Vibrational Lifetimes and Viscoelastic Properties of Ultrastable Glasses

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Vibrational lifetimes and viscoelastic properties of ultrastable glasses

Jan Grießer1 and Lars Pastewka1, 2, ∗


1
Department of Microsystems Engineering, University of Freiburg,
Georges-Köhler-Allee 103, 79110 Freiburg, Germany
2
Cluster of Excellence livMatS, Freiburg Center for Interactive Materials and Bioinspired Technologies,
University of Freiburg, 79110 Freiburg, Germany
(Dated: February 23, 2024)
Amorphous solids are viscoelastic. They dissipate energy when deformed at finite rate and finite
temperature. We here use analytic theory and molecular simulations to demonstrate that linear vis-
coelastic dissipation can be directly related to the static and dynamic properties of the fundamental
vibrational excitations of an amorphous system. We study ultrastable glasses that do not age, i.e.
arXiv:2402.14336v1 [cond-mat.soft] 22 Feb 2024

that remain in stable minima of the potential energy surface at finite temperature. Our simula-
tions show four types of vibrational modes, which differ in spatial localization, similarity to plane
waves and vibrational lifetimes. At frequencies below the Boson peak, the viscoelastic response can
be split into contributions from plane-wave and quasilocalized modes. We derive a parameter-free
expression for the viscoelastic storage and loss moduli for both of these modes. Our results show
that the dynamics of microscopic dissipation, in particular the lifetimes of the modes, determine the
viscoelastic response only at high frequency. Quasilocalized modes dominate the linear viscoelastic
response at intermediate frequencies below the Boson peak.

I. INTRODUCTION of excitations is in the form of vibrational normal modes.


For a three dimensional solid with N atoms and therefore
Solid materials deform in response to external load. In 3N vibrational modes, the canonical view is to regard one
ideal elastic solids, the energy stored in the deformation is of these modes as the system of interest and the rest as
fully recovered once the load is removed and they return the heat bath.
to their original state. Real solids always dissipate en- In a linear system, each vibrational mode is fully inde-
ergy at finite deformation rates. Such viscoelastic effects pendent and there is no coupling between mode and bath.
are strong for soft polymeric (rubbery) materials, but Linear systems do not evolve towards thermodynamic
they are also present in hard materials such as glasses, equilibrium and there is no “dissipation”. Coupling and
alloys or even crystals [1, 2]. Viscoelasticity is responsible hence dissipation emerges solely through nonlinearities in
for limiting the quality factor (line width) of oscillators, the intermode (or interatomic) interactions. The time re-
and the design of micro- or nanoelectromechanical oscil- quired for a vibrational mode to return to the thermody-
lators seeks to minimize this effect [3–9]. The resilience namic equilibrium is given by its vibrational lifetime. Vi-
of glasses towards impact or shock is determined by its brational lifetimes can either be obtained via simulations
ability to absorb energy, which requires materials with or from perturbation theory [16–19]. The mode-coupling
large viscoelasticity [10–13]. Understanding the atom- process is particularly interesting for glassy solids that
istic origins of viscoelastic dissipation in hard materials are intrinsically out-of-equilibrium [20]. Besides excita-
is therefore crucial for the design of mechanical devices tion of vibrational modes, viscoelastic dissipation is trig-
and resilient materials. gered by atomic rearrangements, i.e. transitions between
The microscopic mechanism behind viscoelastic dissi- inherent structures [19, 21–24]. We here focus only on the
pation is thermalization. While the first law of thermo- excitation aspect of viscoelastic dissipation in glasses and
dynamics tells us that energy is conserved, “dissipation” study “ultrastable” glasses, which remain in the same in-
of energy is the result of processes that evolve an out- herent structure even at temperatures comparable to the
of-equilibrium system towards its thermodynamic equi- glass transition temperature [25].
librium [14, 15]. These processes become apparent when In computer glass models, energy dissipation is stud-
we regard only an (open) subsystem of a larger system, ied either by measuring the phase shift between the ap-
e.g. by integrating out part of the microscopic degrees of plied strain and the resulting stress in brute-force simula-
freedom that do not belong to the subsystem: The sepa- tions [26–29], or by relating the dynamics of vibrational
ration into a system of interest and a “heat bath” (whose normal modes to the macroscopic viscoelastic proper-
detailed state we do not know) leads to the emergence of ties [30–35]. In the brute-force approach, a sinusoidal
frictional forces (potentially with memory) in the equa- strain γ(t) = γ0 sin (Ωt) with driving frequency Ω and
tions for excitations of the system of interest [14, 15]. For amplitude γ0 is applied to a representative volume el-
solids, such as crystals or glasses, the natural description ement and the resulting stress σ(t) is measured. The
temperature is kept constant during deformation by us-
ing a thermostat, which typically adds a viscous damp-
ing with a characteristic time constant τBF [36–39]. Fit-
∗ Corresponding author: lars.pastewka@imtek.uni-freiburg.de ting the stress response with σ(t) = σ0 sin (Ωt + δ) yields
2

stress amplitude σ0 and phase shift δ. If the macroscopic power law potential of the form
deformation is pure shear with strain γ(t), the complex  " #
q
shear modulus Ĝ(Ω) = G′ (Ω) + iG′′ (Ω) is given by [2]   β P  2l
ϵ σrIJ
IJ
+ c2l σrIJ , rIJ ≤ rc σIJ

U (rIJ ) = l=0
IJ

G′ = σ0 /γ0 cos δ and G′′ = σ0 /γ0 sin δ,



(1) 
0 , rIJ > rc σIJ

where G′ is the storage modulus (measure of stored elas- where β determines the softness, rc is the cutoff radius,
tic energy) and G′′ is the loss modulus (measure of the σIJ is the pairwise interaction length and ϵ is an energy
amount of dissipated energy). scale. The pairwise interaction length is given by the
nonadditive mixing rule
In the microscopic approach, the movement of atoms
around their inherent structure is modeled as driven, 1
damped harmonic Langevin oscillators [30, 31]. Prior σIJ = (σI + σJ )(1 − na |σI − σJ |) (2)
2
work has used constant microscopic damping with a
wavelength-independent relaxation time τ in such micro- with na = 0.1. The coefficients c2l in the potential are
scopic determination of viscoelastic properties [30, 31, chosen such that the potential is continuous up to the
33, 34]. The value of τ was chosen in such a way that q-th derivative at the cutoff [45]. Requiring that the po-
the temperature remained constant during deformation tential and its first q derivatives vanish at the cutoff rc
or by examining the results for different choices of damp- fixes the coefficients [43],
ing, lacking an intrinsic physical motivation for the spe-
cific choice. Constant damping leads to dissipation that (−1)l+1 (β + 2q)!!
c2l = r−(β+2l) . (3)
is independent of wavelength, and this violates momen- (2q − 2l)!!2l!! (β − 2)!!(β + 2l) c
tum conservation. In other words, fast modes will be
generally underdamped and slow modes will be generally We set β = 10 and smooth the potential up to the q = 3
overdamped, as discussed for example in Ref. [40]. As derivative. Throughout the paper we employ reduced
we will show below, a good choice for τ is a value where Lennard-Jones units, where energyp is measured in units
of ϵ, temperature in ϵ/kB , time in M σmin 2 /ϵ ≡ t∗ , fre-
the whole finite system is underdamped, in which case ∗ ∗ −1 3
the time scales of energy flow to the thermostat and of quency in f ≡ (t ) and stress in ϵ/σ .
coupling between phononic excitations of the glass are Glassy configurations are obtained by equilibrating the
separated. structure at a parent temperature Tp using the swap
Monte-Carlo algorithm [25, 43]. Our ensembles of glasses
In the present work, we lift this restriction and use
consist of 100 individual configurations prepared at either
the actual damping constant (or lifetime) of the respec-
Tp = 0.3 ϵ/kB or Tp = 0.4 ϵ/kB . System size dependence
tive vibrational normal mode. Inspired by the seminal
is studied using configurations with N = 8k, N = 64k
works of Ladd, McGaughey and co-workers [41, 42], we
and N = 256k atoms. The timestep for the molecular
measure the lifetimes from the decay of energy autocor-
dynamics simulations is determined by running equilib-
relation functions. This yields a first-principles theory
rium microcanonical (NVE) simulations and studying the
of viscoelastic dissipation in disordered solids. It allows
change in total energy. A timestep of ∆t = 0.001 t∗ for
us to predict the complex shear modulus Ĝ as a func-
temperatures T ≥ 10−2 ϵ/kB is sufficient to guarantee
tion of driving frequency Ω and temperature T , as well
energy conservation. In order to reduce computational
as to make a statement about the accuracy of assuming
cost, a timestep of ∆t = 0.003 t∗ is used at temperatures
constant damping.
lower than T = 10−2 ϵ/kB . Atomic configurations at fi-
nite temperature are equilibrated in the canonical (NVT)
ensemble using a Nosé-Hoover chain thermostat with a
II. METHODS relaxation time of 1 t∗ [38, 46].
These model glasses interact via purely repulsive
forces. This means that the final configurations are al-
A. Polydisperse model and preparation protocol ways under hydrostatic stress. Therefore care needs to
be taken when interpreting the elastic constants [47]. We
We use the polydisperse model glass described in come back to this in the discussion below.
Ref. [43] and only summarize the key parameters and
the interaction potential here. The system consists of
N particles of equal mass M and varying size σi . The B. Vibrational modes
size σi is drawn from a distribution P (σ) ∼ σ −3 with
σ ∈ [σmin = 1.0, σmax = 2.22], where we use σmin as our The ultimate goal is to model viscoelastic energy dissi-
unit of distance. The density of the solid is fixed at pation triggered by the excitation and subsequent decay
ρ = 0.58 σ −3 and periodic boundary conditions are em- of vibrational normal modes. These vibrational modes
ployed in all three spatial directions. The interaction is are excited because a time-dependent long-wavelength
governed by a pair-wise additive [44] smoothed inverse deformation gives rise to non-affine displacements that
3

have a non-zero projection onto the normal modes [30, and the vector ξ(t) contains independent white noise vari-
48, 49]. This mechanism is not limited to disordered sys- ables, i.e. ⟨ξIα (t)ξJβ (t′ )⟩ = δIJ δαβ δ(t − t′ ). The dynam-
tems but also is relevant even for simple crystals [2, 28]. ics of the Langevin oscillators is driven by the non-affine
In the following we denote 3-vectors by arrows (e.g. forces that emerge during cell deformation.
χ
⃗ i ) and 3N -vectors using bold symbols (e.g. em ). Sim- The natural frequencies and eigenvectors are obtained
ilarly, we will use an underline for a 3 × 3 tensor (e.g. from diagonalization of the Hessian, namely by solving
H IJ ) and a calligraphic symbol for a 3N × 3N tensor 2
(e.g. H). Atoms are identified by capital Roman in- M ωm em = Hem . (9)
dices, modes (see below) by lowercase Roman indices and Since H is symmetric, the eigenvectors are orthogonal
Cartesian components by Greek indices. We consider a and we also assume that they are normalized such that
simulation cell whose shape is described by the cell ma- em · el = δml . Given Q = (e1 , ..., e3N ), we can write
trix h̊ ≡ (⃗a1 , ⃗a2 , ⃗a3 ). The cell is filled with N atoms resid- H = M Q · Ĥ · Q−1 where Ĥ is a diagonal matrix of
ing in a local potential energy minimum at positions ˚ ⃗rI . the corresponding eigenfrequencies ωm 2
. We can use the
We refer to this initial state as “reference” and mark all normal-mode basis em to define modal displacements
quantities which are defined in this reference by a circle. X = M Q−1 · χ. The dynamical Eq. (6) then becomes
For amorphous solids, the set of equilibrium positions
{˚⃗rI } is typically called the inherent structure [20, 50]. Ẍ(t) + F̂ · Ẋ(t) + Ĥ · X(t) = Ξ̂αβ ηαβ (t) + Ŝ · Z(t) (10)
In the athermal limit, we distinguish between affine
and nonaffine mechanical deformation [51]. The ini- with F = M Q · F̂ · Q−1 , S = M Q · Ŝ · Q−1 and Ξαβ =
tial simulation cell h̊ and the atomic positions ˚ ⃗rI are Q · Ξ̂αβ . The random vector Z(t) = Q−1 · ξ(t) is just
remapped by an affine transformation a rotation of independent white noise variables, which
itself is a vector of independent white noise variables.
−1 The core assumption of the viscoelastic model, which
⃗rI = h · h̊ ·˚
⃗rI (4)
is often not spelled-out explicity, is that Q not only diag-
where h and ⃗rI are now the “current” cell and atomic po- onalizes H, but also the friction matrix F, and therefore
sitions. For amorphous solids, such an affine transforma- by virtue of Eq. (8) also S. This means, the time evolu-
tion does not result in an atomic configuration in equi- tion of each mode m is given by an independent, driven
librium. Net “non-affine” forces Ξαβ ηαβ on the atoms Langevin oscillator
remain, where −1
Ẍm +τm 2
Ẋm +ωm −1
Xm = Ξ̂m,αβ ηαβ +2kB T τm Zm , (11)
∂2U with vibrational lifetime τm . The inverses of the life-
Ξαβ = − (5)
∂ηαβ ∂r times are the entries of the diagonal matrix F̂. The
common treatment in the literature assumes τm = τ ,
is a force tangent and η is the Green-Lagrange strain independent of the mode m [30, 31, 33]. We here relax
tensor [30, 47]. Allowing the solid to relax to the closest this specific assumption by using mode-dependent vibra-
local minimum leads to a configuration with vanishing tional lifetimes τm , which implicitly include full anhar-
non-affine forces. The atomic displacements ⃗uI during monic scattering [17, 53].
relaxation are called non-affine displacements. We can
−1
write the current atomic positions as ⃗rI = h · h̊ · ˚ ⃗rI +
⃗uI . The non-affine motion is best visualized by pulling C. Characterization of modes
the current atomic positions back to the reference and
defining pulled-back non-affine displacements as χ ⃗ I ≡ h̊ · 2
We characterize ωm and em by the vibrational density
h−1 · ⃗uI . of states (VDOS) g(ω), the participation ratio Pm and
At small finite temperature, atoms vibrate within their the phonon order parameter Om [54–58]. The VDOS is
inherent structure. We express these dynamics as a set given by
of coupled, driven Langevin oscillators in the reference
3N
configuration, 1 X
g(ω) = δ(ω − ωm ) (12)
3N − 3 m=4
M χ̈(t) + F · χ̇(t) + H · χ(t) = Ξαβ ηαβ (t) + S · ξ(t). (6)
where δ(ω − ωm ) is the Dirac delta function and we ne-
The matrix H is the Hessian with blocks glect the three translational modes.
∂2U A common measure used to characterize the spatial lo-
H KL = , (7) calization of vibrational modes is the participation ratio,
∂⃗rK ∂⃗rL
which is defined as [54–56]
F is a friction matrix and S is a matrix with noise am-  −1
plitudes. The fluctuation-dissipation theorem is [52] XN
N Pm =  (⃗emI · ⃗emI )2  . (13)
S · S T = 2kB T F (8) I=1
4

N Pm is the approximate number of atoms participating with


in the vibration. For extended modes, Pm is of the order N
of unity, independent of N , because all atoms partici-
X ∂2U
αm = : (⃗emI ⊗ ⃗emJ ) (18)
pate equally in the vibration. For localized modes, it ∂⃗rI ∂⃗rJ
I,J=1
scales with 1/N because only a finite number of atoms N
participate in the vibration. X ∂3U ...
βm = (⃗emI ⊗ ⃗emJ ⊗ ⃗emK ) (19)
In an elastic medium with translational symmetry, we ∂⃗rI ∂⃗rJ ∂⃗rK
I,J,K=1
expect a population of plane wave vibrational modes, the N
phonons. We evaluate the similarity of the vibrational ex- X ∂4U ....
γm = (⃗emI ⊗ ⃗emJ ⊗ ⃗emK ⊗ ⃗emL )
citations to plane waves by computing the phonon order ∂⃗rI ∂⃗rJ ∂⃗rK ∂⃗rK
I,J,K,L=1
parameter defined in Ref. [57, 58]. The idea is to expand
(20)
the eigenvectors em in the basis of plane waves wp (⃗q).
. .
The basis expansion is written as where ⊗ is the outer product and :, .. and ... are a double,
triple and quadruple contraction over all indices of the
tensors of second, third and fourth order, respectively,
X
em = Am,p (⃗q)wp (⃗q) (14) 2
q
⃗,σ
on both sides of the symbol. Note that αm = M ωm .

where Am,p (⃗q) are the expansion coefficients. In this D. Measuring the vibrational lifetimes
equation, ⃗q is the wave vector and p denotes the polar-
ization, which is either longitudinal or transversal. The
We measure lifetimes of the vibrational modes in
amplitude of a phonon mode with a certain (⃗q, p) for an
microcanonical (NVE) simulations of our equilibrated
atom I in the amorphous configuration is given by the
glasses. The computation of lifetimes is based on the
Fourier basis
assumption that the glass remains in its initial inher-
⃗np (⃗q)   ent structure, {˚⃗rI (t)} ≡ {˚
⃗rI (t = 0)}. Therefore it is
⃗ I,p (⃗q) = √ exp i⃗q · ˚
w ⃗rI (15) necessary to discard those simulations, in which the sys-
N tem evolves towards a new potential energy minimum.
A transition towards a new energy minimum is detected
where ⃗np (⃗q) is the polarisation vector. The longitudinal from the difference in the inherent structure. We define
polarisation vector is defined as ⃗nL (⃗q) = ⃗q/|⃗q| and the two (see also Ref. [19])
transversal polarisation vectors need to fulfill ⃗nT1 (⃗q)·⃗q =
⃗nT2 (⃗q) · ⃗q = ⃗nT1 (⃗q) · ⃗nT2 (⃗q) = 0. ∆rIS (∆t) = |r̊(t) − r̊(t − ∆t)| (21)
The phonon-order parameter is defined as the sum of and test for jumps in ∆rIS . Atomic rearrangements result
the projection of the eigenvectors em onto the plane-wave in values ∆rIS (∆t) ≈ 1 − 10 σ, independent of tempera-
basis wq⃗p , which are larger than some threshold α: ture. Simulations with ∆rIS > 10−2 σ undergo a config-
urational transition and we discard these simulations.
q
⃗Xmax   For those structures where no atomic rearrangement
X 2 2
Om = em · wp∗ (⃗q) θ em · wp∗ (⃗q) −α occurred during the simulation, we compute the lifetimes
q
⃗=⃗
qmin p of the vibrational modes from the decay of the autocor-
(16) relation of the total energy [41, 42, 62]. We write the
Here θ(x) is the Heaviside step function and the star Hamiltonian of the system without external deformation
P
indicates the complex conjugate. The sum runs over all in the harmonic approximation, Hharmonic = m Hm
wavevectors from the smallest ∼ 2π/L (where L is the lin- with
ear size of the system) up to a largest wavevector, that is 1 2 
2 2
given by the minimal distance between two atoms in the Hm (Xm (t), Ẋm (t)) = Ẋm (t) + ωm Xm (t) , (22)
2
configuration. In our analysis, we use a threshold value
of α = 50/(3N − 3). We compared different thresholds the contribution to the Hamiltonian of mode m. The
but found no influence on the result. value of Hm (t) for a specific mode m at a specific time t
is computed by projecting the atomic configuration r(t)
We characterize the anharmonicity of the vibrational
from the NVE simulation onto the eigenmodes em to
modes using higher order derivatives of the potential en-
obtain Xm . The contribution Hm (t) to the overall en-
ergy as described in Refs. [59–61]. The atomic positions
ergy is then computed from the harmonic approximation,
are displaced from the inherent structure ˚ ⃗rI in the di- Eq. (22).
rection of the eigenvectors em by a distance sem , where The lifetimes τm are given by the decay of the normal-
s is a scale factor of units length. We then expand the ized total energy autocorrelation function Cm [18, 41, 42,
potential energy U ({⃗rI }) in a Taylor series, 62],
1 1 1 ⟨δHm (t)δHm (0)⟩
U (sem ) − U (0) = αm s2 + βm s3 + γm s4 (17) Cm (t) = (23)
2 6 24 ⟨δH(0)δH(0)⟩
5

where δHm (t) = Hm (t) − ⟨Hm ⟩ is the deviation of the Considering only oscillatory deformations, we perform a
current total energy from its expectation value for mode Fourier transformation of Eq. (11) and Eq. (28). This
m. The thermodynamic average ⟨·⟩ has to be interpreted yields
as being taken over the basin of the potential energy land-
scape that belongs to the current inherent structure. We ∆Π̃αβ (Ω) = Dαβµν (Ω)η̃µν (Ω) (29)
sample it by running dynamic simulations and ensuring with
that the inherent structure does not change, as discussed
in the previous paragraph. Π 1 X Ξ̂m,αβ Ξ̂m,µν
Dαβµν = cΠ
αβµν + −1 , (30)
In the high temperature limit, equipartition of en- V̊ m
Ω2 − iΩτm − ωm2

ergy holds and it follows that ⟨H⟩ = T ϵ. Prior


work has shown (numerically and theoretically) that where a tilde indicates a Fourier-transformed quan-
Cm (t) exhibits an exponential decay Cm ∝ exp −t/τm tity. This result is an extension of the theory derived
[18, 19, 42, 62, 63]. For fitting τm , we compute a first by Lemaı̂tre and Maloney [30] that includes a mode-
guess by integrating Cm (t) up to a certain maximal dependent lifetime τm . We can write this equation in

threshold, the more common form for the storage Dαβµν and loss
′′
modulus Dαβµν as
tmax
⟨δHm (t)δHm (0)⟩
Z

τm = dt , (24) 1 X Ξ̂m,αβ Ξ̂m,µν (Ω2 − ωm
2
)
0 ⟨δH(0)δH(0)⟩ Π′
Dαβµν = cΠ
αβµν + (31)
2
2 ) + Ω2 τ −2
V̊ m (Ω2 − ωm m
where tmax is taken as the value where the correlation −1
functions drops below 1/e. The approximated lifetimes Π′′ 1 X Ξ̂m,αβ Ξ̂m,µν τm Ω
Dαβµν = 2 . (32)
′ V̊ 2 ) + Ω2 τ −2
(Ω2 − ωm
τm are used as initial guesses for least-squares fits of τm m m
to Cm (t). The data is fitted again up to the time where
Cm (t) has dropped to 1/e. Lemaı̂tre et al. showed that the values of Ξ̂m,αβ can be
considered as independent random variables which are
self-averaging quantities [30]. Therefore, we define the
E. Viscoelastic moduli non-affine correlator
3N
1 X
In order to derive an expression for the viscoelastic re- Γαβµν (ω) = δ(ω − ωm )Ξ̂m,αβ Ξ̂m,µν . (33)
3N − 3 m=4
sponse, we follow the derivation of Lemaı̂tre et al. [30].
We consider the variation of the stress Π from time- Using the non-affine correlator, we can write the storage
dependent box deformation, described by a Lagrange and loss modulus as
strain η(t). We write
1 ∞ g(ω)Γαβµν (ω)(Ω2 − ω 2 )
Z
Π′
Dαβµν = cΠ
αβµν + dω 2
 
V̊ 0 (Ω − ω 2 )2 + Ω2 τ −2 (ω)
1  ∂U ∂U (34)
∆Παβ (t) = − , (25)

Z ∞ −1
∂ηαβ ∂ηαβ

V̊ η=η(t) η=0 Π′′ 1 g(ω)Γαβµν (ω)τ (ω)Ω
Dαβµν = dω 2 (35)
V̊ 0 (Ω − ω 2 )2 + Ω2 τ −2 (ω)
where Π has the properties of a second Piola-Kirchhoff in the thermodynamic limit of large systems. In the limit
stress. We consider only small fluctuations and expand Ω → 0, the loss modulus vanishes and the storage mod-
Eq. (25) in a Taylor series up to first order, ulus converges to the static elastic constants,
1 ∞
Z
∆Παβ (t) = cΠ −1 g(ω)Γαβµν (ω)
αβµν ηµν (t) − V̊ Ξµν · χµν (t) (26) Π′
Dαβµν → cΠαβµν − dω , (36)
V̊ 0 ω2
where where the last summand is the correction to the elastic
constants from non-affine forces [30, 47]. Conversely, in
1 ∂2U

αβµν = (27) the limit Ω → ∞ the loss modulus also vanishes but
V̊ ∂ηαβ ηµν η=0 the storage modulus converges to DαβµνΠ′
→ cΠ
αβµν , which
are the static elastic constants under the assumption of
is the tensor of elastic constants at constant Piola- purely affine deformation.
Kirchoff stress and without non-affine displacements [44, Equation (30) is valid for arbitrary deformation modes.
47] Since amorphous solids behave isotropically, the stress is
Following the derivation of Sec. II B, we project onto hydrostatic and our interatomic potential is purely repul-
the vibrational normal modes to obtain sive, we consider only the case of simple shear. Therefore,
we define the mean shear modulus,
1 X
∆Παβ (t) = cαβµν ηµν (t) − Ξ̂m,αβ Xm (t). (28)
h i
V̊ GΠ (ω) = D1212
Π Π
(ω) + D1313 Π
(ω) + D2323 (ω) /3, (37)
m
6

which is the susceptibility that we discuss throughout this III. RESULTS


paper. We will also refer to the shear correlator Γ below,
which is the mean shear component of the correlator, A. Spatial structure of vibrational modes
Eq. (33).
In order to classify vibrational modes in amorphous
solids, it is useful to define two characteristic frequencies.
The first one is the frequency of the Boson peak, ωBP ,
which is determined from the maximum in the reduced
F. Brute-force calculations
VDOS g(ω)/ω 2 [58, 66]. The second one is the frequency
of the lowest phonon mode ωPh , which can be obtained
The previous section describes the linear response the- from [67]
ory for viscoelastic dissipation. We also compute vis- s
coelastic properties from “brute-force” calculations of os- 2π G′ (Ω → 0)
cillatory cell deformation. The cell is subjected to time- ωPh = , (40)
L ρ
dependent simple shear with strain γxy (t), which means
we evolve the cell matrix according to where ρ is the density and L is the linear size of the box.
For the ensemble of configurations with N = 8k atoms,
  ⟨ωBP ⟩ ≈ 3 f ∗ and ⟨ωPh ⟩ = 1.23 f ∗ . For the large system
1 γxy (t) 0 size with N = 256k atoms, the lowest phonon frequency
h(t) = 0 1 0 · h̊ (38) is at ⟨ωPh ⟩ = 0.39 f ∗ .
0 0 1 Figure 1 shows the vibrational density of states
(VDOS), the participation ratio Pm and the plane-wave
and affinely remap all atomic positions during each strain order parameter Om . At frequencies higher than the Bo-
increment. We measure the resulting shear stress σxy son peak, we see (Fig. 1a) that the VDOS has one dom-
from the virial theorem [64] and fit amplitude σ0 and inant local maximum at ≈ 10 f ∗ and one shoulder at
phase lag δ to compute the complex shear modulus via ≈ 21 f ∗ . In the mid-frequency range 3 f ∗ ≤ ω ≤ 23 f ∗ ,
Eq. (1) (see also Refs. [26, 28, 29, 31, 33]). For each ex- we find delocalized vibrational modes (large Pm ). In the
citation frequency Ω we simulate 60 periods. We neglect high-frequency end above ω ≥ 23 f ∗ , we see a sudden de-
the first 10 periods during which the system approaches crease of Pm , indicative of spatial localization. The vibra-
a dynamical steady-state and average the results for each tional modes in the mid-frequency as well as in the high-
excitation frequency over the remaining 50 periods and frequency range are poorly described by plane waves, as
ensembles with 50 independent configurations. The sys- shown in Fig. 1c. We call vibrational modes in the mid-
tem is strained to γ0 = 1.0% and the temperature is kept frequency range disordered and the strongly localized vi-
constant at T = 10−2 ϵ/kB using a Langevin thermostat brational modes in the end of the frequency spectrum
with a relaxation time constant of τBF = 0.1 t∗ [39]. This localized [68].
maximal strain is well below the yield point. We tested Vibrational modes with frequencies below the Boson
that using a strain amplitude of 0.5% or 1.5% does not peak show discrete local maxima. The first peak coin-
affect the results. cides with the lowest phonon frequency ωPh . Based on
the participation ratio and the phonon order parameter
Care needs to be taken when comparing the micro- in Fig. 1b and c, two types of vibrational modes coexist
scopic theory to brute-force calculations. In the former, in this limit: There are clusters of delocalized vibrational
the stress is defined as the derivative of the potential modes with similar frequencies ωm , and large Pm and
or free-energy with respect to the Green-Lagrange strain Om . The absolute values of Pm and Om in one of these
tensor η. In the latter, we compute the instantaneous clusters show large fluctuations. Based on their similarity
stress using the virial theorem, which gives the Cauchy to phonons in crystalline materials, we term these vibra-
stress [64]. Elastic constants derived from either expres- tional modes plane-waves [58, 67]. The second type of
sion differ by a constant that depends explicitly on the vibrational modes observed in the low-frequency range
stress. Since our system is (approximately) hydrostatic, are localized and have a low overlap with plane waves.
the difference between the shear modulus at constant In accordance with the literature, we term this type of
Cauchy stress G and at constant second Piola-Kirchhoff vibrational mode quasilocalized [54, 55, 58].
stress GΠ is [47, 65] Larger systems (Fig. 1e-f) show identical behavior. In
the low-frequency limit (Fig. 1d), the number of discrete
G = GΠ − P, (39) maxima increases over the smaller system as lower vi-
brational frequencies are resolved. While the separation
between quasilocalized and plane-wave vibrational modes
where P = −σαα /3 is the hydrostatic pressure. In the becomes clearer as a consequence of the increasing sys-
following we report just G for both microscopic theory tem size, the overlap parameter shows larger fluctuations
and brute-force calculations. within each cluster of plane-wave modes. These observa-
7

N=8k N=256k Each ensemble is prepared at a parent temperature of


Ph BP (a) 0.2 Ph (d) kB Tp = 0.4 ϵ. For N = 8k atoms, we show the full fre-
6
VDOS g( ) (t * )
quency spectrum in Fig. 2a. The high-frequency end,
ω ≥ 23 f ∗ , is dominated by localized modes. Their vi-
4 brational lifetimes are frequency-independent. Decreas-
0.1
ing the frequency to ωBP , increases the vibrational life-
2 times. Below the Boson peak, we identify a splitting of
the vibrational lifetimes. The lifetimes can be separated
0 0.0 in vibrational modes with long lifetimes and vibrational
10
0
(b) 10
0
(e) modes with short lifetimes. Comparing this result with
1 1
Fig. 1, we identify the modes with long lifetimes as the
10 10 plane-wave modes and the latter as the quasilocalized
2 2 modes. This splitting of vibrational lifetimes is more
10 10
Pm

3 3 clearly visible in Fig. 2b, c where we show lifetimes for


10 10 larger system sizes with N = 64k and N = 256k atoms.
4 4
10 10 We can clearly observe that the upper limit of lifetimes
5 5 of plane-wave modes scales as ω −2 and that their value in
10 10 a cluster fluctuates over up to two orders of magnitude.
1.0 (c) 1.0 (f)
Aging effects are more pronounced for larger sample
localized
delocalized sizes and in less stable glasses, which are prepared at
higher parent temperature Tp [19]. To avoid aging, we
Om

0.5 0.5 study the temperature dependence of the vibrational life-


times for temperatures up to kB T = 0.1 ϵ using amor-
phous configurations with N = 2k atoms prepared at a
low parent temperature of kB Tp = 0.3 ϵ. The highest
0.0 0.0 temperature is in the same order of magnitude as their
0.3 1.0 10.0 0.3 1.0 mode-coupling temperature (kB TMCT ≈ 0.56 ϵ [43, 69]).
m (f m (f
*) *)
Third-order perturbation theory predicts the vibrational
lifetimes to scale as τm ∝ (kB T )−1 [16, 17, 19]. Fig-
FIG. 1. The ensemble averaged VDOS, the participation ure 2d shows kB T τm for one exemplary configuration.
ratio Pm and the overlap parameter Om for a subset of con- The temperature-scaled lifetimes collapse onto an uni-
figurations with N = 8k atoms (a)-(c) and N = 256k atoms versal curve.
(d)-(f). Each configuration is prepared at a parent temper-
ature of Tp = 0.4 ϵ/kB . The vertical arrows mark the first
phonon frequency ⟨ωPh ⟩ and the location of the Boson peak
⟨ωBP ⟩. The horizontal dashed lines mark the thresholds in the C. Correlator of non-affine forces
participation ratio to distinguish localized and delocalized vi-
brational modes. Threshold are chosen as Pm = 10−2 and In Figure 3a we show the ensemble-averaged shear cor-
Pm = 10−3 for system with N = 8k and N = 256k atoms. relator ⟨Γ(ω)⟩ defined in Eq. (33) for configurations with
N = 8k, N = 64k and N = 256k atoms. For fre-
quencies above the Boson peak, the shear correlator in-
tions are in agreement with previous works on vibrational creases with increasing vibrational frequency. In the mid-
modes in disordered solids [54, 55, 58]. frequency region of the disordered modes, the increase is
well represented by a power law ∝ ω a . By fitting the
N = 8k data in the frequency range 3 ≤ ω ≤ 24 we
B. Lifetimes of vibrational modes obtain a ≈ 0.6. For vibrational modes with frequen-
cies smaller than the Boson peak frequency, the shear
Amorphous solids are intrinsically out-of-equilibrium correlator splits in two branches. For the first branch
and therefore their atomic structure evolves in time (the we observe a frequency independent shear correlator of
structure ages) even at temperatures far below the glass ⟨Γ(ω)⟩ ≈ 200 σ/ϵ. The shear correlator of the second
transition temperature [19, 21]. In order to suppress branch decreases with a power a ≈ 2.4, computed for
aging during measurement of vibrational lifetimes, sim- the ensemble with N = 256k and vibrational frequen-
ulations are performed at a low temperature of T = cies ω ≤ 1.5 f ∗ . The behavior in both branches is inde-
10−4 ϵ/kB . As described in the methods, we detect aging pendent of system size. Note that Zaccone and Scossa-
by comparing the atomic positions in the inherent struc- Romano [70] derived a value of a = 2 for an effective
ture and discard simulations when these atomic positions medium description of an amorphous solid.
change over time. Figure 3b shows the shear correlator as a function of
Figure 2a-c show the vibrational lifetimes for configu- the corresponding participation ratio ⟨Pm ⟩. Spatially de-
rations with N = 8k, N = 64k and N = 256k atoms. localized vibrational modes (large Pm ) have a small shear
8
3
10 (a) N=8k (b) N=64k (c) N=256k (d) N=2k
BP
1 1 2
2 10 10 10
10
2 ( f *)

0 0
10
1 10 10
1
m kB T

10
0 10
1
10
1 T ( /kB)
10 localized 10 1 10 3
delocalized 10 2 10 4
1 2 2 0
10 10 10 10
1 10 30 0.3 1.0 0.3 1.0 1 10 30
m (f ) m (f ) m (f ) m (f )
* * * *

FIG. 2. Vibrational lifetimes τm for ten configurations with N = 8k atoms (a) and for full ensembles with N = 64k atoms (b)
and N = 256k atoms (c). All configurations are prepared at a parent temperature of kB Tp = 0.4 ϵ. Lifetimes are computed at a
temperature of T = 10−4 ϵ/kB . (d) Rescaled vibrational lifetimes for one configuration prepared at kB Tp = 0.3 ϵ with N = 2k
atoms. The lifetimes are rescaled with kB T ω 2 as expected from a third-order perturbation theory. Symbols in each panel are
used to distinguish localized and delocalized modes, as determined from their participation ratio.

correlator while localized vibrational modes (small Pm ) before dropping to a value of G′ ≈ 21 ϵ/σ 3 at the high-
have large values of the shear correlator. The shear corre- frequency end. Using a constant lifetime of τm = 10 t∗
lator as a measure of non-affinity is hence also a measure matches the full microscopic theory over the range of ex-
of localization. citation frequencies considered here.
Figure 4c shows the low-frequency limit of the stor-
age modulus. We present only results for the ensemble
D. Frequency-dependent viscoelastic properties of structures with N = 8k atoms because the sum in
Eq. (30) requires all eigenmodes up to the Debye fre-
We compare the frequency-dependent viscoelastic quency of the system for convergence, but this is com-
moduli computed by three different approaches. In putationally prohibitive for the larger systems. The con-
the first approach, we compute the shear moduli from vergence of this sum has been studied in Ref. [71]. For
Eqs. (31) and (32) using the mode-dependent vibrational the small system sizes, the storage modulus in the limit
lifetimes τm at a temperature of T = 10−2 ϵ/kB . For the Ω → 0 converges independent of the assumed lifetime τ
second approach, we assume a constant lifetime i.e. we to the static solution. The peaks in the storage mod-
set τm = τ and evaluate again Eqs. (31) and (32). In the ulus at low frequencies occur at the characteristic fre-
last approach we perform brute-force (BF) simulations, quencies of the plane-wave vibrational modes. At the
in which we directly “wiggle” the simulation cell and Boson-peak frequency, the storage modulus has a local
measure the resulting stress (see Methods). Note that minimum which can be traced back to a combination of
in contrast to the simulations from which we extracted many vibrational modes, long vibrational lifetimes and a
the vibrational lifetimes, we kept atomic trajectories in high non-affine correlator.
which atomic rearrangements occurred during mechani- The behavior of the loss modulus Fig. 4b reflects the
cal deformation. The results are summarized in Fig. 4, shape of the VDOS in Fig. 1a, d. For increasing exci-
which shows storage modulus G′ and the loss modulus tation frequencies, we observe a dramatic increase of the
G′′ for oscillating shear deformation as a function of os- loss modulus near the Boson peak. The loss modulus has
cillation frequency Ω. From Fig. 4a and b we see that a broad local maximum at Ω ≈ 10 f ∗ and a ”shoulder” at
the results for the loss modulus and the storage modulus Ω ≈ 21 f ∗ . In the low and the high end of the frequency
match perfectly for brute-force calculations and micro- spectrum, the loss modulus decreases rapidly and con-
scopic theory with a constant τm = τBF . This is an indi- verges to zero. Similar to the VDOS, we observe the
cation that at the temperatures and excitation frequen- occurrence of discrete local maxima in the low-frequency
cies studied here atomic rearrangement do not contribute end of the spectrum at frequencies of the plane-wave like
to the viscoelastic modulus. vibrational modes. Below the smallest vibrational fre-
Having validated the microscopic theory, we now turn quency in the system, the loss modulus scales linearly
to the question of whether an optimal choice of a mode- with the excitation frequency ∝ Ω, see Eq. (32). Examin-
independent lifetime can recover the correct frequency- ing the results for the loss modulus obtained for constant
dependent behavior. In contrast to the calculations with lifetimes, we see that the mid-frequency and the high-
τ = 0.1 t∗ , the storage modulus has a clear minimum at frequency part of the spectrum is in agreement with the
the Boson peak frequency, Ω = ωBP . With increasing ex- correct results as long as τ ≥ 0.1 t∗ . One feature visible
citation frequency, the storage modulus increases rapidly in Fig. 4 is that the Boson-peak frequency marks a turn-
and reaches a maximum of ≈ 24.6 ϵ/σ 3 at Ω ≈ 26 f ∗ ing point at which the loss modulus becomes strongly
9

Non-affine correlator ( ) ( / )
(a) BP
temperature. It appears, that these discrete peaks have
10
3 0.6 the same amplitude as the mean value at higher temper-
atures i.e. the peaks are smeared out. Below the first vi-
brational mode, the loss modulus decays linearly for both
2
10 temperatures. The loss moduli at low frequency differ by
L DL N a constant factor that depends linearly on temperature.
10
1 8k

2.4
64k
256k IV. DISCUSSION
0
10
1 0 1
10 10 10
The physical interpretation of Eq. (30) is that storage
Vibrational frequency (f * ) and loss modulus are determined by the superposition of
0 vibrational modes with eigenfrequencies near the excita-
10 (b) tion frequency. Magnitude as well as frequency depen-
1
10 dency of the viscoelastic properties are directly related to
Participation ratio Pm

2
the properties of the excited vibrational modes. Based
10
3 f* on the excellent agreement in Fig. 4 between brute-force
3 simulations in which atomic rearrangements occur and
10 the theory in Eq. (30), excitation of vibrational modes
4 by non-affine displacements is the dominant mechanism
10
of viscoelasticity in the investigated frequency range.
5 In the frequency range above the Boson peak, vibra-
10
0 1 2 3 tional modes are continuously distributed. This distri-
10 10 10 10
Non-affine correlator ( ) ( / ) bution together with short lifetimes results in a smooth
frequency dependency of the viscoelastic moduli due to
FIG. 3. (a) Mean shear correlator as a function of vibrational
averaging over many vibrational modes. The density of
frequency. The two green dashed lines are power law fits vibrational modes is lower towards the high frequency
∝ cω a . The power law ∝ ω 0.6 is a fit to the N = 8k data end of the spectrum than near the Boson peak. These
in the frequency range 3 ≤ ω ≤ 20 while the powerlaw fit modes have a short lifetime and large non-affine correla-
∝ ω 2.4 is fitted to the N = 256k data for ⟨Γ(ω)⟩ ≤ 102 and tor, but the loss modulus at high frequency is low because
ω ≤ 1.5. (b) Dependence of the shear correlator on the spatial the low density of states.
localization of the vibrational modes. Symbols in each panel Below the Boson peak, both viscoelastic moduli are
are used to distinguish localized and delocalized modes, as dominated by the structure of plane-wave or hybridized
determined from their participation ratio. modes that occur in discrete frequency bands for the
finite-sized systems studied here. The discrete bands
are visible in the loss modulus (Fig. 4b,d and Fig. 5b).
dependent on the choice of lifetime. Using fixed lifetimes They smooth out at high temperature Fig. 5a) as, by
tend to overestimate the loss modulus below the Boson virtue of the temperature-dependence of the lifetimes, the
peak and fails to reproduce the local maxima. This effect line width of each oscillator becomes broader with tem-
is even more pronounced for larger system sizes shown in perature. The storage modulus interpolates smoothly
Fig. 4d. between low-frequency and high-frequency limit (see
Although we only performed a partial diagonalization, Sec. II E) as each oscillator contributes step-like rather
the contribution from vibrational modes with frequen- than a Lorentzian function.
cies ωm away from the excitation frequency Ω do not Vibrational lifetimes can be directly computed using
contribute much. Using mode-dependent lifetimes, we perturbation theory [16, 17, 72]. By comparing theoreti-
observe that the discrete peaks become narrower with de- cal and numerical vibrational lifetimes, previous publica-
creasing frequency. In regions without vibrational modes, tions showed that cubic anharmonicity is the main source
the loss modulus is overestimated up to three orders of of anharmonic coupling between vibrational excitations
magnitude while in regions with vibrational modes it in glasses [17, 19, 63]. Furthermore, perturbation theory
tends to be underestimated. predicts that vibrational lifetimes scale with temperature
The effect of temperature on the loss modulus is intrin- ∝ T −1 which is consistent with our results in Sec. III B
sically included in the vibrational lifetimes. In Fig. 5, we [17]. This scaling is observed to be independent of the
show the loss modulus G′′ for two different temperatures. type of vibrational mode and even holds for temperatures
The shape and the mean magnitude of the loss modulus near the glass transition.
above the Boson peak frequency is independent of tem- To understand whether vibrational modes with shorter
perature. Below the Boson peak, but in the frequency lifetimes are related to a larger cubic anharmonicity as
range where vibrational modes exist, narrower and larger suggested by perturbation theory, we evaluate the third
peaks compared to the surrounding are visible for lower order anharmonicity defined in Eq. (20). In Figure 6
10

24 (a) N=8k BP
10
1
(b) N=8k BP
m
22 BF = 0.1

Storage modulus G 0 ( / 3)
= 0.1

Loss modulus G 00 ( / 3)
0
10
20 = 1.0
= 10.0 1
18 10

16 2
10
14
3
10
12

10 20.93 ( / 3) 10
4

1 0 1 2 1 0 1 2
10 10 10 10 10 10 10 10
0
10
15 (c) N=8k (d) N=256k
= 50.0
Storage modulus G 0 ( / 3)

Loss modulus G 00 ( / 3)
10
14

2
13 10

3
12 10

12.88 ( / 3)
4
11 10
1 0 1 0
10 10 10 10
Excitation frequency (f * ) Excitation frequency (f * )

FIG. 4. Ensemble averaged storage ⟨G′ ⟩ (a) and loss modulus ⟨G′′ ⟩ (b) for the configurations with N = 8k atoms in the full
frequency range. The vertical arrow marks the location of the Boson peak. (c) Detailed view on the storage modulus in the
low-frequency limit. The horizontal dashed-dotted line marks the value of ⟨G′ ⟩ in the static limit Ω → 0. (d) Loss modulus
in the low-frequency limit for the ensemble of configurations with N = 256k atoms. Results for ⟨G′ ⟩ and ⟨G′′ ⟩ indicated
by lines are computed using Eq. 30. We employ the actual vibrational lifetimes τm or we assumed a certain constant and
mode-independent lifetime τm = τ . Results from brute-force simulations are denoted by τBF and marked by points. All lines
represent mean values obtained from averaging over multiple configurations while the shaded areas and the error bars mark
the standard deviation.

we present the anharmonicity of the low-frequency vi- with normalized Lorentzian L(x; ϵ) = ϵ[x2 + ϵ2 ]−1 /π.
brational modes for the ensemble of configurations with Since the lifetime τ is roughly constant for quasilocal-
N = 8k and N = 64k atoms prepared at Tp = 0.4 ϵ/kB . ized modes and scales as ω −2 for plane-waves, i.e. it
The figure shows that vibrational modes with shorter life- does not decrease with frequency ω, we can approximate
times tend to have a larger anharmonicity. Furthermore, the Lorentzian by a δ-function at small Ω. This yields
we observe that quasi-localized modes feature larger an-
π g(Ω)⟨Γ(Ω)⟩
harmonicity and therefore shorter lifetimes. G′′ (Ω) ≈ (42)
In molecular simulations we can only probe the high- 2V̊ Ω
frequency limit of the viscoelastic response, in the regime in the low-frequency limit. This result also means, that
of THz for real solids. While the storage modulus appears using a single relaxation time (as in brute-force calcula-
to converge within the frequency range addressable in the tions) works in the low-frequency response limit because
simulation, the asymptotic behavior of the loss modulus at low enough frequency the dominant modes are always
remains unclear. We now attempt to extrapolate the low- underdamped. However, studying the high-frequency
frequency behavior of the loss modulus from the general viscoelastic response near the Boson peak requires either
trends of our simulations. First, we rewrite Eq. (35) as the use of the correct (mode-dependent) relaxation times
or of a global relaxation time that is low enough such that
Z ∞
′′ π the whole system is underdamped (see also Fig. 4).
G (Ω) = dω g(ω)⟨Γ(ω)⟩L(Ω2 − ω 2 ; Ωτ −1 ) (41)
V̊ 0
Our results and prior work [54, 55, 58, 73–75] show that
11

T = 10 2 ( /kB) T = 10 4 ( /kB) the delocalized modes become dominant at low Ω. Cur-


(a) (b) rent limitations in accessible frequency are of a numeri-
10
0 cal nature, since either the diagonalization of the sparse
Hessian or the accessible time scales in brute-force calcu-
G 00 ( / 3)

2 lations become prohibitive for large system sizes.


10
One question that arises in this context is whether the
4
10 scaling relations in Eqs. (43) and (44) are valid for arbi-
trarily small excitation frequencies in the thermodynamic
6
10 limit N → ∞. This question is directly linked to the exis-
1 0 1 1 0 1
10 10 10 10 10 10 tence of localized modes in the thermodynamic limit. In
(f * ) (f * ) [80] it was shown that localized vibrational modes and
plane wave-like modes can only be clearly identified and
FIG. 5. Temperature dependence of the loss modulus ⟨G′′ ⟩ coexist in a frequency range
for the ensemble of configurations with N = 8k atoms. The
vertical dashed line marks the location of the Boson peak. In ωg ∼ L−3/5 ≲ ω ≲ L−2/5 ∼ ω† (45)
both figures, the straight lines mark the ensemble mean while
the shaded areas mark the standard deviation. This scaling suggests that in the thermodynamic limit,
the two types of mode can no longer be distinguished
10
2 N = 64k, localized due to hybridization with each other. Another limita-
N = 64k, delocalized tion is the existence of two-state processes, which we did
not consider in the present work. As the probability of
1
10
m kB T m ( f )

transition between distinct minima in the potential en-


*

ergy landscape [21, 24] of the glass increases with waiting


2

10
0 time, these transitions will also become dominant in the
low-frequency limit of the loss modulus. We therefore
1
conclude that our derived scaling represent a distinct in-
10 termediate regime, similar to the one observed in wave
scattering in glasses [81].
2
10
8 6 4 2 0 2
10 10 10 10 10 10
Cubic anharmonicity | m| V. SUMMARY AND CONCLUSION

FIG. 6. Scaled vibrational lifetime as a function of the cubic In summary, we performed an extensive characteri-
µm anharmonicity of vibrational modes in the low-frequency zation of vibrational modes in ultrastable glasses. We
limit ω ≤ ⟨ωBP ⟩ for configurations with N = 64k atoms. observe four different types of vibrational modes which
have different static and dynamic properties. In the
low-frequency limit, where quasilocalized and plane-wave
we need to distinguish between delocalized (plane-wave- like modes coexist, we observe different scaling rela-
like) and (quasi-)localized modes in the low-frequency tions for vibrational lifetimes of these modes. While
limit. We therefore split the the loss modulus into these the temperature dependency of both types of vibra-
two contributions. We know that plane-waves have a tional modes agrees with the prediction from perturba-
density of states gDL (ω) ∝ ω 2 [76], while quasilocalized tion theory, the predicted frequency dependency is ob-
modes contribute with gL (ω) ∝ ω 4 [55, 58, 75, 77–79] served only for the plane-wave modes. We used the
From our data, we find that the correlator ⟨Γ(ω)⟩ has mode-dependent vibrational lifetimes to extend the the-
power-law behavior ⟨ΓDL (ω)⟩ ∝ ω a with a ≈ 2.4 for ory of Lemaı̂tre et al. [30], yielding a parameter-free
plane-waves but is roughly constant ⟨ΓL ⟩ = const. for theory of linear viscoelasticity. The resulting expres-
quasilocalized modes. We hence find sion indicates, that quasilocalized excitations may be-
come the dominant dissipation channel at intermediate
G′′DL (Ω) ∝ Ω3.4 (43) frequency, while the low-frequency viscoelastic response
is likely dominated by two-state processes in which the
for the delocalized modes and glass jumps between local minima in the potential energy
landscape [21, 24, 26, 27].
G′′L (Ω) ∝ Ω3 (44)

for the localized modes. This result means that the lo-
calized modes dominate the low-frequency viscoelastic re- ACKNOWLEDGEMENTS
sponse. We note that we cannot rule out from our data
that the correlator for the localized modes has a weak We thank Edan Lerner and Eran Bouchbinder for shar-
power-law dependency ⟨ΓL ⟩ = Ωb . For b > a − 2 ≈ 0.4, ing the configurations of ultrastable glasses used in this
12

work as well as for useful comments on the manuscript. the molecular dynamics simulations. We acknowledge
We thank Wolfram Nöhring and Richard Leute for fruit- funding by the Deutsche Forschungsgemeinschaft (DFG,
ful discussions. We used matscipy [82] for the calcu- grant PA 2023/2). Calculations were carried on NEMO
lation of the Hessian, Slepc [83, 84] for the solution of at the University of Freiburg (DFG grant INST 39/963-
large-scale eigenvalue problems and LAMMPS [85] for 1 FUGG) and JUWELS at the Jülich Supercomputing
Center (grant hka18).

[1] J. D. Ferry, Viscoelastic properties of polymers (John Wi- 3839 (1996).


ley & Sons, 1980). [18] A. J. McGaughey and J. M. Larkin, Predicting phonon
[2] M. S. Blanter, I. S. Golovin, H. Neuhauser, and H. Sin- properties from equilibrium molecular dynamics simula-
ning, Internal friction in metallic materials: A Handbook tions, Annu. Rev. Heat Transfer 17, 49 (2014).
(Springer-Verlag, 2007). [19] H. Mizuno, H. Tong, A. Ikeda, and S. Mossa, Intermit-
[3] R. Lifshitz and M. L. Roukes, Thermoelastic damping tent rearrangements accompanying thermal fluctuations
in micro-and nanomechanical systems, Phys. Rev. B 61, distinguish glasses from crystals, J. Chem. Phys. 153,
5600 (2000). 154501 (2020).
[4] B. Houston, D. Photiadis, M. Marcus, J. Bucaro, X. Liu, [20] P. G. Debenedetti and F. H. Stillinger, Supercooled liq-
and J. Vignola, Thermoelastic loss in microscale oscilla- uids and the glass transition, Nature 410, 259 (2001).
tors, Appl. Phys. Lett. 80, 1300 (2002). [21] C. Oligschleger and H. Schober, Collective jumps in a
[5] O. Auciello, S. Pacheco, A. V. Sumant, C. Gudeman, soft-sphere glass, Phys. Rev. B 59, 811 (1999).
S. Sampath, A. Datta, R. W. Carpick, V. P. Adiga, [22] K. Vollmayr-Lee, Single particle jumps in a binary
P. Zurcher, Z. Ma, H. C. Yuan, J. A. Carlisle, B. Kabius, lennard-jones system below the glass transition, J. Chem.
J. Hiller, and S. Srinivasan, Are diamonds a MEMS’ best Phys. 121, 4781 (2004).
friend?, IEEE Microwave Mag. 8, 61 (2007). [23] M. P. Ciamarra, R. Pastore, and A. Coniglio, Particle
[6] J. K. Luo, Y. Q. Fu, H. R. Le, J. A. Williams, S. M. jumps in structural glasses, Soft Matter 12, 358 (2016).
Spearing, and W. I. Milne, Diamond and diamond-like [24] P. Leishangthem, A. D. Parmar, and S. Sastry, The yield-
carbon MEMS, J. Micromech. Microeng. 17, S147 (2007). ing transition in amorphous solids under oscillatory shear
[7] A. V. Sumant, O. Auciello, M. Liao, and O. A. deformation, Nat. Commun. 8, 1 (2017).
Williams, MEMS/NEMS based on mono-, nano-, and [25] A. Ninarello, L. Berthier, and D. Coslovich, Models and
ultrananocrystalline diamond films, MRS Bull. 39, 511 algorithms for the next generation of glass transition
(2014). studies, Phys. Rev. X 7, 021039 (2017).
[8] B. Arash, J.-W. Jiang, and T. Rabczuk, A review on [26] H.-B. Yu and K. Samwer, Atomic mechanism of inter-
nanomechanical resonators and their applications in sen- nal friction in a model metallic glass, Phys. Rev. B 90,
sors and molecular transportation, Appl. Phys. Rev. 2, 144201 (2014).
021301 (2015). [27] H.-B. Yu, R. Richert, and K. Samwer, Correlation be-
[9] M. Fan, A. Nawano, J. Schroers, M. D. Shattuck, and tween viscoelastic moduli and atomic rearrangements in
C. S. O’Hern, Intrinsic dissipation mechanisms in metal- metallic glasses, J. Chem. Phys. Lett. 7, 3747 (2016),
lic glass resonators, J. Chem. Phys. 151, 144506 (2019). 10.1021/acs.jpclett.6b01738.
[10] S. J. Turneaure, J. Winey, and Y. Gupta, Compressive [28] R. Ranganathan, Y. Shi, and P. Keblinski, Frequency-
shock wave response of a Zr-based bulk amorphous alloy, dependent mechanical damping in alloys, Phys. Rev. B
Appl. Phys. Lett. 84, 1692 (2004). 95, 214112 (2017).
[11] G. R. Khanolkar, M. B. Rauls, J. P. Kelly, O. A. Graeve, [29] O. Adeyemi, S. Zhu, and L. Xi, Equilibrium and non-
A. M. Hodge, and V. Eliasson, Shock wave response of equilibrium molecular dynamics approaches for the linear
iron-based in situ metallic glass matrix composites, Sci. viscoelasticity of polymer melts, Phys. Fluids 34, 053107
Rep. 6, 1 (2016). (2022).
[12] S. Chen, H. Cheng, K. Chan, and G. Wang, Metallic glass [30] A. Lemaı̂tre and C. Maloney, Sum rules for the quasi-
structures for mechanical-energy-dissipation purpose: A static and visco-elastic response of disordered solids at
review, Metals 8, 689 (2018). zero temperature, J. Stat. Phys. 123, 415 (2006).
[13] R. Renou, L. Soulard, E. Lescoute, C. Dereure, D. Loi- [31] T. Damart, A. Tanguy, and D. Rodney, Theory of har-
son, and J.-P. Guin, Silica glass structural properties un- monic dissipation in disordered solids, Phys. Rev. B 95,
der elastic shock compression: Experiments and molecu- 054203 (2017).
lar simulations, J. Phys. Chem. C 121, 13324 (2017). [32] B. Cui, J. Yang, J. Qiao, M. Jiang, L. Dai, Y.-J. Wang,
[14] R. Kubo, M. Toda, and N. Hashitsume, Statistical physics and A. Zaccone, Atomic theory of viscoelastic response
II: nonequilibrium statistical mechanics, Vol. 31 (Springer and memory effects in metallic glasses, Phys. Rev. B 96,
Science & Business Media, 2012). 094203 (2017).
[15] R. W. Zwanzig, Nonequilibrium Statistical Mechanics [33] V. V. Palyulin, C. Ness, R. Milkus, R. M. Elder, T. W.
(Oxford University Press, 2001). Sirk, and A. Zaccone, Parameter-free predictions of the
[16] A. Maradudin and A. Fein, Scattering of neutrons by an viscoelastic response of glassy polymers from non-affine
anharmonic crystal, Phys. Rev. 128, 2589 (1962). lattice dynamics, Soft Matter 14, 8475 (2018).
[17] J. Fabian and P. B. Allen, Anharmonic decay of vibra- [34] I. Kriuchevskyi, V. V. Palyulin, R. Milkus, R. M. Elder,
tional states in amorphous silicon, Phys. Rev. Lett. 77, T. W. Sirk, and A. Zaccone, Scaling up the lattice dy-
13

namics of amorphous materials by orders of magnitude, 66, 636 (1991).


Phys. Rev. B 102, 024108 (2020). [55] E. Lerner, G. Düring, and E. Bouchbinder, Statistics and
[35] D. Conyuh and Y. Beltukov, Ioffe-regel criterion and vis- properties of low-frequency vibrational modes in struc-
coelastic properties of amorphous solids, Phys. Rev. E tural glasses, Phys. Rev. Lett. 117, 035501 (2016).
103, 042608 (2021). [56] V. Mazzacurati, G. Ruocco, and M. Sampoli, Low-
[36] H. J. Berendsen, J. v. Postma, W. F. Van Gunsteren, frequency atomic motion in a model glass, Europhys.
A. DiNola, and J. R. Haak, Molecular dynamics with Lett. 34, 681 (1996).
coupling to an external bath, J. Chem. Phys. 81, 3684 [57] M. Shimada, H. Mizuno, and A. Ikeda, Anomalous vibra-
(1984). tional properties in the continuum limit of glasses, Phys.
[37] W. G. Hoover, Canonical dynamics: Equilibrium phase- Rev. E 97, 022609 (2018).
space distributions, Phys. Rev. A 31, 1695 (1985). [58] H. Mizuno, H. Shiba, and A. Ikeda, Continuum limit
[38] G. J. Martyna, D. J. Tobias, and M. L. Klein, Constant of the vibrational properties of amorphous solids, Proc.
pressure molecular dynamics algorithms, J. Chem. Phys. Natl. Acad. Sci. USA 114, E9767 (2017).
101, 4177 (1994). [59] L. Gartner and E. Lerner, Nonlinear modes disentangle
[39] T. Schneider and E. Stoll, Molecular-dynamics study of glassy and goldstone modes in structural glasses, Scipost
a three-dimensional one-component model for distortive Phys. 1, 016 (2016).
phase transitions, Phys. Rev. B 17, 1302 (1978). [60] L. Gartner and E. Lerner, Nonlinear plastic modes in
[40] J. M. Monti, L. Pastewka, and M. O. Robbins, Green’s disordered solids, Phys. Rev. E 93, 011001 (2016).
function method for dynamic contact calculations, Phys. [61] G. Kapteijns, D. Richard, and E. Lerner, Nonlinear
Rev. E 103, 053305 (2021). quasilocalized excitations in glasses: True representatives
[41] A. J. Ladd, B. Moran, and W. G. Hoover, Lattice ther- of soft spots, Phys. Rev. E 101, 032130 (2020).
mal conductivity: A comparison of molecular dynamics [62] Y. Mishin and J. Hickman, Energy spectrum of a
and anharmonic lattice dynamics, Phys. Rev. B 34, 5058 langevin oscillator, Phys. Rev. E 94, 062151 (2016).
(1986). [63] S. Bickham and J. Feldman, Calculation of vibrational
[42] A. J. McGaughey and M. Kaviany, Quantitative valida- lifetimes in amorphous silicon using molecular dynamics
tion of the boltzmann transport equation phonon ther- simulations, Phys. Rev. B 57, 12234 (1998).
mal conductivity model under the single-mode relaxation [64] N. C. Admal and E. B. Tadmor, A unified interpretation
time approximation, Phys. Rev. B 69, 094303 (2004). of stress in molecular systems, J. Elast. 100, 63 (2010).
[43] E. Lerner, Mechanical properties of simple computer [65] D. C. Wallace, Thermoelasticity of stressed materials and
glasses, J. Non-Cryst. Solids 522, 119570 (2019). comparison of various elastic constants, Phys. Rev. 162,
[44] M. H. Müser, S. V. Sukhomlinov, and L. Pastewka, In- 776 (1967).
teratomic potentials: Achievements and challenges, Adv. [66] W. A. Phillips and A. Anderson, Amorphous solids: low-
Phys. X 8, 2093129 (2023). temperature properties, Vol. 24 (Springer, 1981).
[45] M. H. Müser, S. V. Sukhomlinov, and L. Pastewka, In- [67] N. W. Ashcroft and N. D. Mermin, Solid state physics
teratomic potentials: achievements and challenges, Adv. (Cengage Learning, 2022).
Phys.: X 8, 2093129 (2023). [68] P. W. Anderson, Absence of diffusion in certain random
[46] W. Shinoda, M. Shiga, and M. Mikami, Rapid estimation lattices, Phys. Rev. 109, 1492 (1958).
of elastic constants by molecular dynamics simulation un- [69] D. Richard, C. Rainone, and E. Lerner, Finite-size study
der constant stress, Phys. Rev. B 69, 134103 (2004). of the athermal quasistatic yielding transition in struc-
[47] J. Grießer, L. Frérot, J. A. Oldenstaedt, M. H. Müser, tural glasses, J. Chem. Phys. 155, 056101 (2021).
and L. Pastewka, Analytic elastic coefficients in molec- [70] A. Zaccone and E. Scossa-Romano, Approximate analyt-
ular calculations: Finite strain, nonaffine displacements, ical description of the nonaffine response of amorphous
and many-body interatomic potentials, Phys. Rev. Mat. solids, Phys. Rev. B 83, 184205 (2011).
7, 073603 (2023). [71] J. Ashwin, E. Bouchbinder, and I. Procaccia, Cooling-
[48] A. Tanguy, J. Wittmer, F. Leonforte, and J.-L. Barrat, rate dependence of the shear modulus of amorphous
Continuum limit of amorphous elastic bodies: A finite- solids, Phys. Rev. E 87, 042310 (2013).
size study of low-frequency harmonic vibrations, Phys. [72] J. Turney, E. Landry, A. McGaughey, and C. Amon, Pre-
Rev. B 66, 174205 (2002). dicting phonon properties and thermal conductivity from
[49] B. DiDonna and T. Lubensky, Nonaffine correlations in anharmonic lattice dynamics calculations and molecular
random elastic media, Phys. Rev. E 72, 066619 (2005). dynamics simulations, Phys. Rev. B 79, 064301 (2009).
[50] M. D. Ediger, C. A. Angell, and S. R. Nagel, Supercooled [73] R. Biswas, A. M. Bouchard, W. A. Kamitakahara, G. S.
liquids and glasses, J. Phys. Chem. 100, 13200 (1996). Grest, and C. M. Soukoulis, Vibrational localization in
[51] C. E. Maloney and A. Lemaı̂tre, Amorphous systems in amorphous silicon, Phys. Rev. Lett. 60, 2280 (1988).
athermal, quasistatic shear, Phys. Rev. E 74, 016118 [74] C. Oligschleger and H. R. Schober, Dynamics of Se
(2006). glasses, Physica A 201, 391 (1993).
[52] M. Grmela and H. C. Öttinger, Dynamics and thermo- [75] L. Wang, A. Ninarello, P. Guan, L. Berthier, G. Szamel,
dynamics of complex fluids. I. Development of a general and E. Flenner, Low-frequency vibrational modes of sta-
formalism, Phys. Rev. E 56, 6620 (1997). ble glasses, Nat. Commun. 10, 26 (2019).
[53] H. Mizuno, M. Shimada, and A. Ikeda, Anharmonic [76] P. Debye, Zur theorie der spezifischen wärmen, Annalen
properties of vibrational excitations in amorphous solids, der Physik 344, 789 (1912).
Phys. Rev. Res. 2, 013215 (2020). [77] Y. M. Galperin, V. G. Karpov, and V. I. Kozub, Local-
[54] B. B. Laird and H. Schober, Localized low-frequency vi- ized states in glasses, Adv. Phys. 38, 669 (1989).
brational modes in a simple model glass, Phys. Rev. Lett. [78] U. Buchenau, Y. M. Galperin, V. L. Gurevich, and H. R.
Schober, Anharmonic potentials and vibrational localiza-
14

tion in glasses, Phys. Rev. B 43, 5039 (1991). scale with Python, J. Open Source Softw. 9, 5668 (2024).
[79] M. Baity-Jesi, V. Martı́n-Mayor, G. Parisi, and S. Perez- [83] J. E. Roman, C. Campos, E. Romero, and A. Tomás,
Gaviro, Soft modes, localization, and two-level systems Slepc users manual, D. Sistemes Informatics i Com-
in spin glasses, Phys. Rev. Lett. 115, 267205 (2015). putació, Universitat Politecnica de Valencia, Tech. Rep.
[80] E. Bouchbinder and E. Lerner, Universal disorder- DSIC-II/24/02-Revision 3 (2015).
induced broadening of phonon bands: From disordered [84] V. Hernandez, J. E. Roman, and V. Vidal, Slepc: A scal-
lattices to glasses, New J. Phys. 20, 073022 (2018). able and flexible toolkit for the solution of eigenvalue
[81] A. Moriel, G. Kapteijns, C. Rainone, J. Zylberg, problems, ACM Trans. Math. Software 31, 351 (2005).
E. Lerner, and E. Bouchbinder, Wave attenuation in [85] A. P. Thompson, H. M. Aktulga, R. Berger, D. S. Bolin-
glasses: Rayleigh and generalized-Rayleigh scattering tineanu, W. M. Brown, P. S. Crozier, P. J. in ’t Veld,
scaling, J. Chem. Phys. 151, 104503 (2019). A. Kohlmeyer, S. G. Moore, T. D. Nguyen, R. Shan,
[82] P. Grigorev, L. Frérot, F. Birks, A. Gola, J. Golebiowski, M. J. Stevens, J. Tranchida, C. Trott, and S. J. Plimpton,
J. Grießer, J. L. Hörmann, A. Klemenz, G. Moras, W. G. Lammps - a flexible simulation tool for particle-based
Nöhring, J. A. Oldenstaedt, P. Patel, T. Reichenbach, materials modeling at the atomic, meso, and continuum
L. Shenoy, M. Walter, S. Wengert, J. R. Kermode, and scales, Comput. Phys. Commun. 271, 108171 (2022).
L. Pastewka, matscipy: materials science at the atomic

You might also like