Long ranged stress correlations in the hard sphere liquid

Niklas Grimm Fachbereich Physik, Universität Konstanz, 78457 Konstanz, Germany    Martin von Bischopinck Fachbereich Chemie, Universität Konstanz, 78457 Konstanz, Germany    Andreas Zumbusch Fachbereich Chemie, Universität Konstanz, 78457 Konstanz, Germany    Matthias Fuchs Fachbereich Physik, Universität Konstanz, 78457 Konstanz, Germany
(May 15, 2024)
Abstract

The smooth emergence of shear elasticity is an hallmark of the liquid to glass transition. In a liquid, viscous stresses arise from local structural rearrangements. In the solid, Eshelby has shown that stresses around an inclusion decay as a power law rDsuperscript𝑟𝐷r^{-D}italic_r start_POSTSUPERSCRIPT - italic_D end_POSTSUPERSCRIPT, where D𝐷Ditalic_D is the dimension of the system. We study glass-forming hard sphere fluids by simulation and observe the emergence of the unscreened power-law Eshelby pattern in the stress correlations of the isotropic liquid state. By a detailed tensorial analysis, we show that the fluctuating force field, viz. the divergence of the stress field, relaxes to zero with time in all states, while the shear stress correlations develop spatial power-law structures inside regions that grow with longitudinal and transverse sound speeds; we observe the predicted exponents rDsuperscript𝑟𝐷r^{-D}italic_r start_POSTSUPERSCRIPT - italic_D end_POSTSUPERSCRIPT and rD2superscript𝑟𝐷2r^{-D-2}italic_r start_POSTSUPERSCRIPT - italic_D - 2 end_POSTSUPERSCRIPT. In Brownian systems, shear stresses relax diffusively within these regions, with the diffusion coefficient determined by the shear modulus and the friction coefficient.

preprint: APS/123-QED

I Introduction

During the process of cooling, liquids start to show first traces of elasticity over increasing intermediate time windows. The phenomenon of viscoelasticity becomes especially strong in supercooled liquids when freezing can be prevented and the fluid approaches the transition to a glass [1]. An observable reflecting this behavior is the time dependent shear modulus G(t)𝐺𝑡G(t)italic_G ( italic_t ) [2, 3]. It connects the stress response of a system to the rate of shear strain applied on the bulk (wave vector q=0𝑞0q=0italic_q = 0) in the linear regime, and is a central quantity studied in rheology. Two extreme cases of the stress response are, on the one hand, G(t)=ηδ(t)𝐺𝑡𝜂𝛿𝑡G(t)=\eta\delta(t)italic_G ( italic_t ) = italic_η italic_δ ( italic_t ) (with the Dirac delta δ(t)𝛿𝑡\delta(t)italic_δ ( italic_t )) giving an ideal Newtonian fluid, and on the other hand G(t)=Gpl=const.𝐺𝑡subscript𝐺𝑝𝑙𝑐𝑜𝑛𝑠𝑡G(t)=G_{pl}=const.italic_G ( italic_t ) = italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT = italic_c italic_o italic_n italic_s italic_t . giving an elastic Hookean solid. In the former case, the shear stress σ𝜎\sigmaitalic_σ increases linearly with the shear rate γ˙˙𝛾\dot{\gamma}over˙ start_ARG italic_γ end_ARG, viz. σ=ηγ˙𝜎𝜂˙𝛾\sigma=\eta\dot{\gamma}italic_σ = italic_η over˙ start_ARG italic_γ end_ARG, and η𝜂\etaitalic_η is identified as shear viscosity. In the latter case, the stress equals the constant shear modulus times the applied strain γ𝛾\gammaitalic_γ, σ=Gplγ𝜎subscript𝐺𝑝𝑙𝛾\sigma=G_{pl}\gammaitalic_σ = italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT italic_γ. According to Maxwell, viscosity and shear modulus are connected via the structural relaxation time τ𝜏\tauitalic_τ, viz. η=Gplτ𝜂subscript𝐺𝑝𝑙𝜏\eta=G_{pl}\tauitalic_η = italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT italic_τ. Viscoelastic systems are in between those ideal cases and the time-dependence of the shear modulus G(t)𝐺𝑡G(t)italic_G ( italic_t ) is important [4]. In the process of cooling a supercooled liquid, the G(t)𝐺𝑡G(t)italic_G ( italic_t ) develops a shoulder over increasing time windows until ultimately G(t)𝐺𝑡G(t)italic_G ( italic_t ) is non-zero at laboratory time-scales and the liquid freezes and becomes an amorphous solid.

Via the linear response theory close to equilibrium, G(t)𝐺𝑡G(t)italic_G ( italic_t ) is connected to the decay of thermal fluctuations of the (macroscopic) shear stress. This gives G(t)𝐺𝑡G(t)italic_G ( italic_t ) a vivid interpretation. On the other hand, the local fluctuations of all elements of the stress tensor are the central object of this study. In linear elasticity theory, describing the ideal Hookean solid as mentioned above, the space resolved stress correlations were calculated by Eshelby in order to determine the stress and strain fields around inclusions [5]. The well known result is given by a characteristic octupolar angular dependence and an rDsuperscript𝑟𝐷r^{-D}italic_r start_POSTSUPERSCRIPT - italic_D end_POSTSUPERSCRIPT distance dependence of the shear stress field induced by a force dipole [6].

This leads to the question, if this stress pattern characteristic for elastic behavior can already be found in the shear stress correlation of an viscoelastic supercooled liquid. This asks if there are any correlation lengths in supercooled liquids beyond the trivial mean particle separation [7, 8, 9, 10, 11, 12, 13]. By theory, this was addressed using a generalized hydrodynamic approach, where the transition from local viscous to long-range elastic behavior was connected to momentum transport [14, 15]. A generalized Maxwell model has been derived, and already tested in simulations [16], finding that the Eshelby-pattern of elastic materials emerges in reciprocal space already in the supercooled liquid. Our contribution is concerned with the formation of the far field of the stress correlations in real space. We also address the connection between real and reciprocal space, and test the generalized hydrodynamic approach without Maxwell’s approximations. The emergence of elastic stress correlations in fluids originates from the coupling of stress fluctuations to momentum fluctuations. The development of a long range pattern can be argued as follows: A shear stress fluctuation triggers a transverse momentum fluctuation, a shear wave. Due to the onset of elasticity in the supercooled liquid, this momentum fluctuation propagates and vice versa induces stress fluctuations on its trajectory. They are long lived as can be seen by the increasingly large relaxation times of the shear modulus. Since the momentum current is distributed over surfaces increasing with rD1superscript𝑟𝐷1r^{D-1}italic_r start_POSTSUPERSCRIPT italic_D - 1 end_POSTSUPERSCRIPT the displacements scale analogously. The stress is connected to the strain, i.e. the derivative of these displacements, it therefore scales as rDsuperscript𝑟𝐷r^{-D}italic_r start_POSTSUPERSCRIPT - italic_D end_POSTSUPERSCRIPT [11].

In this study we perform computer-simulations of hard spheres in three dimension (3D) and hard discs in two dimension (2D). The discussion in 2D is included because of the allowance for larger systems and a therefore more convincing measurement of the long ranged correlations. Furthermore the mathematical expressions are more handy, making technical aspects more transparent.
Since the stress correlation tensor is of fourth order it is non trivial to determine all its components in simulations. The computational and mathematical effort becomes manageable after a decomposition into tensseral harmonic tensors (real valued spherical harmonic tensors) as established for studies of the stress correlation tensor in zero temperature glasses [17].
The shown simulations are space and time resolved, which gives the opportunity to directly observe the build-up of the various spatial patterns following the passing of longitudinal and transverse sound waves in Newtonian dynamics [18]. We also perform Brownian dynamics simulations to study the diffusive shear stress transport in colloidal glasses [19]. Hard sphere fluids are an established model for colloidal dispersions, where confocal microscopy can track particles and determine stresses from close contacts [20]. The stress far fields have not been recorded yet in colloidal fluids, which we aim to overcome by develo** fluorescently marked soft droplets as stress senors.

The paper is organized as follows. Section II describes the preparation and the execution of the molecular dynamics (MD) simulations. How the stress fields are measured in these MD simulations is described in Sec. III, and Sec. IV recalls the symmetry based decomposition of the fourth rank tensor of stress correlation functions. There it will become clear, that in two dimension fewer independent functions arise than in three, which causes us to first present stress correlations in two dimensions, Sec. V, before addressing three dimensions in Sec. VI. A discussion section including the presentation of stress sensor droplets ends the manuscript.

II Simulation Preparation

We perform event driven simulations of hard particles interacting only when in contact via an elastic collision rule [21]. To prevent crystallization we use a binary mixture of equal particle numbers Ns/Nl=1subscript𝑁𝑠subscript𝑁𝑙1N_{s}/N_{l}=1italic_N start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_N start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = 1 and a diameter ratio of dl/ds=1.4subscript𝑑𝑙subscript𝑑𝑠1.4d_{l}/d_{s}=1.4italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 1.4, where l𝑙litalic_l and s𝑠sitalic_s denote large and small spheres, respectively. In two dimensions, we work at a density of ϕ=0.79italic-ϕ0.79\phi=0.79italic_ϕ = 0.79, slightly below the estimated glass transition of mode coupling theory (MCT) [22] at ϕc=0.7948superscriptitalic-ϕ𝑐0.7948\phi^{c}=0.7948italic_ϕ start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT = 0.7948 [23, 24]. In three dimension the density is ϕ=0.59italic-ϕ0.59\phi=0.59italic_ϕ = 0.59 at the MCT glass transition [25, 26, 27]. As seen in the various shear moduli below (Figs. 2, 7 12, 16) the macroscopic stresses relax within the observation time. Units are set by the diameter dlsubscript𝑑𝑙d_{l}italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, the thermal velocity v0subscript𝑣0v_{0}italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and the thermal energy density kBTnsubscript𝑘𝐵𝑇𝑛k_{B}Tnitalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T italic_n.

We also perform Brownian dynamics simulations based on the event driven algorithm. The dam** is achieved by a Brownian thermostat re-drawing particle velocities from a Maxwell-Boltzmann distribution at fixed time-steps τbsubscript𝜏𝑏\tau_{b}italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. The Brownian thermostat connects the friction to the Brownian time-step via ζ=2/τb𝜁2subscript𝜏𝑏\zeta=2/\tau_{b}italic_ζ = 2 / italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT [21]. We use τb=5×105dl2/D0subscript𝜏𝑏5superscript105superscriptsubscript𝑑𝑙2subscript𝐷0\tau_{b}=5\times 10^{-5}\,d_{l}^{2}/D_{0}italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 5 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_D start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. With these technique, time-dependent shear moduli were already obtained in 2D [24] and 3D [28], but not related to the other stress correlation functions.

III Stress in hard sphere simulations

The stress σ¯¯(𝒓,t)¯¯𝜎𝒓𝑡\underline{\underline{\sigma}}(\bm{r},t)under¯ start_ARG under¯ start_ARG italic_σ end_ARG end_ARG ( bold_italic_r , italic_t ) describes a momentum current at time t𝑡titalic_t at location 𝒓𝒓\bm{r}bold_italic_r and is therefore a second rank tensor [29]. To extract the stress field in real space, the box of volume V𝑉Vitalic_V is subdivided into boxes of volume VBsubscript𝑉𝐵V_{B}italic_V start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT, which have their center at 𝒓Bsubscript𝒓𝐵\bm{r}_{B}bold_italic_r start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT. The average over collisions happening in a time window Δt=[tδt2,t+δt2]Δ𝑡𝑡𝛿𝑡2𝑡𝛿𝑡2\Delta t=[t-\frac{\delta t}{2},t+\frac{\delta t}{2}]roman_Δ italic_t = [ italic_t - divide start_ARG italic_δ italic_t end_ARG start_ARG 2 end_ARG , italic_t + divide start_ARG italic_δ italic_t end_ARG start_ARG 2 end_ARG ] around t𝑡titalic_t of length δt𝛿𝑡\delta titalic_δ italic_t at points in the box VBsubscript𝑉𝐵V_{B}italic_V start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT around 𝒓Bsubscript𝒓𝐵\bm{r}_{B}bold_italic_r start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is taken to determine the stress tensor at time t𝑡titalic_t and position 𝒓Bsubscript𝒓𝐵\bm{r}_{B}bold_italic_r start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT

σ¯¯(𝒓B,t)=1VBδtτcΔt;𝒓cVBΔ𝒓cΔ𝒑c.¯¯𝜎subscript𝒓𝐵𝑡1subscript𝑉𝐵𝛿𝑡subscriptformulae-sequencesubscript𝜏𝑐Δ𝑡subscript𝒓𝑐subscript𝑉𝐵Δsuperscript𝒓𝑐Δsuperscript𝒑𝑐\displaystyle\underline{\underline{\sigma}}(\bm{r}_{B},t)=\frac{1}{V_{B}\delta t% }\sum_{\begin{subarray}{c}\tau_{c}\in\Delta t;\,\bm{r}_{c}\in V_{B}\end{% subarray}}\Delta\bm{r}^{c}\Delta\bm{p}^{c}.under¯ start_ARG under¯ start_ARG italic_σ end_ARG end_ARG ( bold_italic_r start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT , italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_δ italic_t end_ARG ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_τ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∈ roman_Δ italic_t ; bold_italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∈ italic_V start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_CELL end_ROW end_ARG end_POSTSUBSCRIPT roman_Δ bold_italic_r start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT roman_Δ bold_italic_p start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT . (1)

Here Δ𝒓cΔsuperscript𝒓𝑐\Delta\bm{r}^{c}roman_Δ bold_italic_r start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT denotes the vector connecting the centers of the colliding spheres, Δ𝒑cΔsuperscript𝒑𝑐\Delta\bm{p}^{c}roman_Δ bold_italic_p start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT denotes the transferred momentum, τcsubscript𝜏𝑐\tau_{c}italic_τ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT the time of the collision and 𝒓csubscript𝒓𝑐\bm{r}_{c}bold_italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT its position. Note that the finite duration of our time-window, where many collisions are averaged, smoothens the short-time stress divergences expected for hard sphere interactions [30]. We subdivide the simulation box in a grid width an edge length of 41 equally sized boxes in two and three dimensions, which determines the volume VBsubscript𝑉𝐵V_{B}italic_V start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT. This leads in 2D to LB6.3dlsubscript𝐿𝐵6.3subscript𝑑𝑙L_{B}\approx 6.3\,d_{l}italic_L start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ≈ 6.3 italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and in 3D to Lb1.65dlsubscript𝐿𝑏1.65subscript𝑑𝑙L_{b}\approx 1.65\,d_{l}italic_L start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ≈ 1.65 italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT.

We define the stress field in Fourier space as

σ¯~¯(𝒒,t)=1δtτcΔtΔ𝒓cΔ𝒑cei𝒒𝒓c,¯¯~𝜎𝒒𝑡1𝛿𝑡subscriptsubscript𝜏𝑐Δ𝑡Δsuperscript𝒓𝑐Δsuperscript𝒑𝑐superscript𝑒𝑖𝒒subscript𝒓𝑐\displaystyle\underline{\underline{\tilde{\sigma}}}(\bm{q},t)=\frac{1}{\delta t% }\sum_{\tau_{c}\in\Delta t}\,\Delta\bm{r}^{c}\,\Delta\bm{p}^{c}\,e^{i\,\bm{q}% \cdot\bm{r}_{c}},under¯ start_ARG under¯ start_ARG over~ start_ARG italic_σ end_ARG end_ARG end_ARG ( bold_italic_q , italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_δ italic_t end_ARG ∑ start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∈ roman_Δ italic_t end_POSTSUBSCRIPT roman_Δ bold_italic_r start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT roman_Δ bold_italic_p start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i bold_italic_q ⋅ bold_italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , (2)

where the sum runs over all collisions in the simulation box happening in the respective time window.

The correlation functions in real and reciprocal space are then defined as

𝒞¯¯(𝒓B𝒓B,tt)=σ¯¯(𝒓B,t)σ¯¯(𝒓B,t),¯¯𝒞subscript𝒓𝐵superscriptsubscript𝒓𝐵𝑡superscript𝑡delimited-⟨⟩¯¯𝜎subscript𝒓𝐵𝑡¯¯𝜎superscriptsubscript𝒓𝐵superscript𝑡\displaystyle\underline{\underline{\mathcal{C}}}(\bm{r}_{B}-\bm{r}_{B}^{\prime% },t-t^{\prime})=\left\langle\underline{\underline{\sigma}}(\bm{r}_{B},t)\;% \underline{\underline{\sigma}}(\bm{r}_{B}^{\prime},t^{\prime})\right\rangle,under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG ( bold_italic_r start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT - bold_italic_r start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = ⟨ under¯ start_ARG under¯ start_ARG italic_σ end_ARG end_ARG ( bold_italic_r start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT , italic_t ) under¯ start_ARG under¯ start_ARG italic_σ end_ARG end_ARG ( bold_italic_r start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ⟩ , (3)
𝒞¯~¯(𝒒,tt)=1Vσ¯~¯(𝒒,t)σ¯~¯(𝒒,t).¯¯~𝒞𝒒𝑡superscript𝑡1𝑉delimited-⟨⟩superscript¯¯~𝜎𝒒𝑡¯¯~𝜎𝒒superscript𝑡\displaystyle\underline{\underline{\tilde{\mathcal{C}}}}(\bm{q},t-t^{\prime})=% \frac{1}{V}\left\langle\underline{\underline{\tilde{\sigma}}}^{*}(\bm{q},t)\;% \underline{\underline{\tilde{\sigma}}}(\bm{q},t^{\prime})\right\rangle.under¯ start_ARG under¯ start_ARG over~ start_ARG caligraphic_C end_ARG end_ARG end_ARG ( bold_italic_q , italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = divide start_ARG 1 end_ARG start_ARG italic_V end_ARG ⟨ under¯ start_ARG under¯ start_ARG over~ start_ARG italic_σ end_ARG end_ARG end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_q , italic_t ) under¯ start_ARG under¯ start_ARG over~ start_ARG italic_σ end_ARG end_ARG end_ARG ( bold_italic_q , italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ⟩ . (4)

Note that in real space (Eq. 1), the collisions are first coarse grained over boxes, which then are correlated. In contrast in Fourier space (Eq. 2), every collision individually contributes to the stress field, that is Fourier transformed.

The shear modulus G(t)𝐺𝑡G(t)italic_G ( italic_t ) follows then from the fluctuation dissipation theorem as

G(t)=1kBT𝒞~xyxy(𝒒=0,t),𝐺𝑡1subscript𝑘𝐵𝑇subscript~𝒞𝑥𝑦𝑥𝑦𝒒0𝑡\displaystyle G(t)=\frac{1}{k_{B}T}\,\tilde{\mathcal{C}}_{xyxy}(\bm{q}=0,t),italic_G ( italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG over~ start_ARG caligraphic_C end_ARG start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT ( bold_italic_q = 0 , italic_t ) , (5)

i.e. the off-diagonal stress contributions of all collisions happening in the simulation box are summed over and auto-correlated in time. Throughout this work we assume stationary in time and are therefore free to choose t=0superscript𝑡0t^{\prime}=0italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 0.

An important aspect is collecting a large enough statistical data base since we aim for the, in general, small signals of the power law stress correlations at large distances. Thereto, it is important to use the isotropic symmetry of the viscoelastic fluid in order to perform angular averaging. The formalism that enables us to do so is described in the next section.

IV Tensorial aspects of the stress-AC

The correlation function of the stress tensors is a tensor of fourth order with D4superscript𝐷4D^{4}italic_D start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT components in D𝐷Ditalic_D dimensions. This number can be reduced drastically by exploiting the symmetry of the liquid state and the properties of the stress tensor itself [17, 31]. In these seminal papers, the fourth order tensors of the stress correlations were expressed in spherical coordinates and linearly combined to give real valued spherical harmonic tensors (tesseral decomposition). This has three major advantages: i) it simplifies the identification of the non vanishing tensorial contributions, ii) the basis is orthogonal simplifying the algebra and iii) the individual tensor components are isotropic functions allowing angular averaging. These ideas have as well been employed in Ref. [32], where spatially resolved stress correlations in glasses were investigated by simulations.

Next, the notation will be introduced that will be used thought the work. We assume stationarity in time, and homogeneity and isotropy in space, so without loss of generality we can re-write the stress-autocorrelation tensor (SACT) 𝒞¯¯¯¯𝒞\underline{\underline{\mathcal{C}}}under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG with one time and one space argument

𝒞¯¯(𝒓,t)=σ¯¯(𝒓𝒓,tt)σ¯¯(𝒓,t).¯¯𝒞𝒓𝑡delimited-⟨⟩¯¯𝜎𝒓superscript𝒓𝑡superscript𝑡¯¯𝜎superscript𝒓superscript𝑡\displaystyle\underline{\underline{\mathcal{C}}}(\bm{r},t)=\left\langle% \underline{\underline{\sigma}}(\bm{r}-\bm{r}^{\prime},t-t^{\prime})\;% \underline{\underline{\sigma}}(\bm{r}^{\prime},t^{\prime})\right\rangle.under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG ( bold_italic_r , italic_t ) = ⟨ under¯ start_ARG under¯ start_ARG italic_σ end_ARG end_ARG ( bold_italic_r - bold_italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) under¯ start_ARG under¯ start_ARG italic_σ end_ARG end_ARG ( bold_italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ⟩ . (6)

The system conserves angular momentum in every particle collision, so the σ¯¯(𝒓,t)¯¯𝜎𝒓𝑡\underline{\underline{\sigma}}(\bm{r},t)under¯ start_ARG under¯ start_ARG italic_σ end_ARG end_ARG ( bold_italic_r , italic_t ) is symmetric [33].
This implies the minor symmetries

𝒞αβγδ=𝒞αβδγ=𝒞βαγδ.subscript𝒞𝛼𝛽𝛾𝛿subscript𝒞𝛼𝛽𝛿𝛾subscript𝒞𝛽𝛼𝛾𝛿\displaystyle\mathcal{C}_{\alpha\beta\gamma\delta}=\mathcal{C}_{\alpha\beta% \delta\gamma}=\mathcal{C}_{\beta\alpha\gamma\delta}.caligraphic_C start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT = caligraphic_C start_POSTSUBSCRIPT italic_α italic_β italic_δ italic_γ end_POSTSUBSCRIPT = caligraphic_C start_POSTSUBSCRIPT italic_β italic_α italic_γ italic_δ end_POSTSUBSCRIPT . (7)

Here and throughout, Greek indices label the spatial directions. Symmetry under time-inversion holds, which implies the major symmetry

𝒞αβγδ=𝒞γδαβ.subscript𝒞𝛼𝛽𝛾𝛿subscript𝒞𝛾𝛿𝛼𝛽\displaystyle\mathcal{C}_{\alpha\beta\gamma\delta}=\mathcal{C}_{\gamma\delta% \alpha\beta}.caligraphic_C start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT = caligraphic_C start_POSTSUBSCRIPT italic_γ italic_δ italic_α italic_β end_POSTSUBSCRIPT . (8)

Invariance under inversion and rotation means, that for every orthogonal matrix R¯¯¯¯R\underline{\underline{\mathrm{R}}}under¯ start_ARG under¯ start_ARG roman_R end_ARG end_ARG it holds that

𝒞¯¯(R¯¯𝒓,t)=R¯¯R¯¯:𝒞¯¯(𝒓,t):R¯¯TR¯¯T:¯¯𝒞¯¯R𝒓𝑡¯¯R¯¯R¯¯𝒞𝒓𝑡:superscript¯¯R𝑇superscript¯¯R𝑇\displaystyle\underline{\underline{\mathcal{C}}}(\underline{\underline{\mathrm% {R}}}\cdot\bm{r},t)=\underline{\underline{\mathrm{R}}}\,\underline{\underline{% \mathrm{R}}}:\underline{\underline{\mathcal{C}}}(\bm{r},t):\underline{% \underline{\mathrm{R}}}^{T}\underline{\underline{\mathrm{R}}}^{T}under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG ( under¯ start_ARG under¯ start_ARG roman_R end_ARG end_ARG ⋅ bold_italic_r , italic_t ) = under¯ start_ARG under¯ start_ARG roman_R end_ARG end_ARG under¯ start_ARG under¯ start_ARG roman_R end_ARG end_ARG : under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG ( bold_italic_r , italic_t ) : under¯ start_ARG under¯ start_ARG roman_R end_ARG end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT under¯ start_ARG under¯ start_ARG roman_R end_ARG end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT (9)

The requirements on the tensor field given by Eqs. (7), (8), and (9) constrain the SACT to an extent, where it can be written in terms of only 5 tensorial contributions Q¯¯(i)(𝒓^)superscript¯¯Q𝑖bold-^𝒓\underline{\underline{\mathrm{Q}}}^{(i)}(\bm{\hat{r}})under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( overbold_^ start_ARG bold_italic_r end_ARG ) with scalar valued pre-factors q(i)(r,t)superscript𝑞𝑖𝑟𝑡q^{(i)}(r,t)italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) that are isotropic functions of the scalar distance r=|𝒓|𝑟𝒓r=|\bm{r}|italic_r = | bold_italic_r | only [17]; here 𝒓^=𝒓/rbold-^𝒓𝒓𝑟\bm{\hat{r}}=\bm{r}/roverbold_^ start_ARG bold_italic_r end_ARG = bold_italic_r / italic_r is a direction vector. This decomposition reads

𝒞¯¯(𝒓,t)=i=15q(i)(r,t)Q¯¯(i)(𝒓^).¯¯𝒞𝒓𝑡superscriptsubscript𝑖15superscript𝑞𝑖𝑟𝑡superscript¯¯Q𝑖bold-^𝒓\displaystyle\underline{\underline{\mathcal{C}}}(\bm{r},t)=\sum_{i=1}^{5}q^{(i% )}(r,t)~{}\underline{\underline{\mathrm{Q}}}^{(i)}(\bm{\hat{r}}).under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG ( bold_italic_r , italic_t ) = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( overbold_^ start_ARG bold_italic_r end_ARG ) . (10)

The explicit expressions of the tensors Q¯¯(i)(𝒓^)superscript¯¯Q𝑖bold-^𝒓\underline{\underline{\mathrm{Q}}}^{(i)}(\bm{\hat{r}})under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( overbold_^ start_ARG bold_italic_r end_ARG ) are given in three dimensions by Eqs. (43) in App. A.1.

The same decomposition is valid in Fourier space, where the basis tensors only change their dependencies from unit position vectors 𝒓^bold-^𝒓\bm{\hat{r}}overbold_^ start_ARG bold_italic_r end_ARG to unit wave vectors 𝒒^bold-^𝒒\bm{\hat{q}}overbold_^ start_ARG bold_italic_q end_ARG (where 𝒒^=𝒒/qbold-^𝒒𝒒𝑞\bm{\hat{q}}=\bm{q}/qoverbold_^ start_ARG bold_italic_q end_ARG = bold_italic_q / italic_q), i.e. Q¯¯(i)(𝒒^)superscript¯¯Q𝑖bold-^𝒒\underline{\underline{\mathrm{Q}}}^{(i)}(\bm{\hat{q}})under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( overbold_^ start_ARG bold_italic_q end_ARG ). Concomitantly, the isotropic functions are dependent on the wavenumber q𝑞qitalic_q and will be marked by a tilde, i.e. q~(i)(q,t)superscript~𝑞𝑖𝑞𝑡\tilde{q}^{(i)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_q , italic_t ). As Lemaître has emphasized and will be important later, the isotropic functions are not connected by a one dimensional Fourier transformation, i.e. FT[q(i)(r,t)]q~(q,t)𝐹𝑇delimited-[]superscript𝑞𝑖𝑟𝑡~𝑞𝑞𝑡FT[q^{(i)}(r,t)]\neq\tilde{q}(q,t)italic_F italic_T [ italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) ] ≠ over~ start_ARG italic_q end_ARG ( italic_q , italic_t ). Under Fourier transformation the tensorial structure and the spatial dimension of the problem leads to a mixing of the functions q(i)(r,t)superscript𝑞𝑖𝑟𝑡q^{(i)}(r,t)italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) and q~(i)(q,t)superscript~𝑞𝑖𝑞𝑡\tilde{q}^{(i)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_q , italic_t ).

The connection between the small-𝒒𝒒\bm{q}bold_italic_q structure of the correlations in reciprocal space and the far field spatial power laws depends on the dimension of space. The same considerations in two dimension lead to only four tensorial contributions given by Eqs. (45) in App. A.2 [31]. For ease of presentation, we therefore first discuss stresses in two dimension before moving to the three dimensonal system of relevance to colloidal hard sphere suspensions.

The stress correlations in Cartesian basis are easily recovered from the isotropic functions q(i)(r,t)superscript𝑞𝑖𝑟𝑡q^{(i)}(r,t)italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_r , italic_t ). For example in two dimensions, the correlation function of an off-diagonal element reads

𝒞xyxy(𝒓,t)=q(3)(r,t)(2r^x2r^y2)+q(4)(r,t)(122r^x2r^y2).subscript𝒞𝑥𝑦𝑥𝑦𝒓𝑡superscript𝑞3𝑟𝑡2superscriptsubscript^𝑟𝑥2superscriptsubscript^𝑟𝑦2superscript𝑞4𝑟𝑡122superscriptsubscript^𝑟𝑥2superscriptsubscript^𝑟𝑦2\displaystyle\mathcal{C}_{xyxy}(\bm{r},t)={q}^{(3)}(r,t)\,\left(2\,\hat{r}_{x}% ^{2}\hat{r}_{y}^{2}\right)+{q}^{(4)}(r,t)\left(\frac{1}{2}-2\hat{r}_{x}^{2}% \hat{r}_{y}^{2}\right).caligraphic_C start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT ( bold_italic_r , italic_t ) = italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) ( 2 over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) + italic_q start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG - 2 over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (11)

The corresponding expression in 3D will be given in Eq. (31) below.

The symmetry based decomposition of the stress correlations has a profound impact on the statistical averaging of the simulation data. As already mentioned, this is essential since we are concerned with the small signals of the power law correlations at large distances. The first aspect is the generally reduced number (five in D=3𝐷3D=3italic_D = 3 and four in D=2𝐷2D=2italic_D = 2) of independent functions q(i)superscript𝑞𝑖q^{(i)}italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT and q~(i)superscript~𝑞𝑖\tilde{q}^{(i)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT from the generally D4superscript𝐷4D^{4}italic_D start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT number of independent functions 𝒞αβγδ(𝒓,t)subscript𝒞𝛼𝛽𝛾𝛿𝒓𝑡\mathcal{C}_{\alpha\beta\gamma\delta}(\bm{r},t)caligraphic_C start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( bold_italic_r , italic_t ). But even more important is the isotropic symmetry of the functions q(i)(r,t)superscript𝑞𝑖𝑟𝑡q^{(i)}(r,t)italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) since then angular averaging becomes possible, which increases the statistics. Alongside it simplifies the presentation of the results, since the whole SACT can be represented by scalar valued functions with scalar dependencies.

V Stress correlations in two dimensions

V.1 Shear modulus and zero wavevector limit

Similar to Eq. (11) in real space, the corresponding shear stress auto-correlation in reciprocal space reads

𝒞~xyxy(𝒒,t)=2q^x2q^y2(q~(3)(q,t)q~(4)(q,t))+12q~(4)(q,t).subscript~𝒞𝑥𝑦𝑥𝑦𝒒𝑡2superscriptsubscript^𝑞𝑥2superscriptsubscript^𝑞𝑦2superscript~𝑞3𝑞𝑡superscript~𝑞4𝑞𝑡12superscript~𝑞4𝑞𝑡\displaystyle\tilde{\mathcal{C}}_{xyxy}(\bm{q},t)=2\,\hat{q}_{x}^{2}\hat{q}_{y% }^{2}\,\left(\tilde{q}^{(3)}(q,t)-\tilde{q}^{(4)}(q,t)\right)+\frac{1}{2}% \tilde{q}^{(4)}(q,t).over~ start_ARG caligraphic_C end_ARG start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT ( bold_italic_q , italic_t ) = 2 over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) - over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) . (12)

This suggests the interpretation of q~(4)(q,t)superscript~𝑞4𝑞𝑡\tilde{q}^{(4)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) and q~(3)(q,t)superscript~𝑞3𝑞𝑡\tilde{q}^{(3)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) as the shear stress correlations along the axes and along the diagonals, respectively. If we take the q0𝑞0q\to 0italic_q → 0 limit for finite times, we recover the shear modulus because of isotropy. Therefore the limit q0𝑞0q\to 0italic_q → 0 needs to be approached independently of the direction, leading to the constraint limq0(q~q,t(3)q~q,t(4))=0subscript𝑞0subscriptsuperscript~𝑞3𝑞𝑡subscriptsuperscript~𝑞4𝑞𝑡0\lim_{q\to 0}\left(\tilde{q}^{(3)}_{q,t}-\tilde{q}^{(4)}_{q,t}\right)=0roman_lim start_POSTSUBSCRIPT italic_q → 0 end_POSTSUBSCRIPT ( over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q , italic_t end_POSTSUBSCRIPT - over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q , italic_t end_POSTSUBSCRIPT ) = 0 at finite times. This consideration rules out any Eshelby power-law spatial dependence at finite times and distances, as will become clear later. It would lead to a non-analytic small q𝑞qitalic_q limit. To summarize, in the (isotropic) liquid the q0𝑞0q\to 0italic_q → 0 limit for finite times leads to the shear modulus

q~(3)(0,t)=q~(4)(0,t)=2G(t).superscript~𝑞30𝑡superscript~𝑞40𝑡2𝐺𝑡\displaystyle\tilde{q}^{(3)}(0,t)=\tilde{q}^{(4)}(0,t)=2G(t)~{}.over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( 0 , italic_t ) = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( 0 , italic_t ) = 2 italic_G ( italic_t ) . (13)

V.2 Force correlations

From the tensorial decomposition in reciprocal space, the correlation functions of the 𝒒𝒒\bm{q}bold_italic_q-dependent forces can be calculated. The force 𝑭𝑭\bm{F}bold_italic_F is the divergence of the momentum current, i.e. 𝑭~(𝒒,t)=i𝒒σ¯~¯(𝒒,t)~𝑭𝒒𝑡𝑖𝒒¯¯~𝜎𝒒𝑡\tilde{\bm{F}}(\bm{q},t)=i\bm{q}\cdot\underline{\underline{\tilde{\sigma}}}(% \bm{q},t)over~ start_ARG bold_italic_F end_ARG ( bold_italic_q , italic_t ) = italic_i bold_italic_q ⋅ under¯ start_ARG under¯ start_ARG over~ start_ARG italic_σ end_ARG end_ARG end_ARG ( bold_italic_q , italic_t ). Similar to velocity correlations, the force correlation tensor for isotropic systems can then be decomposed into longitudinal and transverse parts. This yields the (rescaled by q2superscript𝑞2q^{-2}italic_q start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT) force correlation function

Z¯¯(𝒒,t)=𝒒^𝒞¯¯~(𝒒,t)𝒒^=𝒒^𝒒^Z(q,t)+(𝟏𝒒^𝒒^)Z(q,t).¯¯𝑍𝒒𝑡bold-^𝒒~¯¯𝒞𝒒𝑡bold-^𝒒bold-^𝒒bold-^𝒒superscript𝑍parallel-to𝑞𝑡1bold-^𝒒bold-^𝒒superscript𝑍perpendicular-to𝑞𝑡\displaystyle\underline{\underline{Z}}(\bm{q},t)=\bm{\hat{q}}\cdot\tilde{% \underline{\underline{\mathcal{C}}}}(\bm{q},t)\cdot\bm{\hat{q}}=\bm{\hat{q}}\,% \bm{\hat{q}}Z^{\parallel}(q,t)+({\bf 1}-\bm{\hat{q}}\,\bm{\hat{q}})\,Z^{\perp}% (q,t).under¯ start_ARG under¯ start_ARG italic_Z end_ARG end_ARG ( bold_italic_q , italic_t ) = overbold_^ start_ARG bold_italic_q end_ARG ⋅ over~ start_ARG under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG end_ARG ( bold_italic_q , italic_t ) ⋅ overbold_^ start_ARG bold_italic_q end_ARG = overbold_^ start_ARG bold_italic_q end_ARG overbold_^ start_ARG bold_italic_q end_ARG italic_Z start_POSTSUPERSCRIPT ∥ end_POSTSUPERSCRIPT ( italic_q , italic_t ) + ( bold_1 - overbold_^ start_ARG bold_italic_q end_ARG overbold_^ start_ARG bold_italic_q end_ARG ) italic_Z start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) . (14)

Explicitly computing both divergences leads to expressions of the force correlation functions in terms of the q~(i)superscript~𝑞𝑖\tilde{q}^{(i)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT of the SACT. This reads in two dimensions

Z(q,t)=(12q~(1)(q,t)12q~(2)(q,t)+12q~(3)(q,t)),superscript𝑍parallel-to𝑞𝑡12superscript~𝑞1𝑞𝑡12superscript~𝑞2𝑞𝑡12superscript~𝑞3𝑞𝑡\displaystyle Z^{\parallel}(q,t)=\left(\frac{1}{2}\tilde{q}^{(1)}(q,t)-\frac{1% }{\sqrt{2}}\tilde{q}^{(2)}(q,t)+\frac{1}{2}\tilde{q}^{(3)}(q,t)\right),italic_Z start_POSTSUPERSCRIPT ∥ end_POSTSUPERSCRIPT ( italic_q , italic_t ) = ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) - divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) ) , (15)
Z(q,t)=12q~(4)(q,t).superscript𝑍perpendicular-to𝑞𝑡12superscript~𝑞4𝑞𝑡\displaystyle Z^{\perp}(q,t)=\frac{1}{2}\tilde{q}^{(4)}(q,t)~{}.italic_Z start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) . (16)

The global longitudinal modulus is the corresponding limit of the longitudinal force correlation function K(t)=Z(0,t)/kBT𝐾𝑡superscript𝑍parallel-to0𝑡subscript𝑘𝐵𝑇K(t)=Z^{\parallel}(0,t)/k_{B}Titalic_K ( italic_t ) = italic_Z start_POSTSUPERSCRIPT ∥ end_POSTSUPERSCRIPT ( 0 , italic_t ) / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T and the shear modulus (Eq. 13) is connected to G(t)=Z(0,t)/kBT𝐺𝑡superscript𝑍perpendicular-to0𝑡subscript𝑘𝐵𝑇G(t)=Z^{\perp}(0,t)/k_{B}Titalic_G ( italic_t ) = italic_Z start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( 0 , italic_t ) / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T. But again, these identifications are only valid for finite times. The general relations will be given in Eq. (23) below.

V.3 Fourier relations and far fields

The tesseral tensors in Fourier space Q¯¯(i)(q^)superscript¯¯Q𝑖^𝑞\underline{\underline{\mathrm{Q}}}^{(i)}(\hat{q})under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( over^ start_ARG italic_q end_ARG ) can be transformed to real space by inverse Fourier transformation; denoted FT1𝐹superscript𝑇1FT^{-1}italic_F italic_T start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. For considering the spatial far field, integrals like the following arise (with Kronecker delta δαβsubscript𝛿𝛼𝛽\delta_{\alpha\beta}italic_δ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT and r>0𝑟0r>0italic_r > 0):

d2𝒒4π2q^α2q^β2ei𝒒𝒓eϵq|ϵ=0=14π1r2(5δαβ+8r^α2r^β2).evaluated-atsuperscriptd2𝒒4superscript𝜋2subscriptsuperscript^𝑞2𝛼subscriptsuperscript^𝑞2𝛽superscript𝑒𝑖𝒒𝒓superscript𝑒italic-ϵ𝑞italic-ϵ014𝜋1superscript𝑟25subscript𝛿𝛼𝛽8superscriptsubscript^𝑟𝛼2superscriptsubscript^𝑟𝛽2\displaystyle\int\frac{\text{d}^{2}\bm{q}}{4\pi^{2}}\hat{q}^{2}_{\alpha}\hat{q% }^{2}_{\beta}e^{i\bm{q}\cdot\bm{r}}e^{-\epsilon\,q}\big{|}_{\epsilon=0}=\frac{% 1}{4\pi}\frac{1}{r^{2}}\left(5\delta_{\alpha\beta}+8\hat{r}_{\alpha}^{2}\,\hat% {r}_{\beta}^{2}\right).∫ divide start_ARG d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_italic_q end_ARG start_ARG 4 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG over^ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT over^ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i bold_italic_q ⋅ bold_italic_r end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_ϵ italic_q end_POSTSUPERSCRIPT | start_POSTSUBSCRIPT italic_ϵ = 0 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 4 italic_π end_ARG divide start_ARG 1 end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( 5 italic_δ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT + 8 over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (17)

Obviously, the right hand side can be decomposed into a power-law distance dependence multiplied by tesseral tensors in real space. Considering all tensors and collecting all terms, one finds the following relations [31]:

FT1[(Q¯¯1(𝒒^)Q¯¯2(𝒒^)Q¯¯3(𝒒^)Q¯¯4(𝒒^))]=𝐹superscript𝑇1delimited-[]matrixsubscript¯¯Q1bold-^𝒒subscript¯¯Q2bold-^𝒒subscript¯¯Q3bold-^𝒒subscript¯¯Q4bold-^𝒒absent\displaystyle FT^{-1}\left[\begin{pmatrix}\underline{\underline{\mathrm{Q}}}_{% 1}(\bm{\hat{q}})\\ \underline{\underline{\mathrm{Q}}}_{2}(\bm{\hat{q}})\\ \underline{\underline{\mathrm{Q}}}_{3}(\bm{\hat{q}})\\ \underline{\underline{\mathrm{Q}}}_{4}(\bm{\hat{q}})\end{pmatrix}\right]=italic_F italic_T start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ ( start_ARG start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_q end_ARG ) end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_q end_ARG ) end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_q end_ARG ) end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_q end_ARG ) end_CELL end_ROW end_ARG ) ] =
12πr2(0000020000220022)(Q¯¯1(𝒓^)Q¯¯2(𝒓^)Q¯¯3(𝒓^)Q¯¯5(𝒓^)),forr>0.12𝜋superscript𝑟2matrix0000020000220022matrixsubscript¯¯Q1bold-^𝒓subscript¯¯Q2bold-^𝒓subscript¯¯Q3bold-^𝒓subscript¯¯Q5bold-^𝒓for𝑟0\displaystyle\frac{1}{2\pi r^{2}}\begin{pmatrix}0&0&0&0\\ 0&-2&0&0\\ 0&0&2&-2\\ 0&0&-2&2\end{pmatrix}\begin{pmatrix}\underline{\underline{\mathrm{Q}}}_{1}(\bm% {\hat{r}})\\ \underline{\underline{\mathrm{Q}}}_{2}(\bm{\hat{r}})\\ \underline{\underline{\mathrm{Q}}}_{3}(\bm{\hat{r}})\\ \underline{\underline{\mathrm{Q}}}_{5}(\bm{\hat{r}})\end{pmatrix},~{}\text{for% }~{}r>0.divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( start_ARG start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL - 2 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 2 end_CELL start_CELL - 2 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL - 2 end_CELL start_CELL 2 end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_r end_ARG ) end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_r end_ARG ) end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_r end_ARG ) end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_r end_ARG ) end_CELL end_ROW end_ARG ) , for italic_r > 0 . (18)

The amplitudes of the far-fields in real space, viz. q(i)(r,t)rDproportional-tosuperscript𝑞𝑖𝑟𝑡superscript𝑟𝐷{q}^{(i)}(r\to\infty,t)\propto r^{-D}italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_r → ∞ , italic_t ) ∝ italic_r start_POSTSUPERSCRIPT - italic_D end_POSTSUPERSCRIPT, can then be connected to the zero wavenumber limits, q0𝑞0q\to 0italic_q → 0, of the isotropic functions in reciprocal space, i.e. q~(i)(0,t)superscript~𝑞𝑖0𝑡\tilde{q}^{(i)}(0,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( 0 , italic_t ). These relations can be derived by

limr𝒞¯¯(𝒓,t)=limrFT1[𝒞¯¯~(𝒒,t)]subscript𝑟¯¯𝒞𝒓𝑡subscript𝑟𝐹superscript𝑇1delimited-[]~¯¯𝒞𝒒𝑡\displaystyle\lim_{r\to\infty}\,\underline{\underline{\mathcal{C}}}(\bm{r},t)=% \lim_{r\to\infty}\,FT^{-1}\left[\tilde{\underline{\underline{\mathcal{C}}}}(% \bm{q},t)\right]roman_lim start_POSTSUBSCRIPT italic_r → ∞ end_POSTSUBSCRIPT under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG ( bold_italic_r , italic_t ) = roman_lim start_POSTSUBSCRIPT italic_r → ∞ end_POSTSUBSCRIPT italic_F italic_T start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ over~ start_ARG under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG end_ARG ( bold_italic_q , italic_t ) ]
=iq~(i)(0,t)FT1[Q¯¯i(𝒒^)]absentsubscript𝑖superscript~𝑞𝑖0𝑡𝐹superscript𝑇1delimited-[]subscript¯¯Q𝑖bold-^𝒒\displaystyle=\sum_{i}\tilde{q}^{(i)}(0,t)\,FT^{-1}\left[\underline{\underline% {\mathrm{Q}}}_{i}(\bm{\hat{q}})\right]= ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( 0 , italic_t ) italic_F italic_T start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_q end_ARG ) ]
=!iq(i)(,t)Q¯¯i(𝒓^)subscript𝑖superscript𝑞𝑖𝑡subscript¯¯Q𝑖bold-^𝒓\displaystyle\overset{!}{=}\sum_{i}q^{(i)}(\infty,t)\underline{\underline{% \mathrm{Q}}}_{i}(\bm{\hat{r}})over! start_ARG = end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_r end_ARG ) (19)

and since the terms FT1[Q¯¯i(𝒒^)]𝐹superscript𝑇1delimited-[]subscript¯¯Q𝑖bold-^𝒒FT^{-1}\left[\underline{\underline{\mathrm{Q}}}_{i}(\bm{\hat{q}})\right]italic_F italic_T start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_q end_ARG ) ] are power laws following Eq. (17), the q~(i)(0,t)superscript~𝑞𝑖0𝑡\tilde{q}^{(i)}(0,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( 0 , italic_t ) determine their amplitudes. The exclamation mark marks that set of linear equations, where the identification of the pre-factors of the tensors Q¯¯i(𝒓^)subscript¯¯Q𝑖bold-^𝒓\underline{\underline{\mathrm{Q}}}_{i}(\bm{\hat{r}})under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( overbold_^ start_ARG bold_italic_r end_ARG ) can be performed. The result then reads

(q(1)(,t)q(2)(,t)q(3)(,t)q(4)(,t))=12πr2(0000020000220022)(q~(1)(0,t)q~(2)(0,t)q~(3)(0,t)q~(4)(0,t)).matrixsuperscript𝑞1𝑡superscript𝑞2𝑡superscript𝑞3𝑡superscript𝑞4𝑡12𝜋superscript𝑟2matrix0000020000220022matrixsuperscript~𝑞10𝑡superscript~𝑞20𝑡superscript~𝑞30𝑡superscript~𝑞40𝑡\displaystyle\begin{pmatrix}{q}^{(1)}(\infty,t)\\ {q}^{(2)}(\infty,t)\\ {q}^{(3)}(\infty,t)\\ {q}^{(4)}(\infty,t)\\ \end{pmatrix}=\frac{1}{2\pi r^{2}}\begin{pmatrix}0&0&0&0\\ 0&-2&0&0\\ 0&0&2&-2\\ 0&0&-2&2\end{pmatrix}\begin{pmatrix}\tilde{q}^{(1)}(0,t)\\ \tilde{q}^{(2)}(0,t)\\ \tilde{q}^{(3)}(0,t)\\ \tilde{q}^{(4)}(0,t)\\ \end{pmatrix}.( start_ARG start_ROW start_CELL italic_q start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) end_CELL end_ROW start_ROW start_CELL italic_q start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) end_CELL end_ROW start_ROW start_CELL italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) end_CELL end_ROW start_ROW start_CELL italic_q start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) end_CELL end_ROW end_ARG ) = divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( start_ARG start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL - 2 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 2 end_CELL start_CELL - 2 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL - 2 end_CELL start_CELL 2 end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( 0 , italic_t ) end_CELL end_ROW start_ROW start_CELL over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( 0 , italic_t ) end_CELL end_ROW start_ROW start_CELL over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( 0 , italic_t ) end_CELL end_ROW start_ROW start_CELL over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( 0 , italic_t ) end_CELL end_ROW end_ARG ) . (20)

This consideration of Fourier transformation rules alone yields the relation q(3)(,t)=q(4)(,t)superscript𝑞3𝑡superscript𝑞4𝑡q^{(3)}(\infty,t)=-q^{(4)}(\infty,t)italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) = - italic_q start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ). The far field stress correlations have to obey this relation, which will be used as consistency check in the simulations.

V.4 Force free solids and their fluid precursors

V.4.1 Inherent states as minima in potential energy landscape

In stable configurations at long times, the forces on each particle need to balance. Thus correlations of the force fields need to vanish, including in metastable glasses. This has been discussed in most details for inherent states [17].

Inherent states (IS) denote configurations of amorphous elastic solids at zero temperature, i.e. the potential energy minima of non-crystalline particle configurations. In these minima all particles are force-free, i.e. the divergence of the stress tensor vanishes. This means that the force correlations from Eq. (15) vanish for long times in IS. This is consistent with the results derived in Ref. [31] for IS

q~(q0,)(1)=q~(2)(q0,)/2=q~(3)(q0,)~𝑞superscript𝑞01superscript~𝑞2𝑞02superscript~𝑞3𝑞0\displaystyle\tilde{q}(q\to 0,\infty)^{(1)}=\tilde{q}^{(2)}(q\to 0,\infty)/% \sqrt{2}=\,\tilde{q}^{(3)}(q\to 0,\infty)\,over~ start_ARG italic_q end_ARG ( italic_q → 0 , ∞ ) start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_q → 0 , ∞ ) / square-root start_ARG 2 end_ARG = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q → 0 , ∞ ) (21a)
q~(4)(q0,)=0.superscript~𝑞4𝑞00\displaystyle\tilde{q}^{(4)}(q\to 0,\infty)=0.over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q → 0 , ∞ ) = 0 . (21b)

If one uses the results of Eq. (21) in Eq. (20), one finds the IS relations in real space to read

q(3)(,t)=q(2)(,t)/2=q(4)(,t).superscript𝑞3𝑡superscript𝑞2𝑡2superscript𝑞4𝑡\displaystyle q^{(3)}(\infty,t)=-q^{(2)}(\infty,t)/\sqrt{2}=-q^{(4)}(\infty,t).italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) = - italic_q start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) / square-root start_ARG 2 end_ARG = - italic_q start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) . (22)

V.4.2 Force and stress correlations in the liquid

There appears to be an inconsistency in Eqs. (21) and (13). In the force free solid, the long time limit of the transverse force correlations vanishes for small wavevectors, q~(4)(q0,)=0superscript~𝑞4𝑞00\tilde{q}^{(4)}(q\to 0,\infty)=0over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q → 0 , ∞ ) = 0 (Eq. 21b), while the macroscopic limit at q=0𝑞0q=0italic_q = 0 is identified as shear modulus G(t)𝐺𝑡G(t)italic_G ( italic_t ), viz. G(t)=q~(4)(q=0,t)/2𝐺𝑡superscript~𝑞4𝑞0𝑡2G(t)=\tilde{q}^{(4)}(q=0,t)/2italic_G ( italic_t ) = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q = 0 , italic_t ) / 2 in Eq. (13). The inconsistency becomes glaring when one recalls Maxwell’s interpretation of the long time limit of the shear modulus as shear elastic constant of the solid, Gsolid(t)=Gplsuperscript𝐺𝑠𝑜𝑙𝑖𝑑𝑡subscript𝐺𝑝𝑙G^{solid}(t\to\infty)=G_{pl}italic_G start_POSTSUPERSCRIPT italic_s italic_o italic_l italic_i italic_d end_POSTSUPERSCRIPT ( italic_t → ∞ ) = italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT; the plateau shear modulus also gets denoted as μ𝜇\muitalic_μ, i.e. one of the two Lamé constants of an isotropic solid [34]. Apparently, the limits do not interchange.

The glassy solid at non-zero temperature was investigated in [32], where the shear-modulus decays onto the well defined plateau value, viz. G(t)=Gpl𝐺𝑡subscript𝐺𝑝𝑙G(t\to\infty)=G_{pl}italic_G ( italic_t → ∞ ) = italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT holds. In this case it was possible to observe a well defined shear modulus, even though the force correlations were zero, as a solid state was studied. The liquid case considered in the present work is more subtle since the shear modulus is only approximately constant over a finite period of time before structural relaxation sets in. We will therefore carefully select a time-window where the shear modulus can be considered constant before the structural relaxation sets in; it will be marked by horizontal bars in Figs. 2, 7, 12, and 16.

The apparent inconsistency between Eqs. (21) and (13) is resolved via the proper identification of the intrinsic response functions of the viscoelastic fluid to applied inhomogeneous deformations. The SACT gives the response to an external velocity gradient, which induces stresses but also momentum currents in the system. This coupling can be determined via the Zwanzig-Mori projection operator formalism [14, 15, 19]. The intrinsic response functions [35], which are not affected by the hydrodynamic momentum transport, then describe the stress responses to the internal velocity gradients [36]. These response function were identified as the Zwanzig-Mori memory kernels, well familiar from liquid theory [29]. The force auto-correlation functions depend on the the history of the memory kernels, and the limits q0𝑞0q\to 0italic_q → 0 and t𝑡t\to\inftyitalic_t → ∞ do not commute [14, 15]. For Newtonian dynamics, the non-Markovian relations are (in an isothermal setting, neglecting heat transport) [37, 38]

tZi(q,t)+q2mn0t𝑑tGi(q,tt)Zi(q,t)=tGi(q,t).subscript𝑡superscript𝑍𝑖𝑞𝑡superscript𝑞2𝑚𝑛superscriptsubscript0𝑡differential-dsuperscript𝑡superscript𝐺𝑖𝑞𝑡superscript𝑡superscript𝑍𝑖𝑞superscript𝑡subscript𝑡superscript𝐺𝑖𝑞𝑡\displaystyle\partial_{t}\,Z^{i}(q,t)+\frac{q^{2}}{mn}\int_{0}^{t}\!\!dt^{% \prime}\;G^{i}(q,t-t^{\prime})\,Z^{i}(q,t^{\prime})=\partial_{t}\,G^{i}(q,t).∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_Z start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_q , italic_t ) + divide start_ARG italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m italic_n end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_G start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_q , italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_Z start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_q , italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_q , italic_t ) . (23)

The superscript i𝑖iitalic_i stand for longitudinal (\|) and transverse (perpendicular-to\perp) correlation functions, and the Gi(q,t)superscript𝐺𝑖𝑞𝑡G^{i}(q,t)italic_G start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_q , italic_t ) are the corresponding Zwanzig-Mori memory kernels; see Refs. [14, 15, 19] for their definitions. They approach the moduli in the macroscopic limit, G(q,t)K(t)superscript𝐺𝑞𝑡𝐾𝑡G^{\|}(q,t)\to K(t)italic_G start_POSTSUPERSCRIPT ∥ end_POSTSUPERSCRIPT ( italic_q , italic_t ) → italic_K ( italic_t ) and G(q,t)G(t)superscript𝐺perpendicular-to𝑞𝑡𝐺𝑡G^{\perp}(q,t)\to G(t)italic_G start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) → italic_G ( italic_t ) for q0𝑞0q\to 0italic_q → 0. And for t𝑡t\to\inftyitalic_t → ∞, the moduli approach the elastic constants in the solid limit (denoted by subscripts, i.e. Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT and Kpl=Ksolid(t)subscript𝐾𝑝𝑙superscript𝐾𝑠𝑜𝑙𝑖𝑑𝑡K_{pl}=K^{solid}(t\to\infty)italic_K start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT = italic_K start_POSTSUPERSCRIPT italic_s italic_o italic_l italic_i italic_d end_POSTSUPERSCRIPT ( italic_t → ∞ )).
While the transverse force correlation has already been given in Eq. (15), the Zwanzig Mori decomposition gives [15]

G(q,t)=12q~(3)(q,t)+𝒪(q4),superscript𝐺perpendicular-to𝑞𝑡12superscript~𝑞3𝑞𝑡𝒪superscript𝑞4\displaystyle G^{\perp}(q,t)=\frac{1}{2}\tilde{q}^{(3)}(q,t)+{\cal O}(q^{4})\,,italic_G start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) + caligraphic_O ( italic_q start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) , (24)

in the incompressible limit, which holds well for the hard sphere fluids at the considered densities. These functions will be discussed together with the numerical results in the following figures.

V.5 Molecular dynamics simulations

Refer to caption
Figure 1: The build up of the stress correlation function q(3)(r,t)superscript𝑞3𝑟𝑡q^{(3)}(r,t)italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_r , italic_t ), where different colors denote different times. The black dashed line for small distances describes a r2superscript𝑟2r^{-2}italic_r start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT-power law, the two for lager distances describes a r4superscript𝑟4r^{-4}italic_r start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT-power law. The power-laws emerge in the drag of the wave-fronts, since the shear wave triggers stress-fluctuations that decay on a much longer time-scale. The vertical solid lines correspond to r=cTt𝑟subscript𝑐𝑇𝑡r=c_{T}\,titalic_r = italic_c start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT italic_t and the dashed vertical ones to r=cLt𝑟subscript𝑐𝐿𝑡r=c_{L}\,titalic_r = italic_c start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT italic_t from Eq. (25). The colors denote different times, matching the symbols in Figure 2.

For Newtonian dynamics two mechanisms determine the structure of the stress correlation function. A fast longitudinal wave producing a r4superscript𝑟4r^{-4}italic_r start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT power-law followed by a shear wave, with a concomitant r2superscript𝑟2r^{-2}italic_r start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT algebraic decay [18]. For times at the plateau the shear wave propagates with cT=Gpl/(mn)subscript𝑐𝑇subscript𝐺𝑝𝑙𝑚𝑛c_{T}=\sqrt{G_{pl}/(m\,n)}italic_c start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = square-root start_ARG italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT / ( italic_m italic_n ) end_ARG and cL=Kpl/(mn)subscript𝑐𝐿subscript𝐾𝑝𝑙𝑚𝑛c_{L}=\sqrt{K_{pl}/(m\,n)}italic_c start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = square-root start_ARG italic_K start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT / ( italic_m italic_n ) end_ARG. The transverse sound velocity is determined via the marked plateau in Fig. 2 (red bar), and the longitudinal one via the dispersion relation in Fig. 18 in App. B. The results are

cL=31.51v0;cT=5.47v0,formulae-sequencesubscript𝑐𝐿31.51subscript𝑣0subscript𝑐𝑇5.47subscript𝑣0\displaystyle c_{L}=31.51\,v_{0};\qquad c_{T}=5.47\,v_{0},italic_c start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = 31.51 italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ; italic_c start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 5.47 italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , (25)

where v0subscript𝑣0v_{0}italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the thermal velocity. Figure 1 shows the built up of the power law as described in the introduction. The induced momentum flow is distributed over increasing surfaces (rings in two dimensions) scaling as 1/rproportional-toabsent1𝑟\propto 1/r∝ 1 / italic_r. The stress is given by its spatial derivative, leading to a 1/r2proportional-toabsent1superscript𝑟2\propto 1/r^{2}∝ 1 / italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT law.

In the shown time-frame of Figure 1, the amplitude of the r2superscript𝑟2r^{-2}italic_r start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT algebraic law is almost constant. Only the distance over which it can be observed increases. Eq. (20) and Eq. (13) together seem to suggest, that the power law amplitude vanishes. This would indeed be the case for r𝑟r\to\inftyitalic_r → ∞ at finite times as assumed in Eq. (13). But for the finite distances at finite times that are shown in Fig. 1 the IS-relations approximately are valid on these length scales. Therefore we can use the IS-predictions Eq. (21) in Eq. (20), suggesting that the real space power-law is only determined by q~(3)(0,t)superscript~𝑞30𝑡\tilde{q}^{(3)}(0,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( 0 , italic_t ). This is indeed the case, as the dashed black line for the 22-2- 2 power law has the amplitude of q(3)(0,t=4dl/v0)/πsuperscript𝑞30𝑡4subscript𝑑𝑙subscript𝑣0𝜋q^{(3)}(0,t=4\,d_{l}/v_{0})/\piitalic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( 0 , italic_t = 4 italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_π, i.e. the time of the purple slope. The time-points of Fig. 1 are roughly on the plateau in Fig. 2, therefore having an approximately equal amplitude of the 22-2- 2 power law.
The amplitude of the 44-4- 4 power-law shows a pronounced time-dependency that was calculated be 6kBTGpl2t2/πnm6subscript𝑘𝐵𝑇superscriptsubscript𝐺𝑝𝑙2superscript𝑡2𝜋𝑛𝑚6k_{B}TG_{pl}^{2}t^{2}/\pi n\,m6 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_π italic_n italic_m in Ref. [18]. For the smallest and largest times this slope is drawn, where the Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT estimate is denoted as the red bar in Fig. 2.

Figure 3 shows the isotropic functions q~(3)(q,t)superscript~𝑞3𝑞𝑡\tilde{q}^{(3)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) and q~(4)(q,t)superscript~𝑞4𝑞𝑡\tilde{q}^{(4)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) with the same colors code the times as Figure 1. One sees that both functions agree in the macroscopic limit, q~0,t(3)=q~0,t(4)subscriptsuperscript~𝑞30𝑡subscriptsuperscript~𝑞40𝑡\tilde{q}^{(3)}_{0,t}=\tilde{q}^{(4)}_{0,t}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_t end_POSTSUBSCRIPT = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_t end_POSTSUBSCRIPT for q=0𝑞0q=0italic_q = 0 (Eq. (13)). On the other hand, the force-free prediction holds over successively larger ranges of q𝑞qitalic_q for increasing time, q~q,(4)=0subscriptsuperscript~𝑞4𝑞0\tilde{q}^{(4)}_{q,\infty}=0over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q , ∞ end_POSTSUBSCRIPT = 0 (Eq. (21b)). While q~q,t(3)subscriptsuperscript~𝑞3𝑞𝑡\tilde{q}^{(3)}_{q,t}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q , italic_t end_POSTSUBSCRIPT shows little variations with q𝑞qitalic_q.

The subtle time- and q𝑞qitalic_q-dependence can be explained by Eq. (23). Entering a constant modulus Gpli(q)q~(3)(q,t)/2subscriptsuperscript𝐺𝑖𝑝𝑙𝑞superscript~𝑞3𝑞𝑡2G^{i}_{pl}(q)\approx\tilde{q}^{(3)}(q,t)/2italic_G start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT ( italic_q ) ≈ over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) / 2 into Eq. (23), one observes that the neglect of the integral requires smaller and smaller wavevectors for increasing time. Rather, the force correlation Z(q,t)=q~q,t(4)/2superscript𝑍perpendicular-to𝑞𝑡subscriptsuperscript~𝑞4𝑞𝑡2Z^{\perp}(q,t)=\tilde{q}^{(4)}_{q,t}/2italic_Z start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q , italic_t end_POSTSUBSCRIPT / 2 varies with the transverse sound velocity Gpl/mn.subscript𝐺𝑝𝑙𝑚𝑛\sqrt{G_{pl}/mn}.square-root start_ARG italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT / italic_m italic_n end_ARG . In the case of the solid, one finds that q~q,(3)=const.subscriptsuperscript~𝑞3𝑞𝑐𝑜𝑛𝑠𝑡\tilde{q}^{(3)}_{q,\infty}=const.over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q , ∞ end_POSTSUBSCRIPT = italic_c italic_o italic_n italic_s italic_t . while q~q,(4)=0subscriptsuperscript~𝑞4𝑞0\tilde{q}^{(4)}_{q,\infty}=0over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q , ∞ end_POSTSUBSCRIPT = 0. This justifies the argumentation above, that local stress fluctuations approximately decay as slowly as the bulk shear-modulus.

Refer to caption
Figure 2: The bulk shear modulus G(t)𝐺𝑡G(t)italic_G ( italic_t ) is shown by the black crosses. The colored crosses and dots denote the q=0𝑞0q=0italic_q = 0 values of the q~(3)superscript~𝑞3\tilde{q}^{(3)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT and q~(4)superscript~𝑞4\tilde{q}^{(4)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT slopes from Fig. 3. The colors as well coincide with the ones in Figure 1. The triangle at long times denotes the time of Fig 4.
Refer to caption
Figure 3: Dashed lines denote q~(3)superscript~𝑞3\tilde{q}^{(3)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT and full lines q~(4)superscript~𝑞4\tilde{q}^{(4)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT. Different colors denote different times that can be read off from Fig. 2. It shows that q~0,t(3)=q~0,t(4)subscriptsuperscript~𝑞30𝑡subscriptsuperscript~𝑞40𝑡\tilde{q}^{(3)}_{0,t}=\tilde{q}^{(4)}_{0,t}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_t end_POSTSUBSCRIPT = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_t end_POSTSUBSCRIPT, while for q0𝑞0q\neq 0italic_q ≠ 0 the function q~q,t(4)subscriptsuperscript~𝑞4𝑞𝑡\tilde{q}^{(4)}_{q,t}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q , italic_t end_POSTSUBSCRIPT approaches zero for long times and q~q,t(3)subscriptsuperscript~𝑞3𝑞𝑡\tilde{q}^{(3)}_{q,t}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q , italic_t end_POSTSUBSCRIPT stays rather constant, which is the IS-prediction from Eqs. (21).

Following Figure 3 we observe that with increasing times the force-free relations Eq. (21) hold for successively smaller wave vectors. This can be observed in how q~(4)superscript~𝑞4\tilde{q}^{(4)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT vanishes for successively smaller q𝑞qitalic_q with increasing time full-filling Eq. (21b). For q=0𝑞0q=0italic_q = 0 it still approaches the value of q~(3)superscript~𝑞3\tilde{q}^{(3)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT, which confirms Eq. (13).

These results imply that the length scale, over which forces are relaxed, increases with increasing times, while elastic stresses remain in the solid. In isotropic solids this length scale is unbounded but in viscoelastic liquids the whole pattern decays with the decay of the shear modulus, that sets the amplitude of the power law and vanishes for long times in the liquid.

Figure 4 shows the isotropic functions in real and reciprocal space for much longer times (the red triangle in Fig. 2). As the bottom panel shows, for small but non-zero q𝑞qitalic_q’s, we can reproduce Eqs. (21), i.e. the predictions of [31]. The top panel shows the power laws in real space that coincide following Eq. (22). The red bar is a power-law with the amplitude from the Fourier space values according to Eq. (20).

Ref. [32] is concerned with the solid case and finds that the only non-vanishing function is determined by Young’s Modulus E𝐸Eitalic_E. In our case, it is the shear modulus G(t)𝐺𝑡G(t)italic_G ( italic_t ) that defines the q=0𝑞0q=0italic_q = 0 limits. This difference originates in the different order of the limits: the limit q0𝑞0q\to 0italic_q → 0 at finite times (fluid), while results from Ref. [32] considers t𝑡t\to\inftyitalic_t → ∞ at finite q𝑞qitalic_q (solid).

Refer to caption
Figure 4: Top panel: For long times, the far fields show indeed the predicted power law rDsuperscript𝑟𝐷r^{-D}italic_r start_POSTSUPERSCRIPT - italic_D end_POSTSUPERSCRIPT with amplitudes obeying the relations from Eq. (22). The red slope shows a power-law with an amplitude taken from the average value indicated by the black bar in the lower panel, i.e. the connection via Eq. (20).
Bottom panel: In Fourier space, relation Eq. (21) is observed, as well as the fact that q~(3)(0,t)=q~(4)(0,t)superscript~𝑞30𝑡superscript~𝑞40𝑡\tilde{q}^{(3)}(0,t)=\tilde{q}^{(4)}(0,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( 0 , italic_t ) = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( 0 , italic_t ) (smallest data point) from Eq.(13).
Refer to caption
Figure 5: The final decay of q~(3)(q,t)superscript~𝑞3𝑞𝑡\tilde{q}^{(3)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) for the range of small wavevectors also covered in Fig. 3. We do not observe a wave-vector dependency of the relaxation of q~3(q,t)superscript~𝑞3𝑞𝑡\tilde{q}^{3}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_q , italic_t ). Follwoing Eq. (24) this implies a q𝑞qitalic_q-independent viscosity.

Extending the time window to the viscous regime, the wavevector dependent decay of the correlation function of the fluctuating shear stresses can be tested. This gives insight into the wavevector dependence of the shear viscosity, which is given by the time-integral of the shear memory kernel [29], viz.

η(q)=0𝑑tG(q,t).superscript𝜂perpendicular-to𝑞superscriptsubscript0differential-d𝑡superscript𝐺perpendicular-to𝑞𝑡\displaystyle\eta^{\perp}(q)=\int_{0}^{\infty}dt\,G^{\perp}(q,t)~{}.italic_η start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_t italic_G start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) . (26)

Figure 5 reports the final decay of the q~(3)(t)superscript~𝑞3𝑡\tilde{q}^{(3)}(t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_t ) functions, which give the dominant contribution according to Eq. (24). The lack of a q𝑞qitalic_q-dependence of the amplitude and of the viscous relaxation predicts that the shear viscosity does not depended on wave vector for the studied q𝑞qitalic_q-range.

V.6 Brownian dynamics simulations

In Ref. [19] a theory for shear stress correlations in Brownian systems was proposed. Even though frictional forces break the conservation of momentum that was used in the Newtonian theory [15], it was possible to transfer the approach to Brownian systems. It was achieved by considering the transverse displacement field as a slow variable. The relation between force correlation functions and memory kernels from Eq. (23) becomes overdamped:

ζZi(q,t)+q2n0t𝑑tGi(q,tt)Zi(q,t)=ζGi(q,t),𝜁superscript𝑍𝑖𝑞𝑡superscript𝑞2𝑛superscriptsubscript0𝑡differential-dsuperscript𝑡superscript𝐺𝑖𝑞𝑡superscript𝑡superscript𝑍𝑖𝑞superscript𝑡𝜁superscript𝐺𝑖𝑞𝑡\displaystyle\zeta\,Z^{i}(q,t)+\frac{q^{2}}{n}\int_{0}^{t}\!\!dt^{\prime}\;G^{% i}(q,t-t^{\prime})\,Z^{i}(q,t^{\prime})=\zeta\,G^{i}(q,t),italic_ζ italic_Z start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_q , italic_t ) + divide start_ARG italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_n end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_G start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_q , italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_Z start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_q , italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = italic_ζ italic_G start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_q , italic_t ) , (27)

where ζ𝜁\zetaitalic_ζ is the friction coefficient. The resulting shear stress auto-correlation function in the over damped case reads

𝒞xyxy(𝒒,t)=G(t)[4q^x2q^y2+(14q^x2q^y2)eq2Dt].subscript𝒞𝑥𝑦𝑥𝑦𝒒𝑡𝐺𝑡delimited-[]4superscriptsubscript^𝑞𝑥2superscriptsubscript^𝑞𝑦214superscriptsubscript^𝑞𝑥2superscriptsubscript^𝑞𝑦2superscript𝑒superscript𝑞2𝐷𝑡\displaystyle\mathcal{C}_{xyxy}(\bm{q},t)=G(t)\left[4\hat{q}_{x}^{2}\hat{q}_{y% }^{2}+(1-4\hat{q}_{x}^{2}\hat{q}_{y}^{2})e^{-q^{2}\,D\,t}\right].caligraphic_C start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT ( bold_italic_q , italic_t ) = italic_G ( italic_t ) [ 4 over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( 1 - 4 over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT - italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_D italic_t end_POSTSUPERSCRIPT ] . (28)

Here, the generalized hydrodynamic limit was taken. It corresponds to taking the q0𝑞0q\to 0italic_q → 0 limit in the Zwanzig-Mori memory kernel, while its full time-dependence is kept, G(t)=G0(t)𝐺𝑡superscriptsubscript𝐺0perpendicular-to𝑡G(t)=G_{0}^{\perp}(t)italic_G ( italic_t ) = italic_G start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_t ). In Refs. [14, 15, 19] an additional Maxwell approximation was performed, G(t)GgM=Gplet/τ+Γδ(t)𝐺𝑡superscript𝐺𝑔𝑀subscript𝐺𝑝𝑙superscript𝑒𝑡𝜏Γ𝛿𝑡G(t)\approx G^{gM}=G_{pl}\,e^{-t/\tau}+\Gamma\delta(t)italic_G ( italic_t ) ≈ italic_G start_POSTSUPERSCRIPT italic_g italic_M end_POSTSUPERSCRIPT = italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_t / italic_τ end_POSTSUPERSCRIPT + roman_Γ italic_δ ( italic_t ), which we can forego because G(t)𝐺𝑡G(t)italic_G ( italic_t ) is known from the simulations. The diffusion constant is connected to the plateau value of the shear modulus Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT and to the friction constant of the solvent ζ𝜁\zetaitalic_ζ, via D=Gpl/(nζ)𝐷subscript𝐺𝑝𝑙𝑛𝜁D=G_{pl}/(n\,\zeta)italic_D = italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT / ( italic_n italic_ζ ) with n𝑛nitalic_n being the number density. This result can easily be derived from Eq. (27), and identifies the part of 𝒞xyxy(𝒒,t)subscript𝒞𝑥𝑦𝑥𝑦𝒒𝑡\mathcal{C}_{xyxy}(\bm{q},t)caligraphic_C start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT ( bold_italic_q , italic_t ) along the axis as force correlation Zsuperscript𝑍perpendicular-toZ^{\perp}italic_Z start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT. This can easily be confirmed because along the axes Eq. (28) simplifies to

𝒞xyxy(qx^,t)=G(t)eq2Dt,subscript𝒞𝑥𝑦𝑥𝑦𝑞^𝑥𝑡𝐺𝑡superscript𝑒superscript𝑞2𝐷𝑡\displaystyle\mathcal{C}_{xyxy}(q\,\hat{x},t)=G(t)e^{-q^{2}\,D\,t}~{},caligraphic_C start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT ( italic_q over^ start_ARG italic_x end_ARG , italic_t ) = italic_G ( italic_t ) italic_e start_POSTSUPERSCRIPT - italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_D italic_t end_POSTSUPERSCRIPT , (29)

and from the tensorial decomposition Eq. (12) we as well know

𝒞xyxy(qx^,t)=Z(q,t)=12q~(4)(q,t).subscript𝒞𝑥𝑦𝑥𝑦𝑞^𝑥𝑡superscript𝑍perpendicular-to𝑞𝑡12superscript~𝑞4𝑞𝑡\displaystyle\mathcal{C}_{xyxy}(q\hat{x},t)=Z^{\perp}(q,t)=\frac{1}{2}\tilde{q% }^{(4)}(q,t)~{}.caligraphic_C start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT ( italic_q over^ start_ARG italic_x end_ARG , italic_t ) = italic_Z start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) . (30)

From Eq. (29) and Eq. (30), we can read off a Gaussian prediction for q~(4)(q,t)superscript~𝑞4𝑞𝑡\tilde{q}^{(4)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) for different times, where the two fit parameters of amplitude and variance give an estimate for the shear modulus G(t)𝐺𝑡G(t)italic_G ( italic_t ) and for the diffusion coefficient D𝐷Ditalic_D respectively.

To test these predictions, we perform Brownian dynamics simulations, and discuss the results in the following figures. Figure 6 shows the results, simulation data as solid and Gaussian fits as dashed lines. Similar to Figure 3 we can observe how the transverse force correlations decay to zero over larger length-scales for increasing times as predicted by Eqs. (21).

Refer to caption
Figure 6: q~(4)(q,t)superscript~𝑞4𝑞𝑡\tilde{q}^{(4)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ), where different colors show different times. Following Eq. (29), the superimposed smooth lines show Gaussian fits of the form a1eq2b1subscript𝑎1superscript𝑒superscript𝑞2subscript𝑏1a_{1}\,e^{-q^{2}\,b_{1}}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, a1subscript𝑎1a_{1}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT giving an estimate for G(t)𝐺𝑡G(t)italic_G ( italic_t ) and b1subscript𝑏1b_{1}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT giving an estimate of Dt𝐷𝑡D\,titalic_D italic_t, where t𝑡titalic_t is of course known. These relations are verified in Figs. 7, 8.

As an consistency check, Figure 7 shows G(t)𝐺𝑡G(t)italic_G ( italic_t ) measured in a conventional way, Eq. (5), in black, while the colored scatters are the amplitude fit parameters a1(t)subscript𝑎1𝑡a_{1}(t)italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) from Figure 6. The orange bar denotes the mean over these points where the shear modulus G(t)𝐺𝑡G(t)italic_G ( italic_t ) was assumed to have plateaued.

Refer to caption
Figure 7: Shear modulus measured from collisions in the whole box (black) and from the amplitude parameters a1(t)subscript𝑎1𝑡a_{1}(t)italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) fitted in Figure 6 (colors), which are connected via Eqs. (29, 30). We consider G(t)𝐺𝑡G(t)italic_G ( italic_t ) to have an approximate plateau at the last seven data-points, and averaging over these data points results in the value of the orange bar, estimating Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT.

Figure 8 shows the fit parameters b1(t)subscript𝑏1𝑡b_{1}(t)italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) from Figure 6. Since we know Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT from figure 7 and ζ𝜁\zetaitalic_ζ from τbsubscript𝜏𝑏\tau_{b}italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT, we can predict the diffusion. This prediction is plotted by the orange slope in Fig. 8. For a comparison the blue one is just a linear fit, showing the good agreement with the orange slope. The linear fit has an off-set parameter b2subscript𝑏2b_{2}italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT that is as well used for the orange dashed slope. This terms accounts for non-diffusive effects of the shear mode for short times, where the modulus has not plateaued. The agreement with a linear fit is convincing validating the prediction that the correlations propagate diffusively in the regime of the plateau of G(t)𝐺𝑡G(t)italic_G ( italic_t ).

Refer to caption
Figure 8: The crosses show the diffusive exponentials from the similarly colored slopes in Fig. 6. The blue line shows a linear fit through these data points. The orange line gives the theory predictions, where ζ𝜁\zetaitalic_ζ comes from the Brownian thermostat and Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT is taken from Fig. 7.
Refer to caption
Figure 9: q(3)(r,t)superscript𝑞3𝑟𝑡q^{(3)}(r,t)italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) in Brownian dynamics at tD0n=0.33; 1.5; 12.2𝑡subscript𝐷0𝑛0.331.512.2t\,D_{0}\,n=0.33;\,1.5;\,12.2italic_t italic_D start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_n = 0.33 ; 1.5 ; 12.2. It was possible to resolve the shear stress correlations up to a distance of roughly r=50dl𝑟50subscript𝑑𝑙r=50\,d_{l}italic_r = 50 italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT.

To this point the investigation of the overdamped system was concerned with the diffusive propagation of the mode. Figure 9 is concerned with the long ranged correlations that are as well present in the Brownian system. It shows the shear stress correlation along the diagonals, i.e. q(3)superscript𝑞3q^{(3)}italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT for three different times: tD0n=0.33; 1.5; 12.2𝑡subscript𝐷0𝑛0.331.512.2t\,D_{0}\,n=0.33;\,1.5;\,12.2italic_t italic_D start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_n = 0.33 ; 1.5 ; 12.2. In contrast to the Newtonian dynamics the statistics is less accurate. The as well expected r4superscript𝑟4r^{-4}italic_r start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT behavior could not be resolved. The dashed line denotes the r2superscript𝑟2r^{-2}italic_r start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT power law with an amplitude given by the orange Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT estimate from Figure 7. Since the tensorial decomposition is independent of the dynamics those are as well connected via Eqs. (13, 20).

VI Three Dimensions

We repeat the previous analysis concerning two dimensional systems, but now in three dimensions. In three dimensions the decomposition consists of five functions q~(i)superscript~𝑞𝑖\tilde{q}^{(i)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT connected to the tensors Eqs. (43) [39].

For instance the shear stress component reads in the three dimensional tesseral decomposition:

𝒞xyxy(𝒒,t)=122q~(4)(q,t)+subscript𝒞𝑥𝑦𝑥𝑦𝒒𝑡limit-from122superscript~𝑞4𝑞𝑡\displaystyle\mathcal{C}_{xyxy}(\bm{q},t)=\frac{1}{2\sqrt{2}}\,\tilde{q}^{(4)}% (q,t)+caligraphic_C start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT ( bold_italic_q , italic_t ) = divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) +
q^x2q^y2(32q~(3)(q,t)+122q~(5)(q,t)2q~(4)(q,t)).superscriptsubscript^𝑞𝑥2superscriptsubscript^𝑞𝑦232superscript~𝑞3𝑞𝑡122superscript~𝑞5𝑞𝑡2superscript~𝑞4𝑞𝑡\displaystyle\hat{q}_{x}^{2}\,\hat{q}_{y}^{2}\,\left(\frac{3}{2}\tilde{q}^{(3)% }(q,t)+\frac{1}{2\sqrt{2}}\,\tilde{q}^{(5)}(q,t)-\sqrt{2}\tilde{q}^{(4)}(q,t)% \right).over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG 3 end_ARG start_ARG 2 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) + divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 5 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) - square-root start_ARG 2 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) ) . (31)

The anisotropic second term is connected to an r3superscript𝑟3r^{-3}italic_r start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT power law that vanishes for large distances at finite times. Nevertheless Eq. (5) needs to hold, which therefore yields a relation of the shear modulus to the q=0𝑞0q=0italic_q = 0 value of q~(4)(q,t)superscript~𝑞4𝑞𝑡\tilde{q}^{(4)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t )

G(t)=122q~(4)(0,t).𝐺𝑡122superscript~𝑞40𝑡\displaystyle G(t)=\frac{1}{2\sqrt{2}}\tilde{q}^{(4)}(0,t).italic_G ( italic_t ) = divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( 0 , italic_t ) . (32)

Along the diagonal in the xy𝑥𝑦xyitalic_x italic_y-plane the corresponding real space decomposition in Eq. (31) reads

𝒞xyxy(rdiag,t)=38q(3)(r,t)+182q(5)(r,t)subscript𝒞𝑥𝑦𝑥𝑦superscript𝑟𝑑𝑖𝑎𝑔𝑡38superscript𝑞3𝑟𝑡182superscript𝑞5𝑟𝑡\displaystyle\mathcal{C}_{xyxy}(r^{diag},t)=\frac{3}{8}{q}^{(3)}(r,t)+\frac{1}% {8\sqrt{2}}\,{q}^{(5)}(r,t)caligraphic_C start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT ( italic_r start_POSTSUPERSCRIPT italic_d italic_i italic_a italic_g end_POSTSUPERSCRIPT , italic_t ) = divide start_ARG 3 end_ARG start_ARG 8 end_ARG italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) + divide start_ARG 1 end_ARG start_ARG 8 square-root start_ARG 2 end_ARG end_ARG italic_q start_POSTSUPERSCRIPT ( 5 ) end_POSTSUPERSCRIPT ( italic_r , italic_t ) (33)

VI.1 Precursors of IS in the liquid

The force correlation in three dimensions read (cf. Eq.  (14)):

Z(q,t)=(13q~(1)(q,t)23q~(2)(q,t)+23q~(3)(q,t)),superscript𝑍parallel-to𝑞𝑡13superscript~𝑞1𝑞𝑡23superscript~𝑞2𝑞𝑡23superscript~𝑞3𝑞𝑡\displaystyle Z^{\parallel}(q,t)=\left(\frac{1}{3}\tilde{q}^{(1)}(q,t)-\frac{2% }{3}\tilde{q}^{(2)}(q,t)+\frac{2}{3}\tilde{q}^{(3)}(q,t)\right),italic_Z start_POSTSUPERSCRIPT ∥ end_POSTSUPERSCRIPT ( italic_q , italic_t ) = ( divide start_ARG 1 end_ARG start_ARG 3 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) - divide start_ARG 2 end_ARG start_ARG 3 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) + divide start_ARG 2 end_ARG start_ARG 3 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) ) , (34a)
Z(q,t)=q~(4)(q,t).superscript𝑍perpendicular-to𝑞𝑡superscript~𝑞4𝑞𝑡\displaystyle Z^{\perp}(q,t)=\tilde{q}^{(4)}(q,t)~{}.italic_Z start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) . (34b)

Confirming the IS calculations [39] that yield

q~(1)(0,)=q~(2)(0,)=2q~(3)(0,),superscript~𝑞10superscript~𝑞202superscript~𝑞30\displaystyle\tilde{q}^{(1)}(0,\infty)=\tilde{q}^{(2)}(0,\infty)=2\,\tilde{q}^% {(3)}(0,\infty)~{},over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( 0 , ∞ ) = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( 0 , ∞ ) = 2 over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( 0 , ∞ ) , (35a)
q~(4)(0,)=0.superscript~𝑞400\displaystyle\tilde{q}^{(4)}(0,\infty)=0~{}.over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( 0 , ∞ ) = 0 . (35b)

Therefore only two independent functions determine the IS.

From the ZM-decomposition in the incompressible limit the shear modulus can be connected to the reciprocal version of Eq. (33), which follows the analogous considerations as Eq. (24) in 2D but in 3D

G(q,t)=38q~(3)(q,t)+182q~(5)(q,t)+𝒪(q4).superscript𝐺perpendicular-to𝑞𝑡38superscript~𝑞3𝑞𝑡182superscript~𝑞5𝑞𝑡𝒪superscript𝑞4\displaystyle G^{\perp}(q,t)=\frac{3}{8}\tilde{q}^{(3)}(q,t)+\frac{1}{8\sqrt{2% }}\,\tilde{q}^{(5)}(q,t)+\mathcal{O}(q^{4}).italic_G start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) = divide start_ARG 3 end_ARG start_ARG 8 end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) + divide start_ARG 1 end_ARG start_ARG 8 square-root start_ARG 2 end_ARG end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 5 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) + caligraphic_O ( italic_q start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) . (36)

The same considerations leading to Eq. (19) in three dimensions, yield the connection of the power-law amplitudes q(i)(,t)superscript𝑞𝑖𝑡q^{(i)}(\infty,t)italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) to the q0𝑞0q\to 0italic_q → 0 limits in Fourier space q~(i)(q0)=q~0(i)superscript~𝑞𝑖𝑞0subscriptsuperscript~𝑞𝑖0\tilde{q}^{(i)}(q\to 0)=\tilde{q}^{(i)}_{0}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_q → 0 ) = over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT

(q,t(1)q,t(2)q,t(3)q,t(4)q,t(5))=34πr3(0000001000001212002100012012)(q~0,t(1)q~0,t(2)q~0,t(3)q~0,t(4)q~0,t(5)).matrixsubscriptsuperscript𝑞1𝑡subscriptsuperscript𝑞2𝑡subscriptsuperscript𝑞3𝑡subscriptsuperscript𝑞4𝑡subscriptsuperscript𝑞5𝑡34𝜋superscript𝑟3matrix0000001000001212002100012012matrixsubscriptsuperscript~𝑞10𝑡subscriptsuperscript~𝑞20𝑡subscriptsuperscript~𝑞30𝑡subscriptsuperscript~𝑞40𝑡subscriptsuperscript~𝑞50𝑡\displaystyle\begin{pmatrix}{q}^{(1)}_{\infty,t}\\ {q}^{(2)}_{\infty,t}\\ {q}^{(3)}_{\infty,t}\\ {q}^{(4)}_{\infty,t}\\ {q}^{(5)}_{\infty,t}\\ \end{pmatrix}=\frac{3}{4\,\pi r^{3}}\begin{pmatrix}0&0&0&0&0\\ 0&-1&0&0&0\\ 0&0&1&-\sqrt{2}&\frac{1}{\sqrt{2}}\\ 0&0&-\sqrt{2}&1&0\\ 0&0&\frac{1}{\sqrt{2}}&0&-\frac{1}{2}\\ \end{pmatrix}\begin{pmatrix}{\tilde{q}}^{(1)}_{0,t}\\ {\tilde{q}}^{(2)}_{0,t}\\ {\tilde{q}}^{(3)}_{0,t}\\ {\tilde{q}}^{(4)}_{0,t}\\ {\tilde{q}}^{(5)}_{0,t}\\ \end{pmatrix}~{}.( start_ARG start_ROW start_CELL italic_q start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ∞ , italic_t end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_q start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ∞ , italic_t end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ∞ , italic_t end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_q start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ∞ , italic_t end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_q start_POSTSUPERSCRIPT ( 5 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ∞ , italic_t end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) = divide start_ARG 3 end_ARG start_ARG 4 italic_π italic_r start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ( start_ARG start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL - 1 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 1 end_CELL start_CELL - square-root start_ARG 2 end_ARG end_CELL start_CELL divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL - square-root start_ARG 2 end_ARG end_CELL start_CELL 1 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG end_CELL start_CELL 0 end_CELL start_CELL - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_t end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_t end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_t end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_t end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 5 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_t end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) . (37)

This is the three dimensional analog of Eq. (20).

As expected, the pressure correlation is not long ranged which can be read off directly from q,t(1)=0superscriptsubscript𝑞𝑡10q_{\infty,t}^{(1)}=0italic_q start_POSTSUBSCRIPT ∞ , italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = 0. The lines are not linearly independent, from which the relation follows

q(3)(,t)=2(q(4)(,t)+q(5)(,t)).superscript𝑞3𝑡2superscript𝑞4𝑡superscript𝑞5𝑡\displaystyle q^{(3)}(\infty,t)=-\sqrt{2}\left(q^{(4)}(\infty,t)+q^{(5)}(% \infty,t)\right).italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) = - square-root start_ARG 2 end_ARG ( italic_q start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) + italic_q start_POSTSUPERSCRIPT ( 5 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) ) . (38)

If one uses the IS results Eq. (35) in Eq. (37) another relation for the real space power laws emerges

q(2)(,t)=2q(4)(,t).superscript𝑞2𝑡2superscript𝑞4𝑡\displaystyle q^{(2)}(\infty,t)=\sqrt{2}\,q^{(4)}(\infty,t)~{}.italic_q start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) = square-root start_ARG 2 end_ARG italic_q start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( ∞ , italic_t ) . (39)

VI.2 Molecular dynamics simulations

Figure 10 shows the evolution of the shear stress correlations as the two wave fronts travel through the system. The simulation was run with Newtonian dynamics, allowing the appearance of oscillating sound waves as seen in the bottom panel. Similar to two dimensions the momentum current is distributed over spherical surfaces scaling as r2proportional-toabsentsuperscript𝑟2\propto r^{-2}∝ italic_r start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, which leads to a stress tensor scaling proportional to its derivative r3proportional-toabsentsuperscript𝑟3\propto r^{-3}∝ italic_r start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. The top panel shows this build up in real space, where the black bar is a guide to the eye. Furthermore we observe a r5superscript𝑟5r^{-5}italic_r start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT power law in the drag of a longitudinal soundwave. The vertical solid lines denote the position of the shear wave front, calculated from the transverse sound velocity given by the plateau value of the shear modulus Fig. 12 denoted by the horizontal bar. The dashed vertical lines denote the expected position of the soundwave front from the longitudinal sound velocity. The different times correspond to similarly colored dots in Fig. 12. The r5superscript𝑟5r^{-5}italic_r start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT amplitude can be calculated from ZM-considerations on the plateau to read

𝒞xyxydiag(t)=578πkBTmnGpl2t2r5.superscriptsubscript𝒞𝑥𝑦𝑥𝑦𝑑𝑖𝑎𝑔𝑡578𝜋subscript𝑘𝐵𝑇𝑚𝑛superscriptsubscript𝐺𝑝𝑙2superscript𝑡2superscript𝑟5\displaystyle\mathcal{C}_{xyxy}^{diag}(t)=\frac{57}{8\,\pi}\,\frac{k_{B}T}{m\,% n}\,G_{pl}^{2}\,t^{2}\,r^{-5}~{}.caligraphic_C start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d italic_i italic_a italic_g end_POSTSUPERSCRIPT ( italic_t ) = divide start_ARG 57 end_ARG start_ARG 8 italic_π end_ARG divide start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_m italic_n end_ARG italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_r start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT . (40)

Most interestingly this power law amplitude increases in the time-window on the plateau. Fig. 10 shows the explicit results of Eq. (40) for the earliest three slopes, i.e. green, red and purple as black dashed lines. Again Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT was the estimate from the red bar in Fig. 12. For the green slope it is even possible to observe how the curve deviates from the r5superscript𝑟5r^{-5}italic_r start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT law at the position of the sound wave front (green dashed vertical line).

The bottom panel shows q~(4)superscript~𝑞4\tilde{q}^{(4)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT that is proportional to the transverse force correlations, cf. Eq. (34b). Therefore Eq. (23) describes the behavior and simulations indeed show the expected waves. Also, the q~(4)superscript~𝑞4\tilde{q}^{(4)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT vanishes for smaller wave vectors as time increases. The dashed lines show the two q~(i)superscript~𝑞𝑖\tilde{q}^{(i)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT of the rhs. of Eq. (36). Indeed the q𝑞qitalic_q-dependence can be neglected as proposed by the ZM-calculations.

Refer to caption
Figure 10: Top: the evolution of the 𝒞xyxysubscript𝒞𝑥𝑦𝑥𝑦\mathcal{C}_{xyxy}caligraphic_C start_POSTSUBSCRIPT italic_x italic_y italic_x italic_y end_POSTSUBSCRIPT shear stress correlation along the diagonals. The black dashed line at shorter distances denotes the r3superscript𝑟3r^{-3}italic_r start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT power law with a constant amplitude. The black dashed lines at larger distances denote the r5superscript𝑟5r^{-5}italic_r start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT-power law for the three earliest times, which has a time dependent amplitude cf. Eq. (40). Continuous vertical lines denote the expected position of the transverse wavefront, dashed vertical lines the position of the longitudinal wave. Bottom: solid lines show q~4superscript~𝑞4\tilde{q}^{4}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT, denoting transverse force correlations that vanish. Dashed lines show the combination of q~(i)superscript~𝑞𝑖\tilde{q}^{(i)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT given in Eq. (36), verifying the q𝑞qitalic_q-independence of the shear modulus. The colors top and bottom match the times shown in Fig. 12.

The precursors of the IS relations in the liquid given in Sec. VI.1 can be observed in Figure 11. It shows the SACT at t=15dl/v0𝑡15subscript𝑑𝑙subscript𝑣0t=15\,d_{l}/v_{0}italic_t = 15 italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, marked by the red vertical bar in Fig. 12. The top panel shows real space measurements of the q(i)superscript𝑞𝑖q^{(i)}italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT in a combination so that the far field power laws should coincide because of Eq. (38). Following Eq. (37) the far field of q(3)superscript𝑞3q^{(3)}italic_q start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT is determined by an r3superscript𝑟3r^{-3}italic_r start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT power-law with an amplitude given by a combination of several q~(i)(0,t)superscript~𝑞𝑖0𝑡\tilde{q}^{(i)}(0,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( 0 , italic_t ). This is shown by the green line.

The second relation Eq. (39) is shown in the center panel, together with the corresponding power laws determined by the Fourier space values again via Eq. (37).

The bottom panel shows that the IS relations in Fourier space Eqs. (35) are present in the liquid state already. Nevertheless for the smallest q𝑞qitalic_q-values the forces seem to be not fully relaxed, giving rise to small deviations from the IS predictions at small q𝑞qitalic_q.

Refer to caption
Figure 11: The SACF functions at t=15v0/dl𝑡15subscript𝑣0subscript𝑑𝑙t=15\,v_{0}/d_{l}italic_t = 15 italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and Newtonian dynamics. Top: relation Eq. (38) is confirmed as well as the power-law amplitudes predicted by Fourier transformation via Eq. (37). Mid: Relation Eq. (39) is confirmed, as well as the corresponding part of Eq. (37). Bottom: the relations from the IS predictions in Eqs. (35) are confirmed well.
Refer to caption
Figure 12: Shear modulus in Newtonian dynamics in 3D as black crosses. Circles denote the small wave-vector qminsubscript𝑞𝑚𝑖𝑛q_{min}italic_q start_POSTSUBSCRIPT italic_m italic_i italic_n end_POSTSUBSCRIPT values of the similarly colored dashed lines in the bottom panel of Fig. 10. The dots denote the times of the similarly colored slopes of top and bottom panel in Fig. 10. The red vertical bar at t=15v0/dl𝑡15subscript𝑣0subscript𝑑𝑙t=15\,v_{0}/d_{l}italic_t = 15 italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT denotes the time of Fig. 11. The red horizontal bar denotes the estimate of Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT, which determines the transverse sound velocity.
Refer to caption
Figure 13: The three dimensional analog to Fig. 5. Via Eq. (36) the q~(i)superscript~𝑞𝑖\tilde{q}^{(i)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT are connected to the generalized shear modulus G(q,t)superscript𝐺perpendicular-to𝑞𝑡G^{\perp}(q,t)italic_G start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ). No length scale dependent relaxation is observed, leading to a q𝑞qitalic_q-independent viscosity according to Eq. (26).

Figure 13 shows the generalized shear modulus G(q,t)superscript𝐺perpendicular-to𝑞𝑡G^{\perp}(q,t)italic_G start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT ( italic_q , italic_t ) that can be calculated from the q~(i)superscript~𝑞𝑖\tilde{q}^{(i)}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT via Eq. (36). No legth scale dependent relaxation is observed, as already anticipated in the bottom panel of Fig. 10.

VI.3 Brownian dynamics simulations

We perform the equivalent analysis of the Brownian dynamics in three dimensions as was already performed in two dimensions in Sec.V.6.
By again considering only the directions along the axis in Eq. (31) (i.e. the isotropic part) from the ZM-projection operator formalism we find the connection to theory predictions of Ref. [19] to read

122q~(4)(q,t)=G(t)eq2Dt.122superscript~𝑞4𝑞𝑡𝐺𝑡superscript𝑒superscript𝑞2𝐷𝑡\displaystyle\frac{1}{2\sqrt{2}}\tilde{q}^{(4)}(q,t)=G(t)\,e^{-q^{2}\,Dt}.divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) = italic_G ( italic_t ) italic_e start_POSTSUPERSCRIPT - italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_D italic_t end_POSTSUPERSCRIPT . (41)

Analogous to Fig. 6 in two dimensions, the theory in three dimensions predicts a diffusive slope for q~(4)(q,t)superscript~𝑞4𝑞𝑡\tilde{q}^{(4)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) in three dimensions. This is shown in Figure 14. The different colors in Fig. 14 are connected to times of the similarly colored data points in Fig. 16, which shows the shear modulus. The red bar denotes the estimate of Gplsubscript𝐺𝑝𝑙G_{pl}italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT. The time dependent variances of the Gaussian fits, denoted as b1(t)subscript𝑏1𝑡b_{1}(t)italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) are shown by the scatters in Figure 15 together with a fit (blue) through the data points where we consider the shear modulus to have plateaued. The agreement with a linear increase proves the diffusive propagation. The orange bar represents the theory prediction of the diffusive mode with a diffusion coefficient given by D=Gpl/(ζn)𝐷subscript𝐺𝑝𝑙𝜁𝑛D=G_{pl}/(\zeta n)italic_D = italic_G start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT / ( italic_ζ italic_n ).

Refer to caption
Figure 14: Fit of the form a1eq2b1subscript𝑎1superscript𝑒superscript𝑞2subscript𝑏1a_{1}\,e^{-q^{2}\,b_{1}}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT to q~(4)(q,t)superscript~𝑞4𝑞𝑡\tilde{q}^{(4)}(q,t)over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ( 4 ) end_POSTSUPERSCRIPT ( italic_q , italic_t ) where different colors denote different times. Via Eq. (41) we can connect a1(t)subscript𝑎1𝑡a_{1}(t)italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) to the shear modulus (cf. Fig. 16) and b1(t)=Dtsubscript𝑏1𝑡𝐷𝑡b_{1}(t)=D\,titalic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) = italic_D italic_t to diffusive dynamics (cf. Fig. 15).
Refer to caption
Figure 15: Crosses denote fit parameters b1(t)subscript𝑏1𝑡b_{1}(t)italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) from Figure 14 to every slope. The blue dashed line denotes a linear Fit with an offset b2subscript𝑏2b_{2}italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT accounting for non-diffusive short time effects. The orange dashed line denotes the theory prediction Eq.(41), where the non-diffusive parameter b2subscript𝑏2b_{2}italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT was added.
Refer to caption
Figure 16: From Eq. (41) we can connect the amplitudes of the Gaussian fits in Fig. 14 to the shear modulus. The black dashed line is calculated in the conventional way of correlating the stresses in the whole simulation box following Eq. (5).

VII Discussion

In agreement with predictions from non-Markovian Langevin equations approaches derived using Zwanzig-Mori projection operators [14, 15, 19], we could observe the emergence of long-ranged stress correlations in simulations of dense hard sphere liquids. These precursors of the Hookean elastic solid, which build up continuously in the supercooled liquid, result from the conservation law of momentum transport. While in a fluid, shear momentum diffuses, in the elastic solid it propagates via elastic waves. This crossover affects all elements of the tensor of stress correlations, which additionally contains the complexity that the long-time states have to be force free. In Brownian fluids, the analogous crossover is observed, with the difference that the stresses that originate from the interparticle interactions diffuse in the solid state. We used the formalism of spherical tensors introduced in [17] for mechanically rigid and time-stationary (inherent) states to analyze the stress correlations. This formalism proved itself also helpful in the time depended analysis of the fluid as performed in this contribution.

Within the size of the simulation boxes in two and three dimensions convincing power-law correlations are observed as shown in Figs. 1 and 10. Exponential screening as proposed in [11] was not observed within our scope. Ref. [10] reports large non-zero average forces, which we do not find. Rather our finding of spatio-temporal relaxation to a homogeneous force-free state in the liquid seems to align very well with the predictions of the stable, force-free system as considered in the IS approach. The necessary generalization to the viscoelastic liquid state is provided by the non-Markovian Langevin approach following Zwanzig and Mori.

The question about possible non-local contributions in the shear viscosity has been raised. While appreciable wavevector dependences have been observed in simulations of supercooled fluids, including of polymeric ones [7, 8, 9, 10], we find the (generalized) shear modulus and its viscous relaxation to be q𝑞qitalic_q-independent to a very good approximation. This difference warrants additional studies.

Refer to caption
Figure 17: Confocal fluorescence images of soft alginate force probe particles. Stress fields in the surrounding viscoelastic fluid lead to a deformation of the soft particles that could be detected as position changes of the fluorescent markers inside probe particle. Left: Scan in a single plane, right: projection of all image planes. Scale bars: 30 μm𝜇𝑚\mu mitalic_μ italic_m

To test our predictions experimentally, we prepared probe particles serving as sensors for shear stress. The particles consist of large, deformable alginate microgel droplets that contain hundreds of small fluorescent polymer beads (Fig. 17). The diameter of the microgel shell is 100absent100\approx 100≈ 100μ𝜇\muitalic_μm, whereas the fluorescent beads have diameters of 200200200200 nm. Particles of this type have previously been used as force sensors in biologial tissues [40, 41]. For the experiments, probe particles immersed in a dispersion of hard spheres allow recording the 3D positions of the fluorescent beads using confocal fluorescence microscopy. Spatial positioning accuracies of 20 nm in xy and 50 nm in the z-direction are commonly achieved and can be recorded over mesoscopic distances. Shear stress in the fluid leads to a deformation of the probe particles which can be analyzed as a change in the position of the fluorescent beads. The position measurements will be used to calculate shear stress auto-correlations. So far, however, the sensitivity of the measurements has proven to not be high enough to capture the predicted fluctuations. Future work in this direction will aim at increasing the softness of the microgel shells by reducing its degree of crosslinking to improve the sensitivity of the probe particles.

Acknowledgements.
The authors thank Annette Zippelius and Jörg Baschnagel for helpful discussions. MF thanks Peter Daivis for clarifying discussions and him and his colleagues from RMIT, Melbourne, for their hospitality during a stay when this paper was written. The work was supported by the Deutsche Forschungsgemeinschaft (DFG) via SFB 1432 project C07.

Appendix A Explicit expressions of the basis tensors

A.1 Three dimensions

The basis orthogonal tesseral basis tensors Q¯¯(i)superscript¯¯Q𝑖\underline{\underline{\mathrm{Q}}}^{(i)}under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT are linear combinations of the following tensors C¯¯(i)superscript¯¯C𝑖\underline{\underline{\mathrm{C}}}^{(i)}under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT

C¯¯1=δ¯¯δ¯¯subscript¯¯C1¯¯𝛿¯¯𝛿\displaystyle\underline{\underline{\mathrm{C}}}_{1}=\underline{\underline{% \delta}}\;\underline{\underline{\delta}}under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = under¯ start_ARG under¯ start_ARG italic_δ end_ARG end_ARG under¯ start_ARG under¯ start_ARG italic_δ end_ARG end_ARG (42a)
C¯¯2=δαδδγβ+δαγδβδsubscript¯¯C2subscript𝛿𝛼𝛿subscript𝛿𝛾𝛽subscript𝛿𝛼𝛾subscript𝛿𝛽𝛿\displaystyle\underline{\underline{\mathrm{C}}}_{2}=\delta_{\alpha\delta}% \delta_{\gamma\beta}+\delta_{\alpha\gamma}\delta_{\beta\delta}under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_α italic_δ end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT italic_γ italic_β end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT italic_α italic_γ end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT italic_β italic_δ end_POSTSUBSCRIPT (42b)
C¯¯3=δ¯¯𝒓^𝒓^+𝒓^𝒓^δ¯¯subscript¯¯C3¯¯𝛿bold-^𝒓bold-^𝒓bold-^𝒓bold-^𝒓¯¯𝛿\displaystyle\underline{\underline{\mathrm{C}}}_{3}=\underline{\underline{% \delta}}\,\bm{\hat{r}}\,\bm{\hat{r}}+\bm{\hat{r}}\,\bm{\hat{r}}\,\underline{% \underline{\delta}}under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = under¯ start_ARG under¯ start_ARG italic_δ end_ARG end_ARG overbold_^ start_ARG bold_italic_r end_ARG overbold_^ start_ARG bold_italic_r end_ARG + overbold_^ start_ARG bold_italic_r end_ARG overbold_^ start_ARG bold_italic_r end_ARG under¯ start_ARG under¯ start_ARG italic_δ end_ARG end_ARG (42c)
C¯¯4=δαδr^βr^γ+δαγr^βr^δ+δβδr^αr^γ+δβγr^αr^δsubscript¯¯C4subscript𝛿𝛼𝛿subscript^𝑟𝛽subscript^𝑟𝛾subscript𝛿𝛼𝛾subscript^𝑟𝛽subscript^𝑟𝛿subscript𝛿𝛽𝛿subscript^𝑟𝛼subscript^𝑟𝛾subscript𝛿𝛽𝛾subscript^𝑟𝛼subscript^𝑟𝛿\displaystyle\underline{\underline{\mathrm{C}}}_{4}=\delta_{\alpha\delta}\,% \hat{r}_{\beta}\hat{r}_{\gamma}+\delta_{\alpha\gamma}\hat{r}_{\beta}\hat{r}_{% \delta}+\delta_{\beta\delta}\hat{r}_{\alpha}\hat{r}_{\gamma}+\delta_{\beta% \gamma}{\hat{r}}_{\alpha}{\hat{r}}_{\delta}under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_α italic_δ end_POSTSUBSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT italic_α italic_γ end_POSTSUBSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT italic_β italic_δ end_POSTSUBSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT italic_β italic_γ end_POSTSUBSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT (42d)
C¯¯5=𝒓^𝒓^𝒓^𝒓^subscript¯¯C5bold-^𝒓bold-^𝒓bold-^𝒓bold-^𝒓\displaystyle\underline{\underline{\mathrm{C}}}_{5}=\bm{\hat{r}}\,\bm{\hat{r}}% \,\bm{\hat{r}}\,\bm{\hat{r}}under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT = overbold_^ start_ARG bold_italic_r end_ARG overbold_^ start_ARG bold_italic_r end_ARG overbold_^ start_ARG bold_italic_r end_ARG overbold_^ start_ARG bold_italic_r end_ARG (42e)

connected by

(Q¯¯1Q¯¯2Q¯¯3Q¯¯4Q¯¯5)=(1300001301200160120320001222122122122122122)(C¯¯1C¯¯2C¯¯3C¯¯4C¯¯5).matrixsubscript¯¯Q1subscript¯¯Q2subscript¯¯Q3subscript¯¯Q4subscript¯¯Q5matrix1300001301200160120320001222122122122122122matrixsubscript¯¯C1subscript¯¯C2subscript¯¯C3subscript¯¯C4subscript¯¯C5\displaystyle\begin{pmatrix}\underline{\underline{\mathrm{Q}}}_{1}\\ \underline{\underline{\mathrm{Q}}}_{2}\\ \underline{\underline{\mathrm{Q}}}_{3}\\ \underline{\underline{\mathrm{Q}}}_{4}\\ \underline{\underline{\mathrm{Q}}}_{5}\\ \end{pmatrix}=\begin{pmatrix}\frac{1}{3}&0&0&0&0\\ \frac{1}{3}&0&-\frac{1}{2}&0&0\\ \frac{1}{6}&0&-\frac{1}{2}&0&\frac{3}{2}\\ 0&0&0&\frac{1}{2\sqrt{2}}&-\sqrt{2}\\ -\frac{1}{2\sqrt{2}}&\frac{1}{2\sqrt{2}}&\frac{1}{2\sqrt{2}}&-\frac{1}{2\sqrt{% 2}}&\frac{1}{2\sqrt{2}}\end{pmatrix}\begin{pmatrix}\underline{\underline{% \mathrm{C}}}_{1}\\ \underline{\underline{\mathrm{C}}}_{2}\\ \underline{\underline{\mathrm{C}}}_{3}\\ \underline{\underline{\mathrm{C}}}_{4}\\ \underline{\underline{\mathrm{C}}}_{5}\\ \end{pmatrix}.( start_ARG start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) = ( start_ARG start_ROW start_CELL divide start_ARG 1 end_ARG start_ARG 3 end_ARG end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL divide start_ARG 1 end_ARG start_ARG 3 end_ARG end_CELL start_CELL 0 end_CELL start_CELL - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL divide start_ARG 1 end_ARG start_ARG 6 end_ARG end_CELL start_CELL 0 end_CELL start_CELL - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_CELL start_CELL 0 end_CELL start_CELL divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG end_CELL start_CELL - square-root start_ARG 2 end_ARG end_CELL end_ROW start_ROW start_CELL - divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG end_CELL start_CELL divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG end_CELL start_CELL divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG end_CELL start_CELL - divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG end_CELL start_CELL divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG 2 end_ARG end_ARG end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) . (43)

Orthogonality yields Q¯¯i:Q¯¯j=δij:subscript¯¯Q𝑖subscript¯¯Q𝑗subscript𝛿𝑖𝑗\underline{\underline{\mathrm{Q}}}_{i}:\underline{\underline{\mathrm{Q}}}_{j}=% \delta_{ij}under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT : under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT simplifying the access to the isotropic functions by simple projections qr,t(i)=𝒞¯¯(𝒓,t):Q¯¯i(r^):subscriptsuperscript𝑞𝑖𝑟𝑡¯¯𝒞𝒓𝑡subscript¯¯Q𝑖^𝑟q^{(i)}_{r,t}=\underline{\underline{\mathcal{C}}}(\bm{r},t):\underline{% \underline{\mathrm{Q}}}_{i}(\hat{r})italic_q start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r , italic_t end_POSTSUBSCRIPT = under¯ start_ARG under¯ start_ARG caligraphic_C end_ARG end_ARG ( bold_italic_r , italic_t ) : under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( over^ start_ARG italic_r end_ARG ), which as well holds in reciprocal space.

A.2 Two dimensions

In two dimensions some tensors in Eqs. (42) become dependent by the relation

C¯¯2C¯¯42C¯¯1+2C¯¯3=0,subscript¯¯C2subscript¯¯C42subscript¯¯C12subscript¯¯C30\displaystyle\underline{\underline{\mathrm{C}}}_{2}-\underline{\underline{% \mathrm{C}}}_{4}-2\underline{\underline{\mathrm{C}}}_{1}+2\underline{% \underline{\mathrm{C}}}_{3}=0,under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT - 2 under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 2 under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0 , (44)

making it possible to use only four of the tensors above in two dimensions. The basis chosen in Ref. [31] reads

(Q¯¯1Q¯¯2Q¯¯3Q¯¯4)=(1/20001/201/201/201211212)(C¯¯1C¯¯2C¯¯3C¯¯5),matrixsubscript¯¯Q1subscript¯¯Q2subscript¯¯Q3subscript¯¯Q4matrix120001201201201211212matrixsubscript¯¯C1subscript¯¯C2subscript¯¯C3subscript¯¯C5\displaystyle\begin{pmatrix}{\underline{\underline{\mathrm{Q}}}}_{1}\\ {\underline{\underline{\mathrm{Q}}}}_{2}\\ {\underline{\underline{\mathrm{Q}}}}_{3}\\ {\underline{\underline{\mathrm{Q}}}}_{4}\end{pmatrix}=\begin{pmatrix}1/2&0&0&0% \\ 1/\sqrt{2}&0&-1/\sqrt{2}&0\\ 1/2&0&-1&2\\ -1&\frac{1}{2}&1&-2\end{pmatrix}\begin{pmatrix}\underline{\underline{\mathrm{C% }}}_{1}\\ \underline{\underline{\mathrm{C}}}_{2}\\ \underline{\underline{\mathrm{C}}}_{3}\\ \underline{\underline{\mathrm{C}}}_{5}\end{pmatrix},( start_ARG start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_Q end_ARG end_ARG start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) = ( start_ARG start_ROW start_CELL 1 / 2 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 1 / square-root start_ARG 2 end_ARG end_CELL start_CELL 0 end_CELL start_CELL - 1 / square-root start_ARG 2 end_ARG end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 1 / 2 end_CELL start_CELL 0 end_CELL start_CELL - 1 end_CELL start_CELL 2 end_CELL end_ROW start_ROW start_CELL - 1 end_CELL start_CELL divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_CELL start_CELL 1 end_CELL start_CELL - 2 end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) , (45)

with the two dimensional versions C¯¯isubscript¯¯C𝑖\underline{\underline{\mathrm{C}}}_{i}under¯ start_ARG under¯ start_ARG roman_C end_ARG end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of Eqs. (42).

Appendix B Determination of sound velocities

To determine the sound velocity it was most convenient to measure the dispersion relation of the sound waves for a number of wave vectors. Therefore longitudinal velocity correlations were measured for three different wave-vectors. The frequency was determined by a damped harmonic oscillator fit, which then yields the dispersion relation. The sound velocity can then be read off the gradient of a linear fit. Figure 18 shows the dispersion relation and the corresponding linear fit for two (blue) and three (orange) dimensions. The so determined sound velocities are noted in the figure.

Refer to caption
Figure 18: Newtonian system in two dimensions, dispersion relation. A linear fit gives the sound velocity cLsubscript𝑐𝐿c_{L}italic_c start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT. Via cL=Kplnsubscript𝑐𝐿subscript𝐾𝑝𝑙𝑛c_{L}=\sqrt{\frac{K_{pl}}{n}}italic_c start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG italic_K start_POSTSUBSCRIPT italic_p italic_l end_POSTSUBSCRIPT end_ARG start_ARG italic_n end_ARG end_ARG one can determine the compression modulus or vice versa.

References

  • Binder and Kob [2011] K. Binder and W. Kob, Glassy Materials and Disordered Solids, revised ed. (WORLD SCIENTIFIC, 2011).
  • Wagner and Mewis [2021] N. Wagner and J. Mewis, Theory and Applications of Colloidal Suspension Rheology, Cambridge series in chemical engineering (Cambridge University Press, 2021).
  • Larson [1999] R. G. Larson, The structure and rheology of complex fluids (Oxford University Press, New York, 1999).
  • Nicolas et al. [2018] A. Nicolas, E. E. Ferrero, K. Martens, and J.-L. Barrat, Deformation and flow of amorphous solids: Insights from elastoplastic models, Rev. Mod. Phys 90, 045006 (2018).
  • Eshelby [1957] J. D. Eshelby, The determination of the elastic field of an ellipsoidal inclusion, and related problems, Poceedings of the Royal Society A 24110.1098/rspa.1957.0133 (1957).
  • Picard et al. [2004] G. Picard, A. Ajdari, F. Lequeux, and L. Bocquet, Elastic consequences of a single plastic event: A step towards the microscopic modeling of the flow of yield stress fluids, The European Physical Journal E 10.1140/epje/i2004-10054-8 (2004).
  • Todd and Daivis [2007] B. D. Todd and P. J. Daivis, Homogeneous non-equilibrium molecular dynamics simulations of viscous flow: techniques and applications, Mol. Sim. 33, 189 (2007).
  • Puscasu et al. [2010] R. M. Puscasu, B. D. Todd, P. J. Daivis, and J. S. Hansen, Nonlocal viscosity of polymer melts approaching their glassy state, J. Chem. Phys. 133, 144907 (2010).
  • Furukawa and Tanaka [2009] A. Furukawa and H. Tanaka, Nonlocal nature of the viscous transport in supercooled liquids: Complex fluid approach to supercooled liquids, Phys. Rev. Lett. 103, 135703 (2009).
  • Tong et al. [2020] H. Tong, S. Sengupta, and H. Tanaka, Emergent solidity of amorphous materials as a consequence of mechanical self-organisation, Nature Communications 10.1038/s41467-020-18663-7 (2020).
  • Dyre [2024] J. C. Dyre, Solid-that-flows picture of glass-forming liquids, J. Phys. Chem. Lett. 15, 1603–1617 (2024).
  • Szamel and Flenner [2011] G. Szamel and E. Flenner, Emergence of long-range correlations and rigidity at the dynamic glass transition, Phys. Rev. Lett. 107, 105505 (2011).
  • Flenner and Szamel [2015] E. Flenner and G. Szamel, Long-range spatial correlations of particle displacements and the emergence of elasticity, Phys. Rev. Lett. 114, 025501 (2015).
  • Maier et al. [2017] M. Maier, A. Zippelius, and M. Fuchs, Emergence of long-ranged stress correlations at the liquid to glass transition, Phys. Rev. Lett. 119, 265701 (2017).
  • Maier et al. [2018] M. Maier, A. Zippelius, and M. Fuchs, Stress auto-correlation tensor in glass-forming isothermal fluids: From viscous to elastic response, The Journal of Chemical Physics 149, 084502 (2018).
  • Steffen et al. [2022] D. Steffen, L. Schneider, M. Müller, and J. Rottler, Molecular simulations and hydrodynamic theory of nonlocal shear-stress correlations in supercooled fluids, The Journal of Chemical Physics 157, 064501 (2022).
  • Lemaître [2015] A. Lemaître, Tensorial analysis of eshelby stresses in 3d supercooled liquids, The Journal of Chemical Physics 143, 164515 (2015).
  • Klochko et al. [2018] L. Klochko, J. Baschnagel, J. P. Wittmer, and A. N. Semenov, Long-range stress correlations in viscoelastic and glass-forming fluids, Soft Matter 14, 6835–6848 (2018).
  • Vogel et al. [2019] F. Vogel, A. Zippelius, and M. Fuchs, Emergence of goldstone excitations in stress correlations of glass-forming colloidal dispersions, EPL 12510.1209/0295-5075/125/68003 (2019).
  • Lin and Cohen [2016] N. Y. C. Lin and I. Cohen, Relating microstructure and particle-level stress in colloidal crystals under increased confinement, Soft Matter 12, 9058 (2016).
  • Scala et al. [2007] A. Scala, T. Voigtmann, and C. De Michele, Event-driven Brownian dynamics for hard spheres, The Journal of Chemical Physics 126, 134109 (2007).
  • Götze [2009] W. Götze, Complex dynamics of glass-forming liquids: A mode-coupling theory, Vol. 143 (Oxford University Press, 2009).
  • Weysser and Hajnal [2011] F. Weysser and D. Hajnal, Tests of mode-coupling theory in two dimensions, Phys. Rev. E 83, 041503 (2011).
  • Fritschi and Fuchs [2017] S. Fritschi and M. Fuchs, Elastic moduli of a brownian colloidal glass former, J. Phys.: Condens. Matter 30, 024003 (2017).
  • van Megen et al. [1991] W. van Megen, S. M. Underwood, and P. N. Pusey, Nonergodicity parameters of colloidal glasses, Phys. Rev. Lett. 67, 1586 (1991).
  • Brambilla et al. [2009] G. Brambilla, D. El Masri, M. Pierno, L. Berthier, L. Cipelletti, G. Petekidis, and A. B. Schofield, Probing the equilibrium dynamics of colloidal hard spheres above the mode-coupling glass transition, Phys. Rev. Lett. 102, 085703 (2009).
  • Reinhardt et al. [2010] J. Reinhardt, F. Weysser, and M. Fuchs, Comment on “probing the equilibrium dynamics of colloidal hard spheres above the mode-coupling glass transition”, Phys. Rev. Lett. 105, 199604 (2010).
  • Lange et al. [2009] E. Lange, J. B. Caballero, A. M. Puertas, and M. Fuchs, Comparison of structure and transport properties of concentrated hard and soft sphere fluids, J. Chem. Phys. 130,, 174903 (2009).
  • Hansen and McDonald [2009] J. Hansen and J. McDonald, Theory of simple liquids, Vol. 3 (Elsevier Science B.V, 2009).
  • Dufty [2002] J. W. Dufty, Shear stress correlations in hard and soft sphere fluids, Mol. Phys. 100, 2331 (2002).
  • Lemaître [2017] A. Lemaître, Inherent stress correlations in a quiescent two-dimensional liquid: Static analysis including finite-size effects, Phys. Rev. E 96, 052101 (2017).
  • Wittmer et al. [2023] J. P. Wittmer, A. N. Semenov, and J. Baschnagel, Correlations of tensor field components in isotropic systems with an application to stress correlations in elastic bodies, Phys. Rev. E 108, 015002 (2023).
  • Martin et al. [1972] P. C. Martin, O. Parodi, and P. S. Pershan, Unified hydrodynamic theory for crystals, liquid crystals, and normal fluids, Phys. Rev. A 6, 2401 (1972).
  • L.D. Landau and E.M. Lifshitz [1986] L.D. Landau and E.M. Lifshitz, Theory of Elasticity, 3rd ed. (Pergamon Press, 1986).
  • Martin [1965] P. Martin, Non local transport coefficients correlation functions, in Statistical Mechanics of Equilibrium and Non-Equilibrium, edited by J. Meixner (North-Holland Publ., Amsterdam, 1965) p. 100.
  • Vogel and Fuchs [2020] F. Vogel and M. Fuchs, Stress correlation function and linear response of Brownian particles, Eur. Phys. J. E. 43, 70 (2020).
  • Evans [1981] D. Evans, Equilibrium fluctuation expressions for the wave-vector- and frequency-dependent shear viscosity, Phys. Rev. A 23, 2622 (1981).
  • Evans and Morriss [2008] D. J. Evans and G. Morriss, Statistical mechanics of nonequilibrium liquids (Cambridge University Press, London, 2008).
  • Lemaître [2018] A. Lemaître, Stress correlations in glasses, The Journal of Chemical Physics 149, 104107 (2018).
  • Campàs et al. [2014] O. Campàs, T. Mammoto, S. Hasso, R. A. Sperling, D. O’Connell, A. G. Bischof, R. Maas, D. A. Weitz, L. Mahadevan, and D. E. Ingber, Quantifying cell-generated mechanical forces within living embryonic tissues, Nat. Meth. 11, 183 (2014).
  • Mohagheghian et al. [2018] E. Mohagheghian, J. Luo, J. Chen, G. Chaudhary, J. Chen, J. Sun, R. H. Ewoldt, and N. Wang, Quantifying compressive forces between living cell layers and within tissues using elastic round microgels, Nat. Commun. 910.1038/s41467-018-04245-1 (2018).