Skip to main content
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Accessibility Statement
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian
  • Log in
  • My Cart

Main menu

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Accessibility Statement
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian

User menu

  • Log in
  • My Cart

Search

  • Advanced search
Home
Home

Advanced Search

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses

New Research In

Physical Sciences

Featured Portals

  • Physics
  • Chemistry
  • Sustainability Science

Articles by Topic

  • Applied Mathematics
  • Applied Physical Sciences
  • Astronomy
  • Computer Sciences
  • Earth, Atmospheric, and Planetary Sciences
  • Engineering
  • Environmental Sciences
  • Mathematics
  • Statistics

Social Sciences

Featured Portals

  • Anthropology
  • Sustainability Science

Articles by Topic

  • Economic Sciences
  • Environmental Sciences
  • Political Sciences
  • Psychological and Cognitive Sciences
  • Social Sciences

Biological Sciences

Featured Portals

  • Sustainability Science

Articles by Topic

  • Agricultural Sciences
  • Anthropology
  • Applied Biological Sciences
  • Biochemistry
  • Biophysics and Computational Biology
  • Cell Biology
  • Developmental Biology
  • Ecology
  • Environmental Sciences
  • Evolution
  • Genetics
  • Immunology and Inflammation
  • Medical Sciences
  • Microbiology
  • Neuroscience
  • Pharmacology
  • Physiology
  • Plant Biology
  • Population Biology
  • Psychological and Cognitive Sciences
  • Sustainability Science
  • Systems Biology
Research Article

Growing timescales and lengthscales characterizing vibrations of amorphous solids

Ludovic Berthier, Patrick Charbonneau, Yuliang Jin, Giorgio Parisi, View ORCID ProfileBeatriz Seoane, and Francesco Zamponi
PNAS July 26, 2016 113 (30) 8397-8401; first published July 8, 2016; https://doi.org/10.1073/pnas.1607730113
Ludovic Berthier
aLaboratoire Charles Coulomb, UMR 5221, Université de Montpellier and CNRS, 34095 Montpellier, France;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Patrick Charbonneau
bDepartment of Chemistry, Duke University, Durham, NC 27708;
cDepartment of Physics, Duke University, Durham, NC 27708;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Yuliang Jin
bDepartment of Chemistry, Duke University, Durham, NC 27708;
dDipartimento di Fisica, Sapienza Universitá di Roma, Istituto Nazionale di Fisica Nucleare, Sezione di Roma I, Istituto per i Processi Chimico-Fisici–Consiglio Nazionale delle Ricerche, I-00185 Rome, Italy;
eLaboratoire de Physique Théorique, École Normale Supérieure & Université de Recherche Paris Sciences et Lettres, Pierre et Marie Curie & Sorbonne Universités, UMR 8549 CNRS, 75005 Paris, France,
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: jinyuliang@gmail.com giorgio.parisi@roma1.infn.it beatriz.seoane.bartolome@lpt.ens.fr
Giorgio Parisi
dDipartimento di Fisica, Sapienza Universitá di Roma, Istituto Nazionale di Fisica Nucleare, Sezione di Roma I, Istituto per i Processi Chimico-Fisici–Consiglio Nazionale delle Ricerche, I-00185 Rome, Italy;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: jinyuliang@gmail.com giorgio.parisi@roma1.infn.it beatriz.seoane.bartolome@lpt.ens.fr
Beatriz Seoane
eLaboratoire de Physique Théorique, École Normale Supérieure & Université de Recherche Paris Sciences et Lettres, Pierre et Marie Curie & Sorbonne Universités, UMR 8549 CNRS, 75005 Paris, France,
fInstituto de Biocomputación y Física de Sistemas Complejos, 50009 Zaragoza, Spain
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Beatriz Seoane
  • For correspondence: jinyuliang@gmail.com giorgio.parisi@roma1.infn.it beatriz.seoane.bartolome@lpt.ens.fr
Francesco Zamponi
eLaboratoire de Physique Théorique, École Normale Supérieure & Université de Recherche Paris Sciences et Lettres, Pierre et Marie Curie & Sorbonne Universités, UMR 8549 CNRS, 75005 Paris, France,
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  1. Contributed by Giorgio Parisi, May 23, 2016 (sent for review February 2, 2016; reviewed by Kunimasa Miyazaki and Grzegorz Szamel)

  • Article
  • Figures & SI
  • Info & Metrics
  • PDF
Loading

Significance

Amorphous solids constitute most of solid matter but remain poorly understood. The recent solution of the mean-field hard-sphere glass former provides, however, deep insights into their material properties. In particular, this solution predicts a Gardner transition below which the energy landscape of glasses becomes fractal and the solid is marginally stable. Here we provide, to our knowledge, the first direct evidence for the relevance of a Gardner transition in physical systems. This result thus opens the way toward a unified understanding of the low-temperature anomalies of amorphous solids.

Abstract

Low-temperature properties of crystalline solids can be understood using harmonic perturbations around a perfect lattice, as in Debye’s theory. Low-temperature properties of amorphous solids, however, strongly depart from such descriptions, displaying enhanced transport, activated slow dynamics across energy barriers, excess vibrational modes with respect to Debye’s theory (i.e., a boson peak), and complex irreversible responses to small mechanical deformations. These experimental observations indirectly suggest that the dynamics of amorphous solids becomes anomalous at low temperatures. Here, we present direct numerical evidence that vibrations change nature at a well-defined location deep inside the glass phase of a simple glass former. We provide a real-space description of this transition and of the rapidly growing time- and lengthscales that accompany it. Our results provide the seed for a universal understanding of low-temperature glass anomalies within the theoretical framework of the recently discovered Gardner phase transition.

  • glass transition
  • disordered solids
  • Gardner transition
  • computer simulations
  • hard spheres

Understanding the nature of the glass transition, which describes the gradual transformation of a viscous liquid into an amorphous solid, remains an open challenge in condensed matter physics (1, 2). As a result, the glass phase itself is not well understood either. The main challenge is to connect the localized, or “caged,” dynamics that characterizes the glass transition to the low-temperature anomalies that distinguish amorphous solids from their crystalline counterparts (3⇓⇓⇓–7). Recent theoretical advances, building on the random first-order transition approach (8), have led to an exact mathematical description of both the glass transition and the amorphous phases of hard spheres in the mean-field limit of infinite-dimensional space (9). A surprising outcome has been the discovery of a novel phase transition inside the amorphous phase, separating the localized states produced at the glass transition from their inherent structures. This Gardner transition (10), which marks the emergence of a fractal hierarchy of marginally stable glass states, can be viewed as a glass transition deep within a glass, at which vibrational motion dramatically slows down and becomes spatially correlated (11). Although these theoretical findings promise to explain and unify the emergence of low-temperature anomalies in amorphous solids, the gap remains wide between mean-field calculations (9, 11) and experimental work. Here, we provide direct numerical evidence that vibrational motion in a simple 3D glass-former becomes anomalous at a well-defined location inside the glass phase. In particular, we report the rapid growth of a relaxation time related to cooperative vibrations, a nontrivial change in the probability distribution function of a global order parameter, and the rapid growth of a correlation length. We also relate these findings to observed anomalies in low-temperature laboratory glasses. These results provide key support for a universal understanding of the anomalies of glassy materials, as resulting from the diverging length- and timescales associated with the criticality of the Gardner transition.

Preparation of Glass States

Experimentally, glasses are obtained by a slow thermal or compression annealing, the rate of which determines the location of the glass transition (1, 2). We find 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) to study vibrational motion inside the glass without interference from particle diffusion. We thus combine a very simple glass-forming model––a polydisperse mixture of hard spheres––to an efficient Monte Carlo scheme to obtain equilibrium configurations at unprecedentedly high densities, i.e., deep in the supercooled regime. The optimized swap Monte Carlo algorithm (12), which combines standard local Monte Carlo moves with attempts at exchanging pairs of particle diameters, indeed enhances thermalization by several orders of magnitude. Configurations contain either N = 1,000 or N = 8,000 (results in Figs. 1–3 are for N = 1,000; results in Fig. 4 are for N = 8,000) hard spheres with equal unit mass m and diameters independently drawn from a probability distribution Pσ(σ)∼σ−3, for σmin≤σ≤σmin/0.45. We similarly study a 2D bidisperse model glass former and report the main results in SI Appendix.

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Two glass phases. Inverse reduced pressure–packing fraction (1/p v φ) phase diagram for polydisperse hard spheres. The equilibrium simulation results at φg (green squares) are fitted to the liquid EOS (Eq. 1, green line). The dynamical cross-over, φd, is obtained from the liquid dynamics. Compression annealing from φg up to jamming (blue triangles) follows a glass EOS (fit to Eq. 3, dashed lines). At φG (red circles and line) with a finite p, stable glass states transform into marginally stable glasses. Snapshots illustrate spatial heterogeneity above and below φG, with sphere diameters proportional to the linear cage size and colors encoding the relative cage size, ui (see the text).

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Emergence of slow vibrational dynamics. (A) Time evolution of Δ(t,tw) for several tw and φ (from top to bottom, φ=0.645,0.67,0.68,0.684,0.688), following compression from φg=0.643. For φ≳φG=0.684, Δ(t,tw) displays strong aging. (B) Comparison between Δ(t,tw) (points) and ΔAB(t+tw) (lines) for the longest waiting time tw=1,024. For φ<φG both observables converge to the same value within the time window considered, but not for φ>φG. (C) The time evolution of δΔ(t,tw) at tw=2 displays a logarithmic tail, which provides a characteristic relaxation time τ. (D) As φ approaches φG, τ grows rapidly for any tw.

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Global fluctuations of the order parameter. (A) Probability distribution functions for ΔAB and Δ above, at, and below the Gardner cross-over, φG=0.670(2) for φg=0.630. Vertical lines mark 〈Δ〉 (solid) and 〈ΔAB〉 (dashed), which also represent the peak positions. (B) Comparing 〈Δ〉 and 〈ΔAB〉 shows that the average values separate for φ≳φG (Data are multiplied by 5k, where k=0,1,…,5 for φ=0.655,0.643,…,0.598, respectively.) Around φG, (C) the global susceptibility χAB grows very rapidly, and (D) the skewness ΓAB peaks. Numerical estimates for φG are indicated by vertical segments.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Growing correlation length. (A) Spatial correlator GL(r) (Eq. 6) for different φ (from bottom to top, φ=0.65,0.67,0.675,0.677,0.68,0.682,0.687) annealed from φg=0.640, with N=8,000 (larger systems are used here to significantly measure the growth of ξ). (B) Fitting GL(r) to Eq. 7 (lines in A) provides the correlation length ξ, which grows with φ and becomes comparable to the linear system size upon approaching φG=0.682(2) (dashed line).

We mimic slow annealing in two steps (Fig. 1). First, we produce equilibrated liquid configurations at various densities φg using our efficient simulation scheme, concurrently obtaining the liquid equation of state (EOS). The liquid EOS for the reduced pressure p=βP/ρ, where ρ is the number density, β is the inverse temperature, and P is the system pressure, is described bypliquid(φ)=1+f(φ)[pCS(φ)−1],[1]with pCS(φ) from ref. 13pCS(φ)=11−φ+3s1s2s3φ(1−φ)2+s23s32(3−φ)φ2(1−φ)3,[2]where sk is the kth moment of Pσ(σ), and f(φ)=0.005−tanh[14(φ−0.79)] are fitted quantities. The structure of the equilibrium configurations generated by the swap algorithm has been carefully analyzed. Unlike for other glass formers (14, 15), no signs of orientational or crystalline order were observed (16, 17). Following the strategy of ref. 18, we also obtain the mode-coupling theory dynamical cross-over φd=0.594(1) (SI Appendix). We have not analyzed the compression of equilibrium configurations with φg<φd, as done in earlier studies (19, 20), because structural relaxation is not well decoupled from vibrational dynamics, although the obtained jammed states should have equivalent properties.

Second, we use these liquid configurations as starting points for standard molecular dynamics simulations during which the system is compressed out of equilibrium up to various target φ>φg (21). Annealing is achieved by growing spheres following the Lubachevsky–Stillinger algorithm (21) at a constant growth rate γg=10−3 (see SI Appendix for a discussion on the γg -dependence). The average particle diameter, σ¯, serves as unit length, and the simulation time is expressed in units of βmσ¯2. To obtain thermal and disorder averaging, this procedure is repeated over Ns samples (Ns≈150 for N=1,000 and Ns=50 for N=8,000), each with different initial equilibrium configurations at φg, and over Nth=64−19,440 independent thermal (quench) histories for each sample. Quantities reported here are averaged over Ns×Nth quench histories, unless otherwise specified. The nonequilibrium glass EOSs associated with this compression (dashed lines) terminate (at infinite pressure) at inherent structures that correspond, for hard spheres, to jammed configurations (blue triangles). To capture the glass EOSs, we use a free-volume scaling around the corresponding jamming point φJpglass(φ;φg)=CφJ(φg)−φ,[3]where the constant C weakly depends on φg.

Our numerical protocol is analogous to varying the cooling rate––and thus the glass transition temperature––of thermal glasses, and then further annealing the resulting amorphous solid. Each value of φg indeed selects a different glass, ranging from the onset of sluggish liquid dynamics around the dynamical cross-over (1, 2), φd, to the very dense liquid regime where diffusion and vibrations (β-relaxation processes) are fully separated (2). For sufficiently large φg, we thus obtain unimpeded access to the only remaining glass dynamics, i.e., β-relaxation processes (4).

Growing Timescales

A central observable to characterize glass dynamics is the mean-squared displacement (MSD) of particles from position ri(tw)Δ(t,tw)=1N∑i=1N〈|ri(t+tw)−ri(tw)|2〉,[4]averaged over both thermal fluctuations and disorder, where time t starts after waiting time tw when compression has reached the target φ. The MSD plateau height at long times quantifies the average cage size (SI Appendix). Because some of the smaller particles manage to leave their cages, the sum in Eq. 4 is here restricted to the larger half of the particle size distribution (SI Appendix). When φ is not too large, φ≳φg, the plateau emerges quickly, as suggested by the traditional view of caging in glasses (Fig. 2A). When the glass is compressed beyond a certain φG, however, Δ(t,tw) displays both a strong dependence on the waiting time tw, i.e., aging, and a slow dynamics, as captured by the emergence of two plateaus. These effects suggest a complex vibrational dynamics. Aging, in particular, provides a striking signature of a growing timescale associated with vibrations, revealing the existence of a “glass transition” deep within the glass phase.

To determine the timescale associated with this slowdown, we estimate the distance between independent pairs of configurations by first compressing two independent copies, A and B, from the same initial state at φg to the target φ, and then measuring their relative distanceΔAB(t)=1N∑i=1N〈|riA(t)−riB(t)|2〉,[5]so that ΔAB(t→∞)≃Δ(t→∞,tw→∞), as shown in Fig. 2B. The two copies share the same positions of particles at φg, but are assigned different initial velocities drawn from the Maxwell–Boltzmann distribution. The time evolution of the difference δΔ(t,tw)=ΔAB(tw+t)−Δ(t,tw) indicates that whereas the amplitude of particle motion naturally becomes smaller as φ increases, the corresponding dynamics becomes slower (Fig. 2C). In other words, as φ grows particles take longer to explore a smaller region of space. In a crystal, by contrast, δΔ(t,tw) decays faster under similar circumstances. A relaxation timescale, τ, can be extracted from the decay of δΔ(t,tw) at large t, whose logarithmic form, δΔ(t,tw)∼1−ln⁡t/ln⁡τ, is characteristic of the glassiness of vibrations. As φ→φG, we find that τ dramatically increases (Fig. 2D), which provides direct evidence of a marked cross-over characterizing the evolution of the glass upon compression.

Global Fluctuations of the Order Parameter

This sharp dynamical cross-over corresponds to a loss of ergodicity inside the glass, i.e., time and ensemble averages yield different results. To better characterize this cross-over, we define a timescale τcage for the onset of caging [τcage≈O(1); SI Appendix], and the corresponding order parameters ΔAB≡ΔAB(τcage) and Δ≡Δ(τcage,tw=0).

The evolution of the probability distribution functions, P(ΔAB) and P(Δ), as well as their first moments, 〈ΔAB〉 and 〈Δ〉, are presented in Fig. 3 A and B for a range of densities across φG. For φ<φG, dynamics is fast, 〈ΔAB〉 and 〈Δ〉 coincide, and P(ΔAB) and P(Δ) are narrow and Gaussian-like. For φ>φG, however, the MSD does not converge to its long-time limit, 〈Δ〉<〈ΔAB〉, which indicates that configuration space explored by vibrational motion is now broken into mutually inaccessible regions. Interestingly, the slight increase of 〈ΔAB〉 with φ in this regime (Fig. 3B) suggests that states are then pushed further apart in phase space, which is consistent with theoretical predictions (11). When compressing a system across φG, its dynamics explores only a restricted part of phase space. As a result, ΔAB displays pronounced, non-Gaussian fluctuations (Fig. 3A). Repeated compressions from a same initial state at φg may end up in distinct states, which explains why ΔAB is typically much larger and more broadly fluctuating than Δ (Fig. 3A). These results are essentially consistent with theoretical predictions (9, 11), which suggest that for φ>φG, P(ΔAB) should separate into two peaks connected by a wide continuous band with the left-hand peak continuing the single peak of P(Δ). The very broad distribution of ΔAB further suggests that spatial correlations develop as φ→φG, yielding strongly correlated states at larger densities.

To quantify these fluctuations we measure the variance χAB and skewness ΓAB (SI Appendix and ref. 11) of P(ΔAB) (Fig. 3 C and D). The global susceptibility χAB is very small for φ<φG and grows rapidly as φG is approached, increasing by about two decades for the largest φg considered (Fig. 3C). Whereas χAB quantifies the increasing width of the distributions, ΓAB reveals a change in their shapes. For each φg we find that ΓAB is small on both sides of φG with a pronounced maximum at φ=φG (Fig. 3D). This reflects the roughly symmetric shape of P(ΔAB) around 〈ΔAB〉 on both sides of φG and the development of an asymmetric tail for large ΔAB around the cross-over, a known signature of sample-to-sample fluctuations in spin glasses (22) and mean-field glass models (11). Note that because the skewness maximum gives the clearest numerical estimate of φG, we use it to determine the values reported in Fig. 1.

Growing Correlation Length

The rapid growth of χAB in the vicinity of φG suggests the concomitant growth of a spatial correlation length, ξ. Its measurement requires spatial resolution of the fluctuations of ΔAB, hence for each particle i we define ui=(|riA−riB|2/〈ΔAB〉)−1 to capture its contribution to deviations around the average 〈ΔAB〉. A first glimpse of these spatial fluctuations is offered by snapshots of the ui field (Fig. 1), which appear featureless for φ<φG, but highly structured and spatially correlated for φ≳φG. More quantitatively, we define the spatial correlatorGL(r)=〈∑μ=13∑i≠juiujδ(r−|ri,μA−rj,μA|)〉〈∑μ=13∑i≠jδ(r−|ri,μA−rj,μA|)〉,[6]where ri,μ is the projection of the particle position along direction μ. Even for the larger system size considered, measuring GL(r) is challenging because spatial correlations quickly become long-ranged as φ→φG (Fig. 4A). Fitting the results to an empirical form that takes into account the periodic boundary conditions in a system of linear size L,GL(r)∼1rae−(r/ξ)b+1(L−r)ae−[(L−r)/ξ]b,[7]where a and b are fitting parameters, nonetheless confirms that ξ grows rapidly with φ and becomes of the order of the simulation box at φ>φG (Fig. 4B). Note that although probed using a dynamical observable, the spatial correlations captured by GL(r) are conceptually distinct from the dynamical heterogeneity observed in supercooled liquids (23), which is transient and disappears once the diffusive regime is reached.

Experimental Consequences

The system analyzed in this work is a canonical model for colloidal suspensions and granular media. Hence, experiments along the lines presented here could be performed to investigate more closely vibrational dynamics in colloidal and granular glasses, using a series of compressions to extract Δ and ΔAB. Experiments are also possible in molecular and polymeric glasses, for which the natural control parameter is temperature T instead of density. Let us therefore rephrase our findings from this viewpoint. As the system is cooled, the supercooled liquid dynamics is arrested at the laboratory glass transition temperature Tg. As the resulting glass is further cooled its phase space transforms, around a well-defined Gardner temperature TG<Tg, from a simple state (akin to that of a crystal) into a more complex phase composed of a large number of glassy states (see SI Appendix for a discussion of the phase diagram as a function of T).

Around TG, vibrational dynamics becomes increasingly heterogeneous (Fig. 1), slow (Fig. 2), fluctuating from realization to realization (Fig. 3), and spatially correlated (Fig. 4). The β-relaxation dynamics inside the glass thus becomes highly cooperative (24, 25) and ages (26). The fragmentation of phase space below TG also gives rise to a complex response to mechanical perturbations in the form of plastic irreversible events, in which the system jumps from one configuration to another (4, 6, 27). This expectation stems from the theoretical prediction that the complex phase at T<TG is marginally stable (9), which implies that glass states are connected by very low energy barriers, resulting in strong responses to weak perturbations (7).

A key prediction is that the aforementioned anomalies appear simultaneously around a TG that is strongly dependent on the scale Tg selected by the glass preparation protocol. Annealed glasses with lower Tg are expected to present a sharper Gardner-like cross-over, at an increasingly lower temperature. Numerically, we produced a substantial variation of φg by using an efficient Monte Carlo algorithm to bypass the need for a broad range of compression rates. In experiments a similar or even larger range of Tg can be explored (28), using poorly annealed glasses from hyperquenching (29) and ultrastable glasses from vapor deposition (30⇓–32). We expect ultrastable glasses, in particular, to display strongly enhanced glass anomalies, consistent with recent experimental reports (33⇓–35). Interestingly, a Gardner-like regime may also underlie the anomalous aging recently observed in individual proteins (36).

Conclusion

Since its prediction in the mean-field limit, the Gardner transition has been regarded as a key ingredient to understand the physical properties of amorphous solids. Understanding the role of finite-dimensional fluctuations is a difficult theoretical problem (37). Our work shows that clear signs of an apparent critical behavior can be observed in three dimensions, at least in a finite-size system, which shows that the correlation length becomes at least comparable to the system size as φ approaches φG. Although the fate of these findings in the thermodynamic limit remains an open question, the remarkably large signature of the effect strongly suggests that the Gardner phase transition paradigm is a promising theoretical framework for a universal understanding of the anomalies of solid amorphous materials, from granular materials to glasses, foams, and proteins.

Acknowledgments

P.C. acknowledges support from the Alfred P. Sloan Foundation and National Science Foundation (NSF DMR-1055586). B.S. acknowledges the support by Ministerio de Economía y Competitividad (MINECO) (Spain) through Research Contract FIS2012-35719-C02. This project has received funding from the European Union’s Horizon 2020 Research and Innovation Programme under the Marie Skłodowska-Curie Grant Agreement 654971, as well as from the European Research Council (ERC) under the European Union’s Seventh Framework Programme (FP7/2007-2013)/ERC Grant Agreement 306845. This work was granted access to the High-Performance Computing (HPC) resources of Mésocentre de Calcul-Université de Recherche Paris Sciences et Lettres (MesoPSL) financed by the Region Ile de France and the project Equip@Meso (Reference ANR-10-EQPX-29-01) of the program Investissements d’Avenir supervised by the Agence Nationale pour la Recherche. This project has received funding from the ERC under the European Union’s Horizon 2020 Research and Innovation Programme (Grant Agreement 694925).

Footnotes

  • ↵1To whom correspondence may be addressed. Email: jinyuliang{at}gmail.com, giorgio.parisi{at}roma1.infn.it, or beatriz.seoane.bartolome{at}lpt.ens.fr.
  • Author contributions: L.B., P.C., Y.J., G.P., B.S., and F.Z. designed research, performed research, analyzed data, and wrote the paper.

  • Reviewers: K.M., Nagoya University; and G.S., Colorado State University.

  • The authors declare no conflict of interest.

  • Data deposition: Data relevant to this work have been archived and can be accessed at doi.org/10.7924/G8QN64NT.

  • This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1607730113/-/DCSupplemental.

View Abstract

References

  1. ↵
    1. Berthier L,
    2. Biroli G
    (2011) Theoretical perspective on the glass transition and amorphous materials. Rev Mod Phys 83(2):587–645.
    .
    OpenUrlCrossRef
  2. ↵
    1. Cavagna A
    (2009) Supercooled liquids for pedestrians. Phys Rep 476(4):51–124.
    .
    OpenUrlCrossRef
  3. ↵
    1. Phillips WA
    (1987) Two-level states in glasses. Rep Prog Phys 50(12):1657.
    .
    OpenUrlCrossRef
  4. ↵
    1. Goldstein M
    (2010) Communications: Comparison of activation barriers for the Johari-Goldstein and alpha relaxations and its implications. J Chem Phys 132(4):041104.
    .
    OpenUrlCrossRefPubMed
  5. ↵
    1. Malinovsky VK,
    2. Sokolov AP
    (1986) The nature of boson peak in Raman scattering in glasses. Solid State Commun 57(9):757–761.
    .
    OpenUrlCrossRef
  6. ↵
    1. Hentschel HGE,
    2. Karmakar S,
    3. Lerner E,
    4. Procaccia I
    (2011) Do athermal amorphous solids exist? Phys Rev E Stat Nonlin Soft Matter Phys 83(6 Pt 1):061101.
    .
    OpenUrlCrossRefPubMed
  7. ↵
    1. Müller M,
    2. Wyart M
    (2015) Marginal stability in structural, spin, and electron glasses. Annu Rev Condens Matter Phys 6:177–200.
    .
    OpenUrlCrossRef
  8. ↵
    1. Wolynes P,
    2. Lubchenko V
    , eds (2012) Structural Glasses and Supercooled Liquids: Theory, Experiment, and Applications (Wiley, Hoboken, NJ).
    .
  9. ↵
    1. Charbonneau P,
    2. Kurchan J,
    3. Parisi G,
    4. Urbani P,
    5. Zamponi F
    (2014) Fractal free energy landscapes in structural glasses. Nat Commun 5:3725.
    .
    OpenUrlPubMed
  10. ↵
    1. Gardner E
    (1985) Spin glasses with p-spin interactions. Nucl Phys B 257:747–765.
    .
    OpenUrlCrossRef
  11. ↵
    1. Charbonneau P, et al.
    (2015) Numerical detection of the Gardner transition in a mean-field glass former. Phys Rev E Stat Nonlin Soft Matter Phys 92(1):012316.
    .
    OpenUrlCrossRefPubMed
  12. ↵
    1. Grigera TS,
    2. Parisi G
    (2001) Fast Monte Carlo algorithm for supercooled soft spheres. Phys Rev E Stat Nonlin Soft Matter Phys 63(4 Pt 2):045102.
    .
    OpenUrlCrossRefPubMed
  13. ↵
    1. Boublik T
    (1970) Hard sphere equation of state. J Chem Phys 53(1):471.
    .
    OpenUrlCrossRef
  14. ↵
    1. Russo J,
    2. Tanaka H
    (2015) Assessing the role of static length scales behind glassy dynamics in polydisperse hard disks. Proc Natl Acad Sci USA 112(22):6920–6924.
    .
    OpenUrlAbstract/FREE Full Text
  15. ↵
    1. Flenner E,
    2. Szamel G
    (2015) Fundamental differences between glassy dynamics in two and three dimensions. Nat Commun 6:7392.
    .
    OpenUrlCrossRefPubMed
  16. ↵
    1. Berthier L,
    2. Coslovich D,
    3. Ninarello A,
    4. Ozawa M
    (2016) Equilibrium sampling of hard spheres up to the jamming density and beyond. Phys Rev Lett 116(23):238002.
    .
    OpenUrlCrossRefPubMed
  17. ↵
    1. Yaida S,
    2. Berthier L,
    3. Charbonneau P,
    4. Tarjus G
    (2015) Point-to-set lengths, local structure, and glassiness. arXiv:1511.03573.
    .
  18. ↵
    1. Charbonneau P,
    2. Jin Y,
    3. Parisi G,
    4. Zamponi F
    (2014) Hopping and the Stokes-Einstein relation breakdown in simple glass formers. Proc Natl Acad Sci USA 111(42):15025–15030.
    .
    OpenUrlAbstract/FREE Full Text
  19. ↵
    1. Chaudhuri P,
    2. Berthier L,
    3. Sastry S
    (2010) Jamming transitions in amorphous packings of frictionless spheres occur over a continuous range of volume fractions. Phys Rev Lett 104(16):165701.
    .
    OpenUrlCrossRefPubMed
  20. ↵
    1. Ozawa M,
    2. Kuroiwa T,
    3. Ikeda A,
    4. Miyazaki K
    (2012) Jamming transition and inherent structures of hard spheres and disks. Phys Rev Lett 109(20):205701.
    .
    OpenUrlCrossRefPubMed
  21. ↵
    1. Skoge M,
    2. Donev A,
    3. Stillinger FH,
    4. Torquato S
    (2006) Packing hyperspheres in high-dimensional Euclidean spaces. Phys Rev E Stat Nonlin Soft Matter Phys 74(4 Pt 1):041127.
    .
    OpenUrlCrossRefPubMed
  22. ↵
    1. Parisi G,
    2. Rizzo T
    (2013) Critical dynamics in glassy systems. Phys Rev E Stat Nonlin Soft Matter Phys 87(1):012101.
    .
    OpenUrlCrossRefPubMed
  23. ↵
    1. Berthier L,
    2. Biroli G,
    3. Bouchaud JP,
    4. Cipelletti L,
    5. van Saarloos W
    , eds (2011) Dynamical Heterogeneities and Glasses (Oxford Univ Press, Oxford).
    .
  24. ↵
    1. Cohen Y,
    2. Karmakar S,
    3. Procaccia I,
    4. Samwer K
    (2012) The nature of the β-peak in the loss modulus of amorphous solids. Europhys Lett 100(3):36003.
    .
    OpenUrlCrossRef
  25. ↵
    1. Bock D, et al.
    (2013) On the cooperative nature of the β-process in neat and binary glasses: A dielectric and nuclear magnetic resonance spectroscopy study. J Chem Phys 139(6):064508.
    .
    OpenUrlCrossRefPubMed
  26. ↵
    1. Leheny RL,
    2. Nagel SR
    (1998) Frequency-domain study of physical aging in a simple liquid. Phys Rev B 57(9):5154.
    .
    OpenUrl
  27. ↵
    1. Brito C,
    2. Wyart M
    (2009) Geometric interpretation of previtrification in hard sphere liquids. J Chem Phys 131(2):024504.
    .
    OpenUrlCrossRefPubMed
  28. ↵
    1. Rössler E,
    2. Sokolov AP,
    3. Kisliuk A,
    4. Quitmann D
    (1994) Low-frequency Raman scattering on different types of glass formers used to test predictions of mode-coupling theory. Phys Rev B Condens Matter 49(21):14967–14978.
    .
    OpenUrlCrossRefPubMed
  29. ↵
    1. Velikov V,
    2. Borick S,
    3. Angell CA
    (2001) The glass transition of water, based on hyperquenching experiments. Science 294(5550):2335–2338.
    .
    OpenUrlAbstract/FREE Full Text
  30. ↵
    1. Swallen SF, et al.
    (2007) Organic glasses with exceptional thermodynamic and kinetic stability. Science 315(5810):353–356.
    .
    OpenUrlAbstract/FREE Full Text
  31. ↵
    1. Singh S,
    2. Ediger MD,
    3. de Pablo JJ
    (2013) Ultrastable glasses from in silico vapour deposition. Nat Mater 12(2):139–144.
    .
    OpenUrlCrossRefPubMed
  32. ↵
    1. Hocky GM,
    2. Berthier L,
    3. Reichman DR
    (2014) Equilibrium ultrastable glasses produced by random pinning. J Chem Phys 141(22):224503.
    .
    OpenUrlCrossRefPubMed
  33. ↵
    1. Pérez-Castañeda T,
    2. Rodríguez-Tinoco C,
    3. Rodríguez-Viejo J,
    4. Ramos MA
    (2014) Suppression of tunneling two-level systems in ultrastable glasses of indomethacin. Proc Natl Acad Sci USA 111(31):11275–11280.
    .
    OpenUrlAbstract/FREE Full Text
  34. ↵
    1. Liu X,
    2. Queen DR,
    3. Metcalf TH,
    4. Karel JE,
    5. Hellman F
    (2014) Hydrogen-free amorphous silicon with no tunneling states. Phys Rev Lett 113(2):025503.
    .
    OpenUrlCrossRefPubMed
  35. ↵
    1. Yu HB,
    2. Tylinski M,
    3. Guiseppi-Elie A,
    4. Ediger MD,
    5. Richert R
    (2015) Suppression of β relaxation in vapor-deposited ultrastable glasses. Phys Rev Lett 115(18):185501.
    .
    OpenUrlCrossRefPubMed
  36. ↵
    1. Hu X, et al.
    (2015) The dynamics of single protein molecules is non-equilibrium and self-similar over thirteen decades in time. Nat Phys 12:171–174.
    .
    OpenUrlCrossRef
  37. ↵
    1. Urbani P,
    2. Biroli G
    (2015) Gardner transition in finite dimensions. Phys Rev B 91(10):100202.
    .
    OpenUrlCrossRef
PreviousNext
Back to top
Article Alerts
Email Article

Thank you for your interest in spreading the word on PNAS.

NOTE: We only request your email address so that the person you are recommending the page to knows that you wanted them to see it, and that it is not junk mail. We do not capture any email address.

Enter multiple addresses on separate lines or separate them with commas.
Growing timescales and lengthscales characterizing vibrations of amorphous solids
(Your Name) has sent you a message from PNAS
(Your Name) thought you would like to see the PNAS web site.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
Growing time- and lengthscales in amorphous solids
Ludovic Berthier, Patrick Charbonneau, Yuliang Jin, Giorgio Parisi, Beatriz Seoane, Francesco Zamponi
Proceedings of the National Academy of Sciences Jul 2016, 113 (30) 8397-8401; DOI: 10.1073/pnas.1607730113

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Request Permissions
Share
Growing time- and lengthscales in amorphous solids
Ludovic Berthier, Patrick Charbonneau, Yuliang Jin, Giorgio Parisi, Beatriz Seoane, Francesco Zamponi
Proceedings of the National Academy of Sciences Jul 2016, 113 (30) 8397-8401; DOI: 10.1073/pnas.1607730113
Digg logo Reddit logo Twitter logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Mendeley logo Mendeley
Proceedings of the National Academy of Sciences: 113 (30)
Table of Contents

Submit

Sign up for Article Alerts

Article Classifications

  • Physical Sciences
  • Physics

Jump to section

  • Article
    • Abstract
    • Preparation of Glass States
    • Growing Timescales
    • Global Fluctuations of the Order Parameter
    • Growing Correlation Length
    • Experimental Consequences
    • Conclusion
    • Acknowledgments
    • Footnotes
    • References
  • Figures & SI
  • Info & Metrics
  • PDF

You May Also be Interested in

Abstract depiction of a guitar and musical note
Science & Culture: At the nexus of music and medicine, some see disease treatments
Although the evidence is still limited, a growing body of research suggests music may have beneficial effects for diseases such as Parkinson’s.
Image credit: Shutterstock/agsandrew.
Large piece of gold
News Feature: Tracing gold's cosmic origins
Astronomers thought they’d finally figured out where gold and other heavy elements in the universe came from. In light of recent results, they’re not so sure.
Image credit: Science Source/Tom McHugh.
Dancers in red dresses
Journal Club: Friends appear to share patterns of brain activity
Researchers are still trying to understand what causes this strong correlation between neural and social networks.
Image credit: Shutterstock/Yeongsik Im.
Yellow emoticons
Learning the language of facial expressions
Aleix Martinez explains why facial expressions often are not accurate indicators of emotion.
Listen
Past PodcastsSubscribe
Goats standing in a pin
Transplantation of sperm-producing stem cells
CRISPR-Cas9 gene editing can improve the effectiveness of spermatogonial stem cell transplantation in mice and livestock, a study finds.
Image credit: Jon M. Oatley.

Similar Articles

Site Logo
Powered by HighWire
  • Submit Manuscript
  • Twitter
  • Facebook
  • RSS Feeds
  • Email Alerts

Articles

  • Current Issue
  • Special Feature Articles – Most Recent
  • List of Issues

PNAS Portals

  • Anthropology
  • Chemistry
  • Classics
  • Front Matter
  • Physics
  • Sustainability Science
  • Teaching Resources

Information

  • Authors
  • Editorial Board
  • Reviewers
  • Librarians
  • Press
  • Site Map
  • PNAS Updates

Feedback    Privacy/Legal

Copyright © 2021 National Academy of Sciences. Online ISSN 1091-6490