Introduction

Compressing and/or cooling a liquid may lead to its crystalline solidification or to its gas-liquid phase separation1. Gases, liquids and (crystalline) solids are equilibrium phases, in principle understood in terms of molecular interactions thanks to Boltzmann’s fundamental expression for the entropy, S = kB lnW, and to the maximum-entropy principle2. A simple and intuitive use of these principles is illustrated by van der Waals explanation of the origin of gases and liquids and their coexistence3,4. In practice, however, compressing and/or cooling a liquid may also lead to the formation of compact and relatively homogeneous amorphous solids (glasses) and of more tenuous sponge-like amorphous materials (gels), that pervade our daily life5,6,7.

So far, however, we do not know how Boltzmann’s principle operates to govern the formation of these non-equilibrium amorphous solids that do not maximize entropy. Such fundamental knowledge would allow us to systematically build an understanding (even at “van der Waals” level) of the exploding experimental and computational information8,9 on the general phenomenology of these non-equilibrium phases. One of the most illustrative examples is the amazing behavior observed when the ordinary equilibrium gas-liquid phase transition interferes with the non-equilibrium glass transition10, leading to remarkable multistep relaxation processes11,12,13 and puzzling delay (or “latency”) effects14,15,16 during the formation of gels by arrested spinodal decomposition.

The main aim of this work is to demonstrate that a microscopic statistical mechanical framework to understand this interplay between gas-liquid spinodal decomposition, gelation and the glass transition, results from a straightforward application of the non-equilibrium self-consistent generalized Langevin equation (NE-SCGLE) theory of irreversible processes in liquids17,18 to a Lennard-Jones–like (“LJ-like”) simple liquid (pair interaction = harsh repulsion + longer-ranged attraction).

The NE-SCGLE theory originates from a generalization of Onsager’s description of irreversible processes and fluctuations, to genuine non-equilibrium and non-linear conditions. Applied to the description of irreversible processes in liquids, this canonical and abstract formalism becomes a generic first-principles theory of its structure and dynamics, at equilibrium and during the non-stationary processes of equilibration or aging. In the Supplemental Material (SM) we briefly review the genesis and achievements of this non-equilibrium theory, and guide the reader through the pertinent references. There we also write in detail the self-consistent system of equations (Eqs (SM6)–(SM11)) that summarizes the simplest approximate version of the NE-SCGLE theory employed in this and in previous studies18,19,20,21,22.

As recently shown19, the transformation of equilibrium hard-sphere (and soft-sphere) liquids into “repulsive” glasses, provided by the solution of these NE-SCGLE equations, naturally explains some of the most essential nonequilibrium signatures of the glass transition (such as the aging of the structure and dynamics)5,6,7. When applied to LJ-like simple liquids, the same NE-SCGLE equations predict new dynamically-arrested phases, identified with gels and porous glasses20, and provides a kinetic perspective of the irreversible evolution of the structure of the system after being instantaneously quenched to the interior of its spinodal region21,22. The main aim of the following discussion is to illustrate the NE-SCGLE prediction that the seemingly complex interplay between spinodal decomposition, gelation, glass transition, and their combinations11,12,13, may be simply understood in terms of two kinetically competing limiting behaviors, associated with the two underlying dynamic arrest transitions predicted to exist20 in the gas-liquid spinodal region by the very same NE-SCGLE Equation.

Results and Discussion

As in refs.20,21,22, here we also refer for concreteness to a representative Brownian LJ–like model system, namely, the “hard-sphere plus attractive Yukawa” (HSAY) potential

$$u(r)=\{\begin{array}{lc}\infty , & r < \sigma ;\\ -\epsilon \frac{\exp [\,-\,z(r/\sigma -\mathrm{1)]}}{(r/\sigma )}, & r > \sigma ,\end{array}$$
(1)

whose state space is spanned by the number density n and the temperature T, expressed in dimensionless form as [3] and [kBT/\(\epsilon \)] (with kB being Boltzmann’s constant), and denoted simply as n and T. We shall also refer to the hard-sphere volume fraction ϕ ≡ πn/6. Complementing the recent study of the evolution of the non-equilibrium structure factor S(k; tw)21, in this paper we shall discuss the full solution of the referred NE-SCGLE equations at all waiting times tw, but focusing now on the kinetics (i.e., the aging) of the non-equilibrium dynamic properties.

Let us assume that this system, initially in equilibrium at the state point (n, Ti), is instantaneously quenched at time tw = 0 to a final temperature Tf under isochoric conditions and in the absence of applied external fields. Then the NE-SCGLE description of the irreversible evolution of S(k; tw) for waiting times tw > 0 is provided20 by the solution of Eq. SM6, i.e.,

$$\frac{\partial S(k;{t}_{w})}{\partial {t}_{w}}=-\,2{k}^{2}{D}^{0}b({t}_{w})n{ {\mathcal E} }_{f}(k)[S(k;{t}_{w})-1/n{ {\mathcal E} }_{f}(k)]\mathrm{.}$$
(2)

In this equation D0 is the short-time self-diffusion coefficient (see the SM) and \({ {\mathcal E} }_{f}(k)\) is the Fourier transform (FT) of \( {\mathcal E} (r;n,{T}_{f})\), which is the functional derivative \( {\mathcal E} [|r-r^{\prime} |;n,T]\equiv [\delta \beta \mu [r;n,T]/\delta n(r^{\prime} )]\) of the local chemical potential (in units of the thermal energy kBT ≡ 1/β), evaluated at the uniform density and temperature profiles n(r) = n and T(r) = Tf. As in refs20,21, \({ {\mathcal E} }_{f}(k)\) will be approximated by its exact hard-sphere value \({ {\mathcal E} }_{HS}(k;\varphi )\) plus the FT of the attractive Yukawa tail, \({ {\mathcal E} }_{f}(k;n,{T}_{f})={ {\mathcal E} }_{HS}(k;\varphi )+\beta {u}_{A}(k)\).

The dimensionless mobility function b(tw) in Eq. 2 is in reality a state function, determined by the full set of NE-SCGLE equations, constituted by Eq. (2) itself, together with Eqs. (SM7)-(SM11). From the analysis of the long-time asymptotic solutions of these equations, it was shown in ref.20. that the liquid-glass transition of Lennard-Jones–like liquids (solid line of Fig. 1(a) for the HSAY model) indeed intersects the spinodal line, as previously discovered by simulations10. The NE-SCGLE theory, however, goes beyond, to predict in addition that below the intersection point (ϕb, Tb), the liquid-glass transition bifurcates in two dynamic arrest transitions (dotted and dashed lines of Fig. 1(a), explained in detail in ref.20). In what follows we shall demonstrate that, in addition, the other observed fingerprints of the interference between the glass and the gas-liquid transitions are also part of the unified and detailed scenario predicted by the NE-SCGLE theory.

Figure 1
figure 1

(a) Dynamic arrest transitions of the HSAY liquid (z = 2) from ref.20: liquid-glass (solid line) and gel-glass (dashed line) transitions, and “type A” dynamic arrest transition (dotted line), which coincides with the spinodal line T = Ts(ϕ) to the left of the bifurcation point (solid diamond). Downward arrows 1–3 represent instantaneous temperature quenches discussed in the text. (b) Thick solid line: waiting-time dependence of the (inverse) mobility b(t) for quench 1 (final temperature Tf(1) = 0.6), exhibiting the plateau bl(Tf) lasting a latency time tl(Tf). The other lines correspond to other values of Tf in the neighborhood but below (softer solid lines) and above (dashed line) the spinodal. The inset compares the predicted divergence tl(T)  (1 − T/Ts)−2.5 (solid line) with the experimental data of ref.14 (circles), fitted by the same power law (dashed line).

The three downward arrows in Fig. 1(a) represent instantaneous temperature quenches along the isochore ϕ = 0.2 from a common initial temperature Ti = 1.5, to the indicated lower final temperature Tf. Quench 1 represents the regime of shallow quenches. Its final temperature Tf(1) = 0.6 lies near but below the first non-equilibrium transition (dotted line). This “type-A” transition20 falls on top of the gas-liquid spinodal curve T = Ts(ϕ), and for this isochore, Ts(ϕ = 0.2) = 0.764. Quenches 2 and 3 are representative of the regime of deep quenches, where Tf lies in the neighborhood of the lower-temperature (“type-B”) gel-glass transition T = Tc(ϕ)20 (dashed line). This is clearly just the continuation inside the spinodal region, of the “ordinary” liquid-glass transition (solid line of Fig. 1(a)).

The irreversible structural relaxation that follows these quenches manifests itself most dramatically, and with rather unexpected consequences, in the evolution of the non-equilibrium dynamics, which exhibits a complex time-temperature-density dependence, as illustrated and summarized by the mobility b(tw). For example, the thick solid line of Fig. 1(b), labeled Tf = 0.6, corresponds to quench 1, and illustrates the most salient kinetic feature of shallow quenches: the inverse b−1(tw), expected to mimic the structural relaxation time or the viscosity, exhibits a remarkable “latency” period, of duration tl, in which it reaches a pseudo-equilibration plateau with constant value bl−1. Within this latency time, the particles are able to diffuse a distance \({d}_{l}\equiv \sqrt{{b}_{l}{t}_{l}/6}\). At the end of this latency period, b−1(tw) diverges with tw as dynamic arrest now fully develops.

As illustrated by the other (thinner) solid lines of Fig. 1(b) (which represent other shallow quenches), tl, bl and dl depend on the final temperature Tf of the quench, increasing monotonically when Tf approaches Ts from below. A more detailed calculation, illustrated in the inset, reveals that the predicted tl(Tf) actually diverges at the spinodal temperature Ts as tl (1 − Tf/Ts)α, with an exponent α ≈ 2.5. It is natural to expect that this kinetic behavior of b−1(tw) will also be observed more directly in the aging kinetics of the viscosity and of the α-relaxation time. In fact, it is quite remarkable that this predicted latency effect has actually been experimentally observed in rheometry experiments during gel formation in weakly attractive nanocolloid suspensions14. Thus, for quenches along the isochore ϕ = 0.2, the experimental latency time, denoted in14 as tG(Tf), was observed to diverge as tG (1 − Tf/T*)α with α ≈ 2.5 (see Fig. 7(a) of14). The agreement of our theoretical prediction with these results, presented in the inset of Fig. 1(b), and discussed in more detail in the Supplemental Material, allows us to identify the singular temperature T* empirically determined in ref.14, with the spinodal temperature Ts.

In contrast to approaching the spinodal line from below, let us now go in the opposite direction (i.e., decreasing Tf toward Tc). We then observe that tl(Tf) and bl(Tf) decrease with Tf until the latency plateau transforms into a mild and fading inflection point. This is illustrated by the solid curves in the inset of Fig. 2(a). The dashed lines correspond to final temperatures below the gel-glass transition T = Tc(ϕ), and serve to illustrate the complete absence of latency effects below Tc. In fact, the main feature to highlight in Fig. 2 is precisely this striking kinetic difference, predicted to occur when Tf crosses the gel-glass transition line. This is illustrated by quenches 2 and 3, chosen such that Tf lies, respectively, slightly above (Tf(2) = 0.23) and slightly below (Tf(3) = 0.18) this transition (which occurs at Tc(ϕ = 0.2) = 0.22). In Fig. 2(a,b) this kinetic difference is exhibited in more detail in terms of the different pattern of aging of the mean squared displacement W(τ; tw, Tf) ≡ <[R(tw + τ) − R(tw)]2>/6σ2 of these two quenches, plotted as a function of the correlation time τ (solid lines), for the same sequence of values of the waiting time tw (equally-spaced in log tw).

Figure 2
figure 2

Mean square displacement W(τ; tw, Tf) as a function of correlation time τ for waiting times tw = 0 (line with dark circles), tw = 100.5n, with n = −3, −2, −1, … (solid lines), and tw = ∞ (dashed lines) after a quench to final temperature Tf(2) = 0.23 (main panel (a)), Tf(3) = 0.18 (main panel (b)) and Tf(1) = 0.60 (inset (b)). For reference we also include the asymptotic curves Wa(τ; Tf) ≡ W(τ; tw = ∞, Tf) (dotted lines) for Tf = 0.18 in (a) and Tf = 0.18 and 0.23 in the inset of (b). The inset of (a) contrasts the behavior of (the inverse of) b(t; Tf) for Tf slightly above (solid lines) and below (dashed lines) the gel-glass transition line.

As observed in Fig. 2(a), quench 2 presents a remnant of latency effects, observed as the inflection point in the tw-dependence of b−1(tw) in the inset of Fig. 2(a) (thicker solid line), and in the crowding of the solid lines for W(τ; tw, Tf(2)) corresponding to \({t}_{w}\lesssim {t}_{l}\), relative to the kinetics of quench 3, presented in Fig. 2(b). The latter is pretty similar to the more uniform and faster aging pattern predicted and observed in the formation of repulsive glasses (see, e.g., Fig. 15 of23). Such pattern involves an asymptotic localization length λHS of about 0.14 σ, a value reminiscent of Lindemann’s criterion of melting, suggestive of a hard-sphere caging mechanism of arrest24. This is consistent with the formation of compact, but porous, repulsive glasses. In fact, a localization length λHS ≈ 0.14σ is predicted all along the gel-glass and liquid-glass portions of the glass transition line, up to the hard-sphere glass-transition at (ϕ = 0.582, T = ∞).

The other most notorious fingerprint of the interference between gelation and glass transition, predicted by the NE-SCGLE theory, is also illustrated by quench 2 in Fig. 2(a). This is in reality another manifestation of the inflection point exhibited by b−1(tw) for Tf slightly above Tc. We refer to the build-up, at long waiting times, of a two-step long-τ relaxation of the main dynamic properties. Notice, for example, that for waiting times smaller than the latency time (\({t}_{w}\lesssim {t}_{l}\)), W(τ; tw, Tf(2)) bends towards the asymptotic MSD Wa(τ; T(3)) ≡ W(τ; tw = ∞,Tf(3)) of quench 3 (dotted line, with localization length λHS ≈ 0.14σ), thus suggesting an initial tendency to arrest through a hard-sphere caging mechanism.

At waiting times tw much longer than the latency time tl, however, this behavior is “corrected”, now bending toward the “true” predicted asymptotic MSD Wa(τ; T(2)) ≡ W(τ; tw = ∞, Tf(2)) of quench 2 (dashed line), whose localization length λgel is a few times larger, reflecting the localization of the particles within the transient particle network of the gel phase12. This is highly reminiscent of the early proposal25 that colloid gelation is the result of a two-level dynamical arrest process, first at the level of the colloidal particles leading to clusters, and then at the level of clusters undergoing a glass transition. This notion, first formalized using mode coupling theory26, and later framed27 in terms of the Cauchy-Born theory of amorphous solids28, has been recently confirmed by experimental observations using different techniques29. The present NE-SCGLE approach, whose equilibrium version is conceptually closer to MCT (see a recent and detailed discussion in ref.30), provides a complementary kinetic perspective, predicting the two-stage aging of the dynamics without assuming a priori a two-level arrest scenario.

It is important to mention that this complex pattern of aging has been observed in the simulation of the non-equilibrium dynamics of suspensions of HS-like particles transiently bonded by cross-linking polymeric agents, whose net effect is an effective attraction between the particles (see, e.g., Fig. 4 of11 and Fig. 11 of13). Although not shown here (see, however, the SM), the corresponding two-step relaxation is also exhibited by the self-intermediate scattering function (self-ISF) FS(k, τ; tw) ≡ <exp ik · [R(tw + τ) − R(tw)]>.

The main features of the NE-SCGLE predicted scenario of the interference between the gas-liquid and the glass transitions may thus be summarized by the conclusion that the latency time and the double relaxation are in reality the most representative manifestations of two competing and complementary extreme behaviors. The first of them was discussed in Fig. 1(b), and refers to the appearance of latency effects associated with the dynamic arrest transition line T = Ts(ϕ) (at which the latency time tl(Tf) diverges). Thus, while for Tf above Ts the system will always reach a homogeneous equilibrium state, for Tf below Ts no homogeneous equilibration is possible. Instead, for Tf below but sufficiently close to Ts, no signs of arrest may be observed, due to the very long latency time tl(Tf), thus allowing the system to undergo full inhomogeneous gas-liquid equilibrium phase separation.

Notice in the inset of Fig. 2(b) that in this regime of shallow quenches, the two-step relaxation is virtually absent. The occurrence of inhomogeneous phase separation, however, renders this observation rather irrelevant in practice. Nevertheless, it is still interesting to notice that the divergence of tl(Tf) and dl(Tf) when Tf approaches Ts from below, is a non-equilibrium dynamic counterpart of the diverging equilibrium correlation length31 expected when we approach the critical point (or any point along the spinodal curve) from above. Both originate in the thermodynamic instability represented by the spinodal line, information that enters in the NE-SCGLE theory through the thermodynamic input \( {\mathcal E} (k;n,{T}_{f})\).

The second complementary extreme behavior is associated with crossing the gel-glass transition line T = Tc(ϕ). Thus, very deep quenches (Tf below Tc, illustrated by quench 3) lead to the formation of porous structures made of compact “repulsive” glasses (localization length λHS ≈ 0.14σ). In contrast, slightly shallower quenches (Tf immediately above Tc, quench 2), are predicted to form “fluffier” and less rigid (i.e., more viscoelastic) and highly heterogeneous materials, with localization lengths λgel several times larger than λHS. The kinetic and dynamical fingerprint of these materials, which we identify with gels, is the two-step pattern of structural relaxation illustrated in Fig. 2(a) by quench 2.

According to the predicted scenario, quenches whose Tf lies midway between these two extreme regimes, will involve kinetic processes featuring a combination of these two limiting tendencies. However, since near the spinodal line the unrestricted tendency to gas-liquid phase separation is expected to prevail over the predicted non-equilibrium divergence of tl(Tf) and dl(Tf), we must conclude that some form of smooth crossover, from full gas-liquid separation to the formation of gels, must occur somewhere between Ts and Tc. Although this would then imply that there is nothing like a sharp gel line, we might determine the “epicenter” T0(ϕ) of this postulated crossover, as indicative of the rather diffuse boundary between heterogeneous gas-liquid separation and gel formation.

In ref.20 an empirical criterion was suggested to determine T0(ϕ), based on the Tf-dependence of the infinitely “aged” (tw → ∞) mean square displacement, Wa(τ; Tf) ≡ W(τ; tw = ∞, Tf) (illustrated by the dotted and dashed lines in Fig. 2(a,b)). The arrested plateau Wa(τ → ∞, Tf) defines the long-time asymptotic value of the squared localization length, λ2(Tf) ≡ Wa(τ → ∞, Tf), whose Tf-dependence is illustrated in Fig. 3(a) for the isochore ϕ = 0.2. For Tf above but near Tc, λ(Tf) increases exponentially with Tf, whereas for Tf below but near Ts, λ(Tf) diverges as (1 − Tf/Ts)ν with ν = 0.75. As explained in ref.20. (where λ2(Tf) is denoted as γa(Tf)), this allows us to determine a crossover temperature T0 between these two regimes. Applying this procedure at other isochores led to the determination of T0(ϕ) reported in Fig. 10 of ref.20, and reproduced here as the empty circles of Fig. 3(a).

Figure 3
figure 3

(a) Numerical results for the inverse localization length λ−1(Tf) (empty circles) and the inverse latency distance dl−1(Tf) (solid circles) as a function of Tf, along the isochore ϕ = 0.2. As discussed in ref.20, the crossover from exponential to power-law divergence of λ−1(Tf) occurs at T0(ϕ) (vertical solid line). As explained in the SM, the same procedure can be applied to dl−1(Tf), yielding ν = 1.25, and an alternative determination of T0(ϕ). (b) Crossover temperature T0(ϕ) obtained from λ(Tf) (empty circles) and from dl(Tf) (solid circles), represented on the non-Equilibrium phase diagram of Fig. 1(a). Regions I, II, and III correspond, respectively, to full gas-liquid phase separation, to gel formation, and to the formation of porous glasses.

Let us now mention that the same notion, derived in ref.20 from long-time asymptotic properties, also emerges from the Tf-dependence of finite-waiting-time properties, more specifically, of the latency distance dl(Tf). This provides us with a similar but independent determination of T0(ϕ). To see this, in Fig. 3(a) we present the predicted dl(Tf) as a function of Tf along the isochore ϕ = 0.2. As explained in detail in the SM, the Tf-dependence of dl(Tf) also happens to exhibit an exponential to power-law crossover, and hence, we can also determine a corresponding crossover temperature T0(ϕ). The result of this procedure are illustrated by the solid circles of Fig. 3(b).

Clearly, both routes determine essentially the same location of the crossover temperature T0(ϕ). Although we have not demonstrated that these points correspond to the gel line, this coincidence is reassuring, and we may take them as indicative of the diffuse boundary, above which the system is more likely to phase separate, and below which it is more likely to form a gel. With this provision, the coincidence of both methods to determine T0(ϕ), lends additional support to the scenario proposed in ref.20, that regions I (T0(ϕ) < Tf < Ts(ϕ)), II (Tc(ϕ) < Tf < T0(ϕ)), and III (Tf < Tc(ϕ)) correspond, respectively, to full gas-liquid phase separation, to gel formation, and to the formation of porous glasses. This predicted non-equilibrium phase diagram (NEPD) is strongly reminiscent of the experimentally-determined NEPD reported in Fig. 4 of ref.32 and of the theoretically-proposed NEPD presented in Fig. 3 of ref.27.

Conclusion

Let us conclude with a word of cautious optimism. The scenario presented here and in refs20,21,22 illustrates the general features predicted by the NE-SCGLE theory regarding the non-equilibrium structural and dynamical evolution leading to the formation of arrested states after quenching a LJ-like simple liquid inside its spinodal region. Some detailed features, however, may depend on the detailed conditions, whose discussion was left out of the scope of this manuscript. We refer, for example, to the variation of the form and range of the pair potential and of the specific thermal manipulation protocol (beyond the most primitive one considered in this work, namely, the instantaneous homogeneous temperature quench).

The good news is that there seems to be no fundamental obstacles to incorporate these effects in additional applications of the NE-SCGLE theory, as we expect to illustrate in separate communications. For the time being, we may conclude that the qualitative scenario provided by the non-equilibrium SCGLE theory, within its simplest version and assumptions, seems to provide a sound and illuminating perspective to the understanding of the interference between one equilibrium phase transition (the gas-liquid phase separation in the present case) and a non-equilibrium kinetic arrest transition (represented here by the glass transition).