<<

Shear modulus of simulated glass-forming model systems: Effects of boundary condition, temperature, and sampling time J. P. Wittmer, H. Xu, P. Polinska, F. Weysser, J. Baschnagel

To cite this version:

J. P. Wittmer, H. Xu, P. Polinska, F. Weysser, J. Baschnagel. Shear modulus of simulated glass- forming model systems: Effects of boundary condition, temperature, and sampling time. Journal of Chemical Physics, American Institute of Physics, 2013, 138 (12), pp. 12A533. ￿10.1063/1.4790137￿. ￿hal-01517066￿

HAL Id: hal-01517066 https://hal.univ-lorraine.fr/hal-01517066 Submitted on 11 May 2017

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. Shear modulus of simulated glass-forming model systems: Effects of boundary condition, temperature, and sampling time , J. P. Wittmer , H. Xu, P. , F. Weysser, and J. Baschnagel Polińska

Citation: The Journal of Chemical Physics 138, 12A533 (2013); doi: 10.1063/1.4790137 View online: http://dx.doi.org/10.1063/1.4790137 View Table of Contents: http://aip.scitation.org/toc/jcp/138/12 Published by the American Institute of Physics THE JOURNAL OF CHEMICAL PHYSICS 138, 12A533 (2013)

Shear modulus of simulated glass-forming model systems: Effects of boundary condition, temperature, and sampling time J. P. Wittmer,1,a) H. Xu,2 P. P o l i nska,´ 1 F. Weysser,1 and J. Baschnagel1 1Institut Charles Sadron, Université de Strasbourg & CNRS, 23 rue du Loess, 67034 Strasbourg Cedex, France 2LCP-A2MC, Institut Jean Barriol, Université de Lorraine & CNRS, 1 bd Arago, 57078 Metz Cedex 03, France (Received 6 December 2012; accepted 15 January 2013; published online 11 February 2013)

The shear modulus G of two glass-forming colloidal model systems in d = 3 and d = 2 dimensions is investigated by means of, respectively, molecular dynamics and Monte Carlo simulations. Compar- ing ensembles where either the shear strain γ or the conjugated (mean) τ are imposed, we compute G from the respective stress and strain fluctuations as a function of temperature T while keeping a constant normal P. The choice of the ensemble is seen to be highly relevant for the shear stress fluctuations μF(T) which at constant τ decay monotonously with T following the affine shear μA(T), i.e., a simple two-point correlation function. At variance, non-monotonous be- havior with a maximum at the temperature Tg is demonstrated for μF(T) at constant γ . The increase of G below Tg is reasonably fitted for both models by a continuous cusp singular- 1/2 ity, G(T ) ∝ (1 − T/Tg) , in qualitative agreement with recent theoretical predictions. It is argued, however, that longer sampling times may to a sharper transition. © 2013 American Institute of Physics.[http://dx.doi.org/10.1063/1.4790137]

I. INTRODUCTION ish for T → 0. This is due to the fact that the parti- cle displacements need not follow an imposed macroscopic Among the fundamental properties of any solid material strain affinely.14–17, 19, 23 The stress fluctuation term quanti- are its elastic constants.1–4 In their seminal work Squire, Holt fies the extent of these non-affine displacements, whereas and Hoover5 derived an expression for the isothermal elas- the Born term reflects the affine part of the particle dis- tic constants extending the classical Born theory1 to canoni- placements. How important the non-affine motions are, de- cal ensembles of imposed particle number N, volume V , and pends on the system under consideration.19 While the elastic finite (mean) temperature T. They found a correction term properties of crystals with one atom per unit cell are given Cαβγ δ to the Born expression for the elastic constants which F by the Born term only, stress fluctuations are significant for involves the mean-square fluctuations of the stress5–8 crystals with more complex unit cells.6 They become pro- αβγ δ ≡−  −   22 CF βV( σˆαβ σˆγδ σˆαβ σˆγδ )(1)nounced for polymer like soft materials and amorphous solids.8, 12, 14–19, 21, 23, 24 This is revealed by computational with β = 1/k T being the inverse temperature, k Boltz- B B studies of various glass formers,12 such as Lennard-Jones (LJ) mann’s constant,σ ˆ the instantaneous value of a component αβ mixtures,14–17 polymer glasses,8, 21 or silica melts.24 Many of of the stress tensor σ =σˆ , and greek letters denoting αβ αβ these studies concentrate on the mechanical properties deep in the spatial coordinates in d dimension. Perhaps due to the fact the glassy state, exploring for instance correlations between that the demonstration of the “stress fluctuation formalism” the non-affine displacement field and vibrational anomalies in Ref. 5 assumed explicitly a well-defined reference position of the glass (“Boson peak”),17, 24 the mechanical heterogene- and a displacement field for the particles, it has not been ap- ity at the nanoscale,21, 25 or the onset of molecular plastic- preciated that this approach is consistent (at least if the initial ity in the regime where the macroscopic is still pressure is properly taken into account) with the well-known elastic.26 pressure fluctuation formula for the compression modulus K Surprisingly, the temperature dependence of the elas- of isotropic fluids7–10 which has been derived several decades tic constants of glassy materials does not appear to have earlier, presumably for the first time in the late 1940s by been investigated extensively using the stress fluctuation Rowlinson.9 In any case, the stress fluctuation formalism pro- formalism.8, 12, 20 Following the pioneering work by Barrat vides a convenient route to calculating the elastic properties et al.,12 we present in this work such a study where we fo- in computer simulations which has been widely used in the cus on the behavior of the shear modulus G of two isotropic past.8, 11–22 It has been generalized to systems with nonzero colloidal glass-formers in d = 3 and d = 2 dimensions sam- initial stress,11 hard-sphere interactions,13 and arbitrary con- pled, respectively, by means of molecular dynamics (MD) and tinuous potentials,6 or to the calculation of local mechanical Monte Carlo (MC) simulations.7, 10, 27 Since we are interested properties.21 in isotropic systems one can avoid the inconvenient tensorial In particular from Ref. 6 it has become clear that notation of the full stress fluctuation formalism as shown in the stress fluctuation correction does not necessarily van- Fig. 1. Focusing on the shear in the xy-plane we use τ ≡ σ xy for the mean shear stress andτ ˆ ≡ σˆxy for its instantaneous a)Electronic mail: [email protected]. value. As an important contribution to the shear modulus we

0021-9606/2013/138(12)/12A533/18/$30.00138, 12A533-1 © 2013 American Institute of Physics 12A533-2 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

(G = 0) and flows on long time scales as shown by the P τ (a) (c) bottom curve. The shear modulus G can thus serve as an “order parameter” distinguishing and solid states.28–33 V=const τ Upon melting, the shear modulus of crystalline solids is y =const γ = tan (α) α known to vanish discontinuously with T. A natural ques- x tion which arises is that of the behavior of G(T) for amor- Ly phous solids and glasses in the vicinity of the glass transi- N=const tion temperature Tg. Interestingly, qualitative different theo- y T=const (d) P=const retical predictions have been put forward by mode-coupling x 33, 34 32 y γ =const theory (MCT) and some versions of replica theory pre- τ dicting a discontinuous jump at the glass transition whereas P x fixed wall other replica calculations30, 31 and a very recent study of the Lx disorder-assisted melting of amorphous solids focusing on the (b) 23 low−T limit (e) non-affine displacements suggest a continuous transition in G(t) P=const the static limit of large sampling times t. (See Ref. 31 for y τ=const liquid a topical discussion of the replica approach.) The numeri- 0 x cal characterization of G(T) for colloidal glass-former is thus log(t) of high current interest. Since the stress fluctuation formal- ism can readily be added to the standard simulation codes of FIG. 1. Setup, notations, and ensembles investigated numerically: (a) 12, 19, 35 Schematic experimental setup for probing the shear modulus G. (b) Expected colloidal glass-formers at imposed particle number N, shear modulus G(t) as a function of measurement time t at T  Tg (top box volume V , box shape (γ = 0) and temperature T, called  curve), T slightly below Tg (middle curve) and in the liquid limit for T Tg NVγ T-ensembles, this suggests the use of Eq. (3) to tackle (bottom curve). (c) We use periodic simulation boxes with deformable box this issue. Note that the latter relation should also be use- shapes where we impose τ = 0. The shear modulus G may be computed us- ing Eqs. (6) or (9) from the instantaneous strainsγ ˆ and stressesτ ˆ.(d)Using ful for wide range of related systems, beyond the scope of Eq. (3) G may also be computed using the affine shear elasticity μA and the the present work, including the glass transition of polymeric = shear stress fluctuations μF in ensembles of fixed strain (γ 0). (e) Com- ,8 colloidal hard sphere suspensions,36 colloidal gels,37 bined volumetric and shear strain fluctuations allowing the computation of 28 both compression modulus K and shear modulus G from the respective strain and self-assembled networks as observed experimentally, fluctuations. e.g., for bridged networks of telechelic polymers in water-oil emulsions.38 This begs the delicate question of whether the general shall monitor the shear stress fluctuations stress fluctuation formalism and, especially, Eq. (3) for the ≡− xyxy =  2≥ μF CF βV δτˆ 0(2)shear modulus only holds for elastic solids with a well-defined microscopic displacement field, as suggested in the early as a function of temperature T. Quite generally, the shear mod- literature,5–7 or if the approach may actually be also applied ulus G may then be computed according to the stress fluctua- through a solid-liquid transition up to the liquid state9, 22 with tion formula8, 12 μF → μA and thus G = Gττ → 0 as it should for a liquid. By reworking the theoretical arguments leading to Eq. (3) it will Gττ ≡ μA − μF, (3) be seen that this is in principal the case, at least if standard where μA stands for the so-called “affine shear elasticity” cor- thermodynamics can be assumed to apply. The latter point responding to an assumed affine response to an external shear is indeed not self-evident for the colloidal glasses we are in- strain γ . For pairwise interaction potentials it can be shown terested in where (i) some degrees of freedom are frozen on that the time scale probed and (ii) the measured response to an imposed infinitesimal shear τ may depend on the sampling μA ≡ μB − Pex ≥ 0, (4) time t, especially for temperatures T around Tg as shown in i.e., μA comprises the so-called “Born-Lamé coefficient” μB, Fig. 1(b). To be on the safer side, it thus appears to be appro- a simple moment of the first and the second derivative of priate to test numerically the applicability of the stress fluctu- the pair potential with respect to the distance of the inter- ation formula, Eq. (3), using the corresponding strain fluctua- 1, 8, 19 39, 40 acting particles, and the excess contribution Pex to the tion relations which are more fundamental on experimen- total pressure P. These notations will be properly defined tal grounds. in Sec. II B below where a simple derivation will be given. Panel (a) of Fig. 1 shows a schematic setup for probing Since Eq. (3) is not the only operational definition of the experimentally the shear modulus G of isotropic systems con- shear modulus discussed, an index has been added remind- fined between two rigid walls. We assume that the bottom wall ing that the shear stressτ ˆ is used for the determination is fixed and the particle number N, the (mean) normal pres- of G. sure P, and the (mean) temperature T are kept constant. In If an infinitesimal shear stress τ is applied at time linear response either a shear strain γ = tan (α) or a (mean) t = 0, an isotropic solid exhibits, quite generally, an elastic re- shear stress τ may be imposed. The thin solid line indicates sponse quantified in strength by a finite static (t-independent) the unstrained reference system with γ = 0 and τ = 0, the shear modulus (G > 0) as shown by the top curve in dashed line the deformed state. From both the mechanical Fig. 1(b), whereas a liquid has a vanishing modulus and the thermodynamical point of view, the shear modulus is 12A533-3 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013) defined as relations. A simple derivation of the stress fluctuation for-   mula, Eq. (3), for the shear modulus Gττ in constant-γ en- ≡ ∂τ(γ ) G  , (5) sembles is given in Sec. II B. Shear stress fluctuations μF in ∂γ → γ 0 ensembles with different boundary conditions are discussed = where the derivative may also be taken at τ 0. In practice, in Sec. II C. For constant-τ ensembles μF = μF|τ will be one obtains G by fitting a linear slope to the (γ , τ)-data at shown to reduce to the affine shear elasticity μA, i.e., the vanishing strain. Obviously, the noise level is normally large functional Gττ must vanish for all temperatures. As pointed for small strains. From the numerical point of view it is thus out recently,42 the truncation of the interaction potentials used more appropriate to work at imposed mean stress τ = 0, to in computational studies implies an impulsive correction for let the strain freely fluctuate and to sample the instantaneous the affine shear elasticity μA. This is summarized in Sec. II values (γ, ˆ τˆ). The shear modulus G is then simply obtained D. The numerical models used in this study are presented in by linear regression Sec. III. We turn then in Sec. IV to our main numerical re-   sults. Starting with the high-T liquid limit, Sec. IV A shows δγδˆ τˆ Gγτ ≡  , (6) that all our operational definitions correctly yield a vanish- δγˆ 2 τ=0 ing shear modulus. Moving to the opposite low-T limit we where the index indicates that both the instantaneous values present in Sec. IV B the elastic response for a topologically γˆ andτ ˆ are used. The associated dimensionless correlation fixed network of harmonic springs constructed from the “dy- coefficient 14, 15  namical matrix” of our 2D model glass system. As shown  δγδˆ τˆ  in Sec. IV C, the same behavior is found qualitatively for low- cγτ ≡   (7) T glassy bead systems. The temperature dependence of stress δγˆ 2δτˆ2 τ=0 fluctuations and shear moduli in different ensembles is then allows to determine the quality of the fit. We do not know discussed in Sec. IV D. As stated above, one central goal of whether Eqs. (6) and (7) have actually been used in a real this work is to determine the behavior of the shear modulus experiment, however since no difficulty arises to work in a around Tg. We conclude the review in Sec. V. Thermody- computer simulation at constant τ and to record (γ, ˆ τˆ), we namic relations and numerical findings related to the com- use this linear regression as our key operational definition of pression modulus K are regrouped in the Appendix. G which is used to judge the correctness of the stress fluctua- tion relation, Eq. (3), computed at constant γ . To avoid addi- II. THEORETICAL CONSIDERATIONS tional physics at the walls, we use of course periodic boundary conditions10 (the central image being sketched) as shown by A. Reminder of thermodynamic relations the panels on the right-hand side of Fig. 1. As discussed in We assume in the following that standard thermodynam- Sec. II C, it is possible to determine Gγτ at constant volume ics and thermostatistics40 can be applied and remind a few (NVτT-ensemble) as shown by panel (c) or constant normal useful relations. Let us denote an extensive variable by X and pressure (NPτT-ensemble) as shown by panel (e). The lat- its instantaneous value by Xˆ , the conjugated intensive vari- ter ensemble has the advantage that one can easily impose in able by I and its instantaneous value by Iˆ.43 With F(T, X) addition the same normal pressure P for all temperatures T being the free energy at temperature T and imposed X,thein- as it is justified for experimental reasons. If thermodynamics tensive variable I and the associated modulus M are given by holds, it is known (Sec. II A) that

δγδˆ τˆ| = kBT/V (8) ∂F(T,X) τ I ≡ , (11) should hold.41 Using Eq. (6) this implies a second operational ∂X definition  k T/V ∂I ∂2F (T,X) G ≡ B  , (9) M ≡ V = V . (12) γγ  2  ∂X ∂X2 δγˆ τ=0 = which has the advantage that only the instantaneous shear In our case, the extensive variable is X Vγ, the intensive = = strainγ ˆ has to be recorded. The main technical point put for- variable I τ (Ref. 44) and the modulus M G.Weremind ward in this review is that all operational definitions yield sim- that there is an important difference between extensive and in- ˆ ilar results tensive variables: If the extensive variable X is fixed, X cannot   fluctuate. If, on the other hand, the intensive variable I =Iˆ   ≈ | Gγτ τ=0 Gγγ τ=0 Gττ γ =0 , (10) is imposed, not only Xˆ but also Iˆ may fluctuate. In our case ≡ even for temperatures where a strong dependence on the sam- this means, thatγ ˆ γ for the NVγ T-ensemble, while the pling time t is observed. In other words, Eq. (10) states that shear stress fluctuations μF defined above do clearly not van- this time dependence applies similarly (albeit perhaps not ex- ish in the NVτT-ensemble shown in Fig. 1(d). Assuming I to 40 actly) to the various moments computed for the three opera- be imposed, it is well known that G tional definitions of . δXˆ (βδIˆ)|I = 1, (13) The is organized as follows: Several theoretical issues are regrouped in Sec. II. We begin the discussion in ∂X V δXˆ 2| = = (14) Sec. II A by reminding a few useful general thermodynamic I ∂(βI) βM 12A533-4 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013) with β being the inverse temperature. For our variables B. Shear modulus at imposed shear strain X = Vγ and I = τ,Eq.(13) implies Eq. (8) from which it As shown in Fig. 1(a), we demonstrate Eq. (3) from the is seen that Eq. (6) is consistent with Eq. (9). The latter rela- free energy change associated with an imposed arbitrarily tion is also obtained using Eq. (14). small macroscopic shear strain γ in the xy-plane assuming a As discussed in textbooks10, 40 a simple average A =Aˆ constant particle number N, a constant volume V , and a con- of some observable A does not depend on the chosen ensem- stant temperature T (NVγ T-ensemble). It is supposed that the ble, at least not if the system is large enough (N,V →∞). A total system Hamiltonian may be written as the sum of a ki- correlation function δAδˆ Bˆ  of two observables A and B may netic energy and a potential energy. Since for a plain shear differ, however, depending on whether X or I is imposed. As strain at constant volume the ideal free energy contribution demonstrated by Lebowitz, Percus, and Verlet in 1967,45 does not change, i.e., is irrelevant for τ and G, we may focus ∂(βI) ∂A ∂B on the excess free energy contribution δAδˆ Bˆ | =δAδˆ Bˆ | − . (15) X I ∂X ∂(βI) ∂(βI) Fex(T,γ) =−kBT ln(Zex(γ )) (22) If the extensive variable and at least one of the observ- ables are identical, the left-hand side of Eq. (15) vanishes: If due to the conservative interaction energy of the particles. The A = X and B = I, this yields Eq. (13).IfA = B = X,thisim- (excess) partition function Zex(0) of the unperturbed system at plies Eq. (14). More importantly, for A = B = I one obtains γ = 0 is the Boltzmann-weighted sum over all states s of the using Eq. (12) system which are accessible within the measurement time t. The argument (0) is a short-hand for the unstrained reference. =  ˆ2| −  ˆ2| M βV δI I βV δI X. (16) Following the derivation of the compression modulus K9 the If it is possible to probe the intensive variable fluctuations in partition function  both ensembles, this does in principle allow to determine the Z (γ ) = exp(−βU (γ )) (23) modulus M. (The latter equation might actually be taken as the ex s s definition of M in cases where the thermodynamic reasoning becomes dubious.) We emphasize that the modulus M is an of the sheared system is supposed to be the sum over the intrinsic property of the system and does not depend on the same states s, but with a different metric corresponding to ensemble, albeit it is determined using ensemble depending the macroscopic strain which changes the total interaction en- correlation functions. For a thermodynamically stable system, ergy Us(γ ) of state s and, hence, the weight of the sheared M > 0, Eq. (16) implies that configuration for the averages computed. This is the central hypothesis made. Interestingly, it is not necessary to specify  ˆ2|  ˆ2| δI I > δI X (17) explicitly the states of the unperturbed or perturbed system, and that both fluctuations must become similar in the limit e.g., it is irrelevant whether the particles are distinguishable where the modulus M becomes small. For the shear stress or not or whether they have a well-defined reference position. Assuming Eq. (22) we compute now the mean shear fluctuations μF,Eq.(16) corresponds to the transformation stress τ and the shear modulus G for a general interaction potential Us(γ ). We note for later convenience that μ | = μ | − G (18) F Vγ F τ ∂ ln(Z (γ )) Z (γ ) ex = ex , (24) and one thus expects μF to be larger for NVτT-systems than ∂γ Zex(γ ) for NVγ T-systems, while in the liquid limit where G → 0the imposed boundary conditions should not matter. We shall ver-   ∂2 ln(Z (γ )) Z (γ ) Z (γ ) 2 ify Eq. (18) numerically below. A glance at Eq. (3) suggests ex = ex − ex (25) 2 the identity ∂γ Zex(γ ) Zex(γ )

! for the derivatives of the free energy and μ | = μ . (19) F τ A  ∂Zex(γ ) − Assuming Eqs. (8) and (19), it follows for the correlation co- =− βU (γ ) e βUs (γ ), (26) ∂γ s efficient cγτ defined above that s   cγτ = Gγτ/μF|τ = G/μA. (20) 2  ∂ Zex(γ ) − = (βU (γ ))2 e βUs (γ ) If the measured (γ, ˆ τˆ) were indeed perfectly correlated, i.e., ∂γ2 s = = | = s cγτ 1, this implies G μA and thus μF Vγ 0. In fact, for  − − βUs (γ ) all situations studied here we always have (βUs (γ ))e (27) s cγτ < 1, thus G<μA, (21) for the derivatives of the excess partition function where a i.e., the affine shear elasticity sets only an upper bound to the prime denotes the derivative of a function f(x) with respect to shear modulus and a theory which only contains the affine its argument x.UsingEq.(26) and taking the limit γ → 0 one strain response must overpredict G. We show now directly verifies for the shear stress that that Eq. (3) and, hence, Eqs. (19), (20), and (21) indeed hold 1  this under quite general conditions. τ =τˆ withτ ˆ ≡ U (γ ) (28) V s γ =0 12A533-5 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013) defining the instantaneous shear stress.44 Note that the average normalized vector with α = x, y, . . . Stating only the small-γ taken is defined as limit for the second derivative of f(r(γ )) with respect to γ we  1 − note finally that ...= ...e βUs (0) (29)   Zex(0) ∂2f (r(γ ))   s = 2 − 2 2 lim r f (r) rf (r) nx ny using the weights of the unperturbed system. The shear stress γ →0 ∂γ2 thus measures the average change of the total interaction en- +rf (r) ny ny , (38) ergy Us(γ ) taken at γ = 0. The shear modulus G is obtained using in addition Eqs. (25) and (27) and taking finally the where we have dropped the argument (0) for the distance r of γ → 0 limit. Confirming thus the stress fluctuation formula the unperturbed reference system. Eq. (3), this yields We assume now that the interaction energy Us(γ )ofa configuration s is given by a pair interaction potential u(r) G = G ≡ μ − μ | (30) ττ A F γ =0 between the particles7, 10 with r(γ ) being the distance between with μA being the “affine shear elasticity” two particles, i.e.,  1 ≡  |  Us (γ ) = u(rl(γ )), (39) μA Us (γ ) γ =0 (31) V l mentioned above and μ as defined in Eq. (2). The compari- F where the index l labels the interaction between the particles i son of Eqs. (18) and (30) confirms Eq. (19). The affine shear and j with i < j. Using the general result Eq. (28) for the shear elasticity μ corresponds to the change (second derivative) A stress it follows using Eq. (36) that of the total energy which would be obtained if one actually strains affinely in a computer simulation a given state s with- 1  τˆ = r u (r ) n n , (40) out allowing the particles to relax their position. It is for this V l l x,l y,l l reason that the shear-stress fluctuation term μF ≥ 0 must oc- cur in Eq. (30) correcting the overprediction. where we have dropped the argument (0) in the γ → 0 limit. Up to now we have not used the coordinate transforma- We have thus rederived for the shear stress (α = x, β = y) tion (metric change) associated with an affine shear strain. the well-known Kirkwood expression for the general excess This is needed to obtain operational definitions for the instan- stress tensor10 taneous shear stressτ ˆ (and thus for τ and μF) and the affine  1 shear elasticity. As shown in Fig. 1(a), a pure shear strain σ ex = σˆ ex withσ ˆ ex ≡ r u (r )n n . (41) αβ αβ αβ V l l α,l β,l transforms the x-coordinate of a particle position or the dis- l tance between particles as The affine shear elasticity μA is obtained using Eqs. (31) and x(0) ⇒ x(γ ) = x(0) + γy(0) (32) (38). This yields = + ex leaving all other coordinates unchanged. The squared particle μA μB σyy (42) distance r2 thus transforms as with the first contribution μB being the so-called “Born-Lamé 2 2 2 2 2 r (0) ⇒ r (γ ) = r (0) + 2γx(0)y(0) + γ y (0), (33) coefficient”1 where the γ 2-term is required for the calculation of the shear    1 2 2 2 modulus below. This implies μB ≡ r u (rl) − rlu (rl) n n . (43) V l x,l y,l l dr2(γ ) = 2x(0)y(0) + 2γy2(0), (34) dγ It corresponds to the first term in Eq. (38). In this work we focus on isotropic materials under isotropic external loads. The normal excess stresses σ ex for d2r2(γ ) αα = 2y2(0) (35) all α are thus identical, i.e. dγ2 1  for, respectively, the first and the second derivative with re- σ ex =−! P = r u (r ) (44) αα ex dV l l spect to γ . Let us now consider an arbitrary function f(r(γ )). l Using Eq. (34) it is readily seen that its first derivative with ≡ − respect to γ may be written as with Pex P Pid being the excess pressure, P the total pressure, Pid = kBTρ the ideal gas pressure, and d the spa- ∂f (r(γ )) f (r) = (2x(0)y(0) + 2γy2(0)) (36) tial dimension. This implies that the affine shear elasticity is ∂γ 2r(0) given by μA = μB − Pex as already stated in the Introduction, Eq. (4). We note that for isotropic d-dimensional systems it can be shown22, 42 that the Born-Lamé coefficient can be sim- ⇒ rf (r) nx ny for γ → 0. (37) plified as In the second step we have introduced the components nx  = = 1 x(0)/r(0) and ny y(0)/r(0) of the normalized distance vec- μ = r2u (r )−r u (r ) . (45) B d(d + 2)V l l l l tor. More generally, we denote by nα the α-component of a l 12A533-6 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

The d-dependent prefactor stems from the assumed isotropy the particle and fx, i the x-coordinate of the force acting on the of the system and the mathematical formula particle. This contribution indeed vanishes on average as can be seen using Eq. (48), 2 1   (nαnβ ) = (1 + 2δαβ ) (46) d(d + 2) ∂yi yi fx,i=−kBT = 0, (50) 46 ∂xi (δαβ being the Kronecker symbol ) for the components of a unit vector in d dimensions pointing into arbitrary directions. which shows that τˆ=τ as it must. The stress fluctuations By comparing the excess pressure Pex,Eq.(44), with the un- may thus be expressed as derlined contribution to the Born coefficient in Eq. (45) this μ = βV(ˆτ − τ)2 implies F    β −1 = yi fx,i yj fx,j rlu (rl) = Pex/(d + 2). (47) V d(d + 2)V i j l   It is thus inconsistent to neglect the explicit excess pressure in 1 ∂yj fx,j =− yi , (51) Eq. (4) while keeping the second term of the Born-Lamé co- V ∂x i j i efficient μB. This approximation is only justified if the excess where we have used the integration by part, Eq. (49).Thisstep pressure Pex is negligible. requires that all particle positions are unconstrained and inde- pendent (generalized) coordinates. Note that this is possible C. Shear stress fluctuations in different ensembles at imposed τ, but cannot hold at fixed γ . Assuming then pair Being simple averages neither P , μ nor μ depend for interactions and using Eq. (40) one can reformulate Eq. (51) ex A B within a few lines. Without further approximation this yields sufficiently large systems on the chosen ensemble, i.e., irre- spective of whether γ or τ is imposed one expects to obtain 1  ∂f μ | =− y2 x,l , (52) the same values. As already reminded in Sec. II A,thisis F τ V l ∂x different for fluctuations in general.10 For instance, it is ob- l l viously pointless to use the shear-strain fluctuation formulae where the sum runs now over all interactions l between par- Gγτ or Gγγ in a constant-γ ensemble. A more interesting re- ticles i < j. xl and yl refer to components of the distance ≡ − sult is predicted if Gττ μA μF is computed in the “wrong” vector of the interaction l and fx, l to the x-component of the | = constant-τ ensemble. Since μF τ μA,Eq.(18), one actually central force f(rl) =−u (rl) between a particle pair at a dis- = expects to observe Gττ 0 for all temperatures T if our ther- tance rl. Note that in Eq. (52) the stress fluctuations stemming modynamic reasoning holds through the glass transition up to from different interactions are decoupled, i.e., μF|τ character- the liquid state. We will test numerically this non-trivial pre- izes the self- or two-point contributions of directly interacting diction in Sec. IV. beads. Since fx(r) =−u (r)x/r and, hence, Up to now we have only considered for simplicity of the ∂fx (r) presentation NVγ T- and NVτT-ensembles at constant vol- −y2 = (r2u (r) − r2u (r)) n2 n2 ∂x x y ume V . Since on general grounds pure deviatoric and dilata- − tional (volumetric) strains are decoupled (both strains com- ru (r) nx ny , (53) mute), one expects the observable Gττ also to be applicable this is identical to the affine shear elasticity μA derived at the in the NPγ T-ensemble shown in Fig. 1(d) and the observ- end of Sec. II B. We have thus confirmed Eq. (19). Interest- ables Gγτ and Gγγ to be applicable in the NPτT-ensemble as ingly, the argument can be turned around: sketched in Fig. 1(e). In the latter ensemble one also expects | = to find Gττ = 0 and μF|τ = μA. This will also be tested below. (1) The stress fluctuation contribution term μF τ μA is ob- We demonstrate now directly Eq. (19) using the general tained by the simple derivation given in this paragraph. mathematical identity (2) Using the general Legendre transformation, Eq. (18), this implies G = μ − μ for the NVγ T- and the  =−   ττ A F f (x)A(x) kBT A (x) (48) NPγ T-ensemble. with x being an unconstrained coordinate, A(x) some prop- As already stressed, our thermodynamic derivation of erty, f(x) ≡−u (x) a “force” with respect to some “energy” Eq. (3) is rather general, not using in particular a well-defined u(x ) and the average being Boltzmann weighted, i.e., ... displacement field for the particle positions. ∝ dx...e−βu(x).(ItisassumedinEq.(48) that u(x) decays sufficiently fast at the boundaries.) Following Zwanzig,47 the instantaneous shear stressτ ˆ is expressed by the Kirkwood for- D. Impulsive corrections for truncated potentials 10 mula, Eq. (40), by the alternative virial representation It is common practice in computational physics7, 10 to 1  truncate a pair interaction potential u(r), with r being the dis- τˆ = τ − yi fx,i. (49) V tance between two particles i and j, at a convenient cutoff rc. i While this truncation allows to reduce the number of interac- The second term, sometimes called the “inner virial”, stands tions to be computed, it introduces technical difficulties, e.g., for a sum over all particles i with yi being the y-coordinate of instabilities in the numerical solution of differential equations 12A533-7 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

10 as investigated especially for the MD method. Without re- with G˜ ττ being the uncorrected bare shear modulus. As we stricting much the generality of our results, we assume below shall see in Sec. IV A, the correction μB is of importance that for the precise determination of Gττ close to the glass tran- r sition where the shear modulus must vanish. An impulsive the potential scales as u(r, σ l) ≡ u(s) with the reduced dimensionless distance s = r/σ where σ characterizes correction has also to be taken into account for the compres- l l = the length scale of the interaction l and sion modulus K Kpp computed from the normal pressure r fluctuations in a constant volume ensemble as discussed in the same reduced cutoff sc = rc/σ l is set for all inter- actions l. the Appendix. Using the symmetry of isotropic systems, Eq. (A24), one verifies42 For monodisperse particles with constant diameter σ ,as 2 + d for the LJ potential u (s) = 4 (s−12 − s−6), the scaling vari- K = K˜ − μ (60) LJ pp pp d B able becomes s = r/σ and the reduced cutoff sc = rc/σ .The ˜ effect of introducing sc is to replace u(s) by the truncated po- with Kpp being the uncorrected compression modulus. tential We consider below mixtures and polydisperse systems where σ l may differ for each interaction l. For such poten- u (s) = u(s)H(s − s) (54) t c tials u(s), ut(s), and us(s) and their derivatives take in principal with H(s) being the Heaviside function.46 Even if Eq. (54) an explicit index l, i.e., one should write ul(s), ut, l(s), us, l(s), is taken by definition as the system Hamiltonian, impulsive and so on. For example one might wish to consider the gen- corrections at the cutoff have to be taken into account for the eralization of the monodisperse LJ potential to a mixture or polydisperse system with excess pressure Pex and other moments of the first derivatives 7 −12 −6 of the potential as may be seen considering the derivative of ul(s) = 4 l(s − s ) with s = r/σl, (61) the truncated potential where l and σ l are fixed for each interaction l. In practice, = − − − ut(s) u (s)H(sc s) u(s)δ(s sc). (55) each particle i may be characterized by an energy scale Ei and a “diameter” Di. The interaction parameters l(Ei, Ej) These corrections with respect to first derivatives can be 7 and σ l(Di, Dj) are then given in terms of specified functions avoided by considering a properly shifted potential of these properties.15 The extensively studied Kob-Andersen (KA) model for binary mixtures of beads of type A and B,49 us(s) = (u(s) − u(sc)) H(sc − s) (56) is a particular case of Eq. (61) with fixed interaction ranges = − since us(s) u (s)H(sc s). With this choice no impulsive σ AA, σ BB, and σ AB and energy parameters AA, BB, and AB corrections arise for moments of the instantaneous shear characterizing, respectively, AA-, BB-, and AB-contacts. The stressτ ˆ and of other components of the excess stress tensor expressions Eqs. (58), (59), and (60) remain valid for such ex 42 σˆαβ ,Eq.(41). Specifically, if the potential is shifted, all impul- potentials. sive corrections are avoided for the shear stress fluctuations μ ,Eq.(2). F III. ALGORITHMIC AND TECHNICAL ISSUES The standard shifting of a truncated potential is, however, insufficient in general for properties involving higher deriva- The computational data presented in this work have been tives of the potential.42 This is the case for the Born coefficient obtained using two well-known models of colloidal glass- 14, 15, 19, 49 μB,Eq.(45), which is required to compute the affine shear formers. We present first both models and the MD 7, 10, 27 elasticity μA,Eq.(4). The second derivative of the truncated and MC methods used to compute them in different and shifted potential being ensembles. We comment then on the quench protocol. The glass transition temperature T for both models is located via u (s) = u (s)H(s − s) − u (s)δ(s − s ), (57) g s c c dilatometry, as shown in Fig. 2. This is needed as a prerequi- site to our study of the elastic behavior in Sec. IV. As sketched this implies that the Born coefficient reads μB = μ˜ B − μB withμ ˜ being the bare Born coefficient and μ the impul- in Fig. 1, periodic boundary conditions are applied in all spa- B B = =··· sive correction which needs to be taken into account. The lat- tial directions with Lx Ly for the linear dimensions of ter correction may be readily computed numerically from the the d-dimensional simulation box. Time scales are measured = 2 1/210 configuration ensemble using42 in units of the LJ time τ LJ (mσ / ) for our MD simula- tions (m being the particle mass, σ the reference length scale, μB = lim hB(s) and the LJ energy scale) and in units of Monte Carlo Steps → − s sc (MCS) for our MC simulations. LJ units are used throughout with  this work ( = σ = m = 1) and Boltzmann’s constant kB is 1 h (s) ≡ s2u (s ) δ(s − s) (58) also set to unity. B d(d + 2)V l l l l The KA model49 for binary mixtures of LJ beads in = being a pair distribution function which is related to the stan- d 3 dimensions has been investigated by means of MD 7 dard radial pair distribution function g(r).48 Since G = μ simulation taking advantage of the public domain LAMMPS ττ B 35 implementation. We use N = nA + nB = 6912 beads per − Pex − μF,Eq.(58) implies in turn simulation box and molar fractions ca = nA/n = 0.8 and cb Gττ = G˜ ττ − μB (59) = nB/n = 0.2 for both types of beads A and B. Following 12A533-8 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

= 4 0.96 0.92 (1-0.04(T/T 4 -4 (teq 10 ) and various properties as the ones recorded in Ta- pLJ N=10 , P=2, τ=0 R=10 -5 = × 4 R=10 ble I are computed using a sampling time t 5 10 .Note T

-1)) -6 γ 0.94 g R=10 that in the NVγ T-ensemble the equations of motion are in-

-7 NP R=10 10 =0.26

g -8 tegrated using a velocity Verlet algorithm with a time step

T R=10 -7 = 0.92 R=10 for NPτT δt 0.01 and systems are kept at the imposed temperature final best value T using a Langevin thermostat10 with a friction coefficient

(T) 0.5. By redoing for several temperatures around and below ρ 0.90 log(1/ρ(T)) T the tempering and the production runs we have checked 0.92 (1-0.07(T/ g -0.1 KA P=1 that the presented results remain unchanged (within numeri- 0.88 cal accuracy) and that ageing effects are negligible. Averages T Tg=0.41 g -1 T )) are taken over four independent configurations. As one may -0.2 0.86 0.2 0.3 0.4 0.5 0.6 expect from the discussion in Sec. II C, very similar results 0.0 0.1 0.2 0.3 0.4 0.5 have been found for both ensembles for simple averages, for T instance for the affine shear elasticity μA, and for the shear modulus Gττ,Eq.(3). Unfortunately, we are yet unable to FIG. 2. ρ(T) and determination of glass transition temperature Tg. Inset: Plotting log (1/ρ)vs.Tfor the KA model as suggested in Ref. 8 re- present data for the KA model data at imposed shear stress veals a linear low-T (solid line) and a linear high-T regime which confirms τ (NVτT-, NPτT-ensembles). ≈ 49 Tg 0.41. Main panel: Number density ρ for pLJ systems at constant nor- A systematic comparison of constant-γ and constant-τ mal pressure P = 2 as a function of temperature T for different quench rates R in units of MCS. The crosses refer to a quench in the NPτT-ensemble where ensembles has been performed, however, by MC simulation the box shape is allowed to fluctuate at τ = 0. All other data refer to the of a specific case of the generalized LJ potential, Eq. (61), = NPγ T-ensemble with γ 0. The solid line and the dashed line are linear fits where all interaction energies are identical, l = , and the in- to the best density estimates (stars) obtained for, respectively, low and high 48 teraction range is set by the Lorentz rule σ l = (Di + Dj)/2 temperatures. By matching both lines one determines Tg ≈ 0.26. A similar value is obtained if log (1/ρ) is plotted as a function of T. with Di and Dj being the diameters of the interacting particles. Only the strictly two-dimensional (2D) case (d = 2) is consid-

Ref. 49 we set σ AA = 1.0σ, σ BB = 0.88σ, and σ AB = 0.8σ ered. Following Ref. 15, the bead diameters of this polydis- for the interaction range and AA = 1.0 , BB = 0.5 , and AB perse LJ (pLJ) model are uniformly distributed between 0.8σ = 1.5 for the LJ energy scales. Only data for the usual (re- and 1.2σ . For the examples reported in Sec. IV we have used = duced) cutoff sc = rc/σ l = 2.5 are presented. For the NPγ T- N 10 000 beads per box. We only present data for a reduced = = 1/6 ensemble we use the Nosé-Hoover barostat provided by the cutoff sc 2.0s0 with s0 2 being the the minimum of LAMMPS code (“fix npt command”)35 whichisusedtoim- the polydisperse LJ potential, Eq. (61). Various properties ob- pose a constant pressure P = 1 for all temperatures. Starting tained for the pLJ model are summarized in Table II. with an equilibrated configuration at Ti = 0.7 well above the For all ensembles considered here local MC moves are glass transition temperature Tg = 0.41 (as confirmed in the used (albeit not exclusively) where a particle is chosen ran- inset of Fig. 2), the system is slowly quenched in the NPγ T- domly and a displacement δr, uniformly distributed over a ensemble. The imposed mean temperature T(t) varies linearly disk of radius δrmax, from the current position r of the particle −5 as T(t) = Ti − Rt with R = 5 × 10 being the constant is attempted. The corresponding energy change δE is calcu- 4 quench rate. After a tempering time teq = 10 we compute lated and the move is accepted using the standard Metropo- 27 over a sampling time t = 5 × 104 various properties for the lis criterion. The maximum displacement distance δrmax is isobaric system at fixed temperature such as the moduli Gττ chosen such that the acceptance rate A remains reasonable, represented by the filled spheres in Fig. 10. Fixing only then i.e., 0.1 ≤ A ≤ 0.5.27 As may be seen from the inset in the volume (NVγ T-ensemble) the system is again tempered Fig. 3 where the acceptance rate A(T) is shown as a function

TABLE I. Some properties for the KA model at normal pressure P = 1 and shear stress τ = 0: the mean density ρ, the mean energy per volume eρ, the impulsive correction term μB (Sec. II D), the (corrected) Born-Lamé coefficient μB, the affine shear elasticity μA = μB − Pex, the shear modulus Gττ obtained in the NVγ T-ensemble, the hypervirial ηB and the affine dilatation elasticity ηA = 2Pid + ηB + Pex discussed in the Appendix and the compression modulus Kpp obtained for the NVγ T-ensemble. The affine elasticities μA and ηA are the natural scales for, respectively, the shear modulus G and the compression modulus K.

T ρ eρ μB μB μA Gττ ηB ηA Kpp

0.025 1.24 −9.42 0.80 52.1 51.15 21.9 86.2 87.2 82.4 0.100 1.20 −8.85 0.80 50.4 49.51 19.6 83.4 84.5 75.7 0.200 1.21 −8.57 0.78 48.3 47.52 17.3 80.0 81.2 67.7 0.250 1.20 −8.31 0.77 47.2 46.54 13.9 78.2 79.6 63.1 0.310 1.19 −7.99 0.74 45.7 45.10 11.8 75.8 77.2 56.9 0.350 1.18 −7.82 0.73 45.1 44.45 9.3 73.9 76.1 52.6 0.375 1.17 −7.67 0.72 44.6 43.96 6.1 73.9 75.3 50.5 0.400 1.17 −7.49 0.71 43.8 43.24 2.5 72.6 74.1 45.3 0.450 1.15 −7.07 0.68 42.0 41.48 0.03 69.6 71.1 37.9 0.800 1.00 −4.49 0.55 29.7 29.5 ≈0 49.4 51.2 16.6 12A533-9 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

TABLE II. Some properties of the pLJ model at P = 2andτ = 0: the mean density ρ, the mean energy per volume eρ, the impulsive correction term μB, the affine shear elasticity μA, the shear moduli Gγτ and Gγγ obtained in the NPτT-ensemble and Gττ obtained in the NVγ T-ensemble, the affine dilatation elasticity ηA, the compression modulus Kvp obtained in the NPτT-ensemble, and the Rowlinson formula Kpp for the NVγ T-ensemble. All data have been obtained for a sampling time t = 107 MCS.

T ρ eρ μB μA Gγτ Gγγ Gττ ηA Kvp Kpp

0.001 0.96 −2.70 0.31 33.9 15.6 15.7 15.5 71.7 70.8 70.5 0.010 0.96 −2.69 0.38 33.7 14.3 14.2 13.0 71.3 69.7 69.7 0.100 0.94 −2.60 0.39 32.4 10.9 10.9 10.3 68.8 61.2 61.4 0.200 0.93 −2.48 0.38 31.1 7.2 7.2 7.4 66.2 51.2 50.6 0.225 0.93 −2.45 0.38 30.6 4.6 4.6 3.7 65.3 47.9 46.4 0.250 0.92 −2.42 0.37 30.0 1.4 1.4 0.94 64.0 41.0 44.4 0.275 0.92 −2.38 0.37 29.7 0.6 0.6 0.05 63.4 40.0 39.4 0.300 0.91 −2.33 0.35 29.3 0.1 0.1 0.09 62.6 37.3 37.6 0.325 0.91 −2.29 0.39 28.9 ≈0 ≈0 ≈0 61.7 34.0 34.3 0.350 0.90 −2.24 0.33 28.3 ≈0 ≈0 ≈0 60.5 31.6 31.2 1.000 0.73 −1.23 0.22 17.0 ≈0 ≈0 ≈0 38.0 9.3 9.0 of temperature T, we find that temperatures below T = 0.1 mensurate with the simulation box. The random amplitudes are best sampled using δrmax ≈ 0.01, whereas δrmax ≈ 0.1is and phases of the waves are assumed to be uniformly dis- a good choice for the interesting temperature regime between tributed. Note that the maximum amplitude amax(q) of each T = 0.2 and T = 0.3 around the glass transition temperature wave is chosen inversely proportional to the length q =|q| of Tg = 0.26. See the main panel of Fig. 2 for the determination the wavevector in agreement with continuum theory. This al- of Tg. Various data sets are presented in Sec. IV for a fixed lows to keep the same acceptance rate for each global mode if value δrmax = 0.1 which allows a reasonable comparison of q  1. As one expects, the plane-wave displacement attempts the sampling time dependence for temperatures around the becomes inappropriate for large q where continuum elasticity glass transition. This value is not necessarily the best choice breaks down14, 15, 19 and, hence, the acceptance rate A(q) gets for tempering the system and for computing static equilibrium too large to be efficient. These global plane-wave moves have properties at the given temperature, especially below T ≈ 0.2. been added for the tempering of all presented configurations 7 Local MC jumps in the NVγ T-ensemble become obvi- below T = 0.3 with teq = 10 and for production runs over ously inefficient for relaxing large scale structural properties27 sampling times t = 107 for static properties for T < 0.2. for the lowest temperatures computed for the pLJ model Using again the Metropolis criterion for accepting a sug- (T ≤ 0.1). We have thus crosschecked and improved the val- gested change of the overall box volume V or of the shear ues obtained using only local MC moves by adding global strain γ , pure dilatational and pure shear strain fluctuations collective MC move attempts for all particles with longitu- are used to impose a mean normal pressure P = 2 and a dinal and (more importantly) transverse plane waves com- mean shear stress τ = 0. For a shear strain altering MC move one first chooses randomly a strain fluctuation δγ with |δγ| ≤ -4 δγ max. Assuming an imposed shear stress τ the suggested R=10 τ δ pLJ P=2, =0, rmax=0.01 -2.3 -5 δγ is accepted if R=10 T -6 γ R=10 -1)) g -7 NP ξ ≤ exp (−βδE + βδγτ) (62) -2.4 R=10 -8 R=10 -7 with ξ being a uniformly distributed random variable with R=10 for NPτT -2.5 final best value -2.62 (1-0.14(T/T 0 ≤ ξ<1. The energy change δE associated with the affine δ strain may be calculated using Eq. (33). Since τ = 0 is sup- rmax=0.01 y per bead e -2.6 1.0 g posed in this study, the last term in the exponential drops out. 0.8 0.6 NPγT The maximum shear strain step δγ max is fixed such that at the ener A -2.7 0.4 given temperature the acceptance rate remains reasonable.50 -1)) 0.2 δr =0.1 (T/T g max 8 =0.26 As described in Ref. 10, one similarly chooses a relative vol-

g 0.0 0.0 0.1 0.2 0.3 0.4 -2.8 T 0.5 ume change δV/V with V being the current volume. The vol- -2.62 (1-0.0 T 0.0 0.1 0.2 0.3 0.4 0.5 ume altering move is accepted at an imposed pressure P if T ξ ≤ exp (−βδE + N log(1 + δV/V ) − βδVP) , (63) FIG. 3. Energy per particle e (main panel) and acceptance rate A (inset) ob- where δE may be computed using Eq. (A16). The logarith- tained for pLJ systems quenched at constant pressure P = 2 at a given quench rate R as indicated. The crosses refer to a quench at imposed τ = 0. The solid mic contribution corresponds to the change of the transla- line and the dashed line are linear fits to the energy for, respectively, low and tional entropy. Using Eq. (62) after each MCS for local MC ≈ high temperatures confirming the value Tg 0.26 for the pLJ model. Only moves realizes the NVτT-ensemble shown in Fig. 1(c),us- local MC moves with one fixed (non-adaptive) maximum particle displace- ing Eq. (63) the NPγ T-ensemble shown in Fig. 1(d) and us- ment δrmax for all temperatures T have been used for the MC data shown in the main panel. The acceptance rate decreases thus with decreasing T as ing in turn both strain fluctuations the NPτT-ensemble in shown in the inset for δrmax = 0.01 (upper data) and δrmax = 0.1. Fig. 1(e). We remind that NVγ T- and NPγ T-ensembles are 12A533-10 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013) used to determine Gττ and NVτT- and NPτT-ensembles to ten down every 10 MCS and averaged using standard gliding 10 obtain Gγτ and Gγγ. As shown in Fig. 2, we quench the averages, i.e., we compute mean values and fluctuations for configurations starting from Ti = 0.5intheNPγ T- and the a given time interval [t0, t1 = t0 + t] and average over all pos- NPτT-ensemble using a constant quench rate R. Note that sible intervals of length t. The horizontal axis indicates the the smallest quench rate R = 10−8 for one configuration in latter interval length t in units of MCS. The vertical axis is the NPγ T-shown did require alone a run over six months made dimensionless by rescaling the moduli with the (cor- using one core of an Intel Xeon E5410 processor. Interest- rected) affine shear elasticity μA = μB − Pex ≈ 17 from ingly, similar results are obtained with the NPτT-ensemble Table II. (Being a simple average, μA is found to be identi- only using a rate R = 10−7. The configurations created by the cal for all computed ensembles.) NPγ T-quenches are used (after tempering) for the sampling In agreement with Eq. (21), the ratio G(t)/μA for all oper- in the NVγ T- and NPγ T-ensembles, the configurations ob- ational definitions drops rapidly below unity. The dashed line tained by the NPτT-quenches for the sampling in the NVτT- indicates the asymptotic power-law slope −1.22, 42 This can be and NPτT-ensembles. Only two independent configurations understood by noting that at imposed τ the shear strain freely have been sampled following the full protocol for each tem- diffuses in the liquid limit, i.e., δγˆ 2∼t, which implies perature and each of the four ensembles. Instead of increasing that G ≈ Gγτ ∼ 1/t. For both NVτT- and NPτT-ensembles further the number of configurations we have focused in the we compare our key definition Gγτ,Eq.(6), to the observ- present study on long tempering times and production runs able Gγγ,Eq.(9). In both cases we confirm Gγτ(t) ≈ Gγγ(t) over t = 107 MCS which have been redone several times for for larger times as suggested by the thermodynamic relation a few temperatures to check for equilibration problems and Eq. (8) which has been directly checked.41 We have also com- to verify that ageing effects can be ignored for the properties puted the dimensionless correlation coefficient cγτ(t), Eq. (7), considered in this work. which characterize the correlation of the measured instanta- neous shear strainsγ ˆ and shear stressesτ ˆ. As indicated for the NPτT-ensemble (stars), c vanishes rapidly with t, i.e., IV. COMPUTATIONAL RESULTS γτ γˆ andτ ˆ are decorrelated. The dashed-dotted line indicates the A. High-temperature liquid limit power-law exponent −1/2 expected from Eq. (20) and Gγτ(t) ∼ 1/t. Note that by plotting c (t)vs.G (t)/μ ,Eq.(20) can We begin our discussion by focusing on the high tem- γτ γτ A be directly demonstrated (not shown). perature liquid limit where all reasonable operational defini- The stress fluctuation formula, Eq. (3), is represented in tions of the shear modulus G must vanish. The results pre- Fig. 4 by squares and for, respectively, NVγ T- and sented in Fig. 4 have been obtained for the pLJ model at mean NPγ T-ensembles. The bare shear modulus G˜ without the pressure P = 2, mean shear stress τ = 0 and mean temper- ττ impulsive correction μ for the Born coefficient μ dis- ature T = 1, i.e., far above the glass transition temperature B B cussed in Sec. II D is given by small filled symbols. The T ≈ 0.26. Only local MC monomer displacements with a g impulsive correction μ ≈ 0.2 (Table II) computed inde- maximum jump displacement δr = 0.1 are reported here. B max pendently from Eq. (58) is shown by a horizontal line. The Instantaneous properties relevant for the moments are writ- corrected moduli Gττ = G˜ ττ − μB vanish indeed as ex- pected. See Ref. 42 for a systematic characterization of μB 0 4 10 pLJ: T=1, N=10 , ρ=0.727, s =2.0s , μ =17 as a function of the reduced cutoff sc. As also shown there, c 0 A a truncation correction is also needed for the KA model. The μB for different temperatures may be found in Table I for -1 the KA model and in Table II for the pLJ model. We as- 10 cγτ sume from now on that the impulsive corrections are prop-

A ~ G for NVτT μ μ γτ Gττ/ A erly taken into account without stating this technical point -2 τ 10 Gγγ for NV T explicitly. τ slope -1/2 Gγτ for NP T G(t)/ τ Gγγ for NP T -3 γ slope -1 10 Gττ for NV T G for NPγT B. Dynamical matrix harmonic network ττ μ c for NPτT Gττ/ A γτ We turn now to the opposite low-T limit where we fo- -4 10 1 2 3 4 5 6 7 cus on the pLJ model at T = 0.001. Obviously, such deeply 10 10 10 10 10 10 10 t [MCS] quenched colloidal glasses must behave as amorphous solids with a constant shear modulus for large times by the top curve. FIG. 4. Rescaled shear modulus G(t)/μA vs. sampling time t in the high- Before presenting in Sec. IV C the elastic properties of these T liquid limit of the pLJ model comparing various ensembles and observ- glasses, we discuss first conceptually simpler substitute sys- ables. The filled squares show the bare shear modulus G˜ ττ obtained using Eq. (3) for the NVγ T-ensemble without the impulsive correction μB ≈ 0.2 tems formed by permanent spring networks. We do this for (horizontal line). The corrected modulus Gττ = G˜ ττ − μB for the NVγ T- illustration purposes since questions related to ergodicity and ensemble (open squares) and the NPγ T-ensemble (open diamonds) vanishes ageing are by construction irrelevant and the thermodynami- as it should. Note that Gττ decreases slightly faster than Gγτ ≈ Gγγ.As shown by the dashed line, all operational definitions of G decay inversely cal relations reminded in Sec. II should hold rigorously. As- with t. The dashed-dotted line indicates the expected power-law exponent suming ideal harmonic springs the interaction energy reads −1/2 for the correlation coefficient c (stars). = 1 − 2 γτ E 2 l Kl (rl Rl) with Kl being the spring constant and 12A533-11 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

μ (t)/μ μ =33.9 B A 0 A 10 1.0 μ μ τ fixed A(t)/ A

G 0.8

τ G/μ = 0.67 permanent harmonic springs =0 for imposed

A A for P=2,τ=0,T=0.001: μ =33.9 =0.67

A A μ

G for NVτT 0.6 μ A γτ (t)/ G/ μ weak t-dependence ! G for NVτT γγ F G for NPτT μ slope -1 γγ 0.4 μ /μ = 1 - G/μ = 0.32 τ F A A G(t)/ G for NVγT NV T γ ττ NPτT fixed G for NPγT ττ 0.2 NVγT ! τ G for NVτT ττ NPγT ! τ cγτ for NV T t-dependence 0.0 2 3 4 5 6 7 8 9 10 10 10 10 10 10 10 10 dynamical matrix: d=2, P=2, τ=0, T=0.001 -1 t [MCS] 10 1 2 3 4 5 6 7 8 9 10 10 10 10 10 10 10 10 10 = t [MCS] FIG. 6. Shear stress fluctuations μF(t)/μA with μA 33.9 for the same sys- tems as in Fig. 5. We observe μF(t)/μA → 1 (dash-dotted line) for imposed τ, while μ (t)/μ → 1 − G/μ ≈ 0.32 (dashed line) for imposed γ .Note FIG. 5. Rescaled shear modulus G(t)/μ vs. sampling time t for systems F A A A the broad time-dependence of μ (t) in the latter cases. of permanent linear springs corresponding to the dynamical matrix of pLJ F systems at P = 2, τ = 0, and T = 0.001. As indicated by the bold line, we obtain G/μA ≈ 0.67 in the long-time limit. In agreement with Eq. (19) Gττ vanishes with time if it is computed in an ensemble where γ fluctuates as shown for the NVτT-ensemble (filled squares). ensembles or using Gττ in the NVγ T- or NPγ T-ensembles at fixed shear strain γ . Note that even in these cases the shear Rl the reference length of spring l. The sum runs over all modulus decreases first with time albeit the network is per- springs l between topologically connected vertices i and j of fectly at thermodynamic equilibrium and no ageing occurs by the network. We assume here a strongly and homogeneously construction. As shown by the dash-dotted line, G(t)/μA → 1 connected network where every vertex i is in contact with for short times, i.e. in this limit the response is affine. As indi- many neighbors j. Such a network may be constructed from cated by the stars, we have also computed the correlation co- = a pLJ configuration keeping the particle positions for the po- efficient cγτ(t) as a function of time. From√cγτ 1 for small l sitions of the vertices and replacing each LJ interaction by t, it decreases somewhat becoming cγτ ≈ 0.67 ≈ 0.82 for a spring. Both spring constants are chosen such that the “dy- large times in agreement with Eq. (20).IfGττ is computed in namical matrix” of the pLJ configuration and harmonic net- the “wrong” NVτT- or NPτT-ensembles, it is seen to vanish work are identical, i.e., the second derivative of the total in- rapidly with time as shown by the filled squares. This decay 1 teraction potential with respect to the particle positions is the is of course expected from Eq. (19). same for T → 0. This implies To emphasize the latter point we show in Fig. 6 the stress fluctuations μF(t) for different ensembles. This is also done K ≡ u (r)| = , l r rl to have a closer look at the time dependence of various con-  (64)  tributions to G (t). As before, the vertical axis has been u (r)  ττ Rl ≡ rl −  rescaled with μ ≡ 33.9. We also indicate the rescaled values u (r) A r=rl of μB(t) (thin top line) and μA(t) = μB(t) − Pex(t) (dashed- with u(r) = u(s) being the interaction potential for reduced dotted line) computed by a gliding average over time intervals = + distances sl = rl/σ l ≤ sc. (Impulsive corrections at sl = sc are [t0, t1 t0 t]. Being simple averages both properties do neglected here.) Taking advantage of the fact that all interact- not depend on the ensemble probed (not shown) and become ing vertices are known, these networks can be computed us- virtually immediately t-independent as can be seen from the ing global MC moves with longitudinal and transverse planar indicated horizontal lines. This is quite different for the dif- waves as described in Sec. III. No local MC jumps for individ- ferent shear stress fluctuations μF presented which increase ≈ ual vertices have been added. Note that although some Kl and monotonously from μF 0 for short times to their plateau Rl may even be negative, this does not affect the global or lo- value for long times. In agreement with Eq. (19) we find that cal stability of the network. For networks at the same pressure μF(t)|τ → μA for NVτT- and NPτT-ensembles and as pre- P, shear stress τ, and temperature T as the original pLJ con- dicted by Eq. (18) we confirm μF(t)|γ → μA − G for NVγ T- figuration, we obtain, not surprisingly, the same affine shear and NPγ T-ensembles. The asymptotic limit is reached after 4 elasticity μA = 33.9 as for the reference. The same applies for about 10 MCS in the first case for imposed τ, but only after 7 other simple averages. 10 for imposed γ . This is due to the fact that μF|τ corre- Figure 5 presents various measurements of the shear sponds to the fluctuations of the self-contribution of individ- modulus G for different ensembles as a function of sampling ual springs as discussed in Sec. II C, while μF|γ also contains time t. As in Sec. IV A, we use gliding averages and the contributions from stress correlations of different springs. It vertical axis has been rescaled by the affine shear elasticity should be rewarding to analyze in the future finite system- μA. As can be seen, we obtain in the long-time limit G/μA size effects and the time dependence of three-dimensional ≈ 0.67 (bold line) irrespective of whether the shear modu- (3D) systems using the dynamical matrix associated to the KA lus is determined from Gγτ or Gγγ in the NVτT- or NPτT- model. 12A533-12 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

1.0 δ G for NVτT pLJ with rmax=0.01 μ μ γτ B(t)/ A G for NVτT γγ 1.0 τ G for NPτT μ (t)/μ fixed 0.8 γτ A A γ Gττ for NV T

γ =0.46 Gττ for NP T 0.8 A μ

τ A Gττ for NP T A 0.6 G/ 2 μ μ μ μ μ τ F/ A = 1-G/ A=0.54 cγτ for NP T 0.6 γ fixed (t)/ F μ G(t)/ 0.4 μ G/ A = 0.46 0.4 ! NVτT 1/t NPτT t-dependence γ 0.2 0.2 open symbols: T=0.001,μ =33.9 NV T A γ filled squares : T=0.010,μ =33.7 NP T μ t-dependence A pLJ: P=2, T=0.001, A=33.9 0.0 0.0 1 2 3 4 5 6 7 1 2 3 4 5 6 7 8 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 t [MCS] t [MCS]

FIG. 8. Reduced shear stress fluctuation μF(t)/μA for pLJ beads using the FIG. 7. Gγτ(t)/μA, Gγγ(t)/μA and Gττ(t)/μA for pLJ beads in the low- temperature limit for P = 2, τ = 0andT = 0.001 where μ ≈ 33.9. Qual- same representation as in Fig. 6. The small filled squares refer to data ob- A = = itatively, the data compares well with the various shear moduli presented in tained for T 0.010 using the NVγ T-ensemble, all other data to T 0.001. The reduced Born coefficient μ (t)/μ (thin line) and the affine shear elas- Fig. 5. The filled triangles show Gττ for the NPτT-ensemble which is seen B A to vanish rapidly with sampling time t. The dashed line indicates a 1/t-decay. ticity μA(t)/μA (dashed-dotted line) are essentially time independent while μF(t)/μA approaches its large-t limit (dashed line) from below. C. Low-temperature glass limit

We return now to the colloidal glasses where the topol- The rescaled shear stress fluctuations μF(t)/μA for the ogy of the interactions is, of course, not permanently fixed same ensembles are presented in Fig. 8. While the two-point (at least not at finite temperatures) and the shear stresses may correlations μB(t)/μA (thin horizontal line) and μA(t)/μA thus fluctuate more strongly. As an example, we present in (dashed-dotted horizontal line) do not dependent on time, one Fig. 7 the shear modulus G(t) as a function of sampling time observes again that μF(t)/μA increases monotonously from t for pLJ beads at mean pressure P = 2, mean shear stress zero to its long-time asymptotic plateau value. In agreement τ = 0 and mean temperature T = 0.001, i.e., the same condi- with Eq. (19) we find μF(t)|τ → μA for the NVτT and the tions as in the previous subsection. All simple averages, such NPτT ensembles, while μF(t)|γ → μA − G (dashed hori- as the affine shear elasticity μA = 33.9, are the same for all en- zontal line) for the NVγ T and the NPγ T ensembles. (Similar sembles albeit different quench protocols have been used. For behavior has been obtained for the KA model in the NVγ T- ≤ all examples given we use local MC jumps with δrmax = 0.01 ensemble for temperatures T 0.1.) As may be better seen for which allows to keep a small, but still reasonable acceptance the data set obtained for T = 0.010 using the NVγ T-ensemble rate A ≈ 0.161. We have also performed runs with additional (small filled squares) where μA ≈ 33.7, the stress fluctua- global MC moves using longitudinal and transverse plane tions μF(t)|γ do not rigorously become constant, but increase waves. This yields similar data which is shifted horizontally extremely slowly on the logarithmic time scales presented. to the left (not shown). There is, hence, even at low temperatures always some sam- As one would expect, the shear moduli presented in pling time dependence if the interactions are not permanently Fig. 7 are similar to the ones shown in Fig. 5 for the per- fixed as for the topologically fixed networks discussed above. = manent spring network. The filled triangles show Gττ com- However, this effect becomes only sizeable above T 0.2 for puted in the “wrong” NPτT-ensemble. As expected from the KA model and above T = 0.1 for the pLJ model. Eq. (18) these values vanish rapidly with time. If on the other hand G , G , and G are computed in their natural ensem- γτ γγ ττ D. Scaling with temperature bles, all these measures of G are similar and approach, as one expects for such a low temperature, the same finite asymp- We turn now to the characterization of the temperature totic value G/μA ≈ 0.46 indicated by the bold horizontal line. dependence of the shear modulus G and various related prop- 2 The stars indicate the squared correlation coefficient cγτ(t) erties. We focus first on the best values obtained for the which is seen to collapse perfectly on the rescaled modulus longest simulation runs available. Figure 9 presents the stress Gγτ(t)/μA which shows that Eq. (20) even holds for small fluctuations obtained for the KA model using the NVγ T- times before the thermodynamic plateau has been reached. ensemble (open spheres) and for the pLJ model using NVγ T- Similar reduced shear moduli Gττ/μA have been also obtained and the NPτT-ensembles. The affine shear elasticity μA ob- for the KA model in the low-T limit as may be seen from tained for each system is represented by small filled symbols. Table I. These results confirm that about half of the affine As shown for both models by a bold solid line for the low- shear elasticity is released by the shear stress fluctuations in T and by a dashed line for the high-T regime, μA(T) decays agreement with Refs. 14 and 15. Interestingly, this value is roughly linearly with T. Interestingly, the slopes match pre- slightly smaller as the corresponding value for the permanent cisely at T = Tg for both models. As may be seen from the dynamical matrix model presented above. Since some, albeit large triangles for the NPτT-ensemble, we confirm for all very small, rearrangement of the interactions must occur for temperatures that μF|τ ≈ μA as expected from Eq. (19).For the pLJ particles at finite temperature, this is a reasonable systems without box shape fluctuations, it is seen that μF|γ finding. becomes non-monotonous with a clear maximum at T ≈ Tg. 12A533-13 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

0.5 43 (1-0.2(T/T small closed symbols: μ =μ -P KA : P=1,T =0.41 50 A B ex g γ Gττ for KA NV T g -1)) pLJ: P=2,T =0.26 KA NVγT g τ Gγτ for pLJ NV T 45 γ pLJ NV T 0.4 τ Gγτ for pLJ NP T pLJ NPτT τ 40 Gγγ for pLJ NP T

KA: d=3,P=1 (T)

A 0.3 γ 30 (1-0.13(T/T y = 0.45 (1-x) Gττ for pLJ NV T μ 35 G for pLJ NPγT (T) 0.4 g -1)) 43 (1-0.33(T/T ττ F =0.41 g μ 2 continuous ! T 0.2 c 30 γτ 1/2 g -1)) 0.2 y = G(T)/ 25 pLJ: d=2,P=2 0.1 NVτT =0.26 NPτT g 30 (1-0.17(T/T pLJ 20 T 0.0 μ 0.0 0.5 1.0 1.5 maximum of |γ ! g -1)) x F 0.0 15 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.0 0.5 1.0 1.5 T x = T/Tg

FIG. 9. Stress fluctuation μF (open symbols) and affine shear elasticity μA FIG. 11. Rescaled shear modulus y = G/μA vs. reduced temperature (small filled symbols) vs. temperature T for both models. The indicated val- x = T/Tg for the KA model at P = 1(NVγ T-ensemble) and the pLJ model ues have been obtained from the longest simulation runs available for a given at P = 2(NVτT-, NPτT-, NVγ T-, and NPγ T-ensembles). The bold line temperature. μA decays roughly linearly with T as shown by bold solid lines indicates again a continuous cusp-singularity. Inset: Squared correlation co- for the low-T and by dashed line for the high-T regime. For constant-τ sys- 2 efficient cγτ for NVτT- and NPτT-ensembles. tems and for all systems with T ≥ Tg we confirm that μF ≈ μA, while μF|γ (open spheres and squares) is seen to be non-monotonous with a maximum at T . g values obtained by MC simulations of the pLJ model (P = 2, Tg ≈ 0.26) using both local and global moves and different In the liquid regime for T>Tg where the boundary condi- ensembles as indicated. For the pLJ model the values Gττ ob- tions do not matter, we obtain μF|τ ≈ μF|γ ≈ μA, i.e., all tained in the NVγ T ensemble (squares) are seen to be essen- shear stress fluctuations decay with temperature. Below Tg the tially identical for all T to the moduli Gγτ (crosses) and Gγγ boundary constraint (γ = 0) becomes more and more relevant (triangles) obtained in ensembles with strain fluctuations. As with decreasing T which reduces the shear stress fluctuations. shown by the small filled triangles, Gττ vanishes in the NPτT According to Eq. (18), μF|γ must thus decrease with increas- ensemble for all T in agreement with Eq. (19).(Weremind ing G as the system is further cooled down. That μF|γ must that the impulsive correction μB has to be taken into account necessarily be non-monotonous with a maximum at Tg is one if Gττ is used.) Decreasing the temperature further below Tg of the central results of this work. the shear moduli are seen to increase rapidly for both models. The temperature dependence of the (unscaled) shear As shown by the solid and the dashed lines both models are moduli G(T) is presented in Fig. 10. Data obtained by MD well fitted by a cusp-singularity simulation of the KA model (P ≈ 1, T ≈ 0.41) are indi-   g ≈ − 1/2 + cated for the NVγ T-ensemble (open spheres) and the NPγ T- G(T ) g1 1 T/Tg g0 for T

g sufficient to rule out completely a small positive off-set. =0.26 T

τ g 5 pdLJ d=2,P=2, =0 T A slightly different representation of the data is given in Fig. 11 where the reduced shear modulus y = G(T)/μA(T)is plotted as a function of the reduced temperature x = T/Tg 0 for both models. Using the independently determined glass 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 transition temperature T and affine shear elasticity μ (T)as T g A scales to make the axes dimensionless, the rescaled data are FIG. 10. Unscaled shear modulus G as a function of temperature T for both found to collapse ! This is of course a remarkable and rather models. For the pLJ model the values Gττ obtained in the NVγ T ensemble unexpected result considering that two different models in (squares) are essentially identical to the moduli Gγτ (crosses) and Gγγ (tri- two different dimensions have been compared. The bold line angles) obtained for constant τ. As shown by the small filled triangles, Gττ indicates the cusp-singularity y = 0.45(1 − x)1/2 with a pref- vanishes in the NPτT-ensemble for all T in agreement with Eq. (19).The solid and the dashed line indicate the cusp-singularity, Eq. (65), for, respec- actor which is compatible to the ones used in Eq. (65) con- 30,31 tively, the KA model and the pLJ beads. sidering the typical values of μA(T) in both models. Whether 12A533-14 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

pLJ P=2, τ=0, T=1, NPτT-ensemble give a tentative characterization of the sampling time depen- dence. We plot Gγτ(t) as a function of t for different temper- 1 10 atures as indicated. To make the data comparable all simula- tions for T ≥ 0.1 have been performed with local MC jumps T=0.001 of same maximum distance δrmax = 0.1. As we have already T=0.100 T=0.125 seen in Fig. 4, the shear modulus decays inversely with time (t)

γτ T=0.150 in the liquid regime at high temperatures. This limit is indi- G 0 T=0.175 10 T=0.200 cated by the thin line. With decreasing T the (unscaled) shear T=0.225 T=0.250 modulus increases and a shoulder develops. Note that even slope -1 T=0.275 = T=0.300 for T 0.25, i.e., slightly below the glass transition temper- T=0.325 ature Tg = 0.26 (Fig. 2), Gγτ(t) decays strongly with time. A T=0.350 more or less t-independent shoulder (on the logarithmic scales -1 10 2 3 4 5 6 7 used) can only be seen below T ≈ 0.2. We emphasize that we 10 10 10 10 10 10 t [MCS] present here a sampling time effect expressing the fact that the configuration space is more and more explored with increas- 10 FIG. 12. Unscaled shear modulus Gγτ as a function of sampling time t in ing t and not a time correlation function. Please note also MCS obtained for the pLJ model in the NPτT-ensemble for different temper- that this time dependence has nothing to do with equilibra- atures as indicated. The thin line indicates the power-law slope −1 for large times in the liquid regime. A clear shoulder (on the logarithmic scales used) tion problems or ageing effects. Time translational invariance can only be seen below T ≈ 0.2 and only below T = 0.01 one observes a is perfectly obeyed in our simulations as we have checked t-independent plateau over two orders of magnitude. We emphasize that we by rerunning the sampling after additional tempering over at present here a sampling time effect and not a time correlation function. The least 107 MCS. Similar t-dependencies for different tempera- same sampling time effect is observed after additional tempering. tures T have also been observed for Gγτ(t) and Gττ(t)forthe pLJ model and for Gττ(t) for the KA model (not shown). this striking collapse is just due to some lucky coincidence or A different representation of the data for both mod- makes manifest a more general universal scaling is impossi- els is shown in Fig. 13 where the reduced shear modulus ble to decide at present. No rescaling of the vertical axis is y = G(T)/μA(T) is plotted against the reduced temperature 2 = necessary for the squared correlation coefficient cγτ shown in x T/Tg for different sampling times t as indicated. Sim- the inset for two different ensembles of the pLJ model. The ilar behavior is found for both models. The shear modulus same continuous cusp-like decay (bold line) is found as for decreases with t and this the more the larger the temperature the modulus G. This is again consistent with the thermody- T. For smaller temperatures the different data sets appear to namic reasoning put forward at the end of Sec. II A,Eq.(20). approach more readily as one expects from the t-independent A word of caution must be placed here. The results pre- shoulder for the pLJ model in Fig. 12. As a consequence the sented in Figs. 9–11 correspond to the longest simulation runs glass transition is seen to become sharper with increasing t. we have at present been able to perform. Obviously, this does The data approach the continuous cusp-singularity indicated not mean that they correspond to the limit for asymptotically by the bold line. However, it is not clear from the current long sampling times, if the latter limit exists, or at least an in- data if the data converge to an at least intermediately sta- termediate plateau of respectable period of time. Focusing on ble behavior. Longer simulation runs are clearly necessary for the pLJ model in the NPτT-ensemble we attempt in Fig. 12 to both models to clarify this issue. Similar data have also been

KA model with friction constant γ=0.5 (b) δ (a) t=3τ pLJ local MC moves with r =0.1 3 LJ max t=10 0.4 t=30τ 4 LJ 0.5 t=10 t=300τ 5 LJ t=10 t=3000τ 6 0.3 LJ 0.4 t=10 (T) y = 0.45 (1-x) (T) 7 A A t=10 μ μ 0.3 (T)/

(T)/ y = 0.45 (1-x) 0.2 1/2 ττ t-dependence γτ decreases 0.2 1/2 y = G 0.1 y = G 0.1 continuous continuous or jump ? or jump ? 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.2 0.4 0.6 0.8 1.0 1.2 1.4 x = T/Tg x = T/Tg

FIG. 13. Reduced shear modulus G(T)/μA(T) as a function of reduced temperature x = T/Tg for different sampling times t as indicated: (a) Gττ(T)forthe KA model (NVγ T-ensemble), (b) Gγτ(T) for the pLJ model (NPτT-ensemble). For both models the glass transition is seen to become sharper with increasing t. The continuous cusp-singularity is indicated by the bold line. More points in the vicinity of x ≈ 1 and much larger sampling times are needed in the future to decide whether the shear modulus is continuous or discontinuous at the glass transition. 12A533-15 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013) obtained for the pLJ model from the independent determi- sponses. Since μA >μF|γ for T

r f (r) rf (r) η ∂ 2 d2 0.7 50 1 0.6 0.0 0.5 1.0 1.5 2.0 − rf (r) (A20) K for KA, NVγT d y = K / pp x K for pLJ, NPτT 0.5 vp → τ for 0 and dropping the argument (0) on the right-hand Kvv for pLJ, NP T K for pLJ, NVγT sides. According to the general result Eq. (A11) for the nor- pp 0.4 2 c for pLJ, NPτT mal pressure and using Eq. (A19) one confirms for pair inter- vp 9, 10 actions the expected virial relation 0.3 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 1  x = T/T Pˆ =− r u (r ) (A21) g ex dV l l l FIG. 14. Affine dilatational elasticity ηA(T) and compression modulus K(T) = for the instantaneous excess pressure where the interactions for both models. Inset: ηA vs. reduced temperature x T/Tg. The small tri- between pairs of particles i < j are again labeled by an in- angles indicate ηF for the pLJ model using the NPτT-ensemble confirming that Eq. (A7) holds. Main panel: The Rowlinson formula Kpp,Eq.(A14),for dex l.UsingEq.(A18) the excess contribution to the affine the NVγ T-ensemble is given for both the KA model (spheres) and the pLJ dilatational elasticity becomes model (squares). For the pLJ model we indicate in addition Kvp (crosses), 2 Kvv (large triangles) and the rescaled correlation coefficient cvp (stars) which ηA,ex = ηB + Pex, (A22) collapses on the other pLJ data as suggested by Eq. (A8). 12A533-18 Wittmer et al. J. Chem. Phys. 138, 12A533 (2013)

21 with Eq. (A24) relating ηA and μA and the fits given in K. Yoshimoto, T. Jain, K. van Workum, P. Nealey, and J. de Pablo, Phys. Fig. 9 for μ . We emphasize that the affine dilatational elas- Rev. Lett. 93, 175501 (2004). A 22 ticity η (and thus μ ) is not only larger for the KA model N. Schulmann, H. Xu, H. Meyer, P. Polinska,´ J. Baschnagel, and J. P. A A Wittmer, Eur.Phys.J.E35, 93 (2012). but, more importantly that, is also increases more rapidly 23A. Zaccone and E. Terentjev, preprint arXiv:1212.2020 (2013). with decreasing T than the pLJ model. The main panel of 24F. Léonforte, A. Tanguy, J. P. Wittmer, and J.-L. Barrat, Phys. Rev. Lett. Fig. 14 shows the best long-time values of the compression 97, 055501 (2006). 25M. Tsamados, A. Tanguy, C. Goldenberg, and J.-L. Barrat, Phys. Rev. E modulus K for both models. The vertical axis is rescaled by 80, 026112 (2009). ηA. The rescaled stress fluctuation formula, Eq. (A14),for 26G. Papakonstantopoulos, R. Riggleman, J. L. Barrat, and J. J. de Pablo, the NVγ T-ensemble is given for both models. For the pLJ Phys.Rev.E77, 041502 (2008). 27D. P. Landau and K. Binder, A Guide to Monte Carlo Simulations in Statis- model we also indicate Kvp (crosses) and Kvv (large trian- tical Physics (Cambridge University Press, Cambridge, 2000). gles) for the NPτT-ensemble. No rescaling with ηA is neces- 28S. Ulrich, X. Mao, P. Goldbart, and A. Zippelius, Europhys. Lett. 76, 677 2 sary for the squared correlation coefficient cvp (stars) also pre- (2006). sented. A perfect collapse of all pLJ data is found confirming 29G. Parisi and F. Zamponi, Rev. Mod. Phys. 82, 789 (2010). 30H. Yoshino and M. Mézard, Phys. Rev. Lett. 105, 015504 (2010). the equivalence of all operational definitions used. For both 31 ≈ H. Yoshino, J. Chem. Phys. 136, 214108 (2012). models we observe a strong increase of K/ηA around x 1 32G. Szamel and E. Flenner, Phys. Rev. Lett. 107, 105505 (2011). with decreasing temperature. Interestingly, the rescaled data 33C. Klix, F. Ebert, F. Weysser, M. Fuchs, G. Maret, and P. Keim, Phys. Rev. of both models is similar but not identical. The mentioned in- Lett. 109, 178301 (2012). 34 crease around x ≈ 1 is clearly more pronounced for the KA W. Götze, Complex Dynamics of Glass-Forming Liquids: A Mode- Coupling Theory (Oxford University Press, Oxford, 2009). model. Note that if the unscaled K is directly plotted against 35S. J. Plimpton, J. Comput. Phys. 117, 1 (1995). T, this effect becomes even stronger (not shown) consistently 36T. Mason and D. Weitz, Phys. Rev. Lett. 75, 2770 (1995). with the scaling of η presented in the inset. We note finally 37E. del Gado and W. Kob, J. Non-Newtonian Fluid Mech. 149, 28 (2008). A 38 that for very low temperatures we observe for both models A. Zilman, J. Kieffer, F. Molino, G. Porte, and S. A. Safran, Phys. Rev. Lett. 91, 015901 (2003). = 2 that the ratio K/ηA cγτ approaches unity, i.e., the compres- 39S. Senguptal, P. Nielaba, M. Rao, and K. Binder, Phys. Rev. E 61, 1072 sion modulus is essentially dominated by the affine response. (2000). We remind that this is different for the reduced shear mod- 40H. B. Callen, Thermodynamics and an Introduction to Thermostatistics (Wiley, New York, 1985). ulus for both models which is observed to approach G/μA 41 We have computed the static correlation function βVδγδˆ τˆτ for a broad ≈ 1/2inthelow-T limit (Fig. 11), i.e. the shear stress fluc- range of temperatures and for both NVτT- and NPτT-ensembles. tuations μF|γ are more relevant than the pressure fluctuations Converging rapidly to unity, this confirms that Eq. (8) holds to a high accu- racy well up into the liquid regime. ηF| . V 42H. Xu, J. Wittmer, P. Polinska,´ and J. Baschnagel, Phys.Rev.E86, 046705 (2012). 1M. Born and K. Huang, Dynamical Theory of Crystal Lattices (Clarendon, 43For the simplicity of the notation we assume that X is not the inner energy Oxford, 1954). U. For a more thorough systematic theoretical description it is necessary to 2P. M. Chaikin and T. C. Lubensky, Principles of Condensed Matter Physics define the “entropic intensive variable” J ≡ ∂S(X)/∂X with S(X) being the (Cambridge University Press, 1995). entropy.40 If X = U, one has J =−I/T.40 These entropic intensive variables 3D. C. Wallace, in Solid State Physics: Advances in Research and Applica- are used (for good reasons) in Ref. 45. Note that expressing Eq. (15) in tions, edited by H. Ehrenreich, F. Seitz, and D. Turnbull (Academic, New terms of I, rather than in terms of J, changes the signs. York/London, 1970), Vol. 25, p. 300. 44There is considerable freedom for defining Iˆ as long as the average I does 4L. D. Landau and E. M. Lifshitz, Theory of Elasticity (Pergamon, 1959). not change. The experimentally natural choice for Iˆ = τˆ isgivenbythe 5D.R.Squire,A.C.Holt,andW.G.Hoover,Physica 42, 388 (1969). instantaneous horizontal force (per surface) on the top wall in Fig. 1(a). 6J. F. Lutsko, J. Appl. Phys. 65, 2991 (1989). For theory and simulation, Eq. (28) or Kirkwood’s formula, Eq. (40),are 7D. Frenkel and B. Smit, Understanding Molecular Simulation – From Al- more suitable forτ ˆ, but by no means unique. gorithms to Applications, 2nd ed. (Academic, San Diego, 2002). 45J. L. Lebowitz, J. K. Percus, and L. Verlet, Phys. Rev. 153, 250 (1967). 8B. Schnell, H. Meyer, C. Fond, J. Wittmer, and J. Baschnagel, Eur. Phys. J. 46M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions E 34, 97 (2011). (Dover, New York, 1964). 9J. S. Rowlinson, Liquids and Liquid Mixtures (Butterworths Scientific, 47R. Zwanzig and R. Mountain, J. Chem. Phys. 43, 4464 (1965). London, 1959). 48J. Hansen and I. McDonald, Theory of Simple Liquids (Academic, New 10M. Allen and D. Tildesley, Computer Simulation of Liquids (Oxford Uni- York, 1986). versity Press, Oxford, 1994). 49W. Kob and H. C. Andersen, Phys. Rev. E 52, 4134 (1995). 11 50 J. R. Ray, M. C. Moody, and A. Rahman, Phys.Rev.B32, 733 (1985). For the quench of the configurations sufficiently small values for δγ max 12 J.-L. Barrat, J.-N. Roux, J.-P. Hansen, and M. L. Klein, Europhys. Lett. 7, for a shear strain altering MC move and for δVmax for a volume change 707 (1988). have been chosen such that the acceptance rate of both MC steps remain 13O. Farago and Y. Kantor, Phys.Rev.E61, 2478 (2000). reasonable at the lowest temperatures (T < 0.1). 14 51 J. P. Wittmer, A. Tanguy, J.-L. Barrat, and L. Lewis, Europhys. Lett. 57, Similar values for Tg have also been obtained for both models follow- 423 (2002). ing the standard diagnostics described in the literature12,49 of various dy- 15A. Tanguy, J. P. Wittmer, F. Leonforte, and J.-L. Barrat, Phys. Rev. B 66, namic properties such as the scaling of the mean-square displacements, the 174205 (2002). self-diffusion coefficient or the Fourier transformed density autocorrelation 16A. Tanguy, F. Leonforte, J. P. Wittmer, and J.-L. Barrat, Appl. Surf. Sci. function in the low-wavevector limit. From the analysis of these dynami- 226, 282 (2004). cal properties it is seen that, compared to the KA model, the pLJ model 17F. Léonforte, R. Boissière, A. Tanguy, J. P. Wittmer, and J.-L. Barrat, Phys. is a more gradual glass-former qualitatively similar to the purely repulsive Rev. B 72, 224206 (2005). soft-sphere model studied in Ref. 12. 18C. Maloney and A. Lemaître, Phys. Rev. Lett. 93, 195501 (2004). 52As for the shear stress fluctuations at constant τ in Sec. II C, it is possible 19J.-L. Barrat, in Computer Simulations in Condensed Matter Systems: From to demonstrate by integration by parts, Eq. (48), that Eq. (A7) holds with Materials to Chemical Biology, edited by M. Ferrario, G. Ciccotti, and K. ηA being given by Eqs. (A15) and (A23). Using the Legendre transform Binder (Springer, Berlin/Heidelberg, 2006), Vol. 704, pp. 287–307. Eq. (A6) one thus immediately obtains Rowlinson’s pressure fluctuation 20K. van Workum and J. de Pablo, Phys.Rev.E67, 011505 (2003). formula, Eq. (A5).