1. Introduction
Viscoelastic liquids and amorphous materials are characterized by long-lasting memory effects often involving a wide spectrum of relaxation times correlating the flow to the prior external forces and strains [
1,
2,
3,
4]. Examples of such materials include complex fluids like viscoelastic polymer melts and solutions, molten metallic alloys, glass-forming (supercooled) liquids and soft-matter systems [
3,
4,
5,
6,
7,
8,
9,
10,
11,
12,
13,
14,
15,
16,
17,
18,
19]. The central physical quantity governing the dynamics of such materials is the mechanical stress [
20]. Stress-correlation functions can provide important information on the rheological properties of amorphous systems, including the most important rheological functions like shear and longitudinal relaxation moduli (and the corresponding dynamical moduli) [
3,
4,
21]. Moreover, glass-forming liquids are known to be highly heterogeneous (near or below the glass transition temperature
) [
22,
23,
24,
25,
26,
27,
28,
29], leading to a significant wave-vector dependence of their shear viscosity and relaxation moduli [
30,
31,
32]. A similar behavior was observed [
33] and predicted [
34] for polymer liquids, and it is expected to be even more important for high-molecular-weight polymers.
Useful relations between the spatio-temporal stress correlation functions and the generalized (length-scale dependent) relaxation moduli (GRMs) have recently been obtained using the Zwanzig–Mori projection operator formalism [
35,
36] and the fluctuation-dissipation theorem (FDT) [
32,
37]. Based on these theoretical relations, it was established that liquids with long terminal relaxation times are characterized by transiently frozen stress fields, which, moreover, exhibit long-range correlations supporting the dynamically heterogeneous nature of glass-forming systems [
32,
35,
36,
37,
38,
39]. These theoretical predictions reinforce the conclusions of the prior extensive and pioneering simulation studies on stress-correlations in supercooled liquids [
40,
41,
42,
43] and also agree with more recent simulation results [
32,
44,
45].
While the recent theoretical studies show that stress correlations and viscoelastic relaxation moduli are intimately related in complex fluids like polymer and supercooled liquids [
32,
35,
36,
37], the origin of these relations (involving spatially resolved relaxation functions) appears to be non-trivial: a number of relations are exact and follow from the fluctuation-dissipation theorem (FDT), while alternative physical arguments are required to derive other relations [
32,
37]. Generalizing our recent results on two-dimensional (2D) systems [
32,
37], we here obtain and discuss the full set of such stress–fluctuation relations valid for arbitrary space dimension.
In the next two sections, we reprise the relevant classical results on the bulk elastic and viscoelastic properties of amorphous systems, presenting the fully tensorial relations between fluctuations of volume-averaged stress and elastic (relaxation) moduli. The bulk equations are then generalized in
Section 4 to deal with wave-vector
)-dependent stress correlations (characterized by the tensorial correlation function
) and spatially resolved relaxation moduli (elasticity tensor
). The methodologically new point here is that we first present a detailed derivation of the general
tensorial equation linking
C- and
E-tensor fields, which then yields three basic relations between the generalized shear, longitudinal and transverse (mixed) relaxation moduli
,
and
) on the one hand, and the invariant correlation functions on the other hand. The recently discovered
M-relation (Equation (
64)) then follows from the general tensorial equation in exactly the same way as other relations (Equations (
62) and (
63)). Different aspects concerning the definition of the
q-dependent elasticity tensor
E are discussed in
Section 5. In particular, it is highlighted there that not all the components of
E can be unambiguously defined for
based on a stress-to-strain response. In
Section 6, we introduce the concept of stress noise
and propose a new definition of all components of the elasticity tensor
in terms of
. It is also demonstrated there that the new definition is consistent with all the known properties of this tensor. On this basis, we derive the full set of exact relations between the correlation and elasticity tensors,
and
, and establish two approximate relations allowing to obtain the full correlation tensor
in terms of only three material functions,
,
and
, also known as viscoelastic memory functions (VMFs). We also discuss how to improve the accuracy of an approximate relation for two-dimensional systems. The theoretical predictions are then compared with simulation results on 2D polydisperse systems of Lennard–Jones (LJ) particles. Such 2D systems have been recently studied experimentally [
46,
47,
48] and have received a lot of attention in simulation studies [
32,
44]. The main results of the paper are summarized in the last
Section 7. In particular, the most important novel results are highlighted in the last point 11 of this section.
2. Classical Elasticity
For tutorial purposes, we start with the linear (affine) elasticity of a macroscopically uniform isotropic solid body. Its deformation is defined by the (coarse-grained) field of displacements
of its material elements (here,
is the initial position of an element). At equilibrium (
), the mean mechanical stress
is isotropic,
, where
is the external pressure [
49]. (Note that in computer simulations the mean stress tensor for a given configuration can be slightly anisotropic. In this case,
should be considered as the mean stress tensor averaged over a sufficiently large ensemble of configurations). Let us consider a weak affine deformation
where
is the strain tensor which does not change in time, and
,
refer to Cartesian components. It leads to a macroscopically homogeneous stress as a linear response,
, where
is the mean (time-averaged) stress in the deformed state. According to Hooke’s law
where
is the tensor of static elastic moduli. (Here and below, the Einstein convention for summation over repeated indices is assumed). For systems (liquids) with no orientational order, the stress tensor is always symmetric, leading to
To exclude a rotation of the system as a whole (which does not cost any energy), we can consider the symmetric part of
,
The symmetric
will be referred to as the classical strain. Then, Equation (
2) transforms into the classical relation
The reason for the equivalence of the two equations, (
2) and (
5), is the minor symmetry of the
E-tensor:
coming from the assumed isotropy of the system demanding that
E must be an isotropic tensor [
44], which, together with the symmetry relation (
3), leads to a well-known equation
with
and
being the Lamé coefficients [
20,
44,
49]. (Note that Equation (
6) also comes directly from Equation (
2) since rotations of the body as a whole (corresponding to an anti-symmetric
) must not lead to any change in the mean (ensemble-averaged) stress. The usefulness (convenience) of Equation (
2) is also clarified in
Section 5.3 in relation to the wave-vector- dependent elasticity).
The free energy increment
associated with a small strain
is
where
V is the system volume and the higher-order terms in
are omitted. (Note that Equation (
8) remains valid also if
is replaced with
). Suppose the elastic body has a free surface, so it can be deformed by a thermal fluctuation. Then, by virtue of the Boltzmann equipartition principle Equation (
8), it leads to the following correlation properties for thermal fluctuations of the classical strain,
:
where
and
are strain components taken at the same time, the brackets
mean the complete ensemble- and time-averaging,
d is the space dimension, and
T is the temperature in energy units.
It is also possible to relate stress fluctuations with the
E-tensor. Here, it is useful to recall that the latter tensor reflects the long-time (static) stress response (cf Equation (
2)); hence, the relevant stress correlation function
must involve a
quasi-static (rather than instantaneous) stress
(note that
). To define it, one has to assume that the strain fluctuations are sufficiently slow (cf refs. [
50,
51]) as compared to the internal structural relaxation of the system (with terminal relaxation time
); the characteristic time of the strain fluctuations,
, must be much longer than
[
50]. Therefore,
(involved in the definition of the
C-tensor) must be considered as the stress component coarse-grained over a time interval
such that
. In this case,
(note that
is a slow strain fluctuation, hence
, being a stress response to
, cf Equation (
5), is a fluctuation as well; that is why we use the notation
instead of
here) so that using Equations (7) and (9), we obtain
Equation (
9) can be obtained in a different way using the classical fluctuation-dissipation theorem (FDT) [
52,
53]. As before, one can define the mean strain,
(cf Equation (
1)) based on the coarse-grained displacement field. The strain
can be considered as a tensorial variable conjugate to the external ‘force’,
, such that the external potential energy (the contribution of the ‘force’ to the free energy of the system) is
Obviously, the ‘force’ tensor
is the external stress applied to the system [
20]. The symmetric part of the deformation,
, induced by a
weak external stress can be obtained by minimization of the total free energy
, where
is defined in Equation (
8). As a result, we obtain a relation between the mean deformation,
, and
(cf Equation (
2)):
According to the classical FDT [
52,
53], the correlation tensor
must be proportional to the susceptibility,
:
Equation (
9) can then be deduced from Equations (7), (12) and (13).
3. Classical Viscoelasticity
Let us turn to viscoelastic systems, including polymer melts and solutions and glass-forming supercooled liquids, but also, in principle, glassy amorphous solids below the glass transition temperature
. Such systems are characterized by time-dependent relaxation moduli, like the shear relaxation modulus
. The goal is to find relations between the stress-correlation functions and the relaxation moduli. Obviously, the argument leading to Equation (
10) is not applicable in this case due to its static nature. By contrast, it is well-known that the classical FDT can be applied to relaxation processes [
52,
53]. There is, however, a fundamental problem with its application to flows of liquids using the deformation tensor as a variable, which is based on the concept of a necessarily continuous (coarse-grained) displacement field
, cf Equation (
1). The point is that during long relaxation times characteristic of most viscoelastic liquids, the initially neighboring particles can go far away from each other (by self-diffusion), which means that
becomes ill-defined (virtually discontinuous). In other words, here, we arrive at a contradiction between continuum-field and corpuscular views on the fluid dynamics [
20]. To avoid such problems, another version of the FDT [
32,
34,
37] should be employed here. It is outlined below. (Note that this approach is applicable more generally also for networks with transient or permanent bonds and solid amorphous systems where
is well defined).
(i) We use a more precise definition of the elastic moduli. The tensor of relaxation moduli
is defined via the stress tensor response,
, to a small but instantaneous deformation of the system,
, at
:
(cf Equation (
2); note that
is the mean, ensemble-averaged, stress increment at time
t). This deformation must be
affine-canonical [
54,
55,
56]. (The transformation is necessarily canonical since we assume that the system dynamics remains Hamiltonian also with the perturbation), which implies changes of both the coordinates
) and the velocities
) of all particles:
(ii) We assume that before the perturbation (at
), the system was at equilibrium, being characterized by an isothermal-isobaric distribution in the phase space:
where
is the probability density, and
stands for the microstate in the phase space (coordinates and velocities of all particles),
,
is the system Hamiltonian,
, its volume, and
, the imposed pressure. (Note the normalization condition:
.) Importantly, we consider a liquid or a
fully equilibrated amorphous system here, so that their shape variations, which are allowed, do not bring about any correction to the Hamiltonian; the external stress corresponds solely to an
isotropic pressure
.
Right after the instantaneous deformation (at
), the distribution changes to
The microscopic definition of the stress tensor reads [
57]
where
is the interaction energy of a pair of interacting particles
,
r is the distance between them,
is the corresponding displacement vector (from
i to
j),
is the mass of particle
i, and
is the
-component of its velocity. Using Equation (
18), we find that the change in
generated by the transformation of Equation (
15) (for a system initially in microstate
) is
where
is the stress fluctuation, and
is the equilibrium ensemble-averaged stress,
. Taking also into account that the transformation conserves the phase space, we obtain
(iii) From the above equations, it immediately follows that
where
is the stress at time
t under the condition that at
the system was in the microstate
, and
is the ensemble-averaged stress increment due to the deformation (recall that at
, the system was assumed to be fully equilibrated; hence,
). Note that after the deformation, at
, the volume and shape of a system are not allowed to vary. Thus, e.g., the volume varies only within the ensemble, but not in time.
The last integral in Equation (
21) is obviously equal to the stress correlation function
where ‘0’ means
. On using the above equation and Equations (
14) and (
21), we find the FDT relation:
It is important to note that, strictly speaking, Equation (
23) is valid for a perfectly equilibrated isothermal-isobaric ensemble (as reflected in its probability distribution in the phase space at any instant before the perturbation) of either liquid or fully equilibrated amorphous solid systems whose
equilibrium shear modulus
is vanishing,
. (Note that
does not exclude that the static modulus
is positive since, in the case of an amorphous glassy system,
corresponds to the long-time glassy plateau of
, which, however, eventually relaxes to 0 at
.)
It is also noteworthy that the condition of perfect equilibration still allows for dynamical fluctuations of energy for each
individual system, including the case of no such fluctuations, ie an energy-conserving and isochoric dynamics for each system (perhaps also involving periodic boundary conditions useful for simulations) [
58,
59]. In the general case, including isotropic liquids in the canonical or microcanonical isochoric ensembles, systems with canonical (Nosé–Hoover) or non-canonical (Gaussian isokinetic) thermostats, and amorphous systems equilibrated in a glassy state (a metabasin), a constant (time-independent) tensor must be added on the rhs of Equation (
23) [
51,
58,
59].
For isotropic systems, the general structure of
is analogous to the static Equation (
7):
where
and
are the generalized Lamé coefficients,
and
, respectively. At long times, in the quasi-static regime,
, where the relaxation moduli change weakly or vanish (note that at
, this regime corresponds to the glassy plateau), Equation (
24) provides the static response in agreement with Equation (
7):
,
.
Note that in the general case, the stress response depends on whether the deformation was isothermic, adiabatic or, else, an imperfect control of temperature is involved (in numerical studies, it may correspond to an isokinetic thermostat, energy-conserving microcanonical simulation, and Nose–Hoover thermostatting, respectively). The thermostatting issues do not affect the shear modulus
, but can be important for
[
32,
60]. Note also that the instantaneous response reflected in the affine moduli,
and
, is always adiabatic (unless a perfect thermostatting of the system is provided) [
58].
The above approach can be used to find an increment of any variable
upon the affine-canonical deformation. The result is
The variable
in the rhs of the above equation can be replaced with
since
by definition (
, see text below Equation (
19)).
4. Space-Resolved Viscoelasticity
In the previous section, we considered classical relaxation moduli related to a response of the volume-averaged stress to a perturbative affine deformation. Let us turn to the space-resolved viscoelasticity providing, in particular, a position-dependent response to a (possibly) localized perturbation. Inspired by the Boltzmann superposition principle [
61], one can regard a weak continuous deformation of the system (e.g., in the course of its slow flow) as a superposition of small strains
, where (cf Equation (
1))
is the strain rate and
is the flow velocity. Thus, generalizing Equation (
14) (considering the stress field
[
20,
62], instead of the volume-averaged stress,
), we can write (based on the superposition principle due to the adopted linear response approximation and recalling the space-time uniformity of the equilibrium macroscopically homogeneous systems we consider)
where
is the ensemble-averaged stress response to the flow (which was absent at
). (A canonical ensemble is assumed by default in this section focused on space-resolved viscoelasticity for an arbitrary but finite wave-vector,
. By contrast to
, which is sensitive to the thermodynamic boundary conditions (fixed volume or fixed pressure), all fluctuations at a finite
are disentangled from fluctuations of global variables like total energy or volume.) Note that the response relation, Equation (
27), is different in nature from Equation (
21) since the condition of no deformation (no flow) at
was assumed in the relevant part of
Section 3. Note that the response relation, Equation (
27), is also different in nature from Equation (
72) since in the latter case (of Equation (
72)), the flow at
is not prescribed in contrast to Equation (
27), where
is considered as a known field.
It is now instructive to provide a microscopic definition of the velocity field:
where
is the mean mass density (
is the total mass of the system and
N is the total number of particles), and
,
are the position and velocity of particle
i. Equation (
27) can be rewritten in terms of Fourier transforms (FT) of the position-dependent functions. For example, the FT of
is the wave-vector
)-dependent tensor of relaxation moduli:
The FT of any other relevant function,
(where
f may stay for a component of
or
tensors) is generically defined as
This way,
and
) have the same physical dimension (note that here and below, we distinguish the original function from its FT by the argument only,
or
). The inverse Fourier transform is given by
where the sum runs over all
-modes defined by the system size. In particular, the strain rate in Equation (
27) is
where (cf Equation (
26))
Thus, Equation (
27) leads to
The kernel
here is the tensor of the generalized (
-dependent) relaxation moduli. Equation (
33) can be considered as a
-dependent generalization of Equation (
2). As to why we use here the
-strain instead of the classical symmetrized
-strain, see
Section 5.3.
Since physical variables like
,
,
are necessarily real, changing
to
always leads to complex conjugation of a
-dependent variable. Hence, for example,
, where star
) means complex conjugate (cf Equation (
30)). Also, obviously, the tensor
must be real and for all isotropic achiral systems, it is an isotropic tensor field [
44], which (being a 4th-rank tensor) must be even in
:
Hence, by virtue of Equation (
29), the same must be true for its Fourier transform:
Thus, the tensor
is real; it is also obviously symmetric with respect to permutation of
and
(cf Equation (
3)).
Note that for
and
corresponding to instantaneous affine deformation
, Equation (
33) becomes identical to Equation (
14). It is also obvious that the velocity
is proportional to the ‘current’ (momentum density)
:
where
The current in the real space is
Our next step will be to accept Equation (
33) (which is equivalent to Equation (
27)) and to try and find a relation of its kernel
E with the stress correlation functions. To this end, consider a system which was at equilibrium at
(with no flow on the average)
but where the flow was generated at
by a perturbative external force field such that the force on particle
i is
where
is a continuous vector field. Such forces
provide a coherent acceleration
of all particles. The external force density, therefore, is
where
is the microscopic mass density.
To simplify the argument, let us consider a very short perturbation,
where
is the coherent velocity increment at position
. Just like the rapid deformation considered in the previous section, this perturbation,
, must lead to a change in the distribution
in the phase space, from the canonical
to
. As before, this transformation conserves the phase-space measure and leads (for a given initial microstate
) to the energy
) increment
Hence, the transformation leads to the following increment of a variable
X at
(cf Equation (
25)):
where
is defined in Equation (
37). At this point, it is convenient to focus on just a single wave-vector
setting
and choosing
. (Note that
here is a constant vector.) Then, Equation (
41) transforms to
while Equation (
42) reduces to
where
is the cross-correlation function of the stress and current, which does not depend on
since time is uniform (and we consider a stationary well-equilibrated system). Note that
due to time reversibility.
The function
is related to the generalized stress-correlation function (cf Equation (
22))
To establish this relation, we employ the fundamental momentum equation [
21,
37]
which reads in Fourier space:
Note that Equation (
49) does not assume an ensemble averaging; it is valid microscopically for each system. On using Equations (
46) and (
48), it leads to
It is now convenient to deal with time-dependent functions,
, in terms of their modified Laplace transform
-transform) [
32,
44]
The transformed Equation (
50) reads
so that Equation (
45) leads to
(Note that Equation (
47) was taken into account here). Setting
in the general Equation, we obtain the response
where
is the current correlation function [
21,
57] whose
s-transform,
, is related to the stress correlation function [
37]:
The above equation can be derived using the momentum equation just like Equation (
52) (with the only difference that
is nonzero since, as follows from Equations (36), (40) and (43),
).
The
s-transform of Equation (
33) reads:
where
since
, as follows from Equations (32) and (35). Using Equations (
53), (
54) and (
56)–(
58) and taking into account that
is an arbitrary vector, we obtain
The above equation provides the general FDT relation between the stress correlation function
C and the tensor
E of the viscoelastic relaxation moduli. It can be simplified using the naturally rotated coordinate frame (NRC) [
37,
44] with axis 1 parallel to wave-vector
:
Equation (
60) can be solved for
with a given
by first setting
and then treating it as a standard matrix equation. Once
is known,
can be obtained directly from Equation (
60). As follows from the
symmetry of the tensor
(which is obviously also applicable to the tensor
[
37]) and their invariance with respect to rotations around the main axis 1, there are only three independent components (involved in Equation (
60)) in each tensor,
and
. For the elasticity tensor, these components are
known as the longitudinal, shear and mixed (transverse) modulus, respectively (cf ref. [
37]). (Another way to identify the three material functions,
L,
G and
M, is expressed by Equation (79) in the next section.) Note that the functions
L,
G and
M are all real and do not depend on the orientation of
(cf Equation (
34) and ref. [
37]). The relevant three general relations derived from Equation (
60) (for
) in the NRC are:
These relations have already been stated in refs. [
32,
34,
37,
59]. The first Equation (
62) is rather well known [
57,
63,
64]. Noteworthily, the above relations are valid both for liquid systems (above the glass transition) and for amorphous solids (vitrified liquids), provided that they are completely equilibrated thermodynamically. Still, they are also valid for metastable glassy systems (trapped in a metabasin), provided that the lifetime of the metastable state is much longer than
[
32].
The three basic relations (
62)–(
64) allow to obtain all the three GRMs,
,
and
, based on the stress-correlation functions. However, the reverse (to obtain all components of
based on the moduli) is, strictly speaking, impossible. Nevertheless, it is still possible to approximately find the undefined components of
for small
q using the three material functions. This is performed in
Section 6.4. Furthermore, the basic Equation (
59) is rederived and generalized in
Section 6 using a different method (the mesoscopic approach involving the concept of stress noise). This way, we both demonstrate the consistency of our approaches and provide a framework for the approximate hydrodynamic theory.
The main results obtained above are discussed in the next section.
7. Summary
1. In the present paper, we established and discussed a number of general relations between the 4th-rank tensor fields of stress correlations,
, cf Equation (
48), and the tensor of generalized (viscoelastic) relaxation moduli,
, cf Equation (
33). The
C-tensor field is generally characterized by, at most, five independent components (invariant correlation functions, see Equation (
A16) in
Appendix B) as long as the minor and major symmetries of
are taken into account (cf Equations (
A17)) [
37]. By contrast, the
E-tensor involves only three material functions, the generalized relaxation moduli (GRMs),
,
and
(cf Equations (
61) and (79)), that can be measured according to their definition via stress response to a weak strain as given in Equation (
33). It is, therefore, not surprising that there exist only three exact relationships (Equations (
62)–(
64)) linking the independent components of the
C-tensor with the three material functions (since obviously five independent correlation functions cannot be expressed using only three material functions). Noteworthily, all the three GRMs can be obtained based on the correlation tensor using Equations (
62)–(
64) (which follow from the exact tensorial relation, Equation (
59)). These three equations are rigorously derived in
Section 4 based on the FDT. Equations (
62) and (
63) have been established before (cf refs. [
34,
37,
57,
59,
63,
64]. The last relation, Equation (
64), was presented in ref. [
37] and verified numerically in ref. [
32]. It is also noteworthy that Equations (
62) and (
63) have been recently derived using the Zwanzig–Mori formalism [
35].
2. In the case of affine deformations, the strain tensor is normally defined as the symmetric part of the tensor
of particle displacement gradients (cf Equations (
1) and (
4)). However, in the more general case of inhomogeneous deformations (which can be considered as a superposition of harmonic waves), the nonsymmetrized definition of strain, Equation (
1), is more appropriate, as argued in
Section 5.3.
3. We considered two definitions of the viscoelastic memory functions (VMFs): in terms of the stress response to a harmonic canonical strain (Equations (
66)–(
68)) and as a response to a coherent external acceleration field (Equations (
73) and (
38)). Importantly, it is demonstrated (see
Section 5.1) that the two definitions lead to exactly the same response functions
). Remarkably, the approach involving the external force, Equation (
38), appears to be more general than imposing a
q-dependent canonical deformation: the latter can be reproduced with a singularity time-dependence of the external field, Equation (
73).
4. It is also remarkable that the stress response to an arbitrary prescribed deformation of an amorphous system can be
completely defined in terms of the reduced elasticity tensor,
, introduced in
Section 5.2 (cf Equation (
76)). All components of this tensor can be obtained based on just three GRMs
VMFs),
, cf Equation (79). The isotropic nature of the system dictates that these material functions are real and do not depend on the orientation of
(cf Equations (
34), (
76) and (79)). Moreover, as we argue in
Section 6.4, these functions must be generally continuous and, moreover, analytical functions of
. At
, the elastic response is local. It is also likely that the same is virtually true at
(cf
Section 6.4), so that, for example,
at
(with relative error
, where
b is the particle interaction range). Importantly, at low
q (
), the three GRMs are related for any time
t (cf Equation (
121) and refs. [
32,
37]).
5. As mentioned above, in this study, we consider the elasticity tensor in terms of the stress response to a prescribed small strain or to an external force perturbation (in the latter case, the force generally depends on the particle position). Noteworthily, considering another type of perturbation by changing the system Hamiltonian from
to
with
, involving a prescribed weak ‘deformation’ function
, does not make much sense: On the one hand, it allows to employ the classical FDT [
52], but on the other hand, it is unclear how the prescribed
can be possibly linked with the physical strain in the system given that the introduction of
changes the classical relations between particle velocities and momenta leading to an anisotropic and position-dependent particle mass.
6. To uncover new relationships between the stress correlations and the elasticity tensor (cf
Section 6), we employ the concept of the stress noise,
, proposed in our previous paper [
37]. It is defined as
, where
is the deterministic stress due to the flow history in the system (cf Equation (
87)). The stress noise
can, thus, be considered as a genuine stress fluctuation unrelated to deformation and flow. This concept opened up the possibility to define
all components of the generalized
-dependent) elasticity tensor,
, in terms of the stress noise correlation function (cf Equations (
98) and (
105)). It is important that the new definition is totally consistent with the classical linear response way to introduce the elasticity tensor, Equation (
33), and, therefore, leads to exactly the same GRMs,
. The latter statement is valid since Equations (
59) and (
62)–(
64) trivially follow from Equation (
115). On the other hand, the new definition, Equation (
105), implies both minor and major symmetries of
, which are inherent in the classical
bulk elasticity tensor. Moreover, the bulk tensor coincides with
at
since the latter tensor field is continuous and analytical as a function of
(see end of
Section 6.2 and
Section 6.4). The definition of the generalized elasticity tensor, Equation (
105), therefore, combines the best of both worlds (of affine strains,
, and harmonic deformations,
).
7. One may wonder how to obtain the correlation function of stress noise,
. The answer is given in
Section 6.2: it can be done using simulations with arrested flow at wave-vector
implying the condition, Equation (
88). This condition can be imposed using an external force field (cf Equation (
89)) leading to an appropriate coherent harmonic acceleration of particles. With the constrained dynamics, the deterministic stress is always constant (time-independent); it is defined by the ‘quenched’ concentration fluctuation at
. Then,
where
is the total stress correlation function with restricted dynamics (cf Equation (
99)) and
is a time-independent tensor, which, however, generally depends on
(cf Equations (
96) and (
106)). This tensor
) simply equals to
; it is related to the equilibrium elastic moduli (at
), cf Equations (
92) and (
95).
8. In
Section 6.2, we introduced the equilibrium elasticity tensor
defined in Equation (
106). In the liquid regime,
coincides with the static elasticity tensor,
, so that
can be considered as a generalization of the classical static elasticity tensor (cf
Section 2) for nonzero
. However, in the glassy (amorphous solid) state, the two tensors, equilibrium and static, are different since even a very long waiting time,
, may not ensure a complete equilibration of a vitrified liquid (amorphous solid). In particular, the stress noise may include a virtually frozen component leading to an incomplete relaxation. Therefore, the static shear modulus,
), remains finite in this case, while the analogous
equilibrium shear modulus must vanish since a
complete equilibration after a small shear deformation of a glassy system must relax the shear stress due to the amorphous structure of the system [
37]. (Note that we do not consider here a
permanently crosslinked network whose equilibrium shear modulus is, of course, finite.) As a result, the equilibrium elasticity tensor can be expressed in terms of just two material functions: the equilibrium longitudinal,
, and transverse,
, elastic moduli (cf Equation (
106)). These moduli, by their definition, provide a linear stress response (after a complete relaxation of the system) to a weak imposed longitudinal strain.
9. The most general relation between the stress-correlation
) and elasticity
) tensors is given in Equation (
115). It is noteworthy that this equation was derived and is valid at
. It cannot be generally applied for
since the stress-correlation function
C is ensemble-dependent in this case [
32,
37]. It is also remarkable that, based on Equation (
115), we not only arrive at Equations (
62)–(
64) linking the shear, longitudinal and transverse components of
C- and
E-tensors, but also obtain two additional exact relations (
116) involving other components of these tensors. The whole set of these relations then allows to obtain all components of the correlation tensor in terms of the elasticity tensor and vice versa. Strictly speaking, all the relations, Equations (
62)–(
64) and (
116), are valid both for liquid systems (above the glass transition) and for amorphous solids (vitrified liquids), provided that they are completely equilibrated thermodynamically (this condition refers to the fact that the derivation of these relations assumed an equilibrium ensemble). Nevertheless, these relations are also valid for metastable glassy systems (trapped in a metabasin), provided that the lifetime of the metastable state is much longer than
[
32] and with the reservation that some
-dependent constants may have to be added in the rhs of Equations (
116), cf ref. [
32]. These constants are due to the presence of frozen stresses in glassy systems, reflecting their metastable nature (within a given metabasin); they must disappear upon averaging over the full equilibrium ensemble of metastable states.
10. There is a subtle problem associated with the new Equations (
116): they involve two new memory functions,
and
, which cannot be obtained based on the stress response to a deformation, and, therefore, apparently cannot be measured experimentally. One may wonder if these functions can be obtained based on the ‘classical’ relaxation moduli,
. Our view is that while their exact prediction is generally impossible, the new functions
N and
P can be still predicted approximately at low
q. As argued in
Section 6 (see end of
Section 6.2 and the beginning of
Section 6.4), the elasticity tensor becomes nearly isotropic at low
q, so that
and
at
(cf Equations (
122)). Replacing
N with
L in the first Equation (
116) leads to an approximate equation, which was derived and rather thoroughly tested in refs. [
32,
37] using simulation data for a 2D system of polydisperse LJ particles. A very good agreement (with an accuracy of 1–2%) between
and its approximate prediction was observed at
(where
b is the interaction range) [
32]. Here, we devised two more precise approximations for
valid for 2D systems (see Equations (
127) and (
130)). All the approximations have been tested at
for a wide range of
q for the same system. The comparison (between the exact and approximate
) is shown in
Figure 1. It demonstrates that the basic (0th) approximations still work for
, while the new approximations are accurate in much wider
q-regions: the first one is valid at
, the second at
.
11. To summarize, let us highlight the main new results presented in the paper:
(i) We provide a rigorous derivation of Equations (
63) and (
64) using FDT-based arguments (cf
Section 4). These equations have been previously stated in refs. [
32,
37], but their detailed derivation was not worked out (note that Equation (
63) was also stated in ref. [
59]). Importantly, in
Section 4, we provide a derivation of the unique fully tensorial equation (Equation (
59)) from which the general FDT relations, Equation (
62) (which is well-known [
57,
63,
64]) and Equations (
63) and (
64), simply follow in a trivial way.
(ii) We derived approximate Equations (
123) and (
124) (valid for
) using FDT and the concept of stress noise (cf
Section 6.3). Note that the derivation of Equation (
123) was only hinted at previously (in ref. [
37]), while the same equation was simply claimed in ref. [
32].
(iii) Building upon the concept of stress noise, a key result of our work is Equation (
115), relating the tensor of stress correlations
with the tensor of elastic moduli
. Noteworthily, the form of Equation (
115) agrees with Equation (
45) of ref. [
49], which establishes a relation between the memory kernel
from the Zwanzig–Mori projection operator formalism and
for monodisperse Brownian particles. This suggests the intriguing possibility of a deeper connection between
and
, which is an interesting topic for future studies.
(iv) For 2D systems, we, for the first time, derived more precise equations (as compared to Equation (
123)) defining the stress correlation function
in terms of the generalized relaxation moduli (cf the first Equation (
116) and Equations (
127) and (
130)).