Phonon gap and localization lengths in floppy materials

Gustavo Düring , Edan Lerner and Matthieu Wyart *
New York University, Center of Soft Matter Research, 4 Washington Place, New York, NY 10003, USA. E-mail: mw135@nyu.edu

Received 16th April 2012 , Accepted 14th September 2012

First published on 15th October 2012


Abstract

Gels of semi-flexible polymers, network glasses made of low valence elements, softly compressed ellipsoid particles and dense suspensions under flow are examples of floppy materials. These systems present collective motions with almost no restoring force. To study theoretically and numerically the frequency-dependence of the response of these materials and the length scales that characterize their elasticity, we use a model of isotropic floppy elastic networks. We show that such networks present a phonon gap for frequencies smaller than a frequency ω* governed by coordination, and that the elastic response is characterized, and in some cases localized, on a length scale ugraphic, filename = c2sm25878a-t5.gif that diverges as the phonon gap vanishes (with a logarithmic correction in the two dimensional case). lc also characterizes velocity correlations under shear, whereas another length scale l* ∼ 1/ω* controls the effect of pinning boundaries on elasticity. We discuss the implications of our findings for suspension flows, and the correspondence between floppy materials and amorphous solids near unjamming, where lc and l* have also been identified but where their roles are not fully understood.


1 Introduction

In 1864 Maxwell1 showed that, in order to be mechanically stable, the average coordination value z of structures made of points connected by rigid bars must be larger than a threshold value zc. Most common solids satisfy a microscopic version of this constraint.2,3 If the constraint is violated, collective modes with no restoring forces (floppy modes) exist. These materials will be referred to as strictly floppy systems. However, in some cases these modes are stabilized by weak interactions: the bending energy of semi-flexible polymers confers a finite elasticity to gels,4 as do van der Waals interactions in weakly coordinated covalent glasses.5 Floppy modes can also be stabilized by the pre-stress applied on the system, as is the case for gently compressed packings of ellipsoid particles.6–9 Fluids, on the other hand, can display modes that are strictly floppy. For example, in granular systems or suspensions of hard particles, large and sometimes percolating clusters of connected particles can be formed,10,11 and motion within these clusters can only occur along floppy modes where no particles overlap. As the density increases toward jamming, the viscosity12–14 and the length scale15–17 characterizing the correlation of the dynamics diverge, up to the point where floppy modes disappear and the dynamics stops.

In simplified numerical models of ellipsoid particles,7,8 covalent networks,18 gels of semi-flexible polymers19 and suspension flows,20 it has been observed that the density of vibrational modes D(ω) displays a gap at low frequency (in addition to the floppy or nearly floppy modes present at zero-frequency). In suspension flows the amplitude of the gap was shown to affect the divergence of the viscosity near jamming.20 An early work by Garboczi and Thorpe21 supported the idea that a gap of modes exists in floppy materials. However, the dependence of the gap on the microscopic structure and its consequences on the material properties and the different length scales characterizing the elastic response are not understood.

In this manuscript, we study the elastic properties of isotropic harmonic spring networks that are strongly disordered, but where spatial fluctuations of coordination are weak. The model networks we use are presented and analyzed numerically in Section 2. In Section 3, we show, using both mean-field methods21–25 and numerical simulations, that strictly floppy elastic networks present a vibrational gap between floppy modes (zero frequency modes) and a frequency ω*zcz ≡ δz. The absence of low frequency phonons suppresses the elastic propagation of the response at frequencies smaller than ω*. As a consequence, we show that the mean response to a local perturbation (local strain) is localized, displaying an exponential decay with a characteristic length ugraphic, filename = c2sm25878a-t6.gif. We also predict a logarithmic correction to the scaling for two dimensional systems. In Section 4 we go beyond the mean field and estimate the fluctuations around the mean response. We show that fluctuations dominate the amplitude over the mean value, however, the fluctuation around the mean response also decays with the same length scale ∼lc. Even though lc can be considered to be the localization length of floppy modes, we show that these modes cannot exist in a region of typical radius smaller than l* ∼ 1/δz. Surprisingly, floppy modes can be strictly localized (have a compact support) only on the much larger scale l*. In Section 5 we study the floppy networks stabilized by weak interactions, which confer a finite but small restoring force to the material. We model the weak interaction by adding springs whose stiffnesses are very small. We find that localization is lost even when weak interactions have a vanishingly small amplitude, showing that the response of strictly floppy systems is a singular limit. However, lc still characterizes the response near the applied strain. In Section 6 we compare our results with recent observations in the affine solvent model, a simplified model for suspension flows of hard spheres. For this strictly floppy system, the spectrum associated with the evolution operator shares both commonalities and striking differences with the density of states of the isotropic floppy networks considered here, leading to a prediction on some dynamical length scale in flow. Finally, our results raise questions associated with the respective role of lc and l* both in floppy materials and in jammed packings, which are discussed in the last section.

2 Model description and simulation

We consider a network of N point particles of mass m, connected by Nc harmonic un-stretched springs in spatial dimension d. The elastic energy following a deformation vector field {δRi}i=1…N is:
 
ugraphic, filename = c2sm25878a-t7.gif(1)
where the index 〈ij〉 runs over the Nc springs, whose stiffnesses are kij and whose directions are along the unit vectors nij. Floppy modes correspond to displacements for which δE = 0. Since the energy is the sum of Nc positive definite terms, it vanishes only if all these terms are zero. Such modes always exist if the number of degrees of freedom Nd is larger than the number of constraints Nc, or equivalently if z ≡ 2Nc/N < zc = 2d.

The force field ∣F〉 generated by displacements can be obtained by taking the derivative of eqn (1). One obtains ∣F〉 = [scr M, script letter M]∣δR〉, where:

 
ugraphic, filename = c2sm25878a-t8.gif(2)
is the stiffness matrix. We have used the bra-ket notation for which 〈i∣δR〉 = δRi. The normal modes are given by the eigenvectors of [scr M, script letter M], and their frequencies are given by the square root of their associated eigenvalues. Herein, we will investigate the spectrum of [scr M, script letter M] and the spatial properties of its eigenvectors when z < zc.

As a model system we consider isotropic disordered elastic networks, generated by following the method of ref. 26: we prepare amorphous packings of compressed soft elastic particles with a coordination value significantly larger than zc. The centers of the particles are the nodes of our networks, and the contacts between particles are replaced by un-stretched springs of stiffness k. Springs are then removed until the desired coordination z is reached. Removal takes place randomly from the set of most connected pairs of nodes, leading to isotropic networks with low heterogeneity in density and coordination. Such networks are appropriate models of amorphous solids for which large spatial heterogeneities are not energetically favorable. In rigidity percolation,21,23 spring removal is completely random, resulting in networks with large fluctuations that affect the elastic properties.

For these networks, we diagonalize [scr M, script letter M] numerically to compute the density of states D(ω) for various z < zc. As shown in the inset of Fig. 1, we find that a gap in the vibrational spectrum appears below some frequency ω* that decreases as zzc. If spatial fluctuations of coordination were large, one would expect this gap to be filled-up in the thermodynamic limit due to Griffiths-like singularities, but we do not observe this effect in the network considered here. The role of this gap in elasticity can be analyzed by considering the response to a local strain. We change the rest length of one spring and let the system relax to zero energy under over-damped dynamics. Such a response can also be obtained in the absence of damping, by imposing a local oscillatory strain at a vanishing small frequency. As shown in Fig. 2, the elastic information does not propagate: the response is localized on some length scale that appears to diverge as the gap vanishes, i.e. as zzc. The response appears to be very heterogeneous. In what follows we will investigate its mean and its fluctuations.


Rescaled density of states vs. rescaled frequency for z ∈ [3.2, 3.95] using N = 10 000 nodes in two dimensions. The continuous line corresponds to the theoretical prediction of eqn (9). Inset: non-rescaled density of states D(ω) vs. ω. Floppy modes lead to a delta function at ω = 0 and are not presented.
Fig. 1 Rescaled density of states vs. rescaled frequency for z ∈ [3.2, 3.95] using N = 10[thin space (1/6-em)]000 nodes in two dimensions. The continuous line corresponds to the theoretical prediction of eqn (9). Inset: non-rescaled density of states D(ω) vs. ω. Floppy modes lead to a delta function at ω = 0 and are not presented.

Displacement field caused by a local strain in a floppy network. An over-damped relaxation was used following the elongation of one spring for (a) δz = 0.2 and (b) δz = 0.05.
Fig. 2 Displacement field caused by a local strain in a floppy network. An over-damped relaxation was used following the elongation of one spring for (a) δz = 0.2 and (b) δz = 0.05.

3 Effective medium theory (EMT)

3.1 General formalism

We use EMT, also known as coherent potential approximation,22,23,27 to investigate the behavior of floppy networks as the one observed in the previous section. EMT is a mean field approximation and neglects spatial fluctuations in the coordination. Therefore, it does not properly describe rigidity percolation.23 However, it was shown to describe materials for which spatial fluctuations of coordination are small.24 EMT attempts to describe disordered materials as ordered materials with a frequency-dependent effective stiffness [k with combining tilde]eff. We apply this technique to isotropic lattices of coordination zinzc, where bonds are then randomly removed with a probability (1 − p), so that the stiffness coefficients kij take the values 0 or k with a probability (1 − p) or p respectively, and the final coordination is z = pzin. The Green's function G(ω) of the disordered floppy system is defined as [−2 + [scr M, script letter M]]G(ω) = −1. G0 is the Green's function of the effective medium, corresponding to the initial ordered lattice with an undetermined effective stiffness coefficient [k with combining tilde]eff. Standard calculations lead to the Dyson relationship G = G0 + G0[scr T, script letter T]G0,28 where the operator [scr T, script letter T] can be expressed as an infinite series in terms of increasing numbers of interacting contacts:
ugraphic, filename = c2sm25878a-t9.gif

The transfer matrix is found to be:

 
ugraphic, filename = c2sm25878a-t10.gif(3)

We seek an effective stiffness [k with combining tilde]eff that captures the average behavior of the system, i.e.G〉 = G0, where the average is taken over the disorder on the stiffness coefficients kij. This condition leads to 〈[scr T, script letter T]〉 = 0. In the EMT this constraint is approximated by 〈Tij〉 = 0. Using standard identities for the Green's function on isotropic lattices,27 one can express this condition as:23

 
ugraphic, filename = c2sm25878a-t11.gif(4)
where tr[·] stands for the trace, and keff[k with combining tilde]eff/k.

Since we are interested in low frequencies ugraphic, filename = c2sm25878a-t12.gif, we approximate G0 by its continuum limit and use a Debye cut-off qD. We introduce the bulk modulus K and the shear modulus μ of the ordered lattice without an effective stiffness, and use the approximation:

 
ugraphic, filename = c2sm25878a-t13.gif(5)

The sum is taken over all polarization vectors qα, where ugraphic, filename = c2sm25878a-t14.gif for pressure waves and ugraphic, filename = c2sm25878a-t15.gif for shear waves.

3.2 Scaling analysis near z = zc

We now perform a scaling analysis of eqn (4) and (5) as zzc from below. Eqn (5) implies that ugraphic, filename = c2sm25878a-t16.gif, where the function f is independent of z and ω. As ω → 0, the elastic moduli of floppy systems vanish and keff → 0. Eqn (4) then leads to ugraphic, filename = c2sm25878a-t17.gif, where ε satisfies the equation
 
εf(ε) = δz/2.(6)

For spatial dimensions d ≥ 3, eqn (5) implies that f(0) is a positive constant, therefore in the regime δz ≪ 1 one finds

ε ≈ δz/(2f(0)).

Using eqn (4) together with the assumption that ∣mω2/keff∣ ≪ 1 and ∣keff∣ ≪ 1 (which can be shown to be true a posteriori in the limit of δz ≪ 1 and ugraphic, filename = c2sm25878a-t18.gif) one finds:

 
ugraphic, filename = c2sm25878a-t19.gif(7)

For d = 2, eqn (5) leads to a logarithmic divergence in the small ε limit:

 
ugraphic, filename = c2sm25878a-t20.gif(8)
where Γ is a constant that depends on the elastic moduli and the Debye cut-off. The weak divergence of f does not modify the perturbation analysis performed for d ≥ 3, except that f(0) must now be replaced by f(ε) in eqn (7). Only f(ε) remains undetermined.

The asymptotic value for ε in the limit δz → 0 can be obtained from eqn (6), which leads to ugraphic, filename = c2sm25878a-t21.gif and therefore f(ε) ∼ −log(δz). This behavior is only valid for values of |log(δz)| ≫ 1, a difficult limit to observe empirically. Therefore, to compare our theoretical predictions with two dimensional numerical observations we compute εz) by solving eqn (6) numerically using eqn (8). We find that the relationship εz) depends on the initial network via only one parameter Λ(cp,cs,Γ). To compare the theory and numerical simulations, we use Λ(cp,cs,Γ) as a fitting parameter.

The mechanical response of floppy systems is therefore fully determined by the Green's function in eqn (5) and the effective stiffness given by eqn (7). We start by computing the density of states D(ω) using the relationship ugraphic, filename = c2sm25878a-t22.gif. For 0 < ω < ω*, where ugraphic, filename = c2sm25878a-t23.gif, we find that D(ω) = 0. For ω > ω* we obtain:

 
ugraphic, filename = c2sm25878a-t24.gif(9)

Thus the size of the gap scales linearly with δz for d ≥ 3, where f(ε) ≈ f(0), whereas for d = 2, a logarithmic correction exists. According to eqn (9), rescaling the frequency by ω* and D(ω) by f(ε)1/2 should collapse the low-frequency part of the spectrum. Fig. 1 shows that the quality of the collapse is very good. For all considered coordinations, we used the same fitting parameter Λ(cp,cs,Γ) = 1.3, which is fixed by this measurement.

3.3 Linear response to a local strain

The response to a local force G0(ω,rij) diverges as ω → 0 due to the presence of floppy modes. One must rather consider the response field due to an imposed displacement. Changing the rest length of a spring placed between nodes i and k at a fixed frequency ω0 corresponds to imposing a displacement (δRi − δRkn = e0t, with n being the unit vector along the connecting bond. The force required to impose such a displacement is
ugraphic, filename = c2sm25878a-t25.gif

In the small frequency regime, the last expression is found to be |F〉 ∼ keffe0tn(|i〉 − |k〉). The magnitude of the force vanishes as ω0 → 0, which is consistent with the existence of floppy modes.

The response at the zero frequency limit in two dimensions, at distances r = |r| larger than the typical springs length (i.e. rqD ≫ 1), can be calculated using the continuum limit

 
〈δR(r)〉 = g(rri) − g(rrk),(10)
where
ugraphic, filename = c2sm25878a-t26.gif
and K0 is the modified Bessel function of the second kind, that behaves exponentially at long distances. Thus, the mean response to a local strain decays exponentially in floppy systems with a characteristic length ugraphic, filename = c2sm25878a-t27.gif. The mean perturbation induced by changing the rest length of a spring has a quadrupolar symmetry shown in Fig. 3a. After averaging over 6000 realizations, we obtain a good agreement with our theoretical prediction, shown in Fig. 3b.


Mean response following the elongation of a spring placed at the center, along the vertical axis. (a) Theoretical prediction given by eqn (10) using cs and cp of a hexagonal lattice. Arrows show the displacement field and continuous lines are stream lines of the vector field. (b) Average response over 6000 independent realizations of networks with N = 40 000 nodes and δz = 0.05.
Fig. 3 Mean response following the elongation of a spring placed at the center, along the vertical axis. (a) Theoretical prediction given by eqn (10) using cs and cp of a hexagonal lattice. Arrows show the displacement field and continuous lines are stream lines of the vector field. (b) Average response over 6000 independent realizations of networks with N = 40[thin space (1/6-em)]000 nodes and δz = 0.05.

The asymptotic solution in any dimension has an exponential decay. Indeed, taking the angular average one gets

 
ugraphic, filename = c2sm25878a-t28.gif(11)
where the over-line stands for angular average. In two dimensional networks, rescaling the distance by lc and the amplitude by ugraphic, filename = c2sm25878a-t29.gif collapses the response to a local strain at different coordination values into a single curve, as shown in Fig. 4a and b.


(a) Mean response δRmvs. the distance r from the imposed strain with N = 40 000 nodes. (b) Rescaled average displacement of the mean response δRmvs. the rescaled distance r/lc, using Λ(cp,cs,Γ) = 1.3 as extracted from Fig. (1). (c) The fluctuations around the mean response are characterized by δRt, defined in eqn (16). δRt is plotted vs. the distance from a local strain r. (d) δRtvs. the rescaled distance r/lc, using Λ(cp,cs,Γ) = 1.3.
Fig. 4 (a) Mean response δRmvs. the distance r from the imposed strain with N = 40[thin space (1/6-em)]000 nodes. (b) Rescaled average displacement of the mean response δRmvs. the rescaled distance r/lc, using Λ(cp,cs,Γ) = 1.3 as extracted from Fig. (1). (c) The fluctuations around the mean response are characterized by δRt, defined in eqn (16). δRt is plotted vs. the distance from a local strain r. (d) δRtvs. the rescaled distance r/lc, using Λ(cp,cs,Γ) = 1.3.

The zero frequency limit can be extended to finite frequencies ugraphic, filename = c2sm25878a-t30.gif by replacing ε → −2/keff in eqn (10). The asymptotic behavior is then given by log(〈δR(r)〉) ∼ −r/lc(ω) + iωr/v(ω), where

ugraphic, filename = c2sm25878a-t31.gif

From eqn (7) one can prove the existence of two regimes. (i) For ω < ω*, the imaginary part 1/v(ω) = 0: one finds a pure exponential decay, with a characteristic length of order lc. (ii) For ωω*, the decay length decreases as lc(ω) ∼ f(ε)1/4ω−1/2, whereas the velocity of the corresponding vibrations grows as v(ω) ∼ f(ε)1/4ω1/2. A similar behavior above ω* has been predicted by one of us for zzc.24

We can also consider the case of our networks immersed in a viscous fluid with viscosity η0 and without hydrodynamic interactions. The response in the over-damped regime is calculated by replacing the inertial term 2 by the viscous force −0ω in eqn (5) and then in the effective stiffness in eqn (7). In this limit the elastic modulus is proportional to keff, whose real and imaginary parts give the storage and loss moduli G′ and G′′ respectively. Two different regimes are observed. (i) For ωω*2m/η0, the storage moduli follow G′ ∼ lc6ω2/f(ε) and the loss moduli G′′ ∼ lc2ω. The response is given by lc(ω) ∼ lc and v(ω) ∼ f(ε)lc−3, displaying a similar exponential decay to that observed in an undamped system, but with propagating waves. (ii) For ωω*2m/η0, both the storage and loss moduli are proportional to ugraphic, filename = c2sm25878a-t32.gif, as also observed in ref. 29. The decaying length lc(ω) ∼ f(ε)1/4ω1/4, while the velocity v(ω) ∼ f(ε)1/4ω3/4. Our scaling for the characteristic length scale lc(ω) differs from the recent result of Tighe.29 In our opinion, the difference stems from an incorrect definition of the length scale characterizing the elastic response in floppy materials (see footnote).

4 Beyond the mean field

4.1 Fluctuations

The obvious difference between the mean response to a local strain (Fig. 3) and a typical one (Fig. 2) indicates large fluctuations. A priori mean field models as the effective medium do not enable to capture those. However, it is possible to combine the EM results with additional considerations to estimate the amplitude of these fluctuations.

We denote the dipole of forces ∣Fij〉 generated by changing the rest length of the spring ij by a distance one:

 
|Fij〉 = nij(|i〉 − |j〉),(12)
where the stiffness coefficient k is set to unity. From the definition of the stiffness matrix, eqn (2) and (12), we can write:
 
ugraphic, filename = c2sm25878a-t33.gif(13)
where the sum is taken over all the bonds. Note that any floppy mode |δR0〉 has by definition no restoring force [scr M, script letter M]R0〉 = 0, implying that 〈δR0|Fij〉 = 0 for all contacts ij.

Thus, the response |δRij〉 to the elongation of a spring ij, which is equivalent to the response to a force dipole |Fij〉, has no components along floppy modes. Therefore the equation [scr M, script letter M]Rij〉 = [thin space (1/6-em)]|Fij〉 can be inverted. Using the spectral decomposition of [scr M, script letter M], one gets:

 
ugraphic, filename = c2sm25878a-t34.gif(14)
where ω2 and ∣δRω〉 are the non-zero eigenvalues and the corresponding eigenvectors of the stiffness matrix [scr M, script letter M].20,26Eqn (14) implies that the norm of the response follows 〈δRijRij〉 = ∑ω>0〈δRω|Fij2/ω4. Introducing the average amplitude of the response ugraphic, filename = c2sm25878a-t35.gif, one gets using eqn (13):
 
ugraphic, filename = c2sm25878a-t36.gif(15)

For comparison, the total amplitude of the mean response can be calculated from eqn (11), and one obtains for the norm square ugraphic, filename = c2sm25878a-t37.gif. Thus for d ≥ 2, the norm of the mean response vanishes as δz → 0, whereas the norm of the fluctuations diverge. Thus relative fluctuations must diverge at small ε, as observed in our data.

The most simple scenario is that the fluctuations of the response 〈δR(r)2〉 decays with the same characteristic length ∝ lc characterizing the mean response. Making this assumption, which is numerically verified (see below), and using the result of eqn (15), the angular average must read:

 
ugraphic, filename = c2sm25878a-t38.gif(16)
where log(h(x)) ∼ −x for x ≫ 1, and the ε dependence is determined by eqn (15). Rescaling the distance by lc leads to a very good collapse of the response at different coordination values, as shown in Fig. 4c and d. Note that in two dimensions the amplitude does not need to be rescaled.

4.2 Pinning boundaries

We now turn to the spatial properties of floppy modes. The response to the stretch of a spring, exemplified in Fig. (2), is a floppy mode of the network where this spring is removed. We have thus shown that floppy modes can be localized (in the sense of presenting an exponential decay) on a length scale ugraphic, filename = c2sm25878a-t39.gif (with a log correction for d = 2). We now extend a previous counting argument,3,30 tested for packings with z > zc in ref. 31, to floppy networks (see also ref. 29) and show that floppy modes are also characterized by another length scale l* ∼ 1/δzlc. Below l*strict localization is impossible: floppy modes cannot have a smaller compact support. Essentially, l* is the length at which the number of constraints rd−1f that result from freezing the boundary of a system of size rf is equal to the number of floppy modes in the bulk δzrdf. This counting argument suggests that if the boundaries are frozen on a scale l* or smaller, floppy modes must vanish.

To test this prediction, we fix all the nodes outside a circle of radius rf. A frozen boundary induces a rigid region where floppy modes are forbidden. We determine this region using the pebble algorithm,32 as shown in Fig. 5a. As rf decreases, all floppy modes eventually vanish. For any z we can define n*, the number of nodes involved in the last floppy mode, and ugraphic, filename = c2sm25878a-t40.gif for d = 2. Our measurements are shown in Fig. 5b and follow the prediction l* ∼ 1/δz. This result implies that exponentially small displacement at distances r > lc cannot be neglected when the rigid–floppy transition induced by freezing boundary conditions is considered. On the other hand, our results imply that floppy networks that are stabilized by pinning boundaries on the scale l* present soft modes with exponentially small frequencies near the center of the sub-system, since there are floppy modes with displacements of very tiny amplitude near the boundaries, of order exp(−l*/lc) ∼ ugraphic, filename = c2sm25878a-t41.gif.


Rigid regions, shown as red diamonds and magenta square, are induced by freezing the nodes outside an external radius (red) and an internal radius (magenta) respectively. Circular nodes (blue) show the minimal floppy region. Freezing any extra particle rigidifies the entire system. (b) l*, as defined in the text, vs. δz. The red line corresponds to l* ∼ 1/δz.
Fig. 5 Rigid regions, shown as red diamonds and magenta square, are induced by freezing the nodes outside an external radius (red) and an internal radius (magenta) respectively. Circular nodes (blue) show the minimal floppy region. Freezing any extra particle rigidifies the entire system. (b) l*, as defined in the text, vs. δz. The red line corresponds to l* ∼ 1/δz.

5 Weak interactions

As discussed in the Introduction, in some materials, floppy modes are stabilized by weak interaction, as is the case in covalent glasses. In order to model these weak interactions, we consider (see for example ref. 26) floppy networks with k = 1 and add a number of weak springs of stiffness kweak*2, which gives a finite elasticity to the network. Floppy modes then gain finite frequencies, of order the characteristic frequency scale associated with the weak interaction ugraphic, filename = c2sm25878a-t42.gif. A gap in the density of states remain present in the frequency range [ωc,ω*], if the weak interaction is weak enough, i.e.ugraphic, filename = c2sm25878a-t43.gif. In this limit the elastic moduli scale as μweakkweakz,26 where δz is the excess coordination of the network of strong interaction. At long enough wavelengths the material, to good approximation, must behave as a continuous elastic medium, and for ωωc the density of states must follow a Debye behavior D(ω) ∼ ωd−1. Extracting a velocity of sound from the elastic moduli and computing the wavelength at ωωc, one obtains a wavelength of order lc, which thus characterizes as well the length scale above which a continuum description becomes a good approximation.

In Fig. 6a the response to a local strain for a system with weak springs with kweak = 10−7 is shown. The deformation field clearly differs from the response of the same network with kweak = 0 (Fig. 6b). We now argue that the limit kweak → 0 is singular: for any kweak > 0, the response decays as a power-law at large distances rlc, even in the limit kweak → 0. The singularity of this limit can be seen by decomposing the response to a local strain in two parts. First, we consider the displacement of characteristic amplitude δR that would follow such a strain in the absence of weak interactions, i.e. with kweak = 0 (Fig. 6b). Second, we consider the additional displacement induced by the presence of weak springs (Fig. 6c). Indeed due to the weak springs forces are not balanced after step one, and forces of order FweakkweakδR have appeared on the nodes. In the limit kweak → 0, these forces are vanishingly small and lead to no relaxation on the modes of non-vanishing frequency ω > ω*. However the spectrum now presents modes (stemming from the floppy modes that exist when kweak = 0) of characteristic frequency ωc. These modes relax due to the unbalanced weak forces with amplitude x that must satisfy 2cxkweakxFweakkweakδR, implying that x is independent of kweak. The convergence of the response δRt(r,kweak) in the small kweak limit is shown in Fig. 6d. We can define

ugraphic, filename = c2sm25878a-t44.gif
to quantify the difference between the response of the networks with weak springs and the ones strictly floppy. In the inset of Fig. 6d one can clearly see that Cweak converges to a non-zero value showing the singularity of this limit.


(a) Example of the response to a local strain for δz = 0.2 and kweak = 10−7. (b) Response to the same strain in the absence of weak spring (kweak = 0), as studied in the last sections. (c) Difference between the displacement field of (a) and (b). (d) Average displacement δRtvs. the distance R from a local strain for floppy networks (δz = 0.2) with weak springs, as indicated in legend, with N = 90 000 nodes. Inset: Cweakvs. kweak (see text for definition), where L0 has been taken to be 80 spring lengths.
Fig. 6 (a) Example of the response to a local strain for δz = 0.2 and kweak = 10−7. (b) Response to the same strain in the absence of weak spring (kweak = 0), as studied in the last sections. (c) Difference between the displacement field of (a) and (b). (d) Average displacement δRtvs. the distance R from a local strain for floppy networks (δz = 0.2) with weak springs, as indicated in legend, with N = 90[thin space (1/6-em)]000 nodes. Inset: Cweakvs. kweak (see text for definition), where L0 has been taken to be 80 spring lengths.

Although localization is lost, the coordination continues to influence the response to a local strain (see Fig. 7a). In particular, near the imposed strain for rlc, the response is weakly influenced by the weak interaction, and decays with the same characteristic length, as shown in Fig. 7b. Note that we expect this scenario to hold also for gently compressed, hypostatic soft ellipses.§


(a) Average displacement δRtvs. the distance from a local strain r for floppy networks plus weak springs (see text) with N = 90 000 nodes and kweak/k = 10−7. (b) Average displacement δRtvs. the rescaled distance r/lc, using Λ(cp,cs,Γ) = 1.3. The response of strictly floppy networks (filled symbols) has been included for comparison. Inset: the initial decay magnified.
Fig. 7 (a) Average displacement δRtvs. the distance from a local strain r for floppy networks plus weak springs (see text) with N = 90[thin space (1/6-em)]000 nodes and kweak/k = 10−7. (b) Average displacement δRtvs. the rescaled distance r/lc, using Λ(cp,cs,Γ) = 1.3. The response of strictly floppy networks (filled symbols) has been included for comparison. Inset: the initial decay magnified.

6 Rheology of dense suspensions

6.1 Random networks under a global shear

It is interesting to consider the behavior of the present networks under shear, if they were placed in a viscous solvent. This problem is formally related20 to the affine solvent model of suspension flow of hard particles, where hydrodynamic interactions between particles are neglected. We consider that our networks are made of point particles, subjected to a viscous drag F proportional to the particles velocities V in the reference frame of the solvent, which is assumed to follow an affine shear. One can show, from the definition of this model, that the viscosity η is proportional to the non-affine velocity squared of the particles.20,29,33 The viscosity is proportional to the ratio P/[small gamma, Greek, dot above]2, where P is the total power dissipated and [small gamma, Greek, dot above] is the strain rate. The power dissipated follows P ∼ 〈F·V〉 ∼ ‖V2. It is convenient to write the velocity field in terms of the non-affine displacement following an infinitesimal strain δRγ, i.e.V = [small gamma, Greek, dot above]δRγ. One can find that η ∼ 〈(δRγ)2〉, which relates the viscosity to the non-affine response to shear.

For isotropic floppy networks it was found numerically that ugraphic, filename = c2sm25878a-t45.gif, implying η ∼ 1/δz. Using our previous result on the response to a dipole in floppy materials and a simple hypothesis it is straightforward to derive this result, thus extending a previous derivation valid for z > zc,26 see also ref. 29. Here we perform this calculation for completeness, and because the present argument can be extended to predict that the correlation length under shear is lc. We consider an affine shear strain δγ applied on the network. After such a strain, unbalanced forces appear on the nodes: ∣Fδγ〉 = ∑ijγijFij〉 where the sum is taken over all the bonds and ∣Fij〉 correspond to a dipole of force as defined in eqn (12). The coefficients γij are equal to nij·Γ·nij where Γ is the strain tensor, which is linear in δγ at first order. The displacement field can be written as a linear combination of responses to local dipoles ∣δRδγ〉 = ∑ijγij∣δRij〉, where ∣δRij〉 is defined in eqn (14). Two-point displacement correlations in space obey

 
ugraphic, filename = c2sm25878a-t46.gif(17)

We now make the assumption that in random isotropic networks (as those considered here), the response of different dipoles is weakly correlated (this assumption turns out to be incorrect for flow of particles where subtle correlations are present in the structures visited by the dynamics). Using this assumption:

 
ugraphic, filename = c2sm25878a-t47.gif(18)
where we used γij2 ∼ δγ2. Combining with eqn (15) one finds
ugraphic, filename = c2sm25878a-t48.gif
implying that the viscosity diverges as ηCδγ(0)/δγ2ε−1.

Moreover, eqn (18) indicates that the correlation length in eqn (18) is essentially the length scale appearing in the response to a dipole lc. 〈δRij(x + r)·δRij(x)〉 must vanish when r is larger than the length lc where the response to a dipole is localized. On the other hand, correlations are not expected to vanish on a scale much smaller than lc, since the mean dipolar response will already give correlations on that scale:

ugraphic, filename = c2sm25878a-t49.gif

6.2 ASM model of suspension flows

We predict a plateau of modes above a frequency ω* ∼ δz. This plateau also appears in models of suspension flows.20 This observation supports that the plateau is generically present in floppy materials, and that the frequency response above ω* will be well described by the effective medium predictions derived here. There is, however, a crucial difference between the spectra of our networks and those of shear flows: in the latter one mode appears at a frequency ωminω* that strongly couples to a global shear and leads to a divergence of the viscosity η ∼ 1/δz2.85,20 which is sharper than in isotropic networks where η ∼ 1/δz. Therefore, typical configurations visited in dense flows differ qualitatively from our random networks, for reasons explained in ref. 34. Nevertheless, we expect some of our present predictions to hold in flow.

For instance, the counting argument comparing surface effects and bulk degrees of freedom should hold, and l* is expected to characterize the effect of freezing boundaries.

We do not know what the spatial correlations of the velocities should be in flows, because in this case the velocity is dominated by the lowest frequency mode that is not captured by the present analysis.20,34 However the response to a local disturbance (such as the formation of a new contact) must couple predominantly to the modes above ω*, and is thus expected to affect the flow on a length lc. This can be shown using spectral decomposition for the response under a dipole of force ∣Fij〉. Decomposing the spectrum in two parts (above and below ω*) eqn (14) reads

 
ugraphic, filename = c2sm25878a-t50.gif(19)

Numerical results show the appearance of a single mode below ω*,20 however a theoretical estimation in the large N limit supports that the distribution of modes between ωmin and ω* is given by Dmin(ω) ∝ ω(ω2ωmin2)(d−2)/2.34 Following the argument of Section 4.1, the average amplitude of the response is given by

ugraphic, filename = c2sm25878a-t51.gif

The relative contribution follows Imin/I* ∼ δz(d−1) (with a logarithmic correction in two dimensions), indicating that in the limit δz ≪ 1 the contribution of the modes below ω* becomes vanishingly small and can be neglected. The modes above ω* in flow have the same density of states and are expected to have the same properties than those of isotropic networks. Thus the response to a local disturbance in flow must decay exponentially with a typical length lc, as could be tested empirically in two dimensional granular flows where imaging is possible, or perhaps using confocal imaging in slow emulsion flows. This simple argument does not hold for the velocity correlation in flow: the modes in the plateau do not contribute significantly to the response to shear, which is rather dominated by the lowest frequency modes.

7 Discussion and open questions

We have argued that the elasticity of floppy networks is characterized by a gap in the vibrational spectrum, and by two length scales lc and l* that diverge near jamming (i.e. δz → 0). The existence of two lengths raises the question of which properties are governed by which scale. Our work supports that the length scale characterizing the response to imposed forces, and to most standard observations, is lc. On the other hand l* characterizes the effect of pinning boundaries. Such effects are subtle, and can depend on surprising ways on the type of elastic networks considered.35

For the strongly disordered floppy isotropic networks that we consider, one example of a question that remains to be explored is the evolution of elasticity (e.g. the shear modulus G) when boundaries are pinned at a distance L. For L > l*, the system remains floppy and G = 0. For Llc, one expects a mean-field argument to apply: pinning boundaries is equivalent to adding springs, leading to an increase of coordination Δz ∼ 1/L. For z > zc it is known that Gzzc,36 and thus we expect G ∼ 1/L. The behavior of G at intermediate length l*Llc remains to be explored.

Finally we compare our results with previous works in amorphous solids made of soft repulsive particles, for which z > zc. Near the unjamming transition where pressure vanishes the coordination approaches the Maxwell threshold from above (zzc). A vanishing frequency scale ω*zzc was predicted to characterize the low-frequency part of the spectrum,37,38 as confirmed numerically.39 The same theoretical argument indicated that boundaries affect elasticity on a length scale l* ∼ 1/(zzc). It was later argued,40 based on numerical observations, that l* characterizes the response to a point force in packings of particles. Another length scale lc was observed numerically to characterize the response at a frequency ω*39 and to affect transport,41 behaviors that were well-captured by effective medium.24 Our work extends these results to floppy materials with z < zc, where these lengths and frequency scales characterize the phonon gap and the localization of floppy modes. However the comparison underlines an important disparity: for floppy networks we find both numerically and theoretically that lc characterize the response to a local force, whereas l* appears to characterize the response to a point force in packings.40 More numerical and theoretical investigations are needed to understand this difference.

Acknowledgements

We thank Y. Elmatad, C. Falcón, P. Rowghanian, A. Grosberg and P. Chaikin for comments on the manuscript. This work has been supported by the Sloan Fellowship, NSF DMR-1105387, and Petroleum Research Fund #52031-DNI9. This work was also supported partially by the MRSEC Program of the National Science Foundation under Award Number DMR-0820341.

References

  1. J. Maxwell, Philos. Mag., 1864, 27, 294–299.
  2. S. Alexander, Phys. Rep., 1998, 296, 65–236 CrossRef CAS.
  3. M. Wyart, Ann. Phys., 2005, 30(3), 1.
  4. C. Heussinger and E. Frey, Phys. Rev. Lett., 2006, 97, 105501 CrossRef.
  5. J. C. Phillips and M. F. Thorpe, Solid State Commun., 1985, 53, 699–702 CrossRef CAS.
  6. A. Donev, I. Cisse, D. Sachs, E. A. Variano, F. H. Stillinger, R. Connelly, S. Torquato and P. M. Chaikin, Science, 2004, 303, 990–993 CrossRef CAS.
  7. M. Mailman, C. F. Schreck, C. S. O'Hern and B. Chakraborty, Phys. Rev. Lett., 2009, 102, 255501 CrossRef.
  8. Z. Zeravcic, N. Xu, A. J. Liu, S. R. Nagel and W. van Saarloos, Europhys. Lett., 2009, 87, 26001 CrossRef.
  9. C. F. Schreck, M. Mailman, B. Chakraborty and C. S. O'Hern, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2012, 85, 061305 Search PubMed.
  10. X. Cheng, J. H. McCoy, J. N. Israelachvili and I. Cohen, Science, 2011, 333, 1276–1279 CrossRef CAS.
  11. D. Bonamy, F. Daviaud, L. Laurent, M. Bonetti and J. P. Bouchaud, Phys. Rev. Lett., 2002, 89, 034301 Search PubMed.
  12. E. Brown and H. Jaeger, Phys. Rev. Lett., 2009, 103, 086001 CrossRef.
  13. K. N. Nordstrom, E. Verneuil, P. E. Arratia, A. Basu, Z. Zhang, A. G. Yodh, J. P. Gollub and D. J. Durian, Phys. Rev. Lett., 2010, 105, 175701 CrossRef CAS.
  14. F. Boyer, E. Guazzelli and O. Pouliquen, Phys. Rev. Lett., 2011, 107, 188301 CrossRef.
  15. O. Pouliquen, Phys. Rev. Lett., 2004, 93, 248001 CrossRef CAS.
  16. P. Olsson and S. Teitel, Phys. Rev. Lett., 2007, 99, 178001 CrossRef.
  17. R. Lespiat, S. Cohen-Addad and R. Höhler, Phys. Rev. Lett., 2011, 106, 148302 CrossRef.
  18. K. O. Trachenko, M. T. Dove, M. J. Harris and V. Heine, J. Phys.: Condens. Matter, 2000, 12, 8041–8064 Search PubMed.
  19. E. M. Huisman and T. C. Lubensky, Phys. Rev. Lett., 2011, 106, 088301 CrossRef CAS.
  20. E. Lerner, G. Düring and M. Wyart, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 4798–4803 Search PubMed.
  21. E. J. Garboczi and M. F. Thorpe, Phys. Rev. B: Condens. Matter Mater. Phys., 1985, 32, 4513 Search PubMed.
  22. X. Mao, N. Xu and T. C. Lubensky, Phys. Rev. Lett., 2010, 104, 085504 Search PubMed.
  23. S. Feng, M. F. Thorpe and E. Garboczi, Phys. Rev. B: Condens. Matter Mater. Phys., 1985, 31, 276–280 CrossRef CAS.
  24. M. Wyart, Europhys. Lett., 2010, 89, 64001 CrossRef.
  25. C. P. Broedersz, X. Mao, T. C. Lubensky and F. C. MacKintosh, Nat. Phys., 2011, 7, 983–988 CrossRef CAS.
  26. M. Wyart, H. Liang, A. Kabla and L. Mahadevan, Phys. Rev. Lett., 2008, 101, 215501 CrossRef CAS.
  27. S. Kirkpatrick, Rev. Mod. Phys., 1973, 45, 574–588 CrossRef.
  28. I. Webman, Phys. Rev. Lett., 1981, 47, 1496–1499 CrossRef CAS.
  29. B. P. Tighe, ArXiv e-prints, 2012, 1203.3411 Search PubMed.
  30. M. Wyart, S. Nagel and T. Witten, Europhys. Lett., 2005, 72, 486–492 CrossRef CAS.
  31. M. Mailman and B. Chakraborty, J. Stat. Mech.: Theory Exp., 2012, P05001 Search PubMed.
  32. D. J. Jacobs and B. Hendrickson, J. Comput. Phys., 1997, 137, 346–365 Search PubMed.
  33. B. Andreotti, J.-L. Barrat and C. Heussinger, ArXiv e-prints, 2011, 1112.1194 Search PubMed.
  34. E. Lerner, G. Düring and M. Wyart, ArXiv e-prints, 2011, 1201.3650 Search PubMed.
  35. C. F. Moukarzel, Europhys. Lett., 2012, 97, 36008 Search PubMed.
  36. A. J. Liu, S. R. Nagel, W. van Saarloos and M. Wyart, Dynamical Heterogeneities in Glasses, Colloids, and Granular media, Oxford, 2010.
  37. M. Wyart, S. R. Nagel and T. A. Witten, Europhys. Lett., 2005, 72, 486 CrossRef CAS.
  38. M. Wyart, L. E. Silbert, S. R. Nagel and T. A. Witten, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2005, 72, 051306 CrossRef.
  39. L. E. Silbert, A. J. Liu and S. R. Nagel, Phys. Rev. Lett., 2005, 95, 098301 CrossRef.
  40. W. G. Ellenbroek, E. Somfai, M. van Hecke and W. van Saarloos, Phys. Rev. Lett., 2006, 97, 258001 CrossRef.
  41. V. Vitelli, N. Xu, M. Wyart, A. J. Liu and S. R. Nagel, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2010, 81, 021301 CrossRef.

Footnotes

The parameter ugraphic, filename = c2sm25878a-t1.gif
Tighe29 defines a length scale λf = f/F as the ratio between the typical contact force f carried by springs to the viscous forces F exerted by the fluid. F is proportional to the velocity of the particles, and therefore to their displacements time the frequency ω. It is thus readily extractable from our results. The contact forces f are such that the forces are balanced on each node, which implies on average that the spatial derivatives of the contact forces f are the external forces F. If one considers for example the response to a local perturbation, our result implies that 〈f〉/|〈F〉| is of the order of the length scale lc(ω) on which the mean elastic response decays. Thus this definition is consistent with our results. However a different result is obtained if one considers the ratio ugraphic, filename = c2sm25878a-t2.gif, as done by Tighe. These two quantities differ because the mean response is much smaller than the fluctuations around it (see Section 4.1). Therefore, associating the quantity ugraphic, filename = c2sm25878a-t3.gif with a length appears unjustified. In the zero frequency limit we can actually calculate this last expression using the formalism developed in ref. 20. For isotropic random floppy networks one gets 〈f2〉 ∼ ω*−3 and 〈F2〉 ∼ ω*−1, leading to a ratio ugraphic, filename = c2sm25878a-t4.gif consistent with the numerical results of Tighe.
§ In the case of ellipses the pre-stress2 stabilizes the system, shifting the zero frequencies of floppy modes to finite values.9 We believe that the pre-stress plays a similar role to the weak springs of our network model. Accordingly, we expect the response to a local strain to be extended, even when the pre-stress is vanishingly small. Finally, note that our theoretical description of the vibrational spectrum does not capture the singular coupling that occurs between translational and rotational modes when the ellipticity α is small. We thus expect our approach to apply only for rather large values of α.

This journal is © The Royal Society of Chemistry 2013