This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Heterogeneity and Low-Frequency Vibrations in Bidisperse Sphere Packings

Yusuke Hara [email protected] Graduate School of Arts and Sciences, The University of Tokyo, Tokyo 153-8902, Japan    Hideyuki Mizuno Graduate School of Arts and Sciences, The University of Tokyo, Tokyo 153-8902, Japan    Atsushi Ikeda Graduate School of Arts and Sciences, The University of Tokyo, Tokyo 153-8902, Japan Research Center for Complex Systems Biology, Universal Biology Institute, The University of Tokyo, Tokyo 153-8902, Japan
Abstract

In the jamming transition of monodisperse packings, spatial heterogeneity is irrelevant as the transition is described by mean-field theories. Here, we show that this situation drastically changes if the particle-size dispersity is large enough. We use computer simulations to study the structural and vibrational properties of bidisperse sphere packings with a large size ratio. Near the critical point, the small particles tend to form clusters, leading to the emergence of large-scale structural heterogeneity. Concomitantly, the low-frequency vibrations are significantly enhanced compared to those in monodisperse packings, and their density of states follows a linear law with the frequency. We numerically and theoretically demonstrate that these behaviors of the structural heterogeneity and the low-frequency vibrations are intimately connected. The present work suggests that the nature of heterogeneous packings is markedly different from that of homogeneous packings.

Introduction.— Soft matters composed of dense macroscopic particles, such as foams, emulsions, and granular materials, commonly undergo the jamming transition into disordered solid states [1, 2]. This transition has attracted much attention in fundamental physics as it is an unusual critical phenomenon in non-thermal systems [1, 2, 3], and it also gives a fresh insight into thermal, structural glasses [4, 5]. Additionally, dense and active biological systems, such as cell tisues [6] and cytoplasm [7, 8], have been studied concerning the jamming transition.

Recent dramatic advances in jamming studies have benefitted from the simplest model, composed of nearly monodisperse, frictionless spheres. It is now established that mean-field theories can describe the jamming transition. At the transition point, a majority (95%\sim 95\%) of particles in the system get frozen simultaneously, and the system becomes isostatic; the contact number per particle becomes Z=2dZ=2d where dd is the spatial dimension [1]. In the jammed phase, vibrational properties are affected by the criticality of transition. Spatially disordered vibrations become abundant, which can not be described by the classical Debye theory based on phonon vibrations. The vibrational density of states (vDOS) shows a plateau in the frequency range down to the characteristic frequency ω\omega_{\ast} [9]. At lower frequencies ω<ω\omega<\omega_{\ast}, the vDOS follows the scaling law D(ω)(ω/ω)2D(\omega)\sim(\omega/\omega_{\ast})^{2}, which is called the non-Debye scaling [10, 11]. These vibrational properties are explained by a replica theory for spherical particles [3, 12] and an effective medium theory for spring networks [13, 14, 15]. These theories are mean-field, ignoring a possible spatial heterogeneity: The former is constructed for the infinite-dimensional limit, where any spatial correlation does not exist [3], while the latter disregards a heterogeneity in the distribution of local contact numbers [15]. Therefore, the success of these theories implies that spatial heterogeneity in structure is irrelevant to the jamming transition. Instead, an unusual homogeneity, the hyperuniformity, in the density field [16, 17] and the contact number distribution [18, 19] has been debated.

Refer to caption
Figure 1: Phase diagram of bidisperse packings with a size ratio of 6, in the plane of the pressure PP and the fraction of small particles XSX_{\rm S} [20]. The solid line indicates the first-order transition line which terminates at the critical point. The dashed line indicates the crossover line where half of the small particles are jammed. The squares indicate the state points (A-D) on which we focus in the present work.

In this Letter, we demonstrate that this situation drastically changes if the size dispersity of particles is large enough: Large-scale structural heterogeneity emerges, altering the low-frequency vibrations significantly. We study a binary mixture composed of large and small particles with a large size ratio. Previous studies [21, 22, 20, 23, 24] have established the phase diagram of such systems, as shown in Fig. 1. There are two distinct jammed phases: The “L-phase” where only large particles are jammed, and the “LS-phase” where both small and large particles are jammed. These phases are separated by the first-order transition line which terminates at the critical point. Crossing the transition line, most small particles get jammed simultaneously such that the fraction of jammed small particles changes discontinuously, whereas above the critical point, it changes continuously.

In this work, we first show that the packing structure becomes highly heterogeneous in approaching the critical point. The small particles tend to form clusters, and a system-spanning cluster emerges at the critical point. We next show that the vDOS at low-frequency regime follows a linear scaling law D(ω)ω/ωD(\omega)\sim\omega/\omega_{\ast} near the critical point; the low-frequency vibrations are significantly intensified. Finally, we numerically and theoretically demonstrate that these two properties, large-scale heterogeneity in structure and abundance of low-frequency vibrations, are closely related. The present work realizes heterogeneously jammed states, which have markedly different properties from homogeneously jammed states.

Model.— We consider NSN_{\rm S} small and NLN_{\rm L} large spherical particles in a three-dimensional (d=3d=3) box of linear dimension LL. We set the size ratio to be σL/σS=6\sigma_{\rm L}/\sigma_{\rm S}=6, where σS\sigma_{\rm S} and σL\sigma_{\rm L} are diameters of small and large particles, respectively. Particles interact via the harmonic potential vij(rij)=ϵ2(σijrij)2Θ(σijrij)v_{ij}(r_{ij})=\frac{\epsilon}{2}(\sigma_{ij}-r_{ij})^{2}\Theta(\sigma_{ij}-r_{ij}), where rijr_{ij} is distance between particles ii and jj, σij=(σi+σj)/2\sigma_{ij}=(\sigma_{i}+\sigma_{j})/2, and Θ(r)\Theta(r) is Heaviside step function. Length, time, and energy are measured in units of σS\sigma_{\rm S}, mSσS2/ϵ\sqrt{m_{\rm S}\sigma_{\rm S}^{2}/\epsilon}, and ϵ\epsilon, respectively (mSm_{\rm S} is the mass of small particles).

The macroscopic state of the system is specified by two state-variables (P,XS)(P,X_{\rm S}): Pressure P=L3ijfijrijP=L^{-3}\sum_{\expectationvalue{ij}}f_{ij}r_{ij} with contact force fij=vij(rij)f_{ij}=v_{ij}^{\prime}(r_{ij}) (\prime denotes derivative), and fraction of small particles XS=NS/(NS+63NL)X_{\rm S}=N_{\rm S}/(N_{\rm S}+6^{3}N_{\rm L}). We fix NS=8000N_{\rm S}=8000 and tune NLN_{\rm L} to control XSX_{\rm S}. To obtain mechanically stable packings at a given (P,XS)(P,X_{\rm S}), we quenched random configurations by using FIRE algorithm [25], and tuned the system size LL by an iterative compression/decompression algorithm to realize the target PP [20]. We prepared 200200 samples at each (P,XS)(P,X_{\rm S}) and calculated the sample average for various quantities, which is denoted by \expectationvalue{\cdot}.

The jammed particles are identified as particles making at least d+1d+1 contacts, as usually employed [1]. The fraction of the jammed small particles,

RS=1NSiSSi,\displaystyle R_{\rm S}=\frac{1}{N_{\rm S}}\sum_{i\in{\rm S}}S_{i}, (1)

is an order parameter to determine the phase behavior, where S{\rm S} denotes a set of small particles, and Si=1S_{i}=1 for the jammed state while Si=0S_{i}=0 otherwise. To systematically approach the critical point, we focus on four state points as indicated in Fig. 1: (A) (P,XS)=(3×102,0.0536)(P,X_{\rm S})=(3\times 10^{-2},0.0536), (B) (2×102,0.0910)(2\times 10^{-2},0.0910), (C) (102,0.135)(10^{-2},0.135), and (D) (7×103,0.152)(7\times 10^{-3},0.152). These states are located on the crossover line where half of the small particles are jammed while the other half are unjammed, i.e., RS=0.5\expectationvalue{R_{S}}=0.5.

Refer to caption
Figure 2: Typical configurations of jammed small particles at A (left top), B (right top), C (left bottom), and D (right bottom). The five largest clusters are in blue, and the smaller clusters are in gray.

Structure.— Since the phase behavior is controlled by the jammed small particles, we first focus on their spatial structures. We define a cluster as a group of particles that are connected via contacts between the jammed small particles only. We numerically identified such clusters using Hoshen-Kopelman algorithm [26]. In Fig. 2, we show snapshots of representative packings, where the five largest clusters are in blue while the other smaller clusters are in gray. Far from the critical point (state A), small clusters are scattered throughout the system. In approaching the critical point (state D), the size of clusters grows so that the largest cluster spans the entire system. To quantify this percolation behavior, we calculated the number of particles in the largest cluster, Smax\expectationvalue{S_{\text{max}}}, and showed an increasing (most likely diverging) behavior of Smax\expectationvalue{S_{\text{max}}} [see the Supplemental Material (SM)].

Figure 2 shows that the jammed small particles are distributed heterogeneously near the critical point. To discuss this heterogeneity more quantitatively, we introduce two types of radial distribution functions:

G(r)=L3NS2i,jSδ(rrij),\displaystyle G(r)=\frac{\expectationvalue{L^{3}}}{N_{\rm S}^{2}}\expectationvalue{\sum_{i,j\in{\rm S}}\delta(\vec{r}-\vec{r}_{ij})}, (2)
GS(r)=L3iSSi2i,jSSiSjδ(rrij),\displaystyle G_{\rm S}(r)=\frac{\expectationvalue{L^{3}}}{\expectationvalue{\sum_{i\in{\rm S}}S_{i}}^{2}}\expectationvalue{\sum_{i,j\in{\rm S}}S_{i}S_{j}\delta(\vec{r}-\vec{r}_{ij})}, (3)

where ri\vec{r}_{i} is the position vector of particle ii, rij=rirj\vec{r}_{ij}=\vec{r}_{i}-\vec{r}_{j}, and r=|r|r=|\vec{r}|. G(r)G(r) is a distribution function for all the small particles including unjammed ones, while GS(r)G_{\rm S}(r) is for the jammed small particles only. As shown in Fig. 3, G(r)G(r) is almost identical at the states A and D, meaning that the spatial distribution of all the small particles is insensitive to the criticality. In contrast, GS(r)G_{\rm S}(r) varies near the critical point. Far from the critical point (state A), GS(r)G_{\rm S}(r) becomes larger than G(r)G(r) at r6r\lesssim 6, whereas GS(r)G(r)G_{\rm S}(r)\approx G(r) at r6r\gtrsim 6. Notice that the length 66 is the diameter of the large particle, which should be in the same order as the size of pores formed by the jammed large particles. In approaching the critical point (state D), GS(r)G_{\rm S}(r) becomes substantially larger than G(r)G(r), and finite spatial correlation persists even at r6r\gtrsim 6. These results indicate that the jamming of small particles takes place cooperatively in the entire system near the critical point, whereas far from the critical point the cooperativity persists only within each pore. Thus, large-scale heterogeneity emerges in the spatial distribution of the jammed small particles near the critical point. We emphasize that such heterogeneity is totally absent in the monodisperse packings.

Refer to caption
Figure 3: Radial distribution functions of the small particles. G(r)G(r) is for all the small particles in the system, while GS(r)G_{\rm S}(r) is for the jammed small particles only.

Vibrations.— We next study the low-frequency vibrations in bidisperse packings using a standard vibrational-mode analysis [27]. Vibrational modes are obtained as eigenvectors of the dynamical matrix \mathcal{M}. Off-diagonal elements of \mathcal{M} from different particles iji\neq j are

ij=kijrijrijrij2+fijrij(Irijrijrij2),\mathcal{M}_{ij}=-k_{ij}\frac{\vec{r}_{ij}\bigotimes\vec{r}_{ij}}{r_{ij}^{2}}+\frac{f_{ij}}{r_{ij}}\left(I-\frac{\vec{r}_{ij}\bigotimes\vec{r}_{ij}}{r_{ij}^{2}}\right), (4)

where II denotes unit tensor, and kij=vij′′(rij)k_{ij}=v_{ij}^{\prime\prime}(r_{ij}). Diagonal elements from same particle i=ji=j are ii=jiij\mathcal{M}_{ii}=-\sum_{j\in\partial_{i}}\mathcal{M}_{ij}, where i\partial_{i} indicates a set of particles contacting with the particle ii. The vDOS is then calculated as D(ω)=13Nkδ(ωωk)D(\omega)=\frac{1}{3N}\langle\sum_{k}\delta(\omega-\omega_{k})\rangle, where ωk\omega_{k} the kk-th eigenfrequency.

Refer to caption
Refer to caption
Figure 4: The vDOSs of the bidisperse packings at the states (A-D) and that of the monodisperse packing at P=7×103P=7\times 10^{-3} for (a) the original system and (b) the unstressed system. The frequency ω\omega is scaled by the contact number ΔZ\Delta Z. The black line represents the linear dependence on ω\omega, i.e., D(ω)ωD(\omega)\propto\omega.

Figure 4 (a) shows the vDOSs of the bidisperse packings, compared to that of the monodisperse packing. The frequency ω\omega is scaled by the excess contact number ΔZ=Z2d\Delta Z=Z-2d since the characteristic frequency follows ωΔZ\omega_{\ast}\propto\Delta Z in the monodisperse packings [9]. In the high-frequency regime ω/ΔZ101\omega/\Delta Z\gtrsim 10^{-1}, the vDOSs of all cases collapse well and show a plateau. This result indicates that the critical point in the bidisperse packings does not affect anomalous vibrational modes in the plateau regime at ω>ω\omega>\omega_{\ast}. In contrast, in the low-frequency regime ω/ΔZ101\omega/\Delta Z\lesssim 10^{-1}, D(ω)D(\omega) strongly depends on the state; D(ω)D(\omega) gets larger in approaching the critical point, and near the critical point, it shows the linear dependence on ω\omega as D(ω)ω/ωD(\omega)\sim\omega/\omega_{\ast}. This linear dependence is in sharp contrast to the monodisperse packings which show the non-Debye scaling law for the boson peak, D(ω)(ω/ω)2D(\omega)\sim(\omega/\omega_{\ast})^{2}, and the quartic law for the quasi-localized vibrational modes, D(ω)(ω/ω)4D(\omega)\sim(\omega/\omega_{\ast})^{4} [28]. We also calculated the “concentration” on small particles MkSM^{S}_{k} for these modes (see the SM). A remarkable feature is that in the regime of D(ω)ω/ωD(\omega)\sim\omega/\omega_{\ast}, vibrational modes concentrate on small particles only, while the contribution from large particles is negligible. Our results establish that the low-frequency vibrations in near-critical bidisperse packings are significantly intensified compared to monodisperse packings, which are dominated by small particles.

In the monodisperse packings, the non-Debye scaling D(ω)(ω/ω)2D(\omega)\sim(\omega/\omega_{\ast})^{2} is a consequence of the marginal stability [12, 15]. This can be numerically confirmed by analyzing the unstressed version of the system [13]. In this analysis, the contact force fijf_{ij} in Eq. (4) is artificially set to be zero, and we diagonalize the resulting dynamical matrix. Since the contact forces always destabilize the system, this artificial operation improves the stability of the packings. As a result, the low-frequency vibrations (except for phonon vibrations) at ωω\omega\lesssim\omega_{\ast} disappear in the unstressed system [13, 29, 28, 30].

Here, we analyze the unstressed version of the bidisperse packings and the monodisperse packing, and show their vDOSs in Fig. 4 (b). In the monodisperse packing, D(ω)D(\omega) at ωω\omega\lesssim\omega_{\ast} sharply drops as expected. On the contrary, such a drop does not appear for the bidisperse packings, and the vDOS near the critical point still follows the linear law D(ω)ω/ωD(\omega)\sim\omega/\omega_{\ast}. This observation indicates that the origin of the low-frequency vibrations in bidisperse packings differs fundamentally from monodisperse packings: The low-frequency vibrations in bidisperse packings do not originate from the instability due to the contact forces. Instead, as shown below, they are explained by the heterogeneities in the structures. In the SM, we also show that the participation ratio PkP_{k} of each mode kk is unchanged between original and unstressed systems, which further validates the above discussion.

Link between structure and vibrations.— Now, we will demonstrate that the abundance of low-frequency vibrations is linked to structural heterogeneity.

We start with numerical evidence that spatial distributions of local contact numbers and low-frequency vibrations are correlated. We define the “concentration” of the low-frequency vibrations on the jammed small particle ii,

Mi=ωk<ω|ei,k|2,M_{i}=\sum_{\omega_{k}<\omega^{*}}\absolutevalue{\vec{e}_{i,k}}^{2}, (5)

where ei,k\vec{e}_{i,k} is the ii-th particle displacement in the kk-th vibrational mode. We next define the local contact number as follows. In general, a contact between two particles constrains their relative motion. However, if one of the particles is immobile, the contact virtually constrains the motion of the other one only. In the present bidisperse packings, the low-frequency vibrations are concentrated on the jammed small particles, meaning that the large particles virtually act as such immobile obstacles. To take this into account, we introduce an effective contact number for the jammed small particle ii,

Zi=ZiSS+2ZiSL,Z_{i}=Z^{SS}_{i}+2Z^{SL}_{i}, (6)

where ZiSSZ^{SS}_{i} and ZiSLZ^{SL}_{i} are the number of contacts with small and large particles, respectively. Since the large particles are immobile, the isostatic condition is that the average of this effective contact number is equal to 2d2d.

We computed above MiM_{i} and ZiZ_{i}, which are then coarse-grained as Z¯i=Zi+jiZj1+Ni\bar{Z}_{i}=\frac{Z_{i}+\sum_{j\in\partial i}Z_{j}}{1+N_{\partial i}} and M¯i=Mi+jiMj1+Ni\bar{M}_{i}=\frac{M_{i}+\sum_{j\in\partial i}M_{j}}{1+N_{\partial i}} where NiN_{\partial i} is the number of neibouring particles for the particle ii. To look at correlations between Z¯i\bar{Z}_{i} and M¯i\bar{M}_{i}, we calculated a conditional distribution function, P(Z¯|M¯)=iSδ(M¯iM¯)δ(Z¯iZ¯)/iSδ(M¯iM¯)\text{P}(\bar{Z}|\bar{M})=\langle\sum_{i\in{\rm S}}\delta(\bar{M}_{i}-\bar{M})\delta(\bar{Z}_{i}-\bar{Z})\rangle/\langle\sum_{i\in{\rm S}}\delta(\bar{M}_{i}-\bar{M})\rangle. Figure 5 shows P(Z¯|M¯)\text{P}(\bar{Z}|\bar{M}) near the critical point (state D). The distribution shifts to the left with increasing M¯\bar{M}: The more a particle participates in the low-frequency vibrations, the smaller its local contact number is. For the largest M¯\bar{M}, the distribution peaks at Z¯5\bar{Z}\approx 5 which is even smaller than 2d(=6)2d~{}(=6). This result clearly demonstrates that particles with fewer contact numbers contribute more to low-frequency vibrations.

Refer to caption
Figure 5: Conditional distribution function P(Z¯|M¯)\text{P}(\bar{Z}|\bar{M}) for the bidisperse packing near the critical point (state D). P(Z¯|M¯)\text{P}(\bar{Z}|\bar{M}) for M¯=4×10n\bar{M}=4\times 10^{-n} with n=1,2,3,4n=1,2,3,4 is plotted as a function of Z¯\bar{Z}.

The above numerical evidence motivates us to explain the linear law D(ω)ω/ωD(\omega)\sim\omega/\omega_{\ast} by considering the local contact-number fluctuations. Here we generalize the “cutting argument” [13], which was originally proposed to explain the plateau in the vDOS for monodisperse packings. We consider a system with the excess contact number ΔZ\Delta Z and linear dimension LL, and divide it into (L/)3(L/\ell)^{3} subsystems of linear dimension \ell. In the subsystem nn (n=1,2,,(L/)3n=1,2,...,(L/\ell)^{3}) with the excess contact number δzn\delta z_{n}, we estimate the number of the soft modes as 𝒩soft,n()=δzn3+2\mathcal{N}_{{\rm soft},n}(\ell)=-\delta z_{n}\ell^{3}+\ell^{2} for δzn<1/\delta z_{n}<1/\ell while 𝒩soft,n()=0\mathcal{N}_{{\rm soft},n}(\ell)=0 otherwise, where unimportant 𝒪(1)\mathcal{O}(1) coefficients are omitted. Since a soft mode in each subsystem can be used to generate a trial mode in the original system with the frequency ω1/\omega\sim 1/\ell, the cumulative vDOS is bounded as C(ω)L3n𝒩soft,n(1/ω)C(\omega)\gtrsim L^{-3}\sum_{n}\mathcal{N}_{{\rm soft},n}(\ell\sim 1/\omega), thus we obtain

C(ω)ω0ωP(δz)(1δzω)𝑑δz,\displaystyle C(\omega)\gtrsim\omega\int^{\omega}_{0}\text{P}_{\ell}(\delta z)\left(1-\frac{\delta z}{\omega}\right)d\delta z, (7)

where P(δz)\text{P}_{\ell}(\delta z) is probability distribution to find subsystems with δz\delta z. Eq. (7) is a generalization of the original cutting argument, which is embedded by variation in δz\delta z among the subsystems. Note that in the original argument for monodisperse packings [13], we suppose the delta function for P(δz)=δ(δzΔZ)\text{P}_{\ell}(\delta z)=\delta(\delta z-\Delta Z), which provides D(ω)=dC/dωω0D(\omega)=dC/d\omega\gtrsim\omega^{0} for ω>ΔZ\omega>\Delta Z.

Now we assume that spatial fluctuations in δz\delta z are sufficiently large; their standard deviation is in the same order as the average ΔZ\Delta Z, so a certain fraction of subsystems with δz0\delta z\sim 0 exists. In such the case, P(δz)\text{P}_{\ell}(\delta z) can be virtually treated as a constant for ω(1/)<ω=ΔZ\omega~{}(\sim 1/\ell)<\omega_{\ast}=\Delta Z, and the delta function δ(δzΔZ)\delta(\delta z-\Delta Z) for ωω=ΔZ\omega\gg\omega_{\ast}=\Delta Z (see the SM for a specific example with the Gaussian distribution). We thus obtain D(ω)ω/ωD(\omega)\gtrsim\omega/\omega_{\ast} for ω<ω\omega<\omega_{\ast} and D(ω)ω0D(\omega)\gtrsim\omega^{0} for ωω\omega\gg\omega_{\ast}. Assuming that the inequality is saturated as in the original argument, we obtain D(ω)ω/ωD(\omega)\sim\omega/\omega_{\ast} for ω<ω\omega<\omega_{\ast} and D(ω)ω0D(\omega)\sim\omega^{0} for ωω\omega\gg\omega_{\ast}, which totally agree with our numerical results in Fig. 4.

Concluding remark.— In this Letter, we investigated the structural and vibrational properties of bidisperse packings with a large size ratio. In approaching the critical point, the jamming of small particles takes place cooperatively, leading to the emergence of large-scale structural heterogeneity. Concomitantly, the low-frequency vibrations are significantly intensified near the critical point, and the vDOS follows the linear law, D(ω)ω/ωD(\omega)\sim\omega/\omega_{\ast}, at ω<ωΔZ\omega<\omega_{\ast}\propto\Delta Z. We numerically and theoretically demonstrated that these two phenomena of spatial heterogeneity and intensified low-frequency vibrations are ultimately connected. In particular, we generalized the cutting argument to the heterogeneous packings, and successfully explained the linear law of the vDOS in terms of spatial fluctuations in the local contact numbers.

In this work, we also found that contact forces on small particles are significantly small, and interestingly, their distribution follows a power law with a negative exponent (see the SM), which is marked contrast to a positive exponent in monodisperse packings [31]. It is an interesting future work to theoretically explain this behavior by extending the stability analysis [32] or the microscopic replica theory [3].

Finally, our results suggest that the particle-size dispersity can cause a large-scale heterogeneity, exerting a strong impact on the vibrational properties of the packings. Since low-frequency vibrations play important roles in plasticy [33], it is interesting to study the non-linear mechanical properties to elucidate impacts from the low-frequency vibrations found in this work. More broadly, as many soft jammed solids, ranging from granular materials to biophysical systems, quite usually have continuous polydispersities with wide-range distributions [34, 35, 36, 37], it should be very interesting to study such systems along the line of this work.

Acknowledgements.
This work was supported by Hosokawa Powder Technology Foundation Grant No. HPTF21509, JST SPRING Grant No.JPMJSP2108, and JSPS KAKENHI Grants No. 20H01868, 20H00128, 22K03543, 23H04495, 23KJ0368.

References

  • O’Hern et al. [2003] C. S. O’Hern, L. E. Silbert, A. J. Liu, and S. R. Nagel, Jamming at zero temperature and zero applied stress: The epitome of disorder, Phys. Rev. E 68, 011306 (2003).
  • van Hecke [2009] M. van Hecke, Jamming of soft particles: geometry, mechanics, scaling and isostaticity, Journal of Physics: Condensed Matter 22, 033101 (2009).
  • Parisi et al. [2020] G. Parisi, P. Urbani, and F. Zamponi, Theory of Simple Glasses: Exact Solutions in Infinite Dimensions (Cambridge University Press, 2020).
  • Liu and Nagel [1998] A. J. Liu and S. R. Nagel, Jamming is not just cool any more, Nature 396, 21 (1998).
  • Wyart [2005] M. Wyart, On the rigidity of amorphous solids, Annales De Physique 30, 1 (2005).
  • Bi et al. [2015] D. Bi, J. H. Lopez, J. M. Schwarz, and M. L. Manning, A density-independent rigidity transition in biological tissues, Nature Physics 11, 1074 (2015).
  • Parry et al. [2014] B. R. Parry, I. V. Surovtsev, M. T. Cabeen, C. S. O’Hern, E. R. Dufresne, and C. Jacobs-Wagner, The bacterial cytoplasm has glass-like properties and is fluidized by metabolic activity, Cell 156, 183 (2014).
  • Nishizawa et al. [2017] K. Nishizawa, K. Fujiwara, M. Ikenaga, N. Nakajo, M. Yanagisawa, and D. Mizuno, Universal glass-forming behavior of in vitro and living cytoplasm, Scientific Reports 7, 15143 (2017).
  • Silbert et al. [2005] L. E. Silbert, A. J. Liu, and S. R. Nagel, Vibrations and diverging length scales near the unjamming transition, Phys. Rev. Lett. 95, 098301 (2005).
  • Charbonneau et al. [2016] P. Charbonneau, E. I. Corwin, G. Parisi, A. Poncet, and F. Zamponi, Universal non-debye scaling in the density of states of amorphous solids, Phys. Rev. Lett. 117, 045503 (2016).
  • Shimada et al. [2020] M. Shimada, H. Mizuno, L. Berthier, and A. Ikeda, Low-frequency vibrations of jammed packings in large spatial dimensions, Phys. Rev. E 101, 052906 (2020).
  • Franz et al. [2015] S. Franz, G. Parisi, P. Urbani, and F. Zamponi, Universal spectrum of normal modes in low-temperature glasses, Proceedings of the National Academy of Sciences 112, 14539 (2015).
  • Wyart et al. [2005] M. Wyart, L. E. Silbert, S. R. Nagel, and T. A. Witten, Effects of compression on the vibrational modes of marginally jammed solids, Phys. Rev. E 72, 051306 (2005).
  • Wyart [2010] M. Wyart, Scaling of phononic transport with connectivity in amorphous solids, Europhysics Letters 89, 64001 (2010).
  • DeGiuli et al. [2014] E. DeGiuli, A. Laversanne-Finot, G. Düring, E. Lerner, and M. Wyart, Effects of coordination and pressure on sound attenuation, boson peak and elasticity in amorphous solids, Soft Matter 10, 5628 (2014).
  • Donev et al. [2005] A. Donev, F. H. Stillinger, and S. Torquato, Unexpected density fluctuations in jammed disordered sphere packings, Phys. Rev. Lett. 95, 090604 (2005).
  • Ikeda et al. [2017] A. Ikeda, L. Berthier, and G. Parisi, Large-scale structure of randomly jammed spheres, Phys. Rev. E 95, 052125 (2017).
  • Hexner et al. [2018] D. Hexner, A. J. Liu, and S. R. Nagel, Two diverging length scales in the structure of jammed packings, Phys. Rev. Lett. 121, 115501 (2018).
  • Hexner et al. [2019] D. Hexner, P. Urbani, and F. Zamponi, Can a large packing be assembled from smaller ones?, Phys. Rev. Lett. 123, 068003 (2019).
  • Hara et al. [2021] Y. Hara, H. Mizuno, and A. Ikeda, Phase transition in the binary mixture of jammed particles with large size dispersity, Phys. Rev. Res. 3, 023091 (2021).
  • Prasad et al. [2017] I. Prasad, C. Santangelo, and G. Grason, Subjamming transition in binary sphere mixtures, Phys. Rev. E 96, 052905 (2017).
  • Petit et al. [2020] J. C. Petit, N. Kumar, S. Luding, and M. Sperl, Additional transition line in jammed asymmetric bidisperse granular packings, Phys. Rev. Lett. 125, 215501 (2020).
  • Srivastava et al. [2021] I. Srivastava, S. A. Roberts, J. T. Clemmer, L. E. Silbert, J. B. Lechman, and G. S. Grest, Jamming of bidisperse frictional spheres, Phys. Rev. Res. 3, L032042 (2021).
  • Petit and Sperl [2023] J. C. Petit and M. Sperl, Structural transitions in jammed asymmetric bidisperse granular packings, Granular Matter 25, 43 (2023).
  • Bitzek et al. [2006] E. Bitzek, P. Koskinen, F. Gähler, M. Moseler, and P. Gumbsch, Structural relaxation made simple, Phys. Rev. Lett. 97, 170201 (2006).
  • Hoshen and Kopelman [1976] J. Hoshen and R. Kopelman, Percolation and cluster distribution. i. cluster multiple labeling technique and critical concentration algorithm, Phys. Rev. B 14, 3438 (1976).
  • Mizuno and Ikeda [2022] H. Mizuno and A. Ikeda, Computational Simulations of the Vibrational Properties of Glasses, in Low-Temperature Thermal and Vibrational Properties of Disordered Solids, edited by M. A. Ramos (WORLD SCIENTIFIC (EUROPE), 2022) Chap. 10, pp. 375–433.
  • Mizuno et al. [2017] H. Mizuno, H. Shiba, and A. Ikeda, Continuum limit of the vibrational properties of amorphous solids, Proceedings of the National Academy of Sciences 114, E9767 (2017).
  • Xu et al. [2010] N. Xu, V. Vitelli, A. J. Liu, and S. R. Nagel, Anharmonic and quasi-localized vibrations in jammed solids – modes for mechanical failure, Europhysics Letters 90, 56001 (2010).
  • Lerner and Bouchbinder [2018] E. Lerner and E. Bouchbinder, Frustration-induced internal stresses are responsible for quasilocalized modes in structural glasses, Phys. Rev. E 97, 032140 (2018).
  • Charbonneau et al. [2015] P. Charbonneau, E. I. Corwin, G. Parisi, and F. Zamponi, Jamming criticality revealed by removing localized buckling excitations, Physical review letters 114, 125504 (2015).
  • Wyart [2012] M. Wyart, Marginal stability constrains force and pair distributions at random close packing, Physical review letters 109, 125502 (2012).
  • Manning and Liu [2011] M. L. Manning and A. J. Liu, Vibrational modes identify soft spots in a sheared disordered packing, Phys. Rev. Lett. 107, 108302 (2011).
  • Sammis et al. [1987] C. Sammis, G. King, and R. Biegel, The kinematics of gouge deformation, Pure and Applied Geophysics 125, 777 (1987).
  • Huang et al. [2001] X. Huang, Y. Kakuda, and W. Cui, Hydrocolloids in emulsions: particle size distribution and interfacial activity, Food Hydrocolloids 15, 533 (2001), 5th International Hydrocolloids Conference.
  • Kwok et al. [2020] S. Kwok, R. Botet, L. Sharpnack, and B. Cabane, Apollonian packing in polydisperse emulsions, Soft Matter 16, 2426 (2020).
  • Shimamoto and Yanagisawa [2023] D. S. Shimamoto and M. Yanagisawa, Common packing patterns for jammed particles of different power size distributions, Phys. Rev. Res. 5, L012014 (2023).