Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2023 Jun 20;120(26):e2218218120. doi: 10.1073/pnas.2218218120

Dynamic Gardner cross-over in a simple glass

Qinyi Liao a,b,1, Ludovic Berthier c,d,1, Hai-Jun Zhou a,e,1, Ning Xu b,f,1
PMCID: PMC10293817  PMID: 37339213

Significance

The jamming transition is crucial to understand the formation and the nature of ubiquitous amorphous solids. Only recently, jamming has been associated with a thermodynamic Gardner phase of glasses by a replica mean-field theory. However, many realistic amorphous solids are formed far from equilibrium, and the generality of Gardner physics remains elusive. We address this question by simulating a simple glass model of hard disks. We discuss the existence of a dynamic Gardner cross-over for anomalous dynamics and provide the complete glass phase diagram by analyzing the landscape complexity. Our results bridge off-equilibrium experiments and thermodynamic studies and pave the way for constructing a dynamical glass theory that includes marginally stable phases.

Keywords: Gardner cross-over, glass, nonequilibrium dynamics, jamming

Abstract

The criticality of the jamming transition responsible for amorphous solidification has been theoretically linked to the marginal stability of a thermodynamic Gardner phase. While the critical exponents of jamming appear independent of the preparation history, the pertinence of Gardner physics far from equilibrium is an open question. To fill this gap, we numerically study the nonequilibrium dynamics of hard disks compressed toward the jamming transition using a broad variety of protocols. We show that dynamic signatures of Gardner physics can be disentangled from the aging relaxation dynamics. We thus define a generic dynamic Gardner cross-over regardless of the history. Our results show that the jamming transition is always accessed by exploring increasingly complex landscape, resulting in anomalous microscopic relaxation dynamics that remains to be understood theoretically.


The jamming transition describes the formation of amorphous solids in materials composed of repulsive particles (1, 2). It has been the subject of important research activity in the last decades, encompassing statistical mechanics analysis, numerical studies, and experimental investigations of granular and colloidal materials (3, 4). The jamming transition also attracts interest in contexts such as the geometry of sphere packings, the statistical mechanics of liquid and glass states, and is related to mechanical properties of biophysical matter and machine learning in computer science (59). Jammed packings have original physical properties that differ dramatically from crystalline solids. In particular, the application of small perturbations often leads to large-scale responses, showing that jammed materials are marginally stable (10).

Understanding the structure of jammed packings is difficult because the absence of thermal fluctuations prevents the use of a statistical ensemble, as first noted by Edwards (11, 12). This key problem was circumvented by describing jamming as the end point of the compression of dense assemblies of thermalized soft or hard repulsive spheres (13). This approach is meaningful because Brownian particles near jamming belong to dynamically arrested glass states (14). It becomes possible to follow quasi-statically the evolution of glassy states compressed toward jamming which only probes a restricted region of configuration space (15, 16). For this thermodynamic description to be correct, it is crucial that the compression history occurs over timescales that are much shorter than the timescales related to the crossing of barriers in configuration space. In the limit of large dimensions, d → ∞, where the analytic calculations were performed, this timescale separation is guaranteed by the divergence of free energy barriers separating distinct glass states (13).

An unexpected outcome was the discovery that glass states followed quasi-statically upon compression undergo a Gardner phase transition separating two types of glass phases described by distinct sets of solutions (17, 18). The Gardner phase bears many similarities with the replica symmetry broken spin glass phase found at low temperatures in models of disordered magnets (19, 20), which implies the existence of specific aging effects with multiple timescales and lengthscales and a hierarchical free energy landscape (21). The Gardner phase directly impacts the physical properties of jammed materials, as shown by mean-field studies (16, 2224). More broadly, Gardner physics arises from breaking the replica symmetry and its study is of great interest in fields encompassing optimization problems (25, 26), deep learning (27), cell tissues (28), magnetic materials (29), polydisperse crystals (30, 31), etc. The dynamic Gardner cross-over analyzed in this work in the hard disk model might apply to other complex systems with rough landscapes, as studied by these different communities.

It is not known whether the Gardner transition can survive in physical dimensions (32, 33). Yet the measured critical exponents of jamming are independent of the dimension for d ≥ 2 (2). The robustness of jamming criticality is difficult to reconcile with the fragility of the Gardner transition. More questionable is the fact that jamming was studied using different approaches, which cannot always be described using quasi-static thermodynamic descriptions (2, 14). There is therefore a deep theoretical gap between conventional jamming studies and the available thermodynamic approach. Even within mean-field spin glasses, the off-equilibrium dynamics of glasses entering the Gardner phase has not been studied beyond some incomplete attempts (3438). In fact, there are increasing efforts to tackle off-equilibrium glassy dynamics (3943), but the dynamic Gardner cross-over itself remains out of reach of those attempts, and appears as a very difficult task. Our results may help shed light on the expected physical behavior.

Numerical simulations and experiments were performed to detect signs of the Gardner transition (4449), or to explore the physical consequences of the hierarchical free energy landscape for the aging dynamics (5052) and the mechanical properties (49, 53) of dense assemblies of repulsive particles. Although the Gardner transition is inexistent in d = 2, the physics of hard disks is reminiscent of its d = 3 counterpart with clear signatures of a strong Gardner cross-over (50). Many numerical studies mimicked the analytic thermodynamic construction using highly stable configurations to reproduce the analytic state-following construction (44, 46, 5053). Very few studies (46, 54) attempted to characterize the Gardner phase and related microscopic dynamics for the less stable systems studied experimentally with mixed conclusions. On the other hand, despite recent reports on Gardner-like phenomena in the systems of long-ranged interactions (43, 55), the relevance of Gardner physics in soft glasses is debated (48, 51, 56, 57). Our work is helpful to resolve the controversy and understand the stability of those glassy materials.

There is a growing interest in Gardner-related phenomena in a wide range of complex systems but results are not always conclusive because the studied models and protocols are often very complex. Using well-controlled methods and well-studied systems is useful. Hard disks represent a simple and practical choice with quite a venerable history across statistical physics, which represents a fruitful playground to study Gardner physics beyond mean-field theory. Here, we simulate a two-dimensional system of hard disks (50) and use a wide range of preparation protocols from poorly annealed systems to stable ones to explore the nonequilibrium dynamics observed during compression protocols toward the jamming transition. We find that the signs of a Gardner cross-over are robust even in the regime where strong structural relaxations are present and can be disentangled from the aging relaxation dynamics. We thus generalize the thermodynamic Gardner cross-over to a dynamic cross-over for any specific history, as well as propose a complete phase diagram for the emergence of Gardner physics for hard disks, thus going much beyond the purely static approach followed in ref. 50. Our results show that Gardner physics is at play in the approach to jamming, independently of the preparation history.

Results

Sample Preparation.

We simulate dense assemblies of two-dimensional hard disks (50) with a continuous size polydispersity to avoid crystallization (Methods). The mean disk diameter sets the unit length, and the pressure is normalized by kBTρ, with ρ the number density and kBT the thermal energy defining the normalized pressure Z. Since the scale set by the potential energy is infinite for hard disks, we can impose kBT = 1. The simulations are performed using a Monte Carlo (MC) algorithm, and the MC sweep is used as the time unit. The number of particles is N = 1024, chosen to allow exploration of the free energy landscape (50). We remove contributions from the translational Mermin–Wagner fluctuations to the dynamics using cage-relative coordinates when computing physical observables (Methods) (5860).

To search for Gardner physics for the whole glass phase diagram, we devise a two-step sample preparation. The prepared samples will then be employed as starting points of compression protocols devised to probe the underlying free energy landscape. Let us first introduce the preparation. We start by equilibrating the system at fixed volume fractions φg using constant volume MC simulations. To achieve equilibrium for dense configurations, we use the swap MC algorithm to speed up the thermalization (61). This enables us to access equilibrium fluid states at densities much higher than the dynamic mode-coupling theory (MCT) cross-over, φMCT ≈ 0.795. In the second step, we switch off the swap MC moves and perform conventional MC simulations in the NPT ensemble at various reduced pressure ZsZ(φg). Constant pressure MC simulations are convenient to approach the jamming transition, which is obtained in the limit Z → ∞. We let the system relax at pressure Zs for a time ts and store the final configurations, which are characterized by three control parameters (φg, Zs, ts). At the end of this preparation, the configurations are off-equilibrium at state point (Zs, φs). We repeat this recipe starting from independent configurations at ϕg when performing ensemble averages over independent trajectories.

Therefore, this two-step protocol enables us to cover multiple state points and physical regimes, in particular those characterized by aging relaxation dynamics far from equilibrium, which could be related to a large class of experimental situations (45, 4749). The samples drawn from a compression with a large Zs = Z(ϕg) (equivalently, ts = 0) correspond to previous studies (44, 46, 50) mimicking the theoretical state-following construction. Adding the parameter ts for lower ϕg values is an important feature of our study, which allows us to explore in a controlled manner and with minimal ingredients a broad range of nonequilibrium states with different degrees of aging and stability.

We display several examples of the NPT second stage of the preparation in Fig. 1A, where each line represents the evolution of the volume fraction φ(ts) for pairs of (φg, Zs) values. The different symbols indicate representative preparation states characterized by (ϕg, Zs, ts) that will be used to investigate the Gardner cross-over. The pressure Zs for states initialized from ergodic liquids with φg = 0.600 ranges from Zs = 30 to Zs = 104. When starting from denser glass states with φg = 0.810 and φg = 0.860, we only show results for Zs = 102. When compressing from φg = 0.600 ≪ φMCT ≈ 0.795, it can be seen that φ grows with the time ts over the entire window, showing that ts is a relevant control parameter for such nonequilibrium preparation histories. Glasses obtained after longer ts are more stable and, accordingly, display smaller particle diffusion during the densification, see Fig. 1B. The mean-squared displacements reveal the presence of aging dynamics for these systems. We shall call these glassy states “ordinary glasses”.

Fig. 1.

Fig. 1.

Broad range of sample preparation. (A) Evolution of the packing fraction in the second step of the sample preparation, starting from different φg at ts = 0 and compressing at constant pressure Zs. For φg = 0.600 we show Zs = 30, 50, 102, 103, and 104 (from Bottom to Top), while for φg = 0.810 and φg = 0.860 we only show Zs = 102. Representative glass samples characterized by (ϕg, Zs, ts) are marked by symbols. (B) The caged-relative mean-squared displacements Δ(tw = 0, t) for glass samples at pressure Zs = 102 marked in (A) with (ϕg, ts) as indicated. Strong aging relaxation effects are obvious for ordinary glasses prepared from ϕg = 0.600. (C) Sketch of the position of the different types of glasses shown in (A) in the free energy landscape, where the height encodes the degree of stability and the black horizontal lines represent the Gardner cross-over.

The situation is qualitatively different for the case of (φg = 0.860, Zs = 102), where the system rapidly achieves restricted equilibrium inside the metastable glass basin in a time shorter than ts ≈ 105. The mean-squared displacement for that system in Fig. 1B displays a stable plateau. We refer to these glasses as “ultrastable glasses,” and they are roughly similar to systems used in earlier numerical studies (44, 46, 50). Finally, glasses prepared by compressing from φg = 0.810 to Zs = 102 are close to the thermodynamic Gardner cross-over that is estimated below to occur near ZG ≈ 102. They show a mild time dependence of φ(ts) and a larger mean-squared displacement. We call these glasses “marginal glasses”. For such large packing fractions, the physical MC dynamics is nearly arrested.

In Fig. 1C, we offer a sketch of the position in the energy landscape of the three categories of glasses prepared in Fig. 1A, namely, ordinary, marginal, and ultrastable glasses. We are particularly interested in studying the pertinence of the Gardner physics for ordinary glasses, which are not well understood theoretically and are closer to systems studied experimentally. To this end, we will compare their study with results obtained on ultrastable and marginal glasses.

Exploring Complex Landscapes upon Compression.

After the two-step sample preparation described above, we perform various compression histories using NPT MC simulations. The end of the sample preparation at ts corresponds to the waiting time tw = 0, after which the pressure is instantaneously changed to a new value Z > Zs, or kept at Z = Zs as in Fig. 1B. Although this may superficially resemble the state-following construction proposed by thermodynamic theories, our numerical approach can be applied to all types of glass states, and we do not assume that infinitely long-lived states exist.

We instantaneously change the pressure from the preparation pressure Zs to a range of applied pressure Z at time tw = 0 and track the evolution of the system as a function of the time tw spent at the final pressure Z. Following earlier work, we simulate Nc independent trajectories, or clones, for each prepared sample at tw = 0 by using different sequences of random numbers in the MC simulations. We further average over Ns independent samples for each state characterized by (φg, Zs, ts) to average over the disorder, which results in Ns × Nc simulations in total to fully characterize a given preparation history. We use (Ns = 100, Nc = 10) to get statistical properties and increase the number of clones to Nc = 100 when analyzing more finely the landscape structure of individual samples.

It is useful to compare the results obtained for ordinary glasses with more stable glasses. To this end, we select three examples prepared at the same values (Zs = 200, ts = 218 × 100) but for different initial equilibrium states of ϕg = 0.600, 0.820, and 0.860, which respectively represent typical ordinary, marginal, and ultrastable glasses. The pressure Z varies from Z = 200 to Z = 104.

We study the dynamics using the cage-relative mean-squared displacement (MSD) Δ(tw, t) between times tw and tw + t at a series of waiting times tw. For ultrastable glasses with (φg = 0.860, Zs = 200), the dynamics exhibits a cross-over from simple vibrations to anomalous aging at a threshold pressure that coincides with the Gardner cross-over ZG ≈ 103 (SI Appendix) By contrast, we observe an aging behavior for marginal and ordinary glasses in the entire pressure range. For the marginal glass, we have confirmed that there is no diffusion to another glass state within the simulated time by decompressing these systems to a lower pressure Z = 102 where the dynamics was purely vibrational. The aging dynamics for the marginal samples is thus entirely due to the marginal stability of the Gardner phase. Instead, ordinary glasses first age at modest Z because they relax from one state to another, whereas for larger Z aging is instead dominated by Gardner physics, as we establish more precisely in the rest of the paper.

Because of the interplay between structural relaxation and Gardner physics, it is not easy to interpret the physical origin of aging of ordinary glasses. To gain further insight, we resort to the cage-relative mean-squared distances between pairs of clones at time tw, ⟨Δab⟩. In Fig. 2, we show the long-time values of Δ(tw, t) and ⟨Δab⟩ measured at tw = t = 217 × 100 as a function of Z for the three glass states. When the landscape is simple, ergodicity is achieved within metastable basins so that the time average Δ and the ensemble average ⟨Δab⟩ coincide. Otherwise, they will differ from each other. The separation between Δ and Δab is thus considered the primary criterion for the emergence of a hierarchical landscape (44). This is less obvious for marginal and ordinary glasses, which may undergo structural rearrangements and therefore have ⟨Δab⟩ larger than Δ throughout the entire range of pressure as observed in Fig. 2.

Fig. 2.

Fig. 2.

Exploring complex landscapes upon compression. Long-time limit (tw = t = 217 × 100) of cage-relative MSD Δ(tw, t) (full lines) and cage-relative mean-squared distance between clones ⟨Δab(tw)⟩ (dashed lines) for three states prepared from ϕg = 0.600 (red), 0.820 (light blue) and 0.860 (dark blue) and the same (Zs = 200, ts = 218 × 100). The quantities Δ and ⟨Δab⟩ differ at all Z for ordinary and marginal glasses, whereas they are equal up to the Gardner cross-over ZG for ultrastable glasses. However, in all cases, ⟨Δab⟩ exhibits a nonmonotonic dependence with Z as a rough landscape emerges, and the positions of the resulting minima are indicated by the vertical lines.

Nevertheless, the evolution of ⟨Δab⟩ with pressure itself carries interesting information about the free energy landscape. Upon compressing glasses within a hierarchical landscape, ⟨Δab⟩ grows with increasing pressure due to the rapid proliferation of states, as observed for ultrastable and marginal glasses. As shown in Fig. 2, ⟨Δab⟩ is nonmonotonic for all three glasses, suggesting that Gardner physics is relevant to the dynamics of ordinary glasses, despite the presence of aging relaxations.

Dynamic Gardner Cross-Over.

In the state-following analysis of ultrastable glasses, the emerging aging dynamics observed beyond the Gardner cross-over is related to the emergence of free energy barriers preventing the unrestricted exploration of the available phase space. As a result, when a population of clones is simulated one finds that different clones may explore different parts of this phase space, which gives rise to a broad distribution of distances between pairs of clones. Importantly, when the pressure is returned to a value smaller than the Gardner cross-over, this clustering in phase space disappears, and the clones evolve freely again. This represents a memory effect (50, 62). This property is important as it enables us to distinguish aging signatures due to the Gardner cross-over showing memory, from the aging relaxation toward different states, which is fully irreversible.

We first plot the distributions of the cage-relative distances between clones P(Δab) in Fig. 3 for the same three samples as in Fig. 2. As shown in the first column for Z = 200, the distribution P(Δab) for ultrastable glasses remains Gaussian at all times tw, indicating unhindered vibrational motion. By contrast, the distributions P(Δab) for marginal and ordinary glasses broaden toward larger Δab values, which corresponds to some form of aging relaxation. The influence of the initial stability is also clear, as the broadening is much stronger for the ordinary glass than it is for the marginal one and is absent for the ultrastable glass.

Fig. 3.

Fig. 3.

Coexistence of aging relaxation and hierarchical landscape. Evolution of the distributions of cage-relative distances between clones P(Δab) at Z = 200 (A1–C1), 400 (A2–C2), 103 (A3–C3), 104 (A4–C4) (from Left to Right). From top to bottom, we show results for the three samples compressed from ϕg = 0.600 (A1–A4), 0.820 (B1–B4), and 0.860 (C1–C4) to the same (Zs = 200, ts = 218 × 100) as in Fig. 2. The distributions broaden at small Δab when an increasingly complex landscape controls the physics at large pressure for all types of glasses. The broadening at large Δab is pronounced at low Z for less stable glasses and reflects structural relaxation.

When the pressure is increased further, all glasses fail to achieve a complete sampling of the glass basin, and P(Δab) now develops a tail toward small values of Δab as tw increases. It indicates that different clones explore different parts of the available phase space, some pairs being close to one another, whereas other pairs are far from each other. When the pressure is high enough, ordinary glasses behave similarly to marginal glasses, showing that Gardner physics always becomes important upon approaching jamming.

To better understand the interplay between aging dynamics due to structural relaxation or to the dynamic Gardner cross-over, we decompress the clones produced from the ordinary glasses from the pressure Z where they have aged for a long time back to the initial pressure Zs = 200 and compare the distributions to the reference case Z = Zs to investigate the presence of memory. When the non-Gaussian distributions at Z > Zs are solely due to the hierarchical Gardner landscape, the two distributions are expected to be the same. We use a short time after decompression t = 28 × 100 to perform this comparison as it does not allow for significant aging for the reference case, see Fig. 3A1. In Fig. 4 we show the effect of decompressing the systems shown in Fig. 3A2–A4 back to Zs = 200 after tw = 218 × 100, and wait 28 × 100 steps before measuring P(Δab). The difference between the decompressed systems and the reference one at large Δab can be attributed to aging relaxation processes that have taken place in the stages presented in Fig. 3A2–A4. They can be observed in Fig. 4, with an amplitude that decreases with increasing Z.

Fig. 4.

Fig. 4.

Partial memory effect in ordinary glasses. Distributions of cage-relative distances between clones P(Δab) for ordinary glasses first compressed for tw = 218 × 100 at pressure Z and then decompressed back to Zs = 200 for a short time t = 28 × 100. The results are compared to the reference curve (cyan line) which only aged for a short time t = 28 × 100 at Zs = 200. Deviations with the reference distribution at large Δab are due to aging relaxation dynamics which is irreversible.

These results suggest that a dynamic Gardner cross-over can be defined from the distributions P(Δab) by focusing on the emergence of small values of Δab upon compression, which is unrelated to aging relaxation dynamics. For such a criterion to coincide with the previously used separation between the averages values ⟨Δab⟩ and ⟨Δ⟩ which can be used for ultrastable glasses, we introduce the quantity

F(tw)=0ΔP(Δab)Δab, [1]

which represents the enhanced probability that the distance between clones is smaller than the average size of the cage. For ultrastable glasses, F(tw) rapidly converges to a constant close to 0.5 for low pressures but fails to converge to 0.5 above the Gardner cross-over (SI Appendix) For marginal and ordinary glasses, F(tw) decreases at low pressure due to aging relaxation but behaves similarly to ultrastable glasses at larger pressure. For all glasses, upon compressing the systems, F(tw) (SI Appendix) changes from a decreasing function to an increasing one, indicating a generic dynamic cross-over. We thus define a threshold pressure ZG at which F(tw) is nearly constant over long times. For the samples shown in Fig. 3, we can thus estimate the Gardner cross-over to be ZG ≈ 500 for the ordinary glass, ZG ≈ 400 for the marginal glass, and ZG ≈ 103 for the ultrastable glass. The presence of subdiffusive motion(57) is not a unique signature of Gardner physics because it is also observed in glasses with simple aging dynamics.

The overall conclusion is that Gardner physics at large pressure is surprisingly robust against the aging relaxation dynamics taking place when the quasi-static conditions used in the theoretical analysis do not hold (SI Appendix) We confirm this conclusion by additional analysis of individual samples (SI Appendix) Nevertheless, the complicated interplay between the various timescales involved in aging and Gardner dynamics certainly challenges analytic approaches to these nonequilibrium dynamics.

Increasing Susceptibility upon Approaching the Jamming Transition.

Having revealed that Gardner physics is pertinent in all preparation regimes, we quantify whether increasing collective fluctuations can also be detected. Following earlier work, we compute the global susceptibility χAB quantifying the fluctuations of the distance field between clones, Δab(r) (Methods). In the context of thermodynamic theory, the Gardner phase is critical and characterized by full replica symmetry breaking with an infinite correlation length and infinite susceptibility χAB. In the physical dimensions, simulations have shown that there exists a long but finite correlation length in the vicinity of the Gardner cross-over, where χAB increases with pressure (44, 50, 51), compatible with a possible divergence at finite pressure in d = 3 (44, 51).

However, most previous studies focused on ultrastable glass and the situation is unclear in the presence of relaxation dynamics at low pressure, which can affect the behavior of the susceptibility. To this end, we adopt a simpler measurement protocol where we do not change the pressure at the end of the preparation at Zs and directly create Nc clones at tw ≥ 0 and Z = Zs which are then followed for tw > 0. We then monitor the time dependence of χAB(tw), which quantifies collective effects. This procedure applies equally well to stable glasses (SI Appendix)

We compare the probability distributions of clone distances at long times, P(Δab) for tw = 218 × 100, with their short-time counterparts which contain only vibrations (measured using tw = 104 − 106 depending on Zs) in Fig. 5A, using the initial states generated at various Zs for (φg = 0.600, ts = 218 × 100).

Fig. 5.

Fig. 5.

Hierarchical landscape and increasing susceptibilities at larger pressures. (A) Probability distributions of mean-squared cage-relative distances between clones P(Δab) for (ϕg = 0.600, ts = 218 × 100), (Ns = 100, Nc = 10) and various Z = Zs. The solid curves are measured at long time, tw = 218 × 100, the dotted lines at short times tw = 104 − 106. (B) Representative heat map of Δab for a sample with (φg = 0.600, Zs = 103, ts = 218 × 100) using Nc = 100 clones at Z = Zs, tw = 218 × 100. The axes represent the clone index, and the value of Δab is color coded. A clear hierarchical structure of phase space appears. (C) Time evolution of the susceptibility χAB(tw) corresponding to the distributions shown in (A). Fast relaxation at small Z does not result in the large susceptibility that Gardner physics yields at large pressure.

At short times, the distributions P(Δab) are Gaussian, centered at the averaged cage size for Zs = 30 − 103, corresponding to the vibrations within the cages selected by the initial states. With increasing pressure, the time to reach a plateau in the MSD increases rapidly. At Zs = 104, the maximal time tw ∼ 108 is not long enough to see the MSD plateau. Therefore, we do not display P(Δab) for this pressure in Fig. 5A.

At long times, the evolution of the distributions P(Δab) demonstrates that clones gradually move away from each other and explore different parts of the landscape. At the lowest pressure Zs = 30, tw = 218 × 100 is long enough for particles to escape from their cages, and Δab is about one order of magnitude larger than the typical cage size. With increasing the pressure, P(Δab) remains much broader than the distribution of cage sizes, despite the slow dynamics. This suggests an increasing complexity of the landscape at higher pressures. Moreover, as shown by the heat map in Fig. 5C for a randomly chosen sample at Zs = 103 and tw = 218 × 100, the organization of Δab is reminiscent of the ultrametric structure characterizing a Gardner phase, confirming the existence of a hierarchical landscape even in ordinary glasses which age.

Finally, we present in Fig. 5C the time evolution of the susceptibility χAB(tw) for the same ordinary glass states as in Fig. 5A. For Zs = 30, 50, χAB(tw) reaches a peak value χAB* at tw= 107 − 108, which corresponds to the timescale for escaping the initial cage. For higher pressures, χAB(tw) remains small at short times but eventually increases at longer times, revealing increasingly long-ranged spatial correlations and stronger dynamics heterogeneities. In particular, χAB(tw) at Zs = 103 increases by about two orders of magnitude within the simulation time and continues to grow at large times. Interestingly, it can reach significantly larger values than the susceptibilities observed in marginal glasses (46, 50, 51) because the dynamics is faster in ordinary glasses. We have also studied the dependence of χAB(tw) on the packing fraction φs by varying the preparation time ts at fixed Zs and find that χAB* is larger for larger ϕs (SI Appendix) Therefore, we conclude that a large susceptibility χAB* emerges upon approaching jamming.

Discussion

By using a broad variety of preparation histories leading to glasses with various degrees of stability, we have established that specific signatures due to the Gardner cross-over are generically present in glasses compressed toward jamming. A careful choice of protocols and observables allowed us to reveal that the characteristics of a Gardner phase are systematically found in all types of glasses, and ultrastability is unnecessary to analyze the Gardner transition even though it makes the dynamics simpler to analyze. We summarize these observations and determine the complete region of the hard disk phase diagram where Gardner physics is pronounced in Fig. 6, which identifies the regions of state points where the glassy landscape becomes complex for a very broad range of initial conditions.

Fig. 6.

Fig. 6.

Phase diagram of dense hard disks from the fluid to jamming. We summarize our results in the inverse pressure 1/Z versus packing fraction φ representation. Equilibrium fluid states (hexagons) follow the Carnahan–Starling equation of state (blue line). The mode-coupling cross-over is indicated by a green symbol. Black dotted lines represent glass equations of state obtained by compressions of stable equilibrium configurations starting from above ϕMCT. A Gardner cross-over separates simple and hierarchical glasses during these compressions (light blue symbols) before hitting the J-line at Z → ∞ (red line). For ordinary glasses, there is no well-defined metastable state to follow but an isoline of susceptibility χAB* ≈ 60 (orange symbols) locates the emergence of a complex landscape. Gardner physics is relatively pronounced in the triangular region “protecting” the J-line.

Hard spheres in large dimensions undergo a dynamic mode-coupling transition corresponding to the emergence of long-lived metastable states, but no such transition occurs in finite dimensions (13, 63, 64). However, the dynamic cross-over determined from an approximate power law fit inspired by mode-coupling theory to the equilibrium relaxation time remains a useful reference point. It is represented by a green symbol at (φMCT ≈ 0.795, ZMCT ≈ 24) in Fig. 6.

For initial packing fractions φgφMCT, we use the swap Monte Carlo algorithm to equilibrate the system. The measured equilibrium equation of state can be described by the Carnahan–Starling equation of state (65). For these dense states, further compressions using ordinary MC simulations represent a relatively faithful implementation of the theoretical state-following construction, allowing us to measure the glass equations of states shown with black dotted lines in Fig. 6. Finally, using the observable F(tw) introduced in Eq. 1, we locate the Gardner cross-over for these stable glassy states, as depicted by the light blue line in Fig. 6, which shows that ZG increases with φg. We find that ZG ≈ 80 for ϕgφMCT, which is significantly higher than the equilibrium value ZGZMCT ≈ 24. In previous simulations of three-dimensional hard spheres (17, 44), the authors did not consider the structural relaxation when determining the Gardner cross-over in the vicinity of the mode-coupling cross-over despite the lack of well-defined metastable states, similarly to our hard disk system. This might explain why we conclude that the pressures ZG(ϕMCT) and ZMCT are distinct. This issue would be worth revisiting in three dimensions.

When φg < φMCT, there are no well-defined metastable basins in the equilibrium liquid phase. To evaluate the complexity of the free energy landscape, we measure the peak value of susceptibility over time χAB* as in Fig. 5C, which grows upon approaching jamming (occurring on the red line). To determine the relevance of Gardner physics, we empirically define the criterion χAB* ≈ 60, which is large enough to signal the presence of Gardner physics, and low enough to be accessed within the simulation time window. The result is displayed with orange symbols in Fig. 6. This line does not exist below Z ≈ 60 where the susceptibility remains small. The data for larger densities shown in (SI Appendix) suggest that this line moves to larger pressures when Z increases, see Fig. 6.

Combining all measurements, we obtain a triangular zone in the phase diagram of Fig. 6 where Gardner physics is highly relevant. Interestingly, this zone appears to ‘protect’ the approach to the line of jamming points at Z = ∞. We conclude that when approaching the jamming transition using all kinds of protocol by compressing ordinary, marginal, or ultrastable glasses, Gardner physics is inevitably involved. It is characterized by a hierarchically complex free energy landscape, leading to anomalous aging dynamics in the motion of dense assemblies of particles near the jamming transition, as well as specific physical properties related to marginal stability. Our simulations show that these signatures are robustly observed in two-dimensional systems, in the presence of thermal fluctuations, and for preparation protocols and compression histories similar to the ones studied in several experiments. Our study therefore naturally reconciles recent experimental and numerical studies of the Gardner cross-over in dense particle systems and provides a guide for future theoretical studies of their off-equilibrium dynamics in the vicinity of the jamming transition.

Materials and Methods

Model.

We simulate an assembly of N = 1024 two-dimensional hard disks in a squared box of length L with periodic boundary conditions. The interaction between two disks is infinite when they overlap and zero otherwise. The disk diameter σ is drawn randomly from a continuous distribution P(σ)∝σ−3 for σ ∈ [0.45σmax, σmax], and the polydispersity is (σ2¯σ¯2)/σ¯0.23 with ···¯ the average over P(σ). The units for length, energy, and time are the mean diameter σ¯, temperature T, and a MC sweep, respectively. The physical control parameters are the volume fraction φ=πσ2¯/4L2, the reduced pressure Z = p/(ρkBT), and Monte Carlo (MC) sweep t, where ρ = N/L2 is the particle number density. The equilibrium equation of state for this model can be fitted by the following empirical relation,

Z(φ)=11φ+σ¯2σ2¯(1+φ/8)φ(1φ)2. [2]

Observables.

The reduced pressure can be computed by the contact number,

Z=1+12Ni=1Nj>iNδRijσij1+, [3]

where σij = (σi + σj)/2 is the average diameter of particle i and particle j, and Rij=|RiRj| is the separation between two particles with Ri the coordinate of i. To get rid of the influence of the Mermin–Wagner fluctuations, we adopt cage-relative coordinates in our observations,

ri=Ri1NijiNiRj, [4]

where ∂i is the set of neighbors of particle i and Ni = |∂i|. Here, we define j ∈ ∂i if Rij<2σ¯. In the state-following scheme, each of the Ns samples is cloned Nc times. We use the statistics of (Ns = 100, Nc = 10) for general properties and (Ns = 1, Nc = 100) when studying individual samples. The cage-relative distance between clones a and b is

Δab=1Ni=1NΔab,i,Δab,i=|riarib|2. [5]

Averaging over the clone pairs and over samples, one can get the mean-squared clone distance,

Δab=1Nsα=1Ns2Nc(Nc1)a,bα,b>aNcΔab. [6]

Here, ∂α is the set of clones of a sample α, ⟨ ⋅ ⋅ ⋅ ⟩ represent averaging over samples and clone pairs. The susceptibility for the spatial displacement field is,

χAB=NΔab2Δab2Δab,i2Δab,i2¯. [7]

We probe the dynamics with two-time mean-squared displacement (MSD),

Δ(tw,t)=1Nsα=1Ns1NcaαNcΔa(tw,t),Δa(tw,t)=1Ni=1N|ria(tw+t)ria(tw)|2,

with tw the initial time for the observation and t the time window.

Supplementary Material

Appendix 01 (PDF)

Acknowledgments

We thank M. Ozawa and F. Zamponi for helpful discussions. The research has been supported by the National Natural Science Foundation of China Grants (Nos. 11734014, 11975295, and 12047503) and the Chinese Academy of Sciences Grants (Nos. QYZDJ-SSW-SYS018 and XDPD15). L.B. acknowledges support from the Simons Foundation (No. 454933 L.B.). We also thank the Supercomputing Center of University of Science and Technology of China for the computer time.

Author contributions

Q.L., L.B., H.-J.Z., and N.X. designed research; Q.L. performed research; Q.L. and L.B. contributed new reagents/analytic tools; Q.L., L.B., H.-J.Z., and N.X. analyzed data; and Q.L., L.B., H.-J.Z., and N.X. wrote the paper.

Competing interests

The authors declare no competing interest.

Footnotes

This article is a PNAS Direct Submission.

Contributor Information

Qinyi Liao, Email: qinyi.liao.phy@gmail.com.

Ludovic Berthier, Email: ludovic.berthier@umontpellier.fr.

Hai-Jun Zhou, Email: zhouhj@itp.ac.cn.

Ning Xu, Email: ningxu@ustc.edu.cn.

Data, Materials, and Software Availability

All study data are included in the article and/or SI Appendix.

Supporting Information

References

  • 1.Liu A. J., Nagel S. R., Nonlinear dynamics: Jamming is not just cool any more. Nature 396, 21 (1998). [Google Scholar]
  • 2.O’Hern C. S., Silbert L. E., Liu A. J., Nagel S. R., Jamming at zero temperature and zero applied stress: The epitome of disorder. Phys. Rev. E 68, 011306 (2003). [DOI] [PubMed] [Google Scholar]
  • 3.Liu A. J., Nagel S. R., The jamming transition and the marginally jammed solid. Annu. Rev. Condens. Matter Phys. 1, 347–369 (2010). [Google Scholar]
  • 4.van Hecke M., Jamming of soft particles: Geometry, mechanics, scaling and isostaticity. J. Phys.: Condens. Matter 22, 033101 (2009). [DOI] [PubMed] [Google Scholar]
  • 5.Bernal J., Mason J., Knight K., Radial distribution of the random close packing of equal spheres. Nature 194, 957–958 (1962). [Google Scholar]
  • 6.Donev A., Torquato S., Stillinger F. H., Pair correlation function characteristics of nearly jammed disordered and ordered hard-sphere packings. Phys. Rev. E 71, 011105 (2005). [DOI] [PubMed] [Google Scholar]
  • 7.Xu N., Vitelli V., Liu A. J., Nagel S. R., Anharmonic and quasi-localized vibrations in jammed solids-modes for mechanical failure. Europhys. Lett. (EPL) 90, 56001 (2010). [Google Scholar]
  • 8.Bi D., Yang X., Marchetti M. C., Manning M. L., Motility-driven glass and jamming transitions in biological tissues. Phys. Rev. X 6, 021011 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Geiger M., et al. , Jamming transition as a paradigm to understand the loss landscape of deep neural networks. Phys. Rev. E 100, 012115 (2019). [DOI] [PubMed] [Google Scholar]
  • 10.Müller M., Wyart M., Marginal stability in structural, spin, and electron glasses. Annu. Rev. Condens. Matter Phys. 6, 177 (2015). [Google Scholar]
  • 11.Edwards S., Oakeshott R., Theory of powders. Physica A 157, 1080–1090 (1989). [Google Scholar]
  • 12.Henkes S., Chakraborty B., Statistical mechanics framework for static granular matter. Phys. Rev. E 79, 061301 (2009). [DOI] [PubMed] [Google Scholar]
  • 13.Parisi G., Urbani P., Zamponi F., Theory of Simple Glasses: Exact Solutions in Infinite Dimensions (Cambridge University Press, 2020). [Google Scholar]
  • 14.Berthier L., Witten T. A., Glass transition of dense fluids of hard and compressible spheres. Phys. Rev. E 80, 021502 (2009). [DOI] [PubMed] [Google Scholar]
  • 15.Krzakala F., Zdeborová L., Following Gibbs states adiabatically—The energy landscape of mean-field glassy systems. Eur. Lett. 90, 66002 (2010). [Google Scholar]
  • 16.Rainone C., Urbani P., Yoshino H., Zamponi F., Following the evolution of hard sphere glasses in infinite dimensions under external perturbations: Compression and shear strain. Phys. Rev. Lett. 114, 015701 (2015). [DOI] [PubMed] [Google Scholar]
  • 17.Charbonneau P., Kurchan J., Parisi G., Urbani P., Zamponi F., Exact theory of dense amorphous hard spheres in high dimension. III. The full replica symmetry breaking solution. J. Stat. Mech.: Theory Exp. 2014, P10009 (2014). [Google Scholar]
  • 18.Berthier L., et al. , Gardner physics in amorphous solids and beyond. J. Chem. Phys. 151, 010901 (2019). [DOI] [PubMed] [Google Scholar]
  • 19.Gardner E., Spin glasses with p-spin interactions. Nuclear Phys. B 257, 747–765 (1985). [Google Scholar]
  • 20.Gross D. J., Kanter I., Sompolinsky H., Mean-field theory of the Potts glass. Phys. Rev. Lett. 55, 304–307 (1985). [DOI] [PubMed] [Google Scholar]
  • 21.Charbonneau P., Kurchan J., Parisi G., Urbani P., Zamponi F., Fractal free energy landscapes in structural glasses. Nat. Commun. 5, 3725 (2014). [DOI] [PubMed] [Google Scholar]
  • 22.Kurchan J., Parisi G., Urbani P., Zamponi F., Exact theory of dense amorphous hard spheres in high dimension. II. The high density regime and the Gardner transition. J. Phys. Chem. B 117, 12979–12994 (2013). [DOI] [PubMed] [Google Scholar]
  • 23.Yoshino H., Zamponi F., Shear modulus of glasses: Results from the full replica-symmetry-breaking solution. Phys. Rev. E 90, 022302 (2014). [DOI] [PubMed] [Google Scholar]
  • 24.Franz S., Parisi G., Urbani P., Zamponi F., Universal spectrum of normal modes in low-temperature glasses. Proc. Natl. Acad. Sci. U.S.A. 112, 14539–14544 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.A. Montanari, “Optimization of the Sherrington-Kirkpatrick Hamiltonian” in 2019 IEEE 60th Annual Symposium on Foundations of Computer Science (FOCS), (IEEE, 2019) pp. 1417–1433.
  • 26.Alaoui A. E., Montanari A., Sellke M., Optimization of mean-field spin glasses. Ann. Probab. 49, 2922 (2021). [Google Scholar]
  • 27.Yoshino H., From complex to simple: Hierarchical free-energy landscape renormalized in deep neural networks. SciPost Phys. Core 2, 005 (2020). [Google Scholar]
  • 28.Pinto D. E., Sussman D. M., da Gama M. M. T., Araújo N. A., Hierarchical structure of the energy landscape in the Voronoi model of dense tissue. Phys. Rev. Res. 4, 023187 (2022). [Google Scholar]
  • 29.Kumar A., Pandey D., Study of magnetic relaxation, memory and rejuvenation effects in the cluster spin-glass phase of b-site disordered Ca(Fe1/2Nb1/2)O3 perovskite: Experimental evidence for hierarchical model. J. Magnet. Magnet. Mater. 511, 166964 (2020). [Google Scholar]
  • 30.Charbonneau P., Corwin E. I., Fu L., Tsekenis G., van Der Naald M., Glassy, Gardner-like phenomenology in minimally polydisperse crystalline systems. Phys. Rev. E 99, 020901 (2019). [DOI] [PubMed] [Google Scholar]
  • 31.Kool L., Charbonneau P., Daniels K. E., Gardner-like crossover from variable to persistent force contacts in granular crystals. Phys. Rev. E 106, 054901 (2022). [DOI] [PubMed] [Google Scholar]
  • 32.Urbani P., Biroli G., Gardner transition in finite dimensions. Phys. Rev. B 91, 100202 (2015). [Google Scholar]
  • 33.Charbonneau P., Yaida S., Nontrivial critical fixed point for replica-symmetry-breaking transitions. Phys. Rev. Lett. 118, 215701 (2017). [DOI] [PubMed] [Google Scholar]
  • 34.Barrat A., Franz S., Parisi G., Temperature evolution and bifurcations of metastable states in mean-field spin glasses, with connections with structural glasses. J. Phys. A: Math. General 30, 5593 (1997). [Google Scholar]
  • 35.Marinari E., Parisi G., Ricci-Tersenghi F., Ruiz-Lorenzo J. J., Off-equilibrium dynamics at very low temperatures in three-dimensional spin glasses. J. Phys. A: Math. General 33, 2373 (2000). [Google Scholar]
  • 36.Montanari A., Ricci-Tersenghi F., On the nature of the low-temperature phase in discontinuous mean-field spin glasses. Eur. Phys. J. B 33, 339–346 (2003). [Google Scholar]
  • 37.Rizzo T., Replica-symmetry-breaking transitions and off-equilibrium dynamics. Phys. Rev. E 88, 032135 (2013). [DOI] [PubMed] [Google Scholar]
  • 38.Altieri A., Biroli G., Cammarota C., Dynamical mean-field theory and aging dynamics. J. Phys. A: Math. Theor. 53, 375006 (2020). [Google Scholar]
  • 39.Manacorda A., Schehr G., Zamponi F., Numerical solution of the dynamical mean field theory of infinite-dimensional equilibrium liquids. J. Chem. Phys. 152, 164506 (2020). [DOI] [PubMed] [Google Scholar]
  • 40.Verstraten R. C., Ozela R. F., Smith C. M., Time glass: A fractional calculus approach. Phys. Rev. B 103, L180301 (2021). [Google Scholar]
  • 41.Sibani P., Boettcher S., Jensen H. J., Record dynamics of evolving metastable systems: Theory and applications. Eur. Phys. J. B 94, 1–23 (2021). [Google Scholar]
  • 42.Lubchenko V., Wolynes P. G., Aging, jamming, and the limits of stability of amorphous solids. J. Phys. Chem. B 122, 3280–3295 (2017). [DOI] [PubMed] [Google Scholar]
  • 43.N. Oyama, H. Mizuno, A. Ikeda, Shear-induced criticality in glasses shares qualitative similarities with the Gardner phase. arXiv [Preprint] (2021). http://arxiv.org/abs/2109.08849 (Accessed 18 September 2021). [DOI] [PubMed]
  • 44.Berthier L., et al. , Growing timescales and lengthscales characterizing vibrations of amorphous solids. Proc. Natl. Acad. Sci. U.S.A. 113, 8397–8401 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Seguin A., Dauchot O., Experimental evidence of the Gardner phase in a granular glass. Phys. Rev. Lett. 117, 228001 (2016). [DOI] [PubMed] [Google Scholar]
  • 46.Seoane B., Zamponi F., Spin-glass-like aging in colloidal and granular glasses. Soft Matter 14, 5222 (2018). [DOI] [PubMed] [Google Scholar]
  • 47.Xiao H., Liu A. J., Durian D. J., Probing Gardner physics in an active quasithermal pressure-controlled granular system of noncircular particles. Phys. Rev. Lett. 128, 248001 (2022). [DOI] [PubMed] [Google Scholar]
  • 48.Albert S., Biroli G., Ladieu F. M. C., Tourbot R., Urbani P., Searching for the Gardner transition in glassy glycerol. Phys. Rev. Lett. 126, 028001 (2021). [DOI] [PubMed] [Google Scholar]
  • 49.Wang Y., Shang J., Jin Y., Zhang J., Experimental observations of marginal criticality in granular materials. Proc. Natl. Acad. Sci. U.S.A. 119, e2204879119 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Liao Q., Berthier L., Hierarchical landscape of hard disk glasses. Phys. Rev. X 9, 011049 (2019). [Google Scholar]
  • 51.Scalliet C., Berthier L., Zamponi F., Nature of excitations and defects in structural glasses. Nat. Commun. 10, 1–10 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Artiaco C., Baldan P., Parisi G., Exploratory study of the glassy landscape near jamming. Phys. Rev. E 101, 052605 (2020). [DOI] [PubMed] [Google Scholar]
  • 53.Jin Y., Urbani P., Zamponi F., Yoshino H., A stability-reversibility map unifies elasticity, plasticity, yielding and jamming in hard sphere glasses. Sci. Adv. 4, eaat6387 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Charbonneau P., Morse P. K., Memory formation in jammed hard spheres. Phys. Rev. Lett. 126, 088001 (2021). [DOI] [PubMed] [Google Scholar]
  • 55.Shang B., Guan P., Barrat J. L., Elastic avalanches reveal marginal behavior in amorphous solids. Proc. Natl. Acad. Sci. U.S.A. 117, 86–92 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Geirhos K., Lunkenheimer P., Loidl A., Johari–Goldstein relaxation far below Tg: Experimental evidence for the Gardner transition in structural glasses? Phys. Rev. Lett. 120, 085705 (2018). [DOI] [PubMed] [Google Scholar]
  • 57.Hammond A. P., Corwin E. I., Experimental observation of the marginal glass phase in a colloidal glass. Proc. Natl. Acad. Sci. U.S.A. 117, 5714–5718 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Illing B., et al. , Mermin–Wagner fluctuations in 2D amorphous solids. Proc. Natl. Acad. Sci. U.S.A. 114, 1856–1861 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Shiba H., Yamada Y., Kawasaki T., Kim K., Unveiling dimensionality dependence of glassy dynamics: 2D infinite fluctuation eclipses inherent structural relaxation. Phys. Rev. Lett. 117, 245701 (2016). [DOI] [PubMed] [Google Scholar]
  • 60.Vivek S., Kelleher C. P., Chaikin P. M., Weeks E. R., Long-wavelength fluctuations and the glass transition in two dimensions and three dimensions. Proc. Natl. Acad. Sci. U.S.A. 114, 1850–1855 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Berthier L., Coslovich D., Ninarello A., Ozawa M., Equilibrium sampling of hard spheres up to the jamming density and beyond. Phys. Rev. Lett. 116, 238002 (2016). [DOI] [PubMed] [Google Scholar]
  • 62.Scalliet C., Berthier L., Rejuvenation and memory effects in a structural glass. Phys. Rev. Lett. 122, 255502 (2019). [DOI] [PubMed] [Google Scholar]
  • 63.Cugliandolo L. F., Kurchan J., Analytical solution of the off-equilibrium dynamics of a long-range spin-glass model. Phys. Rev. Lett. 71, 173–176 (1993). [DOI] [PubMed] [Google Scholar]
  • 64.Götze W., Recent tests of the mode-coupling theory for glassy dynamics. J. Phys.: Condens. Matter 11, A1 (1999). [Google Scholar]
  • 65.Mansoori G., Carnahan N. F., Starling K., Leland T. Jr., Equilibrium thermodynamic properties of the mixture of hard spheres. J. Chem. Phys. 54, 1523–1525 (1971). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Appendix 01 (PDF)

Data Availability Statement

All study data are included in the article and/or SI Appendix.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES