COMPUTER SIMULATIONS OF LIQUID SILICA: WATER-LIKE THERMODYNAMIC AND DYNAMIC ANOMALIES, AND THE EVIDENCE FOR POLYAMORPHISM

Similar documents
Liquid-Liquid Phase Transitions and Water-Like Anomalies in Liquids

Phase diagram of silica from computer simulation

LIQUID-LIQUID CRITICAL POINT

By definition, in a first-order phase transition, thermodynamic

Relationship between the Potential Energy Landscape and the Dynamic Crossover in a Water-Like Monatomic Liquid with a Liquid-Liquid Phase Transition

Molecular structural order and anomalies in liquid silica

Structural order in glassy water

Structure of the First and Second Neighbor Shells of Water: Quantitative Relation with Translational and Orientational Order.

Structure of the first- and second-neighbor shells of simulated water: Quantitative relation to translational and orientational order

arxiv: v1 [cond-mat.stat-mech] 3 Oct 2016

Monte Carlo simulation of confined water

Free Energy Surface of Supercooled Water

T he ubiquity and biological importance of water, as well as its numerous puzzling anomalous properties, have

Silica through the eyes of colloidal models when glass is a gel

Liquid Liquid Coexistence in NaCl Aqueous Solutions: A Simulation Study of Concentration Effects

Response functions near the liquid-liquid critical point of ST2 water

SUPPLEMENTARY INFORMATION. Erasing no-man s land by thermodynamically stabilizing the liquid liquid transition in tetrahedral particles

Dynamics of water at low temperatures and implications for biomolecules

arxiv:cond-mat/ v1 [cond-mat.mtrl-sci] 17 Dec 1996 silica Abstract

Structure of supercooled and glassy water under pressure

ABSTRACT. Title of Thesis: THERMODYNAMICS OF FLUID POLYAMORPHISM

arxiv:cond-mat/ v1 [cond-mat.dis-nn] 27 Mar 1997

Waterlike hierarchy of anomalies in a continuous spherical shouldered potential

Supplementary Information for. Universal elastic-hardening-driven mechanical instability in α-quartz and quartz. homeotypes under pressure

Fast and slow dynamics of hydrogen bonds in liquid water. Abstract

arxiv:cond-mat/ v2 [cond-mat.soft] 27 Aug 1999

The Rise and Fall of Anomalies in Tetrahedral Liquids.

Understanding the fundamental role that water plays on

Physics 563 Term Paper Liquid-Liquid Phase Transition

COOPERATIVE ORIGIN OF LOW-DENSITY DOMAINS IN LIQUID WATER. Jeffrey R. Errington, Pablo G. Debenedetti *, and Salvatore Torquato

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 8 Oct 1996

arxiv: v1 [cond-mat.soft] 27 Apr 2007

Chapter 11. Intermolecular Forces and Liquids & Solids

Equation of State of the Energy Landscape of SPC/E Water

Water: A Tale of Two Liquids

arxiv: v2 [cond-mat.soft] 24 Sep 2008

Clusters of mobile molecules in supercooled water

arxiv: v1 [cond-mat.stat-mech] 6 Sep 2018

Nucleation rate (m -3 s -1 ) Radius of water nano droplet (Å) 1e+00 1e-64 1e-128 1e-192 1e-256

arxiv: v2 [physics.chem-ph] 1 Apr 2013

Chapter 4. The Effect of Elastic Softening and Cooperativity on the Fragility of

The puzzle of Water Behavior at Low Temperature

3 - Ice and crystalline water

Thermodynamics, structure, and dynamics of water confined between hydrophobic plates

Thermodynamic and structural aspects of the potential energy surface of simulated water

arxiv: v2 [physics.chem-ph] 11 Aug 2017

Supercooled and glassy water: Metastable liquid(s), amorphous solid(s), and a no-man s land

SUPERCRITICAL, SUPERCOOLED: WATER AT LOW TEMPERATURES

Intermolecular Forces and Liquids and Solids

Edited by H. Eugene Stanley, Boston University, Boston, MA, and approved January 9, 2009 (received for review November 6, 2008)

Diversity in liquid supercooling and glass formation phenomena illustrated by a simple model

Search for a liquid-liquid critical point in models of silica

Connecting diffusion and entropy of bulk water at the single particle level

Intermolecular Forces and Liquids and Solids

University of Bristol - Explore Bristol Research. Peer reviewed version. Link to published version (if available): /acs.chemrev.

Name: Class: Date: SHORT ANSWER Answer the following questions in the space provided.

Two-State Thermodynamics and the Possibility of a Liquid-Liquid Phase Transition in Supercooled TIP4P/2005 Water

Liquid liquid phase transition in Stillinger Weber silicon

First-principles simulations of liquid silica: Structural and dynamical behavior at high pressure

Molecular dynamics study of isobaric and isochoric glass transitions in a model amorphous polymer


Chap. 2. Polymers Introduction. - Polymers: synthetic materials <--> natural materials

Dynamic crossover in deeply cooled water confined. in MCM-41 at 4 kbar and its relation to the liquidliquid

Research article Evolution of structure of SiO 2 nanoparticles upon cooling from the melt Nguyen Thi Xuan Huynh* 1, Vo Van Hoang* 2 and Hoang Zung 3

SUPPLEMENTARY INFORMATION

Vitrification and glass transition of water: insights from spin probe ESR

Supporting Information. How Different are the Characteristics of Aqueous Solutions of tert-

Ab initio Rutile-Cristobalite Transitions in Silicon Dioxide and Titanium Dioxide

Silica glass structure generation for ab initio calculations using small samples of amorphous silica

Crossover to potential energy landscape dominated dynamics in a model glass-forming liquid

Phase Diagram and Thermodynamic and Dynamic Anomalies in a Pure Repulsive Model

Correlation between local structure and dynamic heterogeneity in a metallic glass-forming liquid

Amorphous state studies with the Gaussian core model

Web Course Physical Properties of Glass. Range Behavior

5. STRUCTURES AND DEFECTS IN AMORPHOUS SOLIDS

A macroscopic model that connects the molar excess entropy of a supercooled liquid near its glass transition temperature to its viscosity

Chapter 10: Liquids, Solids, and Phase Changes

Glasses and Jamming: lessons from hard disks in a narrow channel

Phenomenological Theories of Nucleation

PHYSICAL REVIEW B 68,

Chapter 19. Chemical Thermodynamics

CHEM-UA 652: Thermodynamics and Kinetics

Chapter 10. Lesson Starter. Why did you not smell the odor of the vapor immediately? Explain this event in terms of the motion of molecules.

Liquid-liquid transition in ST2 water

Density anomaly in core-softened lattice gas

Molecular Dynamics Simulation of the Thermodynamic and Structural Properties for the CaO SiO 2 System

Chapter 19. Chemical Thermodynamics

Aging and energy landscapes: application to liquids and glasses

Prediction of entropy and dynamic properties of water below the homogeneous nucleation temperature

The Gibbs Phase Rule F = 2 + C - P

Intermolecular Forces and Liquids and Solids. Chapter 11. Copyright The McGraw Hill Companies, Inc. Permission required for

Physical transformations of pure substances Boiling, freezing, and the conversion of graphite to diamond examples of phase transitions changes of

Metallic & Ionic Solids. Crystal Lattices. Properties of Solids. Network Solids. Types of Solids. Chapter 13 Solids. Chapter 13

STRONG CONFIGURATIONAL DEPENDENCE OF ELASTIC PROPERTIES OF A CU-ZR BINARY MODEL METALLIC GLASS

Transitions between Inherent Structures in Water

Chapter 19. Chemical Thermodynamics

Dynamics of Supercooled Liquids The Generic Phase Diagram for Glasses

A Molecular Dynamics Simulation of a Homogeneous Organic-Inorganic Hybrid Silica Membrane

Cooling rate effects in amorphous Silica: A Computer Simulation Study. Abstract

Supplementary material for Universal stretched exponential relaxation in nanoconfined water

Transcription:

COMPUTER SIMULATIONS OF LIQUID SILICA: WATER-LIKE THERMODYNAMIC AND DYNAMIC ANOMALIES, AND THE EVIDENCE FOR POLYAMORPHISM IVAN SAIKA-VOIVOD 1 and PETER H. POOLE 2 1 Department of Physics and Physical Oceanography, Memorial University of Newfoundland, St. John s, NL, A1B 3X7, Canada 2 Department of Physics, St. Francis Xavier University, Antigonish, NS, B2G 2W5, Canada CONTENTS I. Introduction II. Simulations: Rigid-Ion Models of Silica III. Amorphous Solid Behavior IV. Water-Like Thermodynamic and Dynamic Anomalies V. Evidence for a Liquid Liquid Phase Transition VI. Outlook References I. INTRODUCTION The qualitative similarities between water and silica have long been appreciated [1,2]. The molecular structure of both substances is dominated by tetrahedral bonding arrangements under ambient pressure, in both the liquid and the solid states. As a consequence, several of the low-pressure ice phases have isomorphic analogs among the crystalline phases of silica (e.g., ice I h and tridymite). The similarities between water and silica also extend to the properties of the liquid and amorphous solid forms. Water and silica are both members of a small set of substances that exhibit a density maximum in the liquid phase [1,2]. Liquid Polymorphism: Advances in Chemical Physics, Volume 152, First Edition. Edited by H. Eugene Stanley. 2013 John Wiley & Sons, Inc. Published 2013 by John Wiley & Sons, Inc. 373

374 ivan saika-voivod and peter h. poole In addition, for both substances, when the amorphous solid formed at ambient pressure is compressed, the material collapses to a higher density amorphous solid over a relatively narrow range of pressure [3 5]. Such polyamorphic behavior of the amorphous solid has been interpreted, in the case of water, as a subglass transition manifestation of an underlying thermodynamic instability in the supercooled liquid, specifically, a liquid liquid phase transition (LLPT) occurring without a change in the chemical composition [6 8]. As a consequence of these similarities, silica was one of the first substances to be examined for an LLPT after the initial proposal that such a phenomenon occurs in supercooled water [9]. Because of the considerable challenges associated with the determination of the equation of state of real molten silica over the required range of pressure P and temperature T, computer simulations of silica-like models have played a central role in addressing the question of whether an LLPT might occur in this system. Here, we review the findings of simulation studies of silica related to the phenomena of polyamorphism and LLPTs. As we discuss below, these simulation studies provide ample evidence that the thermodynamic and dynamical behavior of liquid silica follow the pattern of anomalies that are well known in the case of water. However, simulations have yet to provide unequivocal evidence for an LLPT in silica. Hence, there remains important work to be carried out on this system. II. SIMULATIONS: RIGID-ION MODELS OF SILICA A large number of interaction potentials have been formulated for modeling silica interactions in computer simulations. In this review, we will focus primarily on results obtained using so-called rigid-ion models, in which silicon and oxygen atoms are modeled essentially as charged soft spheres. These interaction potentials consist of a sum of Coulombic interactions and repulsive interactions of the Born Mayer Huggins form [10]. The long-range nature of the Coulombic interactions are typically treated using Ewald summation. The specific parametrizations of the rigid-ion approach that we will discuss below are that of Wookcock, Angell, and Cheeseman (WAC) [11] and van Beest, Kramer, and van Santen (BKS) [12]. The WAC and BKS potentials ignore the polarizability of the O and Si atoms. These potentials also ignore the covalent nature of the bonding occurring in silica, and thus lack, for example, any explicit three-body interactions such as bond-angle bending terms. As a consequence, significant quantitative discrepancies occur between the behavior of these models and the real silica. However, the simple form of these potentials has a practical advantage in that they facilitate the long simulations required to model the supercooled liquid state, especially as T decreases and relaxation times increase dramatically. These models also provide a stark contrast to the models most commonly used to simulate water, in which O and H atoms

computer simulations of liquid silica 375 are connected into rigid molecular units, with a nearly perfect tetrahedral angle built into the shape of the molecule. The WAC and BKS models have no explicit propensity for tetrahedral bonding built into the potential. The ability of these models to form tetrahedral structures arises spontaneously from the combined influences of the relative charge, stoichiometry, and size of the O and Si ions. The fact that rigid molecular models of water and rigid-ion models of silica can both produce a common pattern of thermodynamic and dynamic behavior is a striking demonstration of the robustness of the phenomena common to both systems, and strongly suggests that the common origin of these phenomena is the occurrence of tetrahedral structure, regardless of the details of how this structure arises. III. AMORPHOUS SOLID BEHAVIOR Figure 1 shows the characteristic behavior associated with polyamorphism of a glassy substance. The experimental curve is taken from Ref. [3] and shows the response of the density ρ of silica glass to increasing P at ambient T. In the vicinity of 10 GPa, the density increases relatively quickly over a small range of P, while at both lower and higher P, the compressibility of the material is significantly less. Similar behavior is observed during experimental compression of amorphous ice, and suggests the existence of two distinct forms of the amorphous solid, usually termed simply as the low-density and high-density amorphous forms [5,13]. Figure 1 also shows the results of simulations of the WAC model of silica, in both the liquid and the amorphous solid forms [9]. When the simulated amorphous solid is compressed isothermally, the characteristic-inflected shape of the P-ρ isotherms 30 25 Glass T = 100 K (simulated) Glass (experiment) Liquid T = 6000 K (simulated) 20 P (GPa) 15 10 5 0 2 2.5 3 3.5 ρ g cm 3 Figure 1. Pressure-volume isotherms for WAC and real silica. Experimental data are from Ref. [3]. The liquid data are shifted upward by 8 GPa to facilitate comparison with the other curves. The protocol for generating the simulated data is described in Ref. [9]

376 ivan saika-voivod and peter h. poole found in experiments on silica glass is qualitatively reproduced, demonstrating the ability of simple rigid-ion models to generate the signature of polyamorphism. Other simulation studies have aimed at uncovering the structural processes that accompany the compression of silica glass. Experimentally, densification is reversible for a maximum compression of up to 9 GPa. For compressions beyond 12 GPa, decompression back to ambient P results in glass roughly 20% denser than the original. The experiments of Grimsditch showed that the resulting high-density glass is not only denser but also has a different structure from the low-density form [4]. Using another rigid-ion model [14], Lacks suggested that a kinetically hindered first-order amorphous amorphous transition is responsible for the experimental observations, and that the experiments at higher T may allow direct observation of the first-order transition [15]. Huang and Kieffer, in their detailed structural analysis of a potential incorporating three-body terms and charge transfer [16], found that reversible densification is accomplished by the twisting of Si O rings, similar to changes accompanying the transition in cristobalite [17]. On the other hand, they found that irreversible densification requires bond rearrangements and is aided by compression at higher T ; the P required for the irreversible change is lower for higher T [18]. IV. WATER-LIKE THERMODYNAMIC AND DYNAMIC ANOMALIES Figure 1 also shows a P-ρ isotherm for the liquid phase of WAC silica. Notably, this isotherm also exhibits an inflection, indicating the presence of a compressibility maximum in the liquid state [19]. In simulations of supercooled liquid water, the same thermodynamic feature occurs, and as T decreases, the compressibility maximum in simulated water grows into a divergence at a critical point [20]. Below the temperature of this critical point, two thermodynamically distinct liquid phases of simulated water occur, each with a distinct density, reflecting the occurrence of a first-order LLPT. The thermodynamic behavior of both WAC and BKS silica have been examined in the region of the compressibility maximum, and in both cases, the pattern of thermodynamic anomalies that occur have been found to be the same, and analogous to those found in simulation studies of water. An example is shown in Figure 2, that compares the equation of state features of ST2 water [7] with WAC silica [9] in the P T plane. In both cases, a retracing line of density maxima occurs above a monotonic spinodal boundary. The qualitative similarity of these features, despite the widely different T and P scales, is striking, and certainly suggests that a search for an LLPT in these silica models, of the kind found in water simulations, is justified. This is discussed in the next section. With regard to dynamical properties, the behavior of silica and water models has also been shown to be in correspondence. A useful concept for studying the relationship between thermodynamics and dynamics, and testing the universality

computer simulations of liquid silica 377 300 (a) (b) 2 1 200 0 P (MPa) 100 0 1 2 P (GPa) 100 3 200 4 250 300 350 400 6 8 10 12 T (K) T (K/1000) Figure 2. Equation of state features of (a) ST2 water and (b) WAC silica, projected into the P T plane. Density maxima (dashed lines) and liquid spinodal boundaries (dot-dashed lines) are shown. Isochores of P as a function of T are shown as symbols joined by thin solid lines. Equally spaced isochores are shown from bottom to top in (a) from ρ = 0.8 to 1.1 g/cm 3, and in (b) from ρ = 1.8 to 2.4 g/cm 3. of such relations across liquids, is the potential energy landscape (PEL) formalism [21]. At sufficient supercooling, a liquid system can be thought of as a point hopping between local basins in a 3N-dimensional potential energy hypersurface. Each basin is associated with a local minimum in potential energy, also termed an inherent structure (IS). In Fig. 3a, the average IS energy e IS is plotted as a function of T for three isochores spanning a wide density range in a slightly modified version of BKS silica [22,23]. The shapes of the curves at the two highest densities are typical of fragile liquids, with e IS T 1, behavior that has been shown to result from a Gaussian density of states [24]. A further implication of a Gaussian distribution of IS energies is that the configurational entropy S c, which is essentially the logarithm of the number of ISs of a given energy, becomes a steeply falling function of T and appears to approach zero at some finite T (Fig. 3b). Through the Adam Gibbs (AG) relationship [25] ( D T = A exp B ), (1) TS c a rapidly decreasing S c accounts for the super-arrhenius dynamics of fragile liquids, as observed for high-density silica (Fig. 3c and d).

378 ivan saika-voivod and peter h. poole 1.850 (a) 3 (b) e IS (MJ mol 1 ) 1.855 1.860 1.865 3.90 g/cm 3 3.01 2.36 S c / k B 2 1 1.870 2000 3000 4000 5000 6000 7000 T (K) (c) 10 7 0 0 1000 2000 3000 4000 5000 6000 T (K) (d) 10 7 D/T (cm 2 sk 1 ) 10 8 10 9 10 10 10 11 D/T (cm 2 sk 1 ) 10 8 10 9 10 10 10 11 10 12 0.15 0.2 0.25 0.3 0.35 0.4 1000/T (K 1 ) 10 12 0.05 0.1 1000/TS c (mol J 1 ) Figure 3. The potential energy landscape and its connection to dynamics. Shown are (a) inherent structure energy e IS,(b) configurational entropy S c, and (c) diffusivity D as functions of T for three isochores at ρ = 3.90 g/cm 3 (5.13 cm 3 /mol, triangles), ρ = 3.01 g/cm 3 (6.65 cm 3 /mol, squares), and ρ = 2.36 g/cm 3 (8.50 cm 3 /mol, circles). In panel (c), S c is given per, SiO 2 unit. Panel (d) shows a test of the Adam Gibbs relationship. Molar quantities are per mole ions. For the low-density isochore shown in Fig. 3, e IS exhibits an inflection, arising from a departure from a Gaussian density of states as the liquid approaches a low-energy cutoff [26] associated with a perfect random tetrahedral network. This flattening of the energy produces an inflection in S c (T ), which then presumably approaches a finite value at T = 0. This slower variation of S c with T in turn is responsible for nearly Arrhenius dynamics at low T through Eq. (1). Thus, the inflection in e IS, which is associated with a specific heat peak, is responsible for a crossover in dynamics from super-arrhenius at high T to Arrhenius at low T.In so far as a locus of specific heat maxima emanates from a possible liquid liquid critical point located at lower T, it is thus reasonable to expect that the emergent high- and low-density liquid phases associated with this critical point each exhibit distinct dynamics and landscape properties.

computer simulations of liquid silica 379 While Eq. (1) holds for both water [27] and silica, the density dependence of the diffusivity also tracks that of S c (ρ) for both liquids [23], though perhaps more strikingly so for water. It would be interesting to compare the dynamics of the two liquids in the tetrahedral regime at low T and low density. For water [28], the dynamics appear to be governed by energy excitations (i.e., bond breaking) above an amorphous ground state, just as in the case of a simpler model network-former, for which the determining nature of bond breaking on dynamics was carefully studied [29,30]. While the above illustrates the similarity in the connections among thermodynamic, structural, and dynamic quantities for silica and water, there are also significant differences, as pointed out by Shell et al. [31]. For example, silica s region in the density temperature plane of anomalous diffusive behavior (diffusion coefficient increasing with compression) encompasses the structurally anomalous region (translational and tetrahedral order parameters decreasing with compression). The opposite is the case for water. Measures of structural order in silica do not presage the onset of anomalous dynamics as they do for water. The authors of Ref. [31] attribute this difference to the atomic, rather than molecular nature of the potential for silica, and the inherent flexibility in local geometry it confers. In addition, and despite the similarities in the pattern of thermodynamic and dynamic anomalies displayed by both silica and water models, there remains a significant question relating to the interpretation of the density maximum observed in silica simulations. The density maximum found at ambient P in water simulations is typically within 15% of the experimental value (277 K). However, the temperatures at which a density maximum is observed in silica simulations at ambient P are much higher that the experimental value (1823 K) [1,2]. In the case of BKS silica, the density maximum is approximately 3 times higher in T ; in WAC silica it is nearly 5 times higher [19]. Furthermore, the high temperature of the density maximum in simulated silica places it in a regime far above the melting lines for the crystalline phases [32], where the tetrahedral structure of the liquid is highly disrupted, and where the liquid-state dynamics are fragile in character [22,23]. On the other hand, the experimentally observed density maximum in silica occurs in a regime where tetrahedral structure of the liquid is well established, and where the liquid dynamics are decidedly strong. One explanation of this situation might be that the water-like density maximum of silica is not in fact the density maximum that is observed experimentally, but is instead a thermodynamic feature that exists in the high T regime beyond the range of previous experiments. However, such an explanation then implies that the observed density maximum at 1823 K is a distinct feature associated with a wellstructured random tetrahedral network. There is precedence for such a possibility, in the fact that a low-temperature density maximum occurs in both ice I h and LDA ice at approximately 50 K [33]. In addition, if there are two density maxima in liquid silica, they must necessarily be separated by a density minimum. Again, the

380 ivan saika-voivod and peter h. poole analogous behavior may be found in water, where a density minimum has been reported both in simulations [34] and experiment [35]. Clearly, resolution of these questions will require further investigation. V. EVIDENCE FOR A LIQUID LIQUID PHASE TRANSITION While the thermodynamic and dynamical properties of silica and water models are very similar across wide ranges of P and T, the evidence for an LLPT is more highly developed in the case of water models. The evidence for an LLPT is probably most convincing for the case of the ST2 model. The EOS of liquid ST2 water has been extensively studied, and evidence of nonmonotonic isotherms of P versus V, characteristic of behavior in a liquid liquid coexistence region, have been presented [28]. Most importantly, several evaluations of the free energy surface of ST2 water have confirmed the existence of an LLPT [20,36], although this has recently been challenged [37]. No such direct thermodynamic evidence for a LLPT in silica models has yet been presented, mainly because the temperature range in which a liquid liquid critical point might be expected is below that which can be reached in straightforward simulations, due to the rapid increase of the relaxation time with decreasing T. However, several indirect approaches have shown that an LLPT is plausible for silica. For example, Ref. [19] presents an evaluation of the EOS of liquid silica for both the WAC and the BKS models over a wide range of T and P, and then exploits a physically plausible form for extrapolating the isochoric energy to lower T, into the regime in which the liquid cannot be studied in equilibrium by direct simulations. The results provide estimates for the location of the LLPT in WAC and BKS silica that are consistent with the thermodynamic anomalies that occur at higher T and provide a basis for searching for direct evidence of the LLPT in future work. Another analysis consistent with the occurrence of an LLPT is based on structural properties. Ref. [19] discusses the behavior of g 5 (r), defined as the contribution to the usual Si Si pair distribution function g(r) resulting from Si atoms that are fifth nearest-neighbors (nn) of a given Si atom. In a tetrahedral liquid with an LLPT, the low-density phase is expected to have a local structure in which the fifth nn is excluded from the first coordination shell, while in the high density phase, the fifth nn will be found at distances comparable to the first four nns. An analogous definition for g 5 (r) in terms of the fifth nn O atom of a given Si atom should show similar behavior. Plots of g 5 (r) evaluated for simulations conducted at constant volume should display a bimodal shape when the system is undergoing phase separation into regions of high- and low-density liquid. Such an analysis is shown for the BKS model in Fig. 4 [38]. This plot shows that g 5 (r) does indeed

computer simulations of liquid silica 381 (a) 30 4.63 5.14 5.64 6.15 6.65 7.16 7.67 8.17 8.68 4πr 2 g 5 (r) [Si O] 20 10 0 0.2 0.3 0.4 r (nm) 0.2 0.3 0.4 0.2 0.3 0.4 0.2 0.3 0.4 0.2 0.3 0.4 (b) 5000 4500 L 4000 S S + L C + L Q + L T (K) 3500 3000 2500 S + C C C + Q Q 2000 4 5 6 7 8 9 V (cm 3 mol 1 ) Figure 4. (a) Distributions of distances between a Si atom and its fifth nn O atom at T = 3000 K. Subpanels are labelled with molar volumes (cm 3 /mol). (b) Projection onto the V T plane of the coexistence lines for BKS silica separating the liquid (L), quartz (Q), coesite (C), and stishovite (S) phases, obtained as described in Ref. [32]. Density maxima are given by filled circles, and maxima of the isochoric specific heat by open squares. The star locates the position at which evidence for liquid liquid phase separation was reported in Ref. [19]. (c) A configuration taken from an unequilibrated run at T = 2000 K and V = 6.72 cm 3 /mol [the star in (b)]. Shown are Si atoms, with dark shading indicating a high-density environment where the fifth nn Si atom is within 0.355 nm, the location of the minimum in the Si Si g 5 (r) [19]. become bimodal in the density range that is predicted for the coexistence region of the LLPT. The above analysis of g 5 (r) is not conclusive evidence for an LLPT, because local regions having different structures could form in an isotropic, homogeneous

382 ivan saika-voivod and peter h. poole Figure 4. (Continued ) liquid phase without bulk phase separation. Whether or not these local regions phase separate macroscopically depends on the existence of cooperativity in their formation, and how this cooperativity grows as T decreases. A recent study of ST2 water in the vicinity of its LLPT showed that in this case the behavior of g 5 (r) is consistent with the location of the LLPT in that model, and provides a practical local order parameter for distinguishing the two liquid phases, and for interpreting the behavior of the liquid above the T of the LLPT in terms of a modified regular solution theory [39]. A similar analysis of the structure of BKS should yield an estimate for the location of the LLPT that could then be searched for in future simulations studies. VI. OUTLOOK As discussed above, the behavior of silica as found in simulations provides a rational basis for anticipating the occurrence of an LLPT in this system, and estimates of the location of the associated critical point have been reported for several silica models. At the same time, the LLPT, if it exists, is located in a challenging regime where the liquid-state relaxation time is large compared to the longest time scales studied in current simulations, and so it may be some time until the question can be addressed definitively. In the meantime, related silica-like systems, such as BeF 2, have significant potential for exploring these phenomena, as has been demonstrated by Angell and coworkers [40] and Chakraborty and coworkers [41]. It also

computer simulations of liquid silica 383 remains entirely possible that no LLPT occurs in silica (real or simulated) and that this substance exhibits water-like thermodynamic anomalies but without any underlying discontinuities in the thermodynamics of the deeply supercooled liquid. This case would be a realization of the so-called singularity-free scenario [42] for accounting for water-like anomalies. While not as dramatic as an LLPT, confirming singularity-free behavior in silica would do much to deepen our understanding of the range of behavior possible in materials with tetrahedral structure. REFERENCES 1. C. A. Angell and H. Kanno, Science 193, 1121 (1976) 2. C. A. Angell, P. A. Cheeseman, and S. Tamaddam, Science 218, 885 (1982). 3. H. Sugiura, K.-I. Kondo, and A. Sawaoka, in High-Pressure Research in Geophysics, S. Akimoto and M.H. Manghnani eds., Reidel, Dordrecht, 1982, p. 551. 4. M. Grimsditch, Phys. Rev. Lett. 52, 2379 (1984). 5. O. Mishima, L. D. Calvert, and E. Whalley, Nature 314, 76 (1985). 6. P. H. Poole, T. Grande, C. A. Angell, and P. F. McMillan, Science 275, 322 (1997). 7. P. H. Poole, F. Sciortino, U. Essmann, and H. E. Stanley, Nature (London) 360, 324 (1992). 8. O. Mishima and H. E. Stanley, Nature (London) 396, 329 (1998). 9. P. H. Poole, M. Hemmati, and C. A. Angell, Phys. Rev. Lett. 79, 2281 (1997). 10. P. H. Poole, P. F. McMillan, and G. H. Wolf, Rev. Mineralogy 32, 563 (1995). 11. L. V. Woodcock, C. A. Angell, and P. A. Cheeseman, J. Chem. Phys. 65, 1565 (1976). 12. B. van Beest, G. Kramer, and R. van Santen, Phys. Rev. Lett. 64, 1955 (1990). 13. K. Winkel, E. Mayer, and T. Loerting, J. Phys. Chem. B 115, 14141 (2011). 14. S. Tsunekuyi, M. Tsukada, H. Aoki, and Y. Matsui, Phys. Rev. Lett. 61, 869 (1988). 15. D. J. Lacks, Phys. Rev. Lett. 84, 4629 (2000). 16. L. P. Huang and J. Kieffer, J. Chem. Phys. 118, 1487 (2003). 17. L. P. Huang and J. Kieffer, Phys. Rev. B 69, 224203 (2004). 18. L. P. Huang and J. Kieffer, Phys. Rev. B 69, 224204 (2004). 19. I. Saika-Voivod, F. Sciortino, and P. H. Poole, Phys. Rev. E 63, 011202 (2001). 20. F. Sciortino, I. Saika-Voivod, and P.H. Poole, Phys. Chem. Chem. Phys. 13, 19759, 2011. 21. F. Sciortino, J. Stat. Mech. 2005, P05015 (2005). 22. I. Saika-Voivod, P. H. Poole, and F. Sciortino, Nature (London) 412, 514 (2001). 23. I. Saika-Voivod, F. Sciortino, and P. H. Poole, Phys. Rev. E 69, 041503 (2004). 24. S. Sastry, Nature (London) 409, 164 (2001). 25. G. Adam and J. H. Gibbs, J. Chem. Phys. 43, 139 (1965). 26. A. Saksaengwijit, J. Reinisch, and A. Heuer, Phys. Rev. Lett. 93, 235701 (2004). 27. A. Scala, F. Starr, E. La Nave, F. Sciortino, and H. E. Stanley, Nature (London) 406, 166 (2000). 28. P. H. Poole, S. R. Becker, F. Sciortino, and F. W. Starr, J. Phys. Chem. B 115, 14176 (2011). 29. A. J. Moreno, S. V. Buldyrev, E. La Nave, I. Saika-Voivod, F. Sciortino, P. Tartaglia, and E. Zaccarelli, Phys. Rev. Lett. 95, 157802 (2005).

384 ivan saika-voivod and peter h. poole 30. A. J. Moreno, I. Saika-Voivod, E. Zaccarelli, E. La Nave, S. V. Buldyrev, P. Tartaglia, and F. Sciortino, J. Chem. Phys. 124, 204509 (2006). 31. M. S. Shell, P. G. Debenedetti, and A. Z. Panagiotopoulos, Phys. Rev. E 66, 011202 (2002). 32. I. Saika-Voivod, F. Sciortino, T. Grande, and P. H. Poole, Phys. Rev. E 70, 061507 (2001). 33. C. A. Angell, Annu. Rev. Phys. Chem. 55, 559 (2004). 34. P. H. Poole, I. Saika-Voivod, and F. Sciortino, J. Phys. Condens. Matter 17, L431 (2005). 35. D. Liu, Y. Zhang, C.-C. Chen, C.-Y. Mou, P. H. Poole, and S.-H. Chen, Proc. Natl. Acad. Sci. USA 104, 9570 (2007). 36. Y. Liu, A. Z. Panagiotopoulos, and P. G. Debenedetti, J. Chem. Phys. 131, 104508 (2009). 37. D. T. Limmer and D. Chandler, J. Chem. Phys. 135, 134503 (2011). 38. I. Saika-Voivod, F. Sciortino, T. Grande, and P. H. Poole, Phil. Trans. R. Soc. A 363, 525 (2005). 39. M. J. Cuthbertson and P. H. Poole, Phys. Rev. Lett. 106, 115706 (2011). 40. M. Hemmati, C. T. Moynihan, and C. A. Angell, J. Chem. Phys. 115, 6663 (2001). 41. M. Agarwal, R. Sharma, and C. Chakravarty, J. Chem. Phys. 127, 164502 (2007). 42. S. Sastry, P. G. Debenedetti, F. Sciortino, and H. E. Stanley, Phys. Rev. E 53, 6144 (1996).