-Bohy, B. P. Tighe, M. van Hecke, A. J. Liu, and S. R. Nagel, Phys. Rev. E 9, 2238 (24). [34] B. P. Tighe, Phys. Rev. Lett. 7, 5833 (2). [35] L. E. Silbert, A. J. Liu, and S. R. Nagel, Phys. Rev. Lett. 95, 983 (25). [36] M. Wyart, L. E. Silbert, S. R. Nagel, and T. A. Witten, Phys. Rev. E 72, 536 (25). [37] S. Alexander, Phys. Rep 296, 65 (998). [38] J. Boschan, D. Vågberg, E. Somfai, and B. P. Tighe, Soft Matter 2, 545 (26). [39] S. Dagois-Bohy, E. Somfai, B. P. Tighe, and M. van Hecke, Soft Matter (DOI:.39/C7SM846K). [4] J. A. Åström, J. P. Mäkinen, M. J. Alava, and J. Timonen, Phys. Rev. E 6, 555 (2). [4] J. Wilhelm and E. Frey, Phys. Rev. Lett. 9, 83 (23). [42] D. A. Head, A. J. Levine, and F. MacKintosh, Physical review letters 9, 82 (23). [43] M. Das, D. Quint, and J. Schwarz, PloS One 7, e35939 (22). [44] C. P. Goodrich, A. J. Liu, and S. R. Nagel, Phys. Rev. Lett. 4, 2255 (25). [45] D. R. Reid, N. Pashine, J. M. Wozniak, H. M. Jaeger, A. J. Liu, S. R. Nagel, and J. J. de Pablo, arxiv:7.2493 (27).