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

Computer simulations of the Gardner transition in structural glasses

Yuliang Jin CAS Key Laboratory of Theoretical Physics, Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, China School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China Wenzhou Institute, University of Chinese Academy of Sciences, Wenzhou, Zhejiang 325000, China    Hajime Yoshino Cybermedia Center, Osaka University, Toyonaka, Osaka 560-0043, Japan Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan

I Connections between Gardner and spin-glass transitions

The exact mean-field theory for the simplest glass-forming system - the dense assembly of hard spheres in the large dimensional limit - predicts the existence of a Gardner phase Parisi et al. (2020); Berthier et al. (2019). This transition is characterized by full replica symmetry breaking (RSB) that implies two fascinating physical consequences. (i) A hierarchical free-energy landscape, i.e., the thermal fluctuations are organized hierarchically, meaning that configurations are grouped into meta-basins that are further grouped into meta-meta basins, … (ii) The marginal stability, i.e., the system responds sensitively to infinitesimal perturbations. Here we discuss recent results of numerical simulations to examine these mean-field predictions in physical dimensions.

From the viewpoint of RSB, the Gardner transition in structural glasses belongs to the same full RSB universality class of the spin-glass transition (see Fig. 1(A)). This theoretical ground motivates us to borrow ideas from the extensive research on spin-glasses to study the Gardner transition. To this end, it is useful to review firstly some of the essential results obtained in spin-glass experiments and simulations.

The RSB solution immediately implies a hierarchy of linear responses through the fluctuation-dissipation relation Mézard et al. (1987). One expects short-time, intermediate-time, and long-time linear responses associated with thermal fluctuations inside basins, meta-basins, and meta-meta-basins. A remarkable consequence is the “anomaly” that gives a natural explanation for the protocol-dependent linear responses observed experimentally  Nagata et al. (1979). In one protocol called field cooling (FC), one measures the magnetization mFCm_{\rm FC} of a spin-glass under cooling from a high temperature TmaxT_{\rm max} down to a low temperature TminT_{\rm min} below the spin-glass transition temperature TSGT_{\rm SG} in the presence of a weak external magnetic field δh\delta h; in the other protocol called zero field cooling (ZFC), one cools the spin-glass from TmaxT_{\rm max} down to TminT_{\rm min} without the field (h=0h=0), then switches on the magnetic field δh\delta h and measures the magnetization mZFCm_{\rm ZFC} under heating the spin-glass back to TmaxT_{\rm max} (see Fig. 1(B)). The two susceptibilities χFC=mFC/δh\chi_{\rm FC}=m_{\rm FC}/\delta h and χZFC=mZFC/δh\chi_{\rm ZFC}=m_{\rm ZFC}/\delta h are the same above TSGT_{\rm SG}, but different (χFC>χZFC\chi_{\rm FC}>\chi_{\rm ZFC}) below (see Fig. 2(A)). The fact that χFC\chi_{\rm FC} and χZFC\chi_{\rm ZFC} are different in the spin-glass phase is referred to as an “anomaly”, because the susceptibility is protocol-independent in standard magnetic systems. The RSB theory gives χFCχZFC=β[01𝑑qP(q)qqEA]\chi_{\rm FC}-\chi_{\rm ZFC}=\beta\left[\int_{0}^{1}dqP(q)q-q_{\rm EA}\right], with β\beta the inverse temperature. Here qEAq_{\rm EA} is the Edwards-Anderson (EA) order parameter Edwards and Anderson (1975) representing the strength of the thermal fluctuation within lowest-level basins, while 01𝑑qP(q)q\int_{0}^{1}dqP(q)q represents the integral of thermal fluctuations coming from all levels in the hierarchy. In the replica symmetric (RS) solution, P(q)=δ(qqEA)P(q)=\delta(q-q_{\rm EA}), so that the anomaly vanishes.

Refer to caption

Figure 1: (A) Correspondence between spin and structural glasses. Note that one step RSB (1-RSB) exists in certain spin-glass models such as the spherical p-spin (p>2p>2) model Castellani and Cavagna (2005). (B) Schematic of ZFC/FC protocols. In spin and hard-sphere glasses, the control parameter is the temperature TT and the density (volume fraction) φ\varphi respectively, and the external field is the magnetic field hh and the shear-strain γ\gamma respectively.

The anomaly is known to be not a transient but a long-time effect, as demonstrated by a series of experiments that reveal aging effects in spin-glasses Granberg et al. (1988); Vincent et al. (1997); Nordblad and Svedlindh (1998). To study the dynamical effects, one can generalize the ZFC protocol by introducing a waiting time twt_{\rm w} before switching on the magnetic field, and a measurement time τ\tau elapsed in the presence of the field. By increasing τ\tau, mZFC(τ,tw)m_{\rm ZFC}(\tau,t_{\rm w}) increases passing through mZFCm_{\rm ZFC} and heads toward mFCm_{\rm FC} (see Fig. 2(B)). However, mZFC(τ,tw)m_{\rm ZFC}(\tau,t_{\rm w}) does not reach mFCm_{\rm FC} within finite time, and its time evolution as a function of τ\tau slows down with increasing waiting time twt_{\rm w}, manifested by scaling laws depending on τ/tw\tau/t_{\rm w}. These experimental observations are significant because they reveal the out-of-equilibrium nature of spin-glasses. To describe the aging effects and the anomaly from a purely dynamical point of view, a dynamical mean-field theory on spin-glass models is developed Cugliandolo and Kurchan (1993, 1994). The dynamical theory relates RSB to the notion of effective temperature that characterizes out-of equilibrium glassy dynamics Cugliandolo et al. (1997); Franz et al. (1998). The numerical evidence of effective temperature Marinari et al. (2000) and non-zero anomaly in the long-time limit Yoshino et al. (2002) has been indicated by detailed simulations of finite-dimensional spin-glass models.

The marginal stability of the spin-glass phase may account for various complex non-linear responses, such as the effect of static chaos with respect to an infinitesimal change of temperature, or avalanches with respect to an infinitesimal change of magnetic field. Indeed the equilibrium spin configurations at large length scales are completely reshuffled by infinitesimal perturbations, which is predicted first by the droplet theory Bray and Moore (1987); Fisher and Huse (1988a, b), and later by theories based on RSB Kondor (1989); Rizzo and Crisanti (2003); Rizzo and Yoshino (2006); Yoshino and Rizzo (2008); Parisi and Rizzo (2010); Le Doussal et al. (2010, 2012); Franz and Spigler (2017). The rejuvenation-memory effects observed experimentally Jonason et al. (1998) may be related to such non-linear responses  Yoshino et al. (2001); Jönsson et al. (2004).

Once one is aware of the correspondence between spin and structural glasses (see Fig. 1(A)), it is natural to use strategies inherited from spin-glass studies to explore the physics of Gardner phase in structural glasses. For example, in the ZFC/FC protocols, the role of the magnetic field hh for spin-glasses can be replaced by the shear strain γ\gamma for structural glasses (see Fig. 1(B)), which only changes the boundary condition but not the thermodynamic properties of the bulk. Indeed, the replica theory of structural glasses predicts a hierarchy of shear moduli reflecting RSB Yoshino and Mézard (2010); Yoshino (2012); Yoshino and Zamponi (2014). By adapting the methods developed in spin-glasses, one can examine the aging effects, the protocol-dependent linear responses, and the non-linear responses such as avalanches, in structural glasses with respect to shear deformations. Our discussion focuses on one of the simplest models of structural glasses in three dimensions, hard spheres, where the (reduced) pressure pp (or the volume fraction φ\varphi) plays the role of temperature TT. According to the replica theory Parisi et al. (2020) that is exact in the large dimensional limit, a Gardner transition occurs in hard spheres under both compression and shear, which is examined by simulations at three-dimensions in the following sections. The dynamical mean-field theory for the hard-sphere glass has also been set up Maimbourg et al. (2016); Kurchan et al. (2016); Agoritsas et al. (2019a, b), but a detailed theoretical analysis of the out-of equilibrium dynamics remains challenging. Nonetheless, the analogy to the spin-glass problem outlined above allows us to infer the implications of RSB on the dynamics of hard spheres.

Refer to caption

Figure 2: Experimental results on CuMn spin-glasses. (A) ZFC/FC susceptibilities (adapted from Nagata et al. (1979)), and (B) the time evolution of mZFC(τ,tw)/δhm_{\rm ZFC}(\tau,t_{\rm w})/\delta h in an aging experiment (adapted from Granberg et al. (1988)).

II Gardner transition under compression

II.1 Preparation of ultra-stable glasses

To study the Gardner transition, we must prepare a glass at first. Experimentally, glasses are obtained by a slow thermal or compression annealing, the rate of which determines the location of the glass transition. It is found that a detailed numerical analysis of the Gardner transition requires the preparation of extremely well-relaxed glasses (corresponding to structural relaxation timescales challenging to simulate in standard algorithms), in order to study vibrational motions of particles without interference from diffusion. Such ultra-stable glasses can be numerically generated by applying a swap Monte-Carlo scheme Kranendonk and Frenkel (1991); Grigera and Parisi (2001) to a simple glass-forming model – a polydisperse mixture of NN hard spheres Berthier et al. (2016).

The annealing procedure contains two steps Berthier et al. (2016). First, one produces equilibrated liquid configurations at various densities φg\varphi_{\rm g} with the help of the swap algorithm. Second, starting from these liquid configurations, one switches to standard molecular dynamics simulations Lubachevsky and Stillinger (1990) during which the system is compressed out of equilibrium up to target densities φ>φg\varphi>\varphi_{\rm g}. In order to obtain thermal and disorder averaging, this procedure is repeated over many samples, each corresponding to different initial equilibrium configurations at φg\varphi_{\rm g}, and over many independent quench realizations for each sample. The independent realizations of the same sample have identical particle positions at φg\varphi_{\rm g}, but are assigned to different initial velocities drawn from the Maxwell–Boltzmann distribution.

The above numerical protocol is analogous to thermal annealing with different cooling rates, which results in different glass transition temperatures. Each glass transition density φg\varphi_{\rm g} selects a particular glass state. The value of φg\varphi_{\rm g} ranges from the mode-coupling theory (MCT) density (or the dynamical glass transition density) φd\varphi_{\rm d}, at which the liquid relaxation is slow but affected by activated α\alpha-processes, to φgφd\varphi_{\rm g}\gg\varphi_{\rm d}, where particle diffusion and vibrations are fully separated. For sufficiently large φg\varphi_{\rm g}, the α\alpha-relaxation time becomes larger than the simulation time by many orders of magnitude; one thus obtains unimpeded access to the dynamics within the glass state, i.e., the β\beta-relaxation processes Goldstein (2010).

Refer to caption


Figure 3: Phase diagram of a polydisperse hard-sphere glass in three dimensions (adapted from Berthier et al. (2016)). Solids lines represent the CS liquid EOS and the Gardner line, and dashed lines represent glass EOSs. The insets show typical particle motions in three phases.

The liquid equation of state (EOS) for the reduced pressure p=βP/ρp=\beta P/\rho of the model, where ρ\rho is the number density, and PP the system pressure, can be well described by the Carnahan-Starling (CS) equation Boublik (1970). The dynamical glass transition density φd=0.594(1)\varphi_{\rm d}=0.594(1) was estimated following the strategy in Ref. Charbonneau et al. (2014a). Note that the dynamical glass transition is only rigorous in large dimensions; it becomes a dynamical crossover in three dimensions (see Chapter 16 for a detailed discussion). The non-equilibrium glass EOSs associated with compression terminate at inherent states (where pp\to\infty) that correspond to, for hard spheres, jammed configurations at φJ\varphi_{\rm J}, and can be captured by a free-volume scaling form, pglass(φ)(φJφ)p_{\rm glass}(\varphi)\sim(\varphi_{\rm J}-\varphi) Donev et al. (2005). Figure 3 presents the phase diagram and EOSs of the model.

Along each glass EOS of a given φg\varphi_{\rm g}, a corresponding Gardner transition may exist at density φG\varphi_{\rm G} (or pressure pGp_{\rm G}), as predicted by the mean-field theory. For φg<φ<φG\varphi_{\rm g}<\varphi<\varphi_{\rm G}, the system is in a stable glass phase: each glass state is confined in one of the structureless basins on the free-energy landscape (see Fig. 1(A)), and is stable in response to small mechanical deformations. On the other hand, the regime φG<φ<φJ\varphi_{\rm G}<\varphi<\varphi_{\rm J} corresponds to a marginal glass phase, where each simple glass basin splits into a fractal hierarchy of sub-basins and the glass becomes marginally stable to deformations. The Gardner line and the liquid EOS merge around φd\varphi_{\rm d}, suggesting the mixing of dynamical behavior associated to the Gardner transition and to the glass transition – this is why one needs to focus on ultra-stable glasses in order to explore pure Gardner physics.

II.2 Key observables and protocols

In the glass state, particles vibrate inside their cages (see the insets of Fig. 3). The first approach to study the Gardner transition is based on the direct analysis of caging order parameters, which quantify the caging properties of particles. In the case of stable glasses, the caging order parameter is defined as, ΔEA=limt1Ni=1N|ri(t)ri(0)|2\Delta_{\rm EA}=\lim_{t\to\infty}\frac{1}{N}\sum_{i=1}^{N}\langle|\vec{r}_{i}(t)-\vec{r}_{i}(0)|^{2}\rangle, where ri(t)\vec{r}_{i}(t) is the position of particle ii at time tt. The parameter ΔEA\Delta_{\rm EA}, which decreases with the degree of annealing, corresponds to nothing but the EA parameter qEAq_{\rm EA} in spin-glasses.

Similar to the spin-glass transition, the Gardner transition induces the split of basins on the free-energy landscape and aging effects, which suggests that the order parameter must be generalized: one considers (i) the mean-squared displacement (MSD) Δ(τ,tw)\Delta(\tau,t_{\rm w}) and (ii) the distance between pairs of independently quenched configurations ΔAB(t)\Delta_{AB}(t). Here the MSD is defined as, Δ(τ,tw)=1Ni=1N|ri(τ+tw)ri(tw)|2\Delta(\tau,t_{\rm w})=\frac{1}{N}\sum_{i=1}^{N}\langle|\vec{r}_{i}(\tau+t_{\rm w})-\vec{r}_{i}(t_{\rm w})|^{2}\rangle, averaged over both thermal fluctuations and disorder, at the target φ\varphi reached by compression. A waiting time twt_{\rm w} is introduced in order to explicitly examine the aging effects (the total time tt is the sum of the measurement time τ\tau and twt_{\rm w}). On the other hand, ΔAB(t)=1Ni=1N|riA(t)riB(t)|2\Delta_{AB}(t)=\frac{1}{N}\sum_{i=1}^{N}\langle|\vec{r}_{i}^{A}(t)-\vec{r}_{i}^{B}(t)|^{2}\rangle, where the two copies AA and BB are independent realizations at φ\varphi, compressed from the same initial sample at φg\varphi_{\rm g}.

The large-time limits of these quantities have important physical meanings. The EA order parameter is defined as ΔEAlimτlimtwΔ(τ,tw)\Delta_{\rm EA}\equiv\lim_{\tau\to\infty}\lim_{t_{\rm w}\to\infty}\Delta(\tau,t_{\rm w}). Here the order of time limits is crucial Bouchaud et al. (1998): by reversing the order one can define another parameter, ΔABlimtwlimτΔ(τ,tw)=limtΔAB(t)\Delta_{AB}\equiv\lim_{t_{\rm w}\to\infty}\lim_{\tau\to\infty}\Delta(\tau,t_{\rm w})=\lim_{t\to\infty}\Delta_{AB}(t). The RSB is signaled by ΔAB>ΔEA\Delta_{\rm AB}>\Delta_{\rm EA} (note that ΔAB=ΔEA\Delta_{AB}=\Delta_{\rm EA} in stable glasses). In other words, the two large-time limits cannot be interchanged in the Gardner phase, meaning that the aging effects become persistent.

In the second approach, one studies the response of hard-sphere glasses against a shear strain γ\gamma, analogous to observing magnetic susceptibilities in spin-glasses. The simple strain γ\gamma is applied to the xx-coordinates of all particles (xixi+γzix_{i}\to x_{i}+\gamma z_{i}) after a waiting time twt_{\rm w}, under the constant-volume and Lees-Edwards boundary conditions Lees and Edwards (1972). The strain is increased slowly with a constant shear rate γ˙\dot{\gamma}, and the reduced shear stress σ=βΣ/ρ\sigma=\beta\Sigma/\rho is measured, where Σ\Sigma is the stress (for convenience, some data are presented with the unitless stress rescaled by pressure, σ~=σ/p\tilde{\sigma}=\sigma/p).

As in the spin-glass case, one can consider two types of protocols, namely zero field compression (ZFC) and field compression (FC) (see Fig. 1 (B)). In the ZFC protocol, one compresses the configuration from φg\varphi_{\rm g} to φ\varphi, waits for time twt_{\rm w} before applying a shear strain δγ\delta\gamma instantaneously, and then measures the stress σZFC(τ,tw)\sigma_{\rm ZFC}(\tau,t_{\rm w}) as a function of τ\tau. In the FC protocol, one applies δγ\delta\gamma at the initial density φg\varphi_{\rm g}, and then measures the stress σFC(t)\sigma_{\rm FC}(t) once the configuration is compressed to φ\varphi (tt is reset to zero after compression). Similar to the caging order parameters, two large-time limits can be considered: σZFClimτlimtwσZFC(τ,tw)\sigma_{\rm ZFC}\equiv\lim_{\tau\to\infty}\lim_{t_{\rm w}\to\infty}\sigma_{\rm ZFC}(\tau,t_{\rm w}) and σFClimtwlimτσZFC(τ,tw)\sigma_{\rm FC}\equiv\lim_{t_{\rm w}\to\infty}\lim_{\tau\to\infty}\sigma_{\rm ZFC}(\tau,t_{\rm w}).

Theories have demonstrated that the above two approaches (more specifically, the caging order parameters Δ\Delta and the shear moduli μ=σ/δγ\mu=\sigma/\delta\gamma) are intrinsically related  Yoshino and Zamponi (2014): in the large pressure limit, μZFC1/ΔEA\mu_{\rm ZFC}\sim 1/\Delta_{\rm EA} and μFC/p1/ΔAB\mu_{\rm FC}/p\sim 1/\Delta_{AB}. These relationships are the counterpart of the duality between overlapping order parameters and magnetic susceptibilities in spin-glasses.

II.3 Aging effects

Refer to caption


Figure 4: Time evolutions of (A) caging order parameters (adapted from Berthier et al. (2016)) and (B) shear moduli (adapted from Jin and Yoshino (2017)), in the Gardner phase of the hard-sphere glass model.

In the Gardner phase (φ>φG\varphi>\varphi_{\rm G}), aging effects can be observed in both MSD (without shear deformations) and shear responses. Figure 4(A) shows the simulation data of MSD. After a short time τb1\tau_{\rm b}\sim 1 of ballistic motions, the evolution of Δ(τ,tw)\Delta(\tau,t_{\rm w}), as a function of τ\tau, exhibits a plateau followed by further growth. The switch from the former to the latter happens at longer times with increasing waiting time twt_{\rm w}. The height of the short-time plateau gives ΔEA\Delta_{\rm EA} (practically, we set ΔEA=Δ(τ=τb,tw=0)\Delta_{\rm EA}=\Delta(\tau=\tau_{\rm b},t_{\rm w}=0)). Figure 4(A) also displays ΔAB(t)\Delta_{AB}(t), which is time-independent and should correspond to a long-time plateau of Δ(τ,tw)\Delta(\tau,t_{\rm w}) (this plateau is unfortunately beyond the current simulation time window). The clear separation of the two parameters (ΔAB>ΔEA\Delta_{AB}>\Delta_{\rm EA}) is the first numerical evidence of the ergodicity breaking in the Gardner phase Charbonneau et al. (2015a); Berthier et al. (2016); Seoane and Zamponi (2018).

Figure 4(B) shows the time-dependent (unitless) shear moduli, μ~(τ,tw)\tilde{\mu}(\tau,t_{\rm w}), whose behavior is similar to that of MSD. An important feature is that μ~ZFC(τ,tw)\tilde{\mu}_{\rm ZFC}(\tau,t_{\rm w}) exhibits a plateau suggesting the existence of μ~ZFC\tilde{\mu}_{\rm ZFC}. On the other hand, μ~FC(t)\tilde{\mu}_{\rm FC}(t) is essentially a constant in time tt (for t>τbt>\tau_{\rm b}), which shall be denoted as μ~FC\tilde{\mu}_{\rm FC}. In the proper order of large-time limits, one expects that μ~ZFC(τ,tw)\tilde{\mu}_{\rm ZFC}(\tau,t_{\rm w}) decays to μ~FC\tilde{\mu}_{\rm FC}, as limtwlimτμ~ZFC(τ,tw)=μ~FC\lim_{t_{\rm w}\to\infty}\lim_{\tau\to\infty}\tilde{\mu}_{\rm ZFC}(\tau,t_{\rm w})=\tilde{\mu}_{\rm FC}, but the convergence becomes slower as twt_{\rm w} increases. Apparently μ~ZFC\tilde{\mu}_{\rm ZFC} is larger than μ~FC\tilde{\mu}_{\rm FC}, which parallels ΔAB>ΔEA\Delta_{AB}>\Delta_{\rm EA}.

II.4 Anomalous order parameters and responses

Figure 5 shows the pressure dependence of caging order parameters (ΔEA\Delta_{\rm EA} and ΔAB\Delta_{AB}) and shear moduli (μ~ZFC\tilde{\mu}_{\rm ZFC} and μ~FC\tilde{\mu}_{\rm FC}) obtained through the above-mentioned dynamic measurements. One finds that, in the stable glass phase (p<pGp<p_{\rm G}), ΔEA=ΔAB\Delta_{\rm EA}=\Delta_{\rm AB} and μ~ZFC=μ~FC\tilde{\mu}_{\rm ZFC}=\tilde{\mu}_{\rm FC}, while in the Gardner phase (p>pGp>p_{\rm G}), ΔEA<ΔAB\Delta_{\rm EA}<\Delta_{\rm AB} and μ~ZFC>μ~FC\tilde{\mu}_{\rm ZFC}>\tilde{\mu}_{\rm FC}. In the large pressure limit, mean-field theories predict that ΔEApκ\Delta_{\rm EA}\sim p^{-\kappa} Charbonneau et al. (2014b) and μZFCpκ\mu_{\rm ZFC}\sim p^{\kappa} Yoshino and Zamponi (2014), where κ=1.41574\kappa=1.41574. The former is verified by three-dimensional simulations in Ref. Charbonneau et al. (2014b) and the latter by those in Ref. Jin and Yoshino (2017) (see Fig. 5(B)). The theories also give large-pp predictions μFC/p1/ΔABconstant\mu_{\rm FC}/p\sim 1/\Delta_{AB}\sim\rm{constant}, which are consistent with the simulation results in Fig. 5.

Refer to caption

Figure 5: (A) Bifurcation of caging order parameters ΔEA\Delta_{\rm EA} and ΔAB\Delta_{AB} around the Gardner transition pG2.7×102p_{\rm G}\approx 2.7\times 10^{2} (adapted from Berthier et al. (2016)). (B) Protocol-dependent shear moduli μ~ZFC\tilde{\mu}_{\rm ZFC} and μ~FC\tilde{\mu}_{\rm FC} (adapted from Jin and Yoshino (2017)). The solid line indicates the scaling μZFCp1.41574\mu_{\rm ZFC}\sim p^{1.41574} predicted by the mean-field theory Yoshino and Zamponi (2014).

III Gardner transition under shear

As predicted theoretically Rainone et al. (2015), a Gardner transition at γG\gamma_{\rm G} could occur under shear, before the glass yields at γY\gamma_{\rm Y}. Figure 6(A) shows the stability map of hard sphere glasses under shear and compression/decompression Jin et al. (2018); Altieri and Zamponi (2019). The Gardner transition and yielding give rise to three types of behavior in a typical cyclic shear test (see Fig. 6(B)). (i) The stress-strain curve is reversible in the stable glass phase (γ<γG\gamma<\gamma_{\rm G}). (ii) If the shear strain is reversed at a maximum strain γmax\gamma_{\rm max} between γG\gamma_{\rm G} and γY\gamma_{\rm Y}, a hysteresis loop emerges, which however disappears below γG\gamma_{\rm G}. This partial-reversible phenomenon is a manifestation of the hierarchical free-energy landscape consisting of basins within a common meta-basin. The part of the stress-strain curve in the Gardner phase (γG<γ<γY\gamma_{\rm G}<\gamma<\gamma_{\rm Y}) is jerky due to many small avalanches, reflecting the marginal stability. (iii) If γmax>γY\gamma_{\rm max}>\gamma_{\rm Y}, the cycle becomes strongly irreversible, suggesting the destruction of glass meta-basin after yielding.

Refer to caption

Figure 6: (A) Stability map of hard-sphere glasses, obtained from the initial glass at φ=φg\varphi=\varphi_{\rm g} and γ=0\gamma=0 (square). (B) Stress-strain curves measured in cyclic shear simulations at a constant φ=0.66\varphi=0.66. Black and gray lines correspond to γG<γmax<γY\gamma_{\rm G}<\gamma_{\rm max}<\gamma_{\rm Y} and γmax>γY\gamma_{\rm max}>\gamma_{\rm Y}. Adapted from Jin et al. (2018)).

IV Discussion and outlook

As a second-order phase transition, the fluctuation of order parameters (or the susceptibility) is expected to diverge at the Gardner transition in the thermodynamic limit. Simulations have shown that the caging susceptibility grows orders of magnitude approaching the Gardner transition Berthier et al. (2016). Furthermore, the spatial correlations between local caging order parameters become long-ranged in the Gardner phase, implying the heterogeneity of vibrational dynamics Berthier et al. (2016); Liao and Berthier (2019). However, dynamical activations could possibly turn a mean-field thermodynamic phase transition into a crossover in low dimensions. It remains inconclusive whether a sharp Gardner transition survives in three dimensions, although a renormalization group theory based on loop expansions Charbonneau and Yaida (2017) (see Chapter 3 for details) and a machine-learning facilitated finite-size analysis of simulation data Li et al. (2021) seem to suggest so.

The discussion so far has focused on the hard-sphere model. Hard spheres have a well-defined singularity under compression, the jamming transition, where quantities such as pressure and the length scale of mechanical response diverge (see Chapter 19 for a review on the jamming transition). Because the jamming transition lies in the Gardner phase, the full RSB predictions should also apply to the criticality and marginality of jamming, which are quantified by power-law scalings of weak forces, small interparticle gaps Charbonneau et al. (2014b) and low-frequency vibrational modes Franz et al. (2015). Remarkably, numerical results seemingly agree with the mean-field exponents even in low dimensions Charbonneau et al. (2014b, 2015b, 2016). The evidence of ultrametricity that characterizes the hierarchical energy landscape, has also been demonstrated numerically in jammed packings in three dimensions Dennis and Corwin (2020).

The Gardner transition seems to emerge as a “precursor” of certain singularities (jamming under compression and yielding under shear) in hard particles. The situation is more complicated in cases without such singularities, e.g., cooling soft spheres under the constant density condition. On the one hand, the mean-field theory universally identifies the existence of Gardner transition in soft spheres Scalliet et al. (2019), and simulations have reported a rejuvenation-memory effect Scalliet and Berthier (2019) similar to that found in spin-glasses. On the other hand, however, simulations demonstrate that the Gardner transition could be interfered with by low-dimensional effects such as localized “defects” Scalliet et al. (2017). Separating the Gardner physics from strong low-dimensional effects in soft spheres remains a challenge in simulations.

Finally, experimental efforts to detect the Gardner transition have shown encouraging progress. The caging order parameter approach is applied to vibrated granular discs, providing evidence of the Gardner phase Seguin and Dauchot (2016). In the Gardner phase, one expects a logarithmic growth of the MSD with lag time, which is verified in an experiment of glassy colloidal suspensions Hammond and Corwin (2020). The experimental data of shear modulus and MSD in a hard-sphere colloidal glass are consistent with the scalings μZFC1/ΔEApκ\mu_{\rm ZFC}\sim 1/\Delta_{\rm EA}\sim p^{\kappa} Zargar et al. (2017). The critical scalings of weak forces and small interparticle gaps have also been verified by precise experimental measurements of jammed photo-elastic disks Wang et al. (2022). The evidence of a Gardner-like transition is reported in a two-dimensional bidisperse granular crystal Kool et al. (2022), suggesting that the Gardner physics could be observed with minimum disorder Charbonneau et al. (2019). Examining protocol-dependent shear moduli and complex aging dynamics could provide future directions for the experimental exploration of Gardner physics.

References

  • Parisi et al. (2020) Giorgio Parisi, Pierfrancesco Urbani,  and Francesco Zamponi, Theory of Simple Glasses: Exact Solutions in Infinite Dimensions (Cambridge University Press, 2020).
  • Berthier et al. (2019) Ludovic Berthier, Giulio Biroli, Patrick Charbonneau, Eric I Corwin, Silvio Franz,  and Francesco Zamponi, “Gardner physics in amorphous solids and beyond,” The Journal of chemical physics 151, 010901 (2019).
  • Mézard et al. (1987) M. Mézard, G. Parisi,  and M. A. Virasoro, Spin glass theory and beyond (World Scientific, Singapore, 1987).
  • Nagata et al. (1979) Shoichi Nagata, PH Keesom,  and HR Harrison, “Low-dc-field susceptibility of cu mn spin glass,” Physical Review B 19, 1633 (1979).
  • Edwards and Anderson (1975) Samuel Frederick Edwards and Phil W Anderson, “Theory of spin glasses,” Journal of Physics F: Metal Physics 5, 965 (1975).
  • Castellani and Cavagna (2005) Tommaso Castellani and Andrea Cavagna, “Spin-glass theory for pedestrians,” Journal of Statistical Mechanics: Theory and Experiment 2005, P05012 (2005).
  • Granberg et al. (1988) P Granberg, L Sandlund, P Nordblad, P Svedlindh,  and L Lundgren, “Observation of a time-dependent spatial correlation length in a metallic spin glass,” Physical Review B 38, 7097 (1988).
  • Vincent et al. (1997) Eric Vincent, Jacques Hammann, Miguel Ocio, Jean-Philippe Bouchaud,  and Leticia F Cugliandolo, “Slow dynamics and aging in spin glasses,” in Complex Behaviour of Glassy Systems (Springer, 1997) pp. 184–219.
  • Nordblad and Svedlindh (1998) Per Nordblad and Peter Svedlindh, “Experiments on spin glasses,” in Spin Glasses and Random Fields (World Scientific, 1998) pp. 1–27.
  • Cugliandolo and Kurchan (1993) Leticia F Cugliandolo and Jorge Kurchan, “Analytical solution of the off-equilibrium dynamics of a long-range spin-glass model,” Physical Review Letters 71, 173 (1993).
  • Cugliandolo and Kurchan (1994) Leticia F Cugliandolo and Jorge Kurchan, “On the out-of-equilibrium relaxation of the sherrington-kirkpatrick model,” Journal of Physics A: Mathematical and General 27, 5749 (1994).
  • Cugliandolo et al. (1997) Leticia F Cugliandolo, Jorge Kurchan,  and Luca Peliti, “Energy flow, partial equilibration, and effective temperatures in systems with slow dynamics,” Physical Review E 55, 3898 (1997).
  • Franz et al. (1998) Silvio Franz, Marc Mézard, Giorgio Parisi,  and Luca Peliti, “Measuring equilibrium properties in aging systems,” Physical Review Letters 81, 1758 (1998).
  • Marinari et al. (2000) Enzo Marinari, Giorgio Parisi, Federico Ricci-Tersenghi,  and Juan J Ruiz-Lorenzo, “Off-equilibrium dynamics at very low temperatures in three-dimensional spin glasses,” Journal of Physics A: Mathematical and General 33, 2373 (2000).
  • Yoshino et al. (2002) Hajime Yoshino, Koji Hukushima,  and Hajime Takayama, “Extended droplet theory for aging in short-range spin glasses and a numerical examination,” Physical Review B 66, 064431 (2002).
  • Bray and Moore (1987) Alan J Bray and Michael A Moore, “Chaotic nature of the spin-glass phase,” Physical review letters 58, 57 (1987).
  • Fisher and Huse (1988a) Daniel S Fisher and David A Huse, “Equilibrium behavior of the spin-glass ordered phase,” Physical Review B 38, 386 (1988a).
  • Fisher and Huse (1988b) Daniel S Fisher and David A Huse, “Nonequilibrium dynamics of spin glasses,” Physical Review B 38, 373 (1988b).
  • Kondor (1989) I Kondor, “On chaos in spin glasses,” Journal of Physics A: Mathematical and General 22, L163 (1989).
  • Rizzo and Crisanti (2003) Tommaso Rizzo and Andrea Crisanti, “Chaos in temperature in the sherrington-kirkpatrick model,” Physical review letters 90, 137201 (2003).
  • Rizzo and Yoshino (2006) Tommaso Rizzo and Hajime Yoshino, “Chaos in glassy systems from a thouless-anderson-palmer perspective,” Physical Review B 73, 064416 (2006).
  • Yoshino and Rizzo (2008) Hajime Yoshino and Tommaso Rizzo, “Stepwise responses in mesoscopic glassy systems: A mean-field approach,” Physical Review B 77, 104429 (2008).
  • Parisi and Rizzo (2010) Giorgio Parisi and Tommaso Rizzo, “Chaos in temperature in diluted mean-field spin-glass,” Journal of Physics A: Mathematical and Theoretical 43, 235003 (2010).
  • Le Doussal et al. (2010) Pierre Le Doussal, Markus Müller,  and Kay Jörg Wiese, “Avalanches in mean-field models and the barkhausen noise in spin-glasses,” EPL (Europhysics Letters) 91, 57004 (2010).
  • Le Doussal et al. (2012) Pierre Le Doussal, Markus Müller,  and Kay Jörg Wiese, “Equilibrium avalanches in spin glasses,” Physical Review B 85, 214402 (2012).
  • Franz and Spigler (2017) Silvio Franz and Stefano Spigler, “Mean-field avalanches in jammed spheres,” Physical Review E 95, 022139 (2017).
  • Jonason et al. (1998) K Jonason, E Vincent, J Hammann, JP Bouchaud,  and P Nordblad, “Memory and chaos effects in spin glasses,” Physical Review Letters 81, 3243 (1998).
  • Yoshino et al. (2001) Hajime Yoshino, Anaël Lemaıtre,  and J-P Bouchaud, “Multiple domain growth and memory in the droplet model for spin-glasses,” The European Physical Journal B-Condensed Matter and Complex Systems 20, 367–395 (2001).
  • Jönsson et al. (2004) PE Jönsson, R Mathieu, Per Nordblad, H Yoshino, H Aruga Katori,  and A Ito, “Nonequilibrium dynamics of spin glasses: Examination of the ghost domain scenario,” Physical Review B 70, 174402 (2004).
  • Yoshino and Mézard (2010) Hajime Yoshino and Marc Mézard, “Emergence of rigidity at the structural glass transition: A first-principles computation,” Physical review letters 105, 015504 (2010).
  • Yoshino (2012) Hajime Yoshino, “Replica theory of the rigidity of structural glasses,” The Journal of Chemical Physics 136, 214108 (2012).
  • Yoshino and Zamponi (2014) Hajime Yoshino and Francesco Zamponi, “Shear modulus of glasses: Results from the full replica-symmetry-breaking solution,” Physical Review E 90, 022302 (2014).
  • Maimbourg et al. (2016) Thibaud Maimbourg, Jorge Kurchan,  and Francesco Zamponi, “Solution of the dynamics of liquids in the large-dimensional limit,” Physical review letters 116, 015902 (2016).
  • Kurchan et al. (2016) Jorge Kurchan, Thibaud Maimbourg,  and Francesco Zamponi, “Statics and dynamics of infinite-dimensional liquids and glasses: a parallel and compact derivation,” Journal of Statistical Mechanics: Theory and Experiment 2016, 033210 (2016).
  • Agoritsas et al. (2019a) Elisabeth Agoritsas, Thibaud Maimbourg,  and Francesco Zamponi, “Out-of-equilibrium dynamical equations of infinite-dimensional particle systems i. the isotropic case,” Journal of Physics A: Mathematical and Theoretical 52, 144002 (2019a).
  • Agoritsas et al. (2019b) Elisabeth Agoritsas, Thibaud Maimbourg,  and Francesco Zamponi, “Out-of-equilibrium dynamical equations of infinite-dimensional particle systems. ii. the anisotropic case under shear strain,” Journal of Physics A: Mathematical and Theoretical 52, 334001 (2019b).
  • Kranendonk and Frenkel (1991) WGT Kranendonk and D Frenkel, “Computer simulation of solid-liquid coexistence in binary hard sphere mixtures,” Molecular physics 72, 679–697 (1991).
  • Grigera and Parisi (2001) Tomás S Grigera and Giorgio Parisi, “Fast monte carlo algorithm for supercooled soft spheres,” Physical Review E 63, 045102 (2001).
  • Berthier et al. (2016) Ludovic Berthier, Patrick Charbonneau, Yuliang Jin, Giorgio Parisi, Beatriz Seoane,  and Francesco Zamponi, “Growing timescales and lengthscales characterizing vibrations of amorphous solids,” Proceedings of the National Academy of Sciences 113, 8397–8401 (2016).
  • Lubachevsky and Stillinger (1990) Boris D Lubachevsky and Frank H Stillinger, “Geometric properties of random disk packings,” Journal of statistical Physics 60, 561–583 (1990).
  • Goldstein (2010) Martin Goldstein, “Communications: Comparison of activation barriers for the Johari–Goldstein and alpha relaxations and its implications,” J. Chem. Phys. 132, 041104 (2010).
  • Boublik (1970) T. Boublik, “Hard sphere equation of state,” J. Chem. Phys. 53, 471 (1970).
  • Charbonneau et al. (2014a) Patrick Charbonneau, Yuliang Jin, Giorgio Parisi,  and Francesco Zamponi, “Hopping and the Stokes–Einstein relation breakdown in simple glass formers,” Proc. Natl. Acad. Sci. U.S.A. 111, 15025–15030 (2014a).
  • Donev et al. (2005) Aleksandar Donev, Salvatore Torquato,  and Frank H Stillinger, “Pair correlation function characteristics of nearly jammed disordered and ordered hard-sphere packings,” Physical Review E 71, 011105 (2005).
  • Bouchaud et al. (1998) Jean-Philippe Bouchaud, Leticia F Cugliandolo, Jorge Kurchan,  and Marc Mézard, “Out of equilibrium dynamics in spin-glasses and other glassy systems,” Spin glasses and random fields 12, 161 (1998).
  • Lees and Edwards (1972) AW Lees and SF Edwards, “The computer study of transport processes under extreme conditions,” J. Phys. Condens. Matter 5, 1921 (1972).
  • Jin and Yoshino (2017) Yuliang Jin and Hajime Yoshino, “Exploring the complex free-energy landscape of the simplest glass by rheology,” Nature communications 8, 1–8 (2017).
  • Charbonneau et al. (2015a) Patrick Charbonneau, Yuliang Jin, Giorgio Parisi, Corrado Rainone, Beatriz Seoane,  and Francesco Zamponi, “Numerical detection of the gardner transition in a mean-field glass former,” Physical Review E 92, 012316 (2015a).
  • Seoane and Zamponi (2018) Beatriz Seoane and Francesco Zamponi, “Spin-glass-like aging in colloidal and granular glasses,” Soft Matter 14, 5222–5234 (2018).
  • Charbonneau et al. (2014b) Patrick Charbonneau, Jorge Kurchan, Giorgio Parisi, Pierfrancesco Urbani,  and Francesco Zamponi, “Fractal free energy landscapes in structural glasses,” Nature communications 5, 1–6 (2014b).
  • Rainone et al. (2015) Corrado Rainone, Pierfrancesco Urbani, Hajime Yoshino,  and Francesco Zamponi, “Following the evolution of hard sphere glasses in infinite dimensions under external perturbations: Compression and shear strain,” Physical review letters 114, 015701 (2015).
  • Jin et al. (2018) Yuliang Jin, Pierfrancesco Urbani, Francesco Zamponi,  and Hajime Yoshino, “A stability-reversibility map unifies elasticity, plasticity, yielding, and jamming in hard sphere glasses,” Science advances 4, eaat6387 (2018).
  • Altieri and Zamponi (2019) Ada Altieri and Francesco Zamponi, “Mean-field stability map of hard-sphere glasses,” Physical Review E 100, 032140 (2019).
  • Liao and Berthier (2019) Qinyi Liao and Ludovic Berthier, “Hierarchical landscape of hard disk glasses,” Physical Review X 9, 011049 (2019).
  • Charbonneau and Yaida (2017) Patrick Charbonneau and Sho Yaida, “Nontrivial critical fixed point for replica-symmetry-breaking transitions,” Physical review letters 118, 215701 (2017).
  • Li et al. (2021) Huaping Li, Yuliang Jin, Ying Jiang,  and Jeff ZY Chen, “Determining the nonequilibrium criticality of a gardner transition via a hybrid study of molecular simulations and machine learning,” Proceedings of the National Academy of Sciences 118 (2021).
  • Franz et al. (2015) Silvio Franz, Giorgio Parisi, Pierfrancesco Urbani,  and Francesco Zamponi, “Universal spectrum of normal modes in low-temperature glasses,” Proceedings of the National Academy of Sciences 112, 14539–14544 (2015).
  • Charbonneau et al. (2015b) Patrick Charbonneau, Eric I Corwin, Giorgio Parisi,  and Francesco Zamponi, “Jamming criticality revealed by removing localized buckling excitations,” Physical review letters 114, 125504 (2015b).
  • Charbonneau et al. (2016) Patrick Charbonneau, Eric I Corwin, Giorgio Parisi, Alexis Poncet,  and Francesco Zamponi, “Universal non-debye scaling in the density of states of amorphous solids,” Physical review letters 117, 045503 (2016).
  • Dennis and Corwin (2020) RC Dennis and EI Corwin, “Jamming energy landscape is hierarchical and ultrametric,” Physical review letters 124, 078002 (2020).
  • Scalliet et al. (2019) Camille Scalliet, Ludovic Berthier,  and Francesco Zamponi, “Marginally stable phases in mean-field structural glasses,” Physical Review E 99, 012107 (2019).
  • Scalliet and Berthier (2019) Camille Scalliet and Ludovic Berthier, “Rejuvenation and memory effects in a structural glass,” Physical review letters 122, 255502 (2019).
  • Scalliet et al. (2017) Camille Scalliet, Ludovic Berthier,  and Francesco Zamponi, “Absence of marginal stability in a structural glass,” Physical review letters 119, 205501 (2017).
  • Seguin and Dauchot (2016) Antoine Seguin and Olivier Dauchot, “Experimental evidence of the gardner phase in a granular glass,” Physical review letters 117, 228001 (2016).
  • Hammond and Corwin (2020) Andrew P Hammond and Eric I Corwin, “Experimental observation of the marginal glass phase in a colloidal glass,” Proceedings of the National Academy of Sciences 117, 5714–5718 (2020).
  • Zargar et al. (2017) Rojman Zargar, Eric DeGiuli,  and Daniel Bonn, “Scaling for hard-sphere colloidal glasses near jamming,” EPL (Europhysics Letters) 116, 68004 (2017).
  • Wang et al. (2022) Yinqiao Wang, Jin Shang, Yuliang Jin,  and Jie Zhang, “Experimental observations of marginal criticality in granular materials,” Proceedings of the National Academy of Sciences 119, e2204879119 (2022).
  • Kool et al. (2022) Lars Kool, Patrick Charbonneau,  and Karen E Daniels, “Gardner-like transition from variable to persistent force contacts in granular crystals,” arXiv preprint arXiv:2205.06794  (2022).
  • Charbonneau et al. (2019) Patrick Charbonneau, Eric I Corwin, Lin Fu, Georgios Tsekenis,  and Michael van Der Naald, “Glassy, gardner-like phenomenology in minimally polydisperse crystalline systems,” Physical Review E 99, 020901 (2019).