<<

PHYSICAL REVIEW MATERIALS 2, 095602 (2018)

Phase separation of stable colloidal clusters

Thomas Petersen,1,* Martin Z. Bazant,2,3,† Roland J. M. Pellenq,1,4,‡ and Franz-Josef Ulm1,§ 1Department of Civil and Environmental Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA 2Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA 3Department of Mathematics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA 4MultiScale Material Science for Energy and Environment, MIT-CNRS Joint Laboratory at Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA

(Received 5 June 2018; revised manuscript received 28 August 2018; published 25 September 2018)

This article presents a nonequilibrium thermodynamic theory for the mean-field precipitation, aggregation, and pattern formation of colloidal clusters. A variable gradient energy coefficient and the arrest of particle diffusion upon “jamming” of cluster aggregates in the spinodal region predicts observable gel patterns that, at high intercluster attraction, form system-spanning, out-of-equilibrium networks with glasslike, quasistatic structural relaxation. For reactive systems, we incorporate the free energy landscape of stable prenucleation clusters into the Allen-Cahn reaction equation. We show that pattern formation is dominantly controlled by the Damköhler number and the stability of the clusters, which modifies the autocatalytic rate of precipitation. As clusters individually become more stable, bulk separation is suppressed.

DOI: 10.1103/PhysRevMaterials.2.095602

I. INTRODUCTION In this article, we examine the mean-field nonequilibrium thermodynamics of reactive that form mesoscale A is a collection of nanometer- to micron-sized structures by aggregation and precipitation of stable precur- particles interacting in a fluid or . There has been sors. Before introducing the reaction rate, it is shown that much interest in studying colloids due to their ability to mimic a convex gradient energy penalty reproduces characteristics atomic systems inaccessible to microscopy [1] and configure akin to viscoelastic phase separation [24,25], where contrast- into functional, self-assembling structures [2–4]. For instance, ing entropic driving forces rather than differing constitutive the colloidal nature of cement paste, a material of vast societal behaviors summon a rich set of gel-patterns also observable importance, has only recently been exploited to gain insight in nature. Dynamic asymmetry between the low-density gas into the characteristics that lend to its exceptional mechanical and high-density gel phases is imposed by arresting particle properties [5–7]. Likewise, the discharge products of lithium- diffusion at local percolation in the spinodal region, allowing ion batteries are being engineered to maximize ion transport a quasistatic system-spanning gel to form. Next, we show and increase energy storage [8,9], and magnetic nanoparticles that once reactive kinetics are included, the heterogeneity are being functionalized as drug delivery vehicles, sealants, of the system is principally controlled by the Damköhler and separation aids [10–12]. number — the ratio between the reaction rate and the cluster Recent experiments on the thermodynamics of reactive diffusion rate — and the stability of the cluster intermediates. colloids have demonstrated pathways toward amorphous or In particular, the thermodynamic landscape of clusters fully crystalline bulk structures via precipitation of stable prenu- parameterizes a generalized Eyring reaction rate that enhances cleation clusters and reconciled these findings with classical or suppresses bulk nucleation from solution [12,26]. nucleation theory [12]. In fact, two-step nucleation from stable precursors has been demonstrated in a host of particu- late and biomineral systems [13–18]. Stabilizing mechanisms II. NONEQUILIBRIUM THERMODYNAMICS such as long-range electrostatic forces [19–22], favorable ion OF ATTRACTIVE COLLOIDS coordination [15], and polar or micellar association [23] allow Experimental observations and mode-coupling theory have persistent intermediates to form that settle into bulk structures demonstrated that two length scales dominate the physics of upon supersaturation. Yet no physically consistent, systematic attractive colloids [27–29]: Colloids aggregate into clusters study has been brought forth to investigate the influences that of characteristic size, which further assemble into an arrested control phase separation in these systems. mesoscopic network. Upon quenching — that is, rapidly increasing the relative attractive strength between particles, for instance, by decreasing the temperature or changing the constitution of the — these systems undergo glasslike *[email protected] dynamic arrest where cluster-cluster aggregation exhibits lim- † [email protected] ited bond breakage and the structure factor Sq does not ‡[email protected] significantly change on observational timescales. Akin to an §Corresponding author: [email protected] athermal granular medium, the colloid-rich phase undergoes a

2475-9953/2018/2(9)/095602(10) 095602-1 ©2018 American Physical Society PETERSEN, BAZANT, PELLENQ, AND ULM PHYSICAL REVIEW MATERIALS 2, 095602 (2018)

2.0 (b) (c) 2 dynamic arrest (a) binodal spinodal 0.3 diffusion reaction 1.8 φg controlled controlled φ˜g-local percolation ˜ φ μ˜res 0 0.2 0 ˜ ˜ Ω μ 1

1.6 h − + ˜

fluid μ ˜ Ω | − e gel spinodal region 0 h 1.4 0.1 μ ˜ g Δ˜ − ˜ glass −2 | μ˜0 Ω 1.2 0.0 gap 1.0 0.2 0.4 0.6 0.8 1.0 0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00 ˜ φ˜ φ φ˜

FIG. 1. (a) Stability diagram depicting the binodal, spinodal, and gelation lines at varying quench depths; the gelation line corresponds to Eq. (6), and the black line indicates ˜ = 4.0 selected for panels (b) and (c), and all simulation results of nonconserved systems. (b) Homogeneous part of Gibbs free energy of the clusters g˜h, where the linear term (˜ − μ˜ 0 )φ˜ has been added for clarity. (c) Homogeneous part of the diffusional chemical potentialμ ˜ h, withμ ˜ 0 − ˜ and reservoirμ ˜ res indicated by dashed, horizontal lines. The red star indicates the packing fraction at the gelation transition. jamming transition due to the local crowding of cluster aggre- Specifically, the Allen-Cahn reaction equation, which was first gates, as opposed to grains, which presents a “hidden” binodal introduced into electrochemistry to model phase separation at densities far below the thermodynamically predicted dense in lithium-ion batteries [35,36], utilizes the thermodynamic equilibrium [30–32]. As an example, this gelation line is landscape of the stabilized clusters to measure the net rate of drawn in red in the stability diagram shown in Fig. 1(a).For insertion. In other words, we model bulk nucleation as a two- mechanically unperturbed systems, thermal fluctuations pro- step process. Throughout this article we refer to clusters as mote densification toward equilibrium so slowly that further stabilized base units that assemble into an out-of-equilibrium, advances are made principally by reactive precipitation into mesoscopic gel network. the local structure. In particle systems with purely attractive interactions, er- A. Dynamic equation for density patterns godicity breaking at the scale of clusters creates polydisperse cluster sizes. However, if weak, long-range repulsive forces We posit the internal chemical potential of a cluster μ(φ,∇φ) as nonuniform, depending principally on the local are introduced, simulations and experiments have shown col- ∇ loids to form dilute suspensions of stable clusters with a packing density φ and its gradient φ [37]. Thus, local cluster narrow size distribution that persists over days [20,21]. Only rearrangements are driven by spatial variations in μ, and the after further increasing the mean packing density or attractive evolution of the system is modeled by a general reaction- strength does further aggregation and phase separation pro- diffusion equation for nonequilibrium thermodynamic mix- tures [35,37,38], ceed, leading to eventual arrest at the scale of cluster aggre-   gates. To model the physics of these two length scales, we ∂φ˜ Dφ˜ = ∇ · ∇μ + R(φ,μ,μ˜ res ), (1) introduce a dynamic reaction-diffusion equation for colloidal ∂t kBT clusters. Clusters are treated as renormalized particles of ˜ characteristic size a, whose local packing density is advanced where kBT sets the thermal energy scale, D(φ) is the tracer by a field variable φ and diffusion is driven by gradients in diffusivity, R is a reaction rate controlling insertion or deletion  ˜ =  the chemical potential of the clusters μ. The ability of the of clusters, and 0 φ φ/φm 1 is the filling fraction with dynamic equation to evolve realistic density patterns of col- φm the maximum packing density ( ˜ is henceforth used to loidal gels is first demonstrated on conserved systems, where signify nondimensionalized and normalized quantities). The an initially homogeneous density field is assigned, and no first, conserved term is a Cahn-Hilliard kernel that tracks additional insertion or deletion of clusters is permitted. These cluster diffusion within the domain [37], and the second, systems phase separate into high- and low-density regions, nonconserved term is an Allen-Cahn reaction rate that acts where clusters, initially disconnected from one another, form as a cluster source or sink [35,38]. While diffusion depends cluster aggregates whose diffusivity is exponentially reduced only on the local chemical potential, the reaction rate in this at local percolation. Though we admit that polydispersity in open system depends also on the external reservoir potential sizes of cluster aggregates are relevant to the dynamics [33,34] μres. The explicit expression for R will be derived in a section and that the dynamics are a history-dependent function of below. φ, the present study aims to reduce model complexity by To calculate μ, the free-energy landscape of a system of volume V is measured using a Ginzburg-Landau functional, focusing on the mesoscopic parameters that predict instability     and pattern formation. Thus, we do not explicitly track the size κ(φ˜ ) G = g(φ,˜ ∇φ˜ )dV = gh(φ˜ ) + |∇φ˜|2 dV, distribution of the cluster aggregates; we reserve a forthcom- V V 2 ing study to add an additional microscopic order parameter (2) to investigate the elasticity of colloids in a continuum setting. We continue our study by deriving a general reaction rate to where the Gibbs energy density g is expressed as a sum of form clusters that are individually stable but may collectively homogeneous and inhomogeneous contributions. Because the phase separate if a mesoscopic energy barrier is crossed. energy demanded in separating monomers quenched into

095602-2 PHASE SEPARATION OF STABLE COLLOIDAL CLUSTERS PHYSICAL REVIEW MATERIALS 2, 095602 (2018) clusters far exceeds that needed to separate clusters φg acts as the local packing at which the number of kinetic themselves, and we aim to model stabilized clusters that pathways toward a lower energy state drastically reduces. form at a characteristic size, the homogeneous free energy Approaching this threshold leads to a rapid increase of the density is expressed as a regular solution of cluster-occupied mean size of the cluster aggregates and a sharp decrease in sites and vacancies, local particle diffusion. A plethora of details are relevant to h ˜ accurately model particle motion, including the size distribu- g (φ) 2 = φ˜ ln(φ˜ ) + (1 − φ˜ )ln(1− φ˜ ) − ˜ φ˜ + φ˜μ˜ 0, (3) tion of the aggregates, its hydrodynamic interactions with the n k T s B solvent [42], the history dependence of the aggregation pro- with site density ns [35]. Thus clusters act as renormalized cess, and dynamical heterogeneity that varies across orders of  =  = particles where ˜ /kBT andμ ˜0 μ0/kBT are, respec- magnitude [33]. Because this article investigates aggregation tively, the interaction parameter between clusters and chem- that is largely unidirectional, favoring net densification, and ical potential of a cluster in a dilute system, each measured we wish to maintain model parsimony, we opt to introduce an with respect to the thermal energy scale. The double-well expression for the tracer diffusivity that captures the general shape of gh is plotted in Fig. 1(b), where two minima define physics of motion, though admittedly needs to be parameter- low- and high-density thermodynamic equilibria, and the ized for the system at hand. concave spinodal region indicates packing densities at which On approach of the gelation line from low densities, bulk homogeneous base states in conserved systems are unstable to measurements on attractive colloids show power-law diver- density fluctuations. The variable gradient energy coefficient gence in viscosity [31], observations expected to be the result for the inhomogeneous free-energy contribution is chosen as of particle motion that is increasingly collective [43–45]. As ∼ κ(φ˜ ) 2 κ˜0 a consequence, we assume the diffusivity to scale as D = , (4) γ/ν −γ 2 (ξ/a) ∼|φ − φg| for φ<φg, where γ and ν are critical nsa kBT 9 φ(1 − φ) √ exponents for the diffusivity and correlation length ξ, respec- and encodes the interface width l  (2/3) κ0/ and surface tively. Though diffusion drastically reduces once the percola- energy of the density field. De Gennes derived the expression tion cluster forms, continued motion in physical systems per- in Eq. (4) by relating the response of fluctuations in the sists due to the finite size of the system [46], cluster aggregates entropic portion of a free energy density similar to Eq. (3)to that remain disconnected from the percolation cluster [47], the static structure factor and radius of gyration of polymer and activated events caused by internal stresses and thermal coils [39,40]. Here, we repurpose his result for our attractive fluctuations [48]. Hence, we adjust the power-law relation for colloids, where ergodicity breaking at the scale of the clusters ξ to its characteristic length at gelation and infer φ = φg as the sets the length scale of the density fluctuations. With Gibbs packing density at which the cluster aggregates percolate their free energy defined, the diffusional chemical potential is cage of size ξg. Specifically, we relate φ˜g and ξ˜g = (ξg/a)in calculated using the Euler-Lagrange equation [35]:   our expression for the tracer diffusivity as follows: μ ∂g˜ ∂g˜ φ˜   = − ∇ · = ln − 2˜ φ˜ + μ˜ ˜ 1/ν γ 0 −1/ν ξg kBT ∂φ ∂∇φ˜ 1 − φ˜ D = D H( ) + ξ˜ exp − | | . (7) 0 g 2 κ˜ (φ˜ )   + |∇˜ φ˜|2 − ∇˜ · κ˜ (φ˜ )∇˜ φ˜ . (5) 2 Above, D0 ≈ kBT/3πaη is the tracer diffusivity in a dilute = ˜ − ˜ ˜ This is the continuum analog to the standard definition of system, with η being the solvent viscosity, (φg φ)/φg the chemical potential in particle-resolved systems. Above, is the reduced density, and H is the Heaviside function.  denotes ordinary differentiation, and the gradient operator The form of Eq. (7) imparts the following characteristics on ∇=˜ L ∇ has been normalized by the linear size of the D: (i) a power-law dependence on the packing density at sys ˜ system Lsys. The shape of the homogeneous part of the low φ, (ii) ξg as the relevant length scale at φg, and (iii) a chemical potentialμ ˜ h = dg˜h/dφ˜ is plotted in Fig. 1(c) and stretched exponential relaxation upon arrest [48]. Similar to will be discussed in relation to the reactive system’s stability Vogel-Fulcher-Tammann relaxation in glasses [49], Eq. (6) in greater depth below. ascribes dynamic arrest to the loss of thermally activated rearrangement, and Eq. (7) imposes an exponential decay in the collective diffusion of cluster aggregates upon jamming. B. Glasslike arrest of particle diffusion Several experimental investigations of phase-separating colloids have shown that dynamic arrest is initiated when C. Results for conserved systems cluster aggregates crowd their local volume and the rate of Before deriving the reaction rate R, we exemplify the bond formation exceeds the rate of bond breakage at a gelation flexibility of our model in resolving pattern formation of line aptly described by [31,41]  conserved fields. Simulations were run by implementing the  weak form of Eq. (1) in the open source Multiphysics Object- φ = φ0 exp − . (6) g g χk T Oriented Simulation Environment (MOOSE)[50], a finite- B element analysis software. Trials were run by initializing the 0 Above, χ is a constant of order unity, and φg is a packing density field at a homogeneous base state with mean filling ˜ = ˜ = density near the glass transition of hard spheres, here chosen fraction ( V φdV)/V 1/3 and adding Langevin noise 0 = as φg 1.5 φm. As stated in Ref. [31], the exponential form of with variance scaled by D to account for thermal fluctua- Eq. (6) is suggestive of a thermally activated process whereby tions [51]. Figures 2(a)–2(f) compare snapshots of simulated

095602-3 PETERSEN, BAZANT, PELLENQ, AND ULM PHYSICAL REVIEW MATERIALS 2, 095602 (2018)

(a) ˜ . (b) ˜ . (c) ˜ . Ω=25 Ω=35 Ω=45 0 10 ˜ ˜ −(t/τ˜ )β (g) Ω=2.5 Ω=3.0 (h)1.0 r˜ch ∼−e Ω=3˜ .5 Ω=4˜ .0 ¯ ˜ ˜ r β

Ω=4.5 Ω=5.0 / 0.75 ) 1/3

˜ 0.50 ∼ ˜ ch0

ch t r ˜

r τ 0.5 0

− 10 (d) (e) (f) −1 −1 ch 10 r 0 1 10 (˜ Ω˜ − Ω˜ 0 −3 −1 1 0.0 − − 10 10 10 10 3 10 1 101 ˜− ˜ t t0 (t˜− t˜0)/τ

˜ = 2 = ˜ FIG. 2. (a)–(c) Snapshots of φ(x) for conserved colloidal systems at t˜ tLsys/D0 1.0 for varying quench depths ; darkness is proportional to φ˜. (d)–(e) Experimental images for (d) the demixing of milk protein [52], (e) polystyrene-poly(vinyl methyl ether) [53], and (f) polystyrene-diethyl malonate [54]. (g) Time evolution of the characteristic domain size r˜ch,wheret˜0 denotes the time of 0 maximum interface area. (h) Evolution of the characteristic domain size rescaled to show its stretched exponential behavior; r˜ch corresponds to the domain size at t˜0, and the insets show β  C0( − ˜ 0 ) and ln(τ )  C1( − ˜ 0 )/[C2 + ( − ˜ 0 )] with fitting constants Ci . Simulations × = 2 = = = = were run at numerical resolution 512 512 and system size Lsys 1 with parameters κ0/nskBTLsys 0.001, χ 3.0, γ 1.2, ν 0.88, ˜ = 0 = ξg 50, and φg /φm 0.4. and observed colloidal quenched into varying states the first moment in q˜ as of intercluster attraction. As displayed, adjusting intercluster attractive strength ˜ predicts dramatic changes to the texture  −1 qS˜ q dq˜ and dynamics of the resulting fields, despite fixing the mean r˜ch = , (9) filling fraction. The fields predict droplet nucleation at low Iq ˜ , metastable filaments at moderate ˜ , and a spanning hon- eycomb network at high ˜ [Figs. 2(a)–2(c)], similarly seen in experimental analogs [Figs. 2(d)–2(e)]. For the relevant system, our model reproduces several features also observed (a) in viscoelastic phase separation [24,25]: (i) nucleation of a colloid-rich phase, (ii) volume shrinking of the colloid-rich 4 − 10 ∼ q˜ 4 phase and nucleation of holes therein, and (iii) the formation q

of a spanning network. However, instead of attributing the S kinetic asymmetry to differences in elasticity, here the asym- 10−5 metry results from differences in the entropic penalty from ˜− ˜ density fluctuations. The convex shape of κ and arrest of the thick lines: (t t0)=1.0 ˜− ˜ colloid-rich phase in the spinodal region dramatically reduces −14 thin lines: (t t0)=10.0 10 0 1 2 its interfacial tension, whence the field in Fig. 2(b) evolved 10 10 10 into its quasistable nonspherical shape. For arrest at lower q˜ φg, the aggregates become increasingly cohesive, invading the (b) Ω=2˜ .5 Ω=3˜ .0 solvent to form long-range bridges as seen in Fig. 2(c). Unlike 0.015 Ω=3˜ .5 Ω=4˜ .0 viscoelastic phase separation, spanning systems of colloidal Ω=4˜ .5 Ω=5˜ .0 aggregates become quiescent, showing little structural evolu- 0.010 q ˜ tion over decades [42,55], and phase inversion is not observed; I simulations for ˜ ={4.5, 5.0} were run for a decade beyond 0.005 the results shown in Fig. 2 without observing phase inversion. For an isotropic density field, the structure factor is calcu- lated as 0.0 2.5 5.0 7.5 10.0 t˜− t˜0 φˆ(q˜ )φˆ ∗(q˜ ) S (q˜ ) = , (8) q ˜ FIG. 3. (a) Structure factor of φ˜(x) for various ˜ at t˜ − t˜0 = 1.0 (thick lines) and t˜ − t˜ = 10.0 (thin lines), where t˜ denotes the time ∗ 0 0 where φˆ(q˜ ) is the Fourier transform of φ˜(x˜ ), indicates com- at which the interface area is at its maximum; curves are vertically plex conjugation, and brackets denote spherical averaging. offset for clarity. (b) Time evolution of the peak integrated intensity Increased quench depths shift the peak frequency in Sq toward I˜q = Sq dq˜; the legend indicating ˜ also corresponds to the curves shorter wavelengths, and a decline, rather than growth, of in (a). Dissolution and redeposition of small “droplets” onto adjacent long wavelengths (low q˜) signifies a system-spanning gel of larger structures — Ostwald ripening — is marked by step changes in increasingly rigidifying, thinning ligaments as discerned in Iq for ˜ = 2.5and˜ = 3.0. Time and length scales are normalized Fig. 3(a). This finding is verified in Fig. 2(g), which depicts by the dilute limit self-diffusivity of a cluster D0 and the system size = 2 = the evolution of the characteristic domain size measured from Lsys as follows: t˜ tD0/Lsys and q˜ qLsys.

095602-4 PHASE SEPARATION OF STABLE COLLOIDAL CLUSTERS PHYSICAL REVIEW MATERIALS 2, 095602 (2018) where Iq = Sq dq˜ is the integrated peak intensity. For low bulk nucleation. The free energy of formation of such a stable  =   3 ˜ , coarsening proceeds via Ostwald ripening [e.g., ˜ = 2.5; cluster is denoted by μ˜ 0 μ˜ c(N πa /6). Fig. 2(a)] and contraction of gel filaments [e.g., ˜ = 3.5; In a consistent description of reaction kinetics of nonequi- Fig. 2(b)], adhering to early-stage diffusive t˜1/3 power-law librium thermodynamic mixtures, the reaction complex ex- scaling for r˜ch [56–58]. In fact, loss of step changes in the plores the excess chemical potential landscape between local ex evolution of Iq in Fig. 3(b) indicates a clear distinction in minima of cluster-vacant sites μ◦ and cluster-occupied sites ex ex ex coarsening dynamics. As ˜ is increased, diffusive scaling in μ• . If expressions for μ◦ and μ• are known and the ac- ex r˜ch is abandoned and domain growth arrests into an out-of- tivation barrier of the transition state μ‡ can be estimated, equilibrium gel with anticipated bulk elasticity [59,60]; here, the net reaction rate is calculated from the probabilities of percolation of the bulk volume is a prerequisite for arrest. precipitating and dissolving stable clusters from a lattice as Denoting ˜ 0 as the minimal quench depth to form a system- follows [35]:   spanning gel, Fig. 2(h) demonstrates stretched exponential ex ex μ‡ − μ◦ relaxation in the evolution of rch — typical of glass-forming R = k0 (1 − φ˜ )exp − colloids — where the Vogel-Fulcher-Tammann timescale τ kBT  ex ex  and stretching exponent β are functions of the deviation from μ‡ − μ• ˜ − ˜ − φ˜ exp − , (11) the mesoscopic percolation threshold 0 [61]. As low k T stretching exponents β<1 are the result of spatially het- B erogenous dynamics and β increases with ˜ , heterogeneous where the likelihood of cluster insertion or deletion is scaled dynamics are most prominent in weakly attractive systems by the fraction of vacant or occupied sites, respectively, and close to the gelation threshold. the attempt frequency, denoted by k0, is assumed equal in both directions. D. Reaction rate governed by stable cluster precursors With reference to Appendix A, a thermodynamically con- sistent choice for the homogeneous parts of the chemical Our reaction rate R adapts to evidence that many chemi- potentials of cluster-occupied and cluster-vacant sites read cal systems demonstrate pathways to nucleation from stable prenucleation clusters [14–18,62]. These intermediates accel- h = − ˜ + ex erate (decelerate) nucleation by decreasing (increasing) the μ◦ kBT ln(1 φ) μ◦  = − 2 change in free energy μ μ μres upon being added to = kBT ln(1 − φ˜ ) + φ˜ + μres, (12a) the bulk structure [12]. Specifically, we denote μ as the μh = k T ln(φ˜ ) + μex difference in Gibbs free energy between a cluster precipi- • B • tated into the local volume μ and its constituent monomers = kBT ln(φ˜ ) + (φ˜ − 2)φ˜ + μ0, (12b) dissolved in the external reservoir, μ : It is the diffusional res ex chemical potential of a cluster in an open system [35]. In where μ◦ is set to the chemical potential of the reservoir the following, we demonstrate nucleation of precursors that plus the mean change in interaction energy upon adding a  ˜ 2 ex are stabilized by long-range electrostatic forces as is seen in, vacancy, φ , and μ• is set to the chemical potential of a for instance, the early stages of cement paste setting [34,63] cluster in a dilute solution plus the mean change in interaction  ˜ ˜ − and sedimentation of charged nanoparticles [19], though we energy upon inserting a cluster, φ(φ 2). It is readily remark that the generality of the framework is readily adapted observed that the extensive property of the chemical potentials h = h + − ˜ h to alternative stabilization mechanisms. recovers Gibbs energy density, g φμ• (1 φ)μ◦, and For weakly screened particles with a surface charge, the that replacing a vacancy by an occupancy is equivalent to  = − classical energy contributions that scale with cluster volume inserting a particle from an external source, μ μ• μ◦. and surface area are supplemented by a higher-order term that If the transition state excludes one site during precipitation results from Coulomb interactions [64]. The Gibbs free energy and dissolution reactions [35], and we further estimate its of formation of such a cluster of N monomers is approximated excess chemical potential from the landscape of Gibbs energy by [64] of formation of a cluster in Eq. (10), we can write ex = − ˜ + ex + − ex +  μ˜ c = μ˜ 0(N ) − μ˜ res μ‡ kBT ln(1 φ) αμ• (1 α)μ◦ μ‡, (13)

m 2/3 5/3 3 2 2 where α is a symmetry factor measuring the fractional aggre- = Nμ˜ + N σ˜ + N λBρ v , (10) 0 2a c gation of monomers required to reach the transition state [66],  m  =  +  = ‡ − where μ˜ 0 is the change in energy upon placing a free and μcr α μ0 μ‡ μ0 μres is the energy barrier monomer into the interior of the cluster normalized by kBT , with respect to the reservoir potential — otherwise termed σ˜ is the surface energy, and  is a constant shape factor. The the energy of formation of a critical-sized cluster, for which a electrostatic self-energy due to the particles’ surface charge geometric interpretation is given in Fig. 4(a). Hence, Eq. (13) scales with the Bjerrum length λB and the square of the assumes the excess chemical potential of the transition state, 2 ex charge density ρc [65], and v is the characteristic volume μ‡ , to be estimated from a weighted average of the ex- of a monomer. The N 5/3 dependence of this term derives cess chemical potentials of cluster-inserted and cluster-vacant 2 from the number of Coulomb interactions measuring ∼N , states that are separated by a barrier of average height μ‡; 1/3 while their mean separation distance measures ∼N .This lastly, kBT measures the energetic cost to the system in occu- choice of the free-energy landscape, displayed in Fig. 4(a), pying a site during the transition. With the help of expressions allows a local minimum to form stable clusters en route to in Eqs. (12a), (12b), and (13), the reaction rate in Eq. (11) can

095602-5 PETERSEN, BAZANT, PELLENQ, AND ULM PHYSICAL REVIEW MATERIALS 2, 095602 (2018)

(a)4 (b) 4 (c) ‡ 3 Da=0.1 Da=0.4 μ˜ 3 0 1.5 3 μ˜ 0 ˜ 0 0 φ

c 2 ˙ ˙ n d n

μ 2

1.0 /

2 ˜ ω

Δ˜ Δ˜μcr R/ Δ˜μ A 1 0 α 1 − α 0.5 1 1 μ˜ 0 res 0 0 0 20 40 60 0.00 0.25 0.50 0.75 1.00 0.0 0.5 1.0 N φ˜ q˜

 m FIG. 4. (a) Thermodynamic landscape of the transition state to form stable cluster precursors for varying μ0 [see Eq. (10)]. (b) Reaction rate (left axis) and autocatalytic rate (right axis) in function φ˜. (c) Integrated growth rate of the q˜ Fourier mode as calculated from the l2 norm of a perturbation for differing Da. All plots correspond to ˜ = 4.0 and line colors correspond to trials listed in Table I: blue – T1, orange – T2, green – T3, and red – T4. be rewritten as a nonlinear function of μ, ration under permitting characteristics of the reaction rate.     These characteristics are explored next. μ μ R = R0 exp −α − exp (1 − α) , (14) kBT kBT E. Results for nonconserved systems for which the reaction rate coefficient obeys In this section, we explore pattern formation in reactive-   diffusive systems by adjusting the thermodynamic landscape  α 1−α μ‡ of clusters, while maintaining a constant reservoir potential R0 = k0φ˜ (1 − φ˜ ) exp − . (15) kBT and intercluster attraction. We elect to modify the landscape  m  as follows: Decreasing μ0 lowers μ0 and shifts the sym- Equation (14) is the Allen-Cahn reaction rate for colloidal metry of the reaction rate toward lower α [see Fig. 4(a)]. cluster formation, for which we highlight several salient fea- Figure 4(b) plots four sample reaction rates, whose param- tures: (i) It adheres to the De Donder relation requiring reac- eters are listed in Table I. To make predictions about sta- tion rates to proceed in the direction of the chemical affinity, bility and dominant length scales, a linear stability analysis =− ˜ → sgn(R) μ. (ii) In the limit φ 0, it recovers the dilute of the reaction-diffusion equation, Eq. (1), is performed in = − cluster nucleation rate n˙0 k0 exp( μ˜ cr ), an expression Appendix B. We show that stability is ensured by a negative consistent with classical nucleation theory. (iii) The reaction growth rate coefficient, ω(φ˜ h,q) < 0, which measures the rate coefficient R0 adjusts the autocatalytic behavior of R rate of change of small sinusoidal fluctuations δφ˜ with wave ˜ through its dependence on the filling fraction, φ. number q from a homogeneous base state φ˜ = φ˜ h.Fora At this point, it is helpful to recapitulate how physics at purely reactive system, where D = 0, the autocatalytic rate both the scale of the cluster and the mesoscale inform R.The A predicts stability if [68] formation barrier and size of a cluster is dictated by μc(N )   h in Eq. (10) and sets the baseline relation between internal A| = ∂R + ∂R dμ μres < 0. (16) and external potentials μ0 as graphically represented in ∂φ˜ ∂μ dφ˜ Fig. 1(c). Clusters form more readily if μm decreases, 0 The autocatalytic rates are plotted on the secondary axis in the surface energyσ ˜ reduces, or the electrostatic repulsion Fig. 4(b), where it is shown that decreasing μ makes R subsides. Importantly, clusters with attractive interactions can 0 more autoinhibitory. Thus, increasingly stable clusters sup- be metastable with respect to the reservoir potential, where press preferential precipitation onto existing bulk phases. To μ > 0, yet lead to stable bulk nuclei at the mesoscale. 0 further analyze the implications of an autoinhibitory reaction In theory, the energy barrier to nucleate a high-density bulk rate, Fig. 4(c) displays the integrated growth rate of unsta- phase from stable clusters is made up of the stabilizing ble modes in course of reaction ωR˜ dt˜, whereω ˜ is the intercluster repulsive forces and the kinetic energy of the clusters that transition from translational motion in the gas phase to vibrational motion in the jammed phase. In practice, TABLE I. Parameters defining the free energy landscapes of the  =  m + 2/3 + 5/3 the free energy density is fit to experimental observations of stable cluster precursors, where μ˜ c μ˜ 0 N BN CN , the and limits by adjusting ˜ and and B = 0.023 and C = 2.608 are left constant. κ˜0 [67]; these quantities are otherwise difficult to measure.  m  As a result, diluting effects of the electrostatic repulsion Trial μ˜ 0 μ˜ 0 α upon crowding of clusters and size-dependent scaling of the T1 −0.890 2.350 0.324 cluster-cluster interaction energy are subsumed into these T2 −0.900 2.240 0.273 mesoscopic parameters. Bulk nuclei form once the chemical T3 −0.915 1.636 0.219 driving force μ is sufficient to drive φ˜ into the spinodal T4 −0.950 0.000 0.144 region, where Cahn-Hilliard dynamics promote phase sepa-

095602-6 PHASE SEPARATION OF STABLE COLLOIDAL CLUSTERS PHYSICAL REVIEW MATERIALS 2, 095602 (2018)

(a) Da = 0.1 Da = 0.1 Da = 0.1 0 0 . 0 . D D . Da = 0 4 Da = 0 4 D Da = 0 4 increasing increasing Da = 1.0 Da = 1.0 increasing Da = 1.0

T4 T2 T1 T4 T2 T1 T4 T2 T1

increasing Δµ0 increasing Δµ0 increasing Δµ0

FIG. 5. Snapshots of reactive systems for varying precursor landscapes Ti and Da at (a) ˜ = 0.4, (b) ˜ = 0.6, and (c) ˜ = 0.8. linearized growth rate of the q˜ Fourier mode [68]. Growth tude and move the location of the second peak toward closer rates are plotted for different μ0 and Damköhler number separation distances r˜. That is, as A becomes less positive, the 2 Da = n˙0l /D0, where D0 = kBT/3πηa and n˙0 = R(φ˜ → mean packing fraction of the gel phase ˜ g behaves increas- 0) = k0 exp(−μ˜ cr ) are the diffusivity and cluster nucleation ingly linearly with the overall reaction extent (see inset of rate in a dilute , respectively. Increasing Da damp- Fig. 7) and density fluctuations form at smaller wavelengths. ens the peak in ωR˜ dt˜ caused by the Cahn-Hilliard kernel, Lastly, we calculate the mean pore size mz from the pore- and simultaneously decreasing μ0 — that is, increasing the chord length probability density function with threshold φ˜g stability of the prenucleation clusters — allows near complete between gas and gel phases using the following relation [70]: suppression of mode growth.  ∞ These predictions are verified by the pattern evolution mz = zp (z)dz, (17) depicted in Figs. 5(a)–5(c) for increasing reaction extents ˜ = 0 {0.4, 0.6, 0.8}. The characteristic size of the emergent bulk where p(z) is the pore-chord length probability density func- nuclei scale as ∼[κ/(dμh/dφ˜ )]1/2 and depend critically on the tion and z measures the length of a sampled chord. As seen location within the spinodal region at which fluctuations grow. in Fig. 7, the location of onset of bulk nucleation predicts m˜ z Strongly diffusive clusters (low Da) phase separate readily prior to entering Avrami-like growth, where precipitation at near the spinodal, amplifying large wavelengths, whereas lower μ0 universally produces smaller pores. This reduction reaction-controlled dynamics (high Da) delay the growth in pore size is facilitated both by accessing smaller wave- of instabilities toward higher ˜ , forming smaller nuclei in lengths at phase separation, and more uniformly precipitating larger abundance or proceeding by spinodal decomposition. stable precursors once the cluster aggregates have dynami- Importantly, dynamic arrest in diffusion at φ˜g halts Cahn- cally arrested. Mesoscale texture is critical in predicting a Hilliard mode growth, indicating that microtextural patterns host of important material properties — elasticity, fracture in attractive colloids are determined in the low-density portion toughness, fluid and electrical conductivity, to name a few — of the spinodal region. As μ0 is decreased, further cluster and Fig. 7 shows that pattern formation can be controlled if insertion shifts from surface growth of bulk nuclei toward the reaction rate is adjusted within the mobile portion of the homogeneous densification; similar observations were made in molecular-dynamics simulations of crystallizing Lennard- 1.0 Jones fluids in Ref. [69]. These physics are further reflected 100 in the two-point density correlation function C˜ 2 plotted in g ˜ Fig. 6(b), where auto-inhibitory kinetics moderate its ampli- Φ 0.5

z −1 0.5 1.0 ˜

m 10 Φ˜

0.2 T1 T3 T2 T4 −2 2 10

˜ mobile kinetically arrested C 0.0 0.2 0.4 0.6 0.8 1.0 ˜ Da=0.1 φg Φ˜ 0.0 0.2 0.4 r˜ FIG. 7. Mean pore size m˜ z in function of the overall packing fraction. The shaded domain corresponds to the spinodal region, red FIG. 6. The two-point density correlation functions for Da = 0.1 and blue lines correspond to T1 and T4, respectively, and solid, as measured by the inverse Fourier transform of Sq for ˜ = 0.6and striped, and dotted lines correspond to Da = 0.1, 0.4, and 1.0, ˜ = 4.0. respectively.

095602-7 PETERSEN, BAZANT, PELLENQ, AND ULM PHYSICAL REVIEW MATERIALS 2, 095602 (2018) spinodal region. The most important finding of our model is the pressure P (which we have not explicitly defined), and that the thermodynamic landscape of the prenucleation clus- number of occupancies (clusters) N• constant, ters is instrumental in suppressing or enhancing mesoscopic       ∂G ∂G ∂V patterns. By increasing the stability of the clusters, which μ◦ = = ◦ ◦ concurrently lowers α, the range of the unstable growth modes ∂N N•,T ∂V φ,T ∂N N•,T of the Allen-Cahn reaction kernel narrows and shifts outside     δG ∂φ˜ the spinodal region of the Cahn-Hilliard diffusion kernel. + . (A1) ˜ ◦ nsδφ V,T ∂N N•,T Above, V is the system volume, and an equivalent expres- III. CONCLUSIONS sion can be written for μ• by replacing N◦ with N•.Using In summary, we have developed a mean-field, nonequi- Eq. (2)forG and noting the total number of sites as Ntot = librium thermodynamic model for interacting colloids that Vns = N◦ + N• with packing density φ˜ = N•/Ntot,Eq.(A1) respects experimental observations made at two essential readily yields length scales: Clusters nucleate as stable building blocks h h h 2 at the microscale [71], and aggregate into an arrested, out- μ◦ = g − φμ = kBT ln(1 − φ˜ ) + φ˜ + μres, (A2a) of-equilibrium structure at the mesoscale [7,55]. Including h = h + − h = ˜ +  ˜ − ˜ + a variable, entropic gradient energy penalty in Gibbs free μ• g (1 φ)μ kBT ln(φ) (φ 2)φ μ0. energy provides a physics-grounded approach to simulate the (A2b) evolution of mean-field gel patterns. Previous studies have 2 proposed control of the reaction rate to modulate dominant Here, φ˜ is the mean change in interaction energy at- wavelengths during phase separation [36,68,72]. But the ex- tributed to adding a new site into the lattice, while −2φ˜ pressions derived in Eqs. (14) and (15), which marry the measures the mean change in interaction energy upon placing reactive landscape of stable intermediates into the Allen-Cahn a cluster into that site. reaction equation [35], lend precision. They provide a tem- plate to manipulate bulk colloidal structures that emerge from APPENDIX B: LINEAR STABILITY ANALYSIS  stable prenucleation clusters by adjusting μ0 upon entering OF THE REACTION-DIFFUSION EQUATION the mobile portion the spinodal region. While, the present study focused on the response of homogeneous systems to This Appendix outlines the criterion for stability of our sudden changes in the normalized attractive strength and reaction-diffusion equation that describes the nonequilibrium  m thermodynamics of attractive colloids, Eq. (1). The derivation supersaturation of the solution (i.e., μ˜ 0 ), future work should assess how adjustments in the reservoir potential in course follows closely the procedure outlined by one of the authors of reaction can manipulate ensuing colloidal patterns. This in Ref. [68], which may be consulted for additional details. h could advance design of tailor-made colloidal structures that Starting from a homogeneous base state where φ˜(x) = φ˜ , enhance selected mechanical properties. the field is perturbed by small fluctuations δφ˜. We measure the aggregate strength of the fluctuations by the l2 norm of the perturbation field,  ACKNOWLEDGMENTS 1 L = (δφ˜ )2dV. (B1) The authors would like to thank Amir Pahlavan and 2 V Thibaut Divoux for insightful discussions and references. For a base state to be stable with respect to the dynamics Financial support was provided by the National Science Foun- L dation Graduate Research Fellowship, and the Concrete Sus- imposed by Eq. (1), must be a decreasing function of time. tainability Hub at the Massachusetts Institute of Technology That is,    (CSHub@MIT) with sponsorship provided by the Portland dL ∂δφ˜ Cement Association (PCA) and the Ready Mixed Concrete Stable if: = δφ˜ dV dt ∂t (RMC) Research and Education Foundation. Additional sup- V port was provided by the ICoME2 Labex (ANR-11-LABX- = δφ˜(−D(∇δφ˜ )2 + A(δφ˜ )2 )dV < 0, 0053) and the A*MIDEX projects (ANR-11-IDEX-0001-02) V cofunded by the French program “Investissements d’Avenir” (B2) managed by the French National Research Agency. where the chemical diffusion D and autocatalytic rate A,

APPENDIX A: CHEMICAL POTENTIAL δμ˜ D = L , (B3a) OF VACANCIES AND OCCUPANCIES δφ˜ We briefly outline relations for the chemical potentials of a δR ∂R ∂R δμ˜ ∂R ∂μ˜ A = = + + res , (B3b) ◦ • vacancy μ and an occupancy μ . Using the classical thermo- δφ˜ ∂φ˜ ∂μ˜ δφ˜ ∂μ˜ res ∂φ˜ dynamic definition of a chemical potential, μ◦ quantifies the h change in Gibbs free energy due to a change in the number are evaluated at φ˜(x) = φ˜ and L = Dφ/kBT is the Onsager of vacancies (voids) N◦ while keeping the temperature T , coefficient. Next, the variational derivative of the internal

095602-8 PHASE SEPARATION OF STABLE COLLOIDAL CLUSTERS PHYSICAL REVIEW MATERIALS 2, 095602 (2018) chemical potentialμ ˜ simplifies to can be rewritten as  dL = ˜ 2 δμ˜ δ2G dμ˜ h Stable if: ω (δφ) dV < 0, (B5) = = + κ˜ (∇δφ˜ )2, (B4) dt V ˜ ˜ 2 ˜ δφ nskBT (δφ) dφ with   ∂R ∂R ∂μ˜ once insignificant terms dependent on ∇φ˜ are removed. ω(φ˜ h,q) = + res ∂φ˜ ∂μ˜ res ∂φ˜ Lastly, we choose the perturbation to be sinusoidal with    ∇ ˜ = ˜ h Fourier frequency q,forwhich δφ qδφ. If the growth rate ∂R 2 dμ˜ 2 ˜ = ˜ + − Lq + κq˜ . (B6) of the perturbation is linearized as [∂(δφ)/∂t] ωδφ with ∂μ˜ dφ˜ growth rate coefficient ω, the stability criterion in Eq. (B2)

[1]P.J.LuandD.A.Weitz,Annu. Rev. Condens. Matter Phys. 4, [23] J. N. Israelachvili, Intermolecular and Surface Forces 217 (2013). (Academic Press, New York, 2011). [2] D. J. Kraft, R. Ni, F. Smallenburg, M. Hermes, K. Yoon, D. [24] H. Tanaka, Phys.Rev.Lett.76, 787 (1996). A. Weitz, A. van Blaaderen, J. Groenewold, M. Dijkstra, and [25] H. Tanaka and T. Araki, Phys. Rev. Lett. 78, 4966 (1997). W. K. Kegel, Proc. Natl. Acad. Sci. USA 109, 10787 (2012). [26] D. Gebauer, M. Kellermeier, J. D. Gale, L. Bergström, and H. [3] M. Grzelczak, J. Vermant, E. M. Furst, and L. M. Liz-Marzán, Cölfen, Chem. Soc. Rev. 43, 2348 (2014). ACS Nano 4, 3591 (2010). [27] J. Bergenholtz and M. Fuchs, Phys. Rev. E 59, 5706 (1999). [4] X. Zhou, Y.-X. Yin, L.-J. Wan, and Y.-G. Guo, Adv. Energy [28] K. Kroy, M. E. Cates, and W. C. K. Poon, Phys. Rev. Lett. 92, Mater. 2, 1086 (2012). 148302 (2004). [5] H. M. Jennings, Cem. Concr. Res. 30, 101 (2000). [29] M. E. Cates, M. Fuchs, K. Kroy, W. C. Poon, and A. M. Puertas, [6] E. Masoero, E. Del Gado, R. J.-M. Pellenq, F.-J. Ulm, and S. J. Phys.: Condens. Matter 16, S4861 (2004). Yip, Phys. Rev. Lett. 109, 155503 (2012). [30] F. Cardinaux, T. Gibaud, A. Stradner, and P. Schurtenberger, [7] K. Ioannidou, K. J. Krakowiak, M. Bauchy, C. G. Hoover, E. Phys. Rev. Lett. 99, 118301 (2007). Masoero, S. Yip, F.-J. Ulm, P. Levitz, R. J.-M. Pellenq, and E. [31] P. N. Segrè, V. Prasad, A. B. Schofield, and D. A. Weitz, Phys. Del Gado, Proc. Natl. Acad. Sci. USA 113, 2029 (2016). Rev. Lett. 86, 6042 (2001). [8] G. Girishkumar, B. McCloskey, A. Luntz, S. Swanson, and W. [32] U. Gasser, E. R. Weeks, A. Schofield, P. Pusey, and D. Weitz, Wilcke, J. Phys. Chem. Lett. 1, 2193 (2010). Science 292, 258 (2001). [9] B. Horstmann, T. Danner, and W. G. Bessler, Energy Environ. [33] E. Zaccarelli, J. Phys.: Condens. Matter 19, 323101 (2007). Sci. 6, 1299 (2013). [34] K. Ioannidou, M. Kanduc,ˇ L. Li, D. Frenkel, J. Dobnikar, and [10] M. Lattuada and T. A. Hatton, Langmuir 23, 2158 (2007). E. Del Gado, Nat. Commun. 7, 12106 (2016). [11] A. Ditsch, P. E. Laibinis, D. I. Wang, and T. A. Hatton, [35] M. Z. Bazant, Acc. Chem. Res. 46, 1144 (2013). Langmuir 21, 6006 (2005). [36] P. Bai, D. A. Cogswell, and M. Z. Bazant, Nano Lett. 11, 4890 [12] J. Baumgartner, A. Dey, P. H. Bomans, C. Le Coadou, P. Fratzl, (2011). N. A. Sommerdijk, and D. Faivre, Nat. Mater. 12, 310 (2013). [37] J. W. Cahn and J. E. Hilliard, J. Chem. Phys. 28, 258 [13] V. N. Manoharan, M. T. Elsesser, and D. J. Pine, Science 301, (1958). 483 (2003). [38] S. M. Allen and J. W. Cahn, Acta Metall. 27, 1085 (1979). [14] C. C. Carcouët, M. W. van de Put, B. Mezari, P. C. Magusin, [39] P. De Gennes, J. Chem. Phys. 72, 4756 (1980). J. Laven, P. H. Bomans, H. Friedrich, A. C. C. Esteves, N. A. [40] P.-G. De Gennes, Scaling Concepts in Polymer Physics (Cornell Sommerdijk, R. A. van Benthem et al., Nano Lett. 14, 1433 University Press, Ithaca, NY, 1979). (2014). [41] T. Gibaud, F. Cardinaux, J. Bergenholtz, A. Stradner, and P. [15] D. Gebauer, A. Völkel, and H. Cölfen, Science 322, 1819 Schurtenberger, Soft Matter 7, 857 (2011). (2008). [42] Z. Varga and J. Swan, Soft Matter 12, 7670 (2016). [16] P. Tan, N. Xu, and L. Xu, Nat. Phys. 10, 73 (2014). [43] T. A. Witten, Jr. and L. M. Sander, Phys. Rev. Lett. 47, 1400 [17] T. H. Zhang and X. Y. Liu, J. Am. Chem. Soc. 129, 13520 (1981). (2007). [44] P. Wiltzius, Phys. Rev. Lett. 58, 710 (1987). [18] V. Privman, D. V. Goia, J. Park, and E. Matijevic,´ J. Colloid [45] W. Hess, H. L. Frisch, and R. Klein, Z. Phys. B: Condens. Interface Sci. 213, 36 (1999). Matter 64, 65 (1986). [19] T. H. Zhang, J. Klok, R. H. Tromp, J. Groenewold, and W. K. [46] D. Stauffer, Phys. Rep. 54, 1 (1979). Kegel, Soft Matter 8, 667 (2012). [47] P. Meakin, Phys. Rev. Lett. 51, 1119 (1983). [20] A. I. Campbell, V. J. Anderson, J. S. van Duijneveldt, and P. [48] L. Cipelletti, S. Manley, R. C. Ball, and D. A. Weitz, Phys. Rev. Bartlett, Phys. Rev. Lett. 94, 208301 (2005). Lett. 84, 2275 (2000). [21] F. Sciortino, S. Mossa, E. Zaccarelli, and P. Tartaglia, Phys. Rev. [49] G. Adam and J. H. Gibbs, J. Chem. Phys. 43, 139 (1965). Lett. 93, 055701 (2004). [50] Idaho National Laboratory, USA, “Multiphysics object-oriented [22] A. Stradner, H. Sedgwick, F. Cardinaux, W. C. Poon, S. U. simulation environment”, Accessed: 08-20-2017. Egelhaaf, and P. Schurtenberger, Nature (London) 432, 492 [51] P. C. Hohenberg and B. I. Halperin, Rev. Mod. Phys. 49, 435 (2004). (1977).

095602-9 PETERSEN, BAZANT, PELLENQ, AND ULM PHYSICAL REVIEW MATERIALS 2, 095602 (2018)

[52] P. W. de Bont, G. M. van Kempen, and R. Vreeker, Food [63] R. J.-M. Pellenq and H. Van Damme, MRS Bull. 29, 319 (2004). Hydrocolloids 16, 127 (2002). [64] J. Groenewold and W. K. Kegel, J. Phys. Chem. B 105, 11702 [53] K. El-Mabrouk, M. Belaiche, and M. Bousmina, J. Colloid (2001). 2 ˜ Interface Sci. 306, 354 (2007). [65] Though ρc is inversely related to φ in closed systems, its [54] H. Tanaka, Phys. Rev. Lett. 71, 3158 (1993). dependence is less certain in open systems where ion exchange [55] P. J. Lu, E. Zaccarelli, F. Ciulla, A. B. Schofield, F. Sciortino, is possible and we hold it constant here. and D. A. Weitz, Nature (London) 453, 499 (2008). [66] A. M. Kuznetsov and J. Ulstrup, Electron Transfer in Chemistry [56] I. M. Lifshitz and V. V. Slyozov, J. Phys. Chem. Solids 19, 35 and Biology: An Introduction to the Theory (John Wiley & Sons, (1961). Ltd., New York, 1999). [57] E. D. Siggia, Phys. Rev. A 20, 595 (1979). [67] D. A. Cogswell and M. Z. Bazant, ACS Nano 6, 2215 (2012). [58] A. Bailey, W. C. K. Poon, R. J. Christianson, A. B. Schofield, [68] M. Z. Bazant, Faraday Discuss. 199, 423 (2017). U. Gasser, V. Prasad, S. Manley, P. N. Segre, L. Cipelletti, W. [69] F. Trudu, D. Donadio, and M. Parrinello, Phys. Rev. Lett. 97, V. M eyer et al., Phys. Rev. Lett. 99, 205701 (2007). 105701 (2006). [59] V. Testard, L. Berthier, and W. Kob, Phys.Rev.Lett.106, [70] S. Torquato, Random Heterogeneous Materials: Microstructure 125702 (2011). and Macroscopic Properties (Springer Science & Business [60] M. C. Grant and W. B. Russel, Phys.Rev.E47, 2606 (1993). Media, New York, NY, 2013), Vol. 16. [61] M. L. Williams, R. F. Landel, and J. D. Ferry, J. Am. Chem. [71] D. A. Weitz and M. Oliveria, Phys. Rev. Lett. 52, 1433 Soc. 77, 3701 (1955). (1984). [62] J. F. Banfield, S. A. Welch, H. Zhang, T. T. Ebert, and R. L. [72] S. C. Glotzer, E. A. Di Marzio, and M. Muthukumar, Phys. Rev. Penn, Science 289, 751 (2000). Lett. 74, 2034 (1995).

095602-10