ABSTRACT Title of Dissertation: PHASE TRANSITIONS AFFECTED BY MOLECULAR INTERCONVERSION Thomas Joseph Longo Doctor of Philosophy, 2023 Dissertation Directed by: Professor Mikhail A. Anisimov Institute for Physical Science and Technology & Department of Chemical and Biomolecular Engineering Typically, pure substances may be found with only one gaseous or liquid state, while their solid state may exist in various polymorphic states. The existence of two distinct liquid forms in a single component substance is more unusual since liquids lack the long-range order common to crystals. Yet, the existence of multiple amorphous states in a single component substance, a phenomenon known as “liquid polyamorphism,” has been observed or predicted in a wide variety of substances. In contrast to standard phase transitions, it has been suggested that polyamor- phic liquid-liquid transitions are caused by the interconversion of molecular or supramolecular states. To investigate this phenomenon, a nonequilibrium thermodynamic model was developed to quantitatively describe the interplay between the dynamics of molecular interconversion and fluid-phase separation. The theory has been compared to a variety of interconverting systems, and has demonstrated a quantitative agreement with the results of Monte Carlo and Molecular Dynamics simulations. In this thesis, it is shown that there are two major effects of molecular interconversion on the thermodynamics and the kinetics of fluid-phase separation: if the system evolves to an equi- librium state, then the growth of one of the alternative phases may result in the destruction of phase coexistence - a phenomenon referred to as “phase amplification.” It is demonstrated that depending on the experimental or simulation conditions, either phase separation or phase am- plification would be observed. Previous studies of polyamorphic substances report conflicting observations of phase formation, which may be explained by the possibility of phase amplifica- tion occurring. Alternatively, if the system evolves to a nonequilibrium steady state, the phase domain growth could be restricted at a mesoscopic length scale. This phenomenon (referred to as “microphase separation”) is one of the simplest examples of steady-state dissipative structures, and may be applicable to active matter systems, hydrodynamic instabilities, and bifurcations in chemical reactions, in which the nonequilibrium conditions could be imposed by an external flux of matter or energy. PHASE TRANSITIONS AFFECTED BY MOLECULAR INTERCONVERSION by Thomas J. Longo Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2023 Advisory Committee: Professor Mikhail Anisimov, Chair Professor Christopher Jarzynski Professor Jeffery Klauda Professor Pratyush Tiwary Professor John Weeks © Copyright by Thomas J. Longo 2023 Preface This thesis consists of results from my original research performed in collaboration with my co-authors. The contents of this thesis have been adopted with permission from our publica- tions, of which I am either the first or second author. The investigation of the fluid-fluid phase transition in high-pressure hydrogen (Section 2.2), high-density sulfur (Section 2.3), and fluids exhibiting water-like anomalies (Section 2.4) was conducted in Refs. [1–3], respectively. The phenomenological theory that describes the dynamics of phase transitions affected by molec- ular interconversion (Chapter 3) was developed in Refs. [4, 5]. This theory was verified via Monte Carlo (MC) and molecular dynamics (MD) simulations of three physically distinct micro- scopic models of mixtures with species interconversion (Chapter 4), which were investigated in Refs. [6–8]. Throughout my Ph.D. project, I developed the theory, participated in the computer simula- tions of the microscopic models, performed the data analysis, made the figures (except Figs. 4.10- 4.13, which were created by Betül Uralcan), and was a major contributor to the text of the pub- lished manuscripts. I would like to ascribe appropriate credit to my research advisor, who con- ceived the project and supervised the research, and to our collaborators, whose contribution was crucial for the success of the project. The two-state thermodynamic approach (Section 2.1) was originally developed by my advi- sor, Mikhail Anisimov, and his co-workers [9]. The equation-of-state of high-pressure hydrogen (Section 2.2) was developed in collaboration with Nathaniel Fried [1]. The maximum-valence ii model (Section 2.3), the HL model (Section 4.1), and the HCS model (Section 4.3) were de- veloped and simulated by Sergey Buldyrev and Nikolay Shumovskyi [2, 5, 6, 8]. The blinking- checkers lattice model (Section 2.4) was originally developed by Frédéric Caupin and Mikhail Anisimov [10]; I described the interfacial properties of this model [3]. My major contribution to the research project was the generalization of the Cahn-Hilliard theory of spinodal decomposition and the development of the generalized Cook-Binder theory of the time-evolution of the structure factor in fluids with interconversion of species (Chapter 3) [4,5,8]. Numerical calculations of the time evolution of the order parameter and structure factor (Section 3.3) were performed in col- laboration with Salim Asadov, Nikolay Shumovskyi, and Sergey Buldyrev [5]. The simulations of the CM model (Section 4.2) were performed by Betül Uralcan, Frank H. Stillinger, and Pablo Debenedetti [7, 8]. Thomas J. Longo April 15, 2023 iii Acknowledgments I would like to express my sincere gratitude to my research advisor, Professor Mikhail Anisimov. His constant tutelage, endless support (often working, literally, every day with me), and insightful suggestions not only benefited the success of this project, but also enabled me to grow as a researcher. I am grateful for the collaborators that I had the privilege to work closely with, namely: Salim Asadov (University of Maryland, College Park and Azerbaijan National Academy of Sci- ences), Sergey Buldyrev (Boston University and Yeshiva University), Frédéric Caupin (Université Claude Bernard Lyon 1 and Institut Universitaire de France), Pablo Debenedetti (Princeton Uni- versity), Nathaniel Fried (University of Maryland, College Park), Nikolay Shumovskyi (Boston University), and Betül Uralcan (Princeton University and Bogazici University). This endeavor would not be possible without their contributions. I am also thankful to the Chemical Physics Program at the Institute for Physical Science and Technology as well as the Department of Chemical and Biomolecular Engineering for the continuous encouragements and financial support during my time at the University of Maryland, College Park. This dissertation is a product of the research collaboration between the University of Mary- land (the Lead institution), Princeton University, Boston University, and Arizona State University, supported by the National Science Foundation (NSF). I acknowledge the financial support of the iv NSF (award No. CHE-1856479) as well as the partial support of the Chateaubriand Fellowship of the Office for Science & Technology of the Embassy of France in the United States. I would be remiss if I did not mention my family, in particular, my parents, my siblings, and my fiancé. Thank you all for your endless support and encouragements throughout my studies. v Table of Contents Preface ii Acknowledgements iv Table of Contents vi List of Tables viii List of Figures ix List of Abbreviations xxiv Chapter 1: Introduction 1 Chapter 2: Two-State Thermodynamics of Fluid Polyamorphism 8 2.1 Interconversion of Two States . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 2.2 Fluid-Fluid Phase Transition in High-Pressure Hydrogen . . . . . . . . . . . . . 13 2.2.1 Phase Behavior of Hydrogen . . . . . . . . . . . . . . . . . . . . . . . . 13 2.2.2 Phase Separation Coupled with Dimerization . . . . . . . . . . . . . . . 15 2.3 Liquid-Liquid Phase Transition in High-Density Sulfur . . . . . . . . . . . . . . 23 2.3.1 Maximum-Valence Model . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.3.2 Phase Separation Coupled with Polymerization . . . . . . . . . . . . . . 28 2.4 Liquid-Liquid Phase Transition in Fluids Exhibiting Water-Like Anomalies . . . 32 2.4.1 Blinking-Checkers Lattice Model . . . . . . . . . . . . . . . . . . . . . 33 2.4.2 Virtual Critical Line in Interconverting Mixtures . . . . . . . . . . . . . . 37 2.4.3 Anomalies of Interfacial Properties . . . . . . . . . . . . . . . . . . . . . 42 2.5 Conclusion of Chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 Chapter 3: Phase Formation Affected by Species Interconversion 63 3.1 Generalizing Cahn-Hilliard Theory . . . . . . . . . . . . . . . . . . . . . . . . . 63 3.1.1 Conserved and Nonconserved Order-Parameter Dynamics . . . . . . . . 64 3.1.2 Phase Amplification vs. Microphase Separation . . . . . . . . . . . . . . 73 3.2 Temporal Evolution of the Structure Factor . . . . . . . . . . . . . . . . . . . . . 78 3.2.1 Generalized Cook-Binder Theory . . . . . . . . . . . . . . . . . . . . . 78 3.2.2 Characteristic Length Scales . . . . . . . . . . . . . . . . . . . . . . . . 81 3.3 Temporal Evolution of the Order Parameter . . . . . . . . . . . . . . . . . . . . 85 3.4 Effects of Heat and Volume Change of Interconversion . . . . . . . . . . . . . . 89 vi 3.5 Effects of Critical Order-Parameter Fluctuations . . . . . . . . . . . . . . . . . . 93 3.6 Conclusion of Chapter 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 Chapter 4: Application of Theory to Microscopic Models 99 4.1 Hybrid Ising / Binary Lattice (HL Model) . . . . . . . . . . . . . . . . . . . . . 99 4.1.1 HL Model Description . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 4.1.2 Phase Amplification in the HL Model . . . . . . . . . . . . . . . . . . . 103 4.1.3 Nonequilibrium Spatially-Modulated Stripes . . . . . . . . . . . . . . . . 115 4.2 Chiral-Mixture of Interconverting Enantiomers (CM Model) . . . . . . . . . . . 122 4.2.1 CM Model Description . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 4.2.2 Phase Amplification in the CM Model . . . . . . . . . . . . . . . . . . . 127 4.2.3 Formation of Dissipative Structures . . . . . . . . . . . . . . . . . . . . 130 4.2.4 Characteristic Time Scales in the CM Model . . . . . . . . . . . . . . . . 137 4.2.5 Finite-Size Restrictions in the CM Model . . . . . . . . . . . . . . . . . 139 4.3 Coarse-Grained Hard-Core-Shoulder Model (HCS Model) . . . . . . . . . . . . 144 4.3.1 HCS Model Description . . . . . . . . . . . . . . . . . . . . . . . . . . 144 4.3.2 Nonequilibrium Bicontinuous Microemulsions . . . . . . . . . . . . . . 147 4.4 Sources of Microphase Separation . . . . . . . . . . . . . . . . . . . . . . . . . 150 4.4.1 Comparison of the Microscopic Models . . . . . . . . . . . . . . . . . . 150 4.4.2 Onset and Termination of Microphase Separation . . . . . . . . . . . . . 152 4.5 Conclusion of Chapter 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156 Chapter 5: Overall Conclusion 158 Appendix A: Comparison of Exact Solution with Phenomenological Ansatzes 160 Bibliography 165 vii List of Tables 2.1 The suggested locations of the FFCP and the SFF-TP in hydrogen. . . . . . . . . 15 2.2 Liquid-vapor critical points of interconverting systems, referred to as “actual” critical points, for the seven systems considered in this section (with εAA = 1.6, εBB = 2.0, ê = 3, and ŝ = 4). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 2.3 Liquid-liquid critical points (LLCP) for the three systems exhibiting liquid polyamor- phism and a LLCP (with εAA = 1.6, εBB = 2.0, ê = 3, and ŝ = 4). . . . . . . . . 40 2.4 Asymptotic amplitudes of the liquid-liquid interfacial tension and liquid-liquid correlation length of concentration fluctuations for the three systems exhibiting liquid polyamorphism and a liquid-liquid critical point. The asymptotic mean- field behavior is illustrated in Fig. 2.21. . . . . . . . . . . . . . . . . . . . . . . 58 2.5 The surface tension σ, normalized interfacial thickness ζ̂ = ζ/`, and normalized shift δ̂ = δ/` for the three coexisting phases (liquid A, liquid B, and vapor) for the system with εBA = 1.00 at the triple point temperature (T = 0.6843). . . . . . 58 3.1 Limiting cases of the interplay between diffusion, natural interconversion, and forced interconversion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 3.2 Conditions for phase amplification and microphase separation as illustrated in Fig. 3.1. The left column corresponds to the solid lines and the right column corresponds to the dashed lines. . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 4.1 Fit parameters of the theoretical prediction in Fig. 4.8a (top, pr = 1/32) and Fig. 4.8b (bottom, pr = 1/256) of the inverse steady-state domain size, q−. . . . . 118 4.2 Parameters for the CM model in real and reduced units . . . . . . . . . . . . . . 124 4.3 Fit parameters used for the theoretical predictions in Fig. 4.14a of the inverse steady-state domain size, q−. The mobility is given by the Stokes-Einstein rela- tion with amplitude b0 and characteristic temperature T0 = 1.2. Two additional parameters, the amplitude, A, and the characteristic size of the inhomogeneities (q0 ∼ 1/R0) are included. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 4.4 Fit parameters used for the theoretical prediction illustrated in Fig. 4.25a for the HL model, with pr = 1/128. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 A.1 The degree of asymmetry for each profile of the system with εBA = 1.04 at T = 0.6943, determined from the symmetric ansatz (sym), the Fisher-Wortis ansatz (FW), or the exact solution as calculated from Eq. (A.3). . . . . . . . . . . . . . 163 viii List of Figures 1.1 Hypothesized global pressure-temperature phase diagram for water. The black curves indicate the solid-solid and solid-liquid-gas phase transitions experimen- tally reported for water [11], while the blue curve indicates the hypothesized liquid-liquid phase transition. The dotted black curve represents the line of spon- taneous ice nucleation [12]. The red circle is the liquid-vapor critical point (LVCP) and the red star is the hypothesized liquid-liquid critical point (LLCP) around 220 K and 60 MPa as suggested in Ref. [13]. . . . . . . . . . . . . . . . . . . . . 2 2.1 A symmetric binary mixture is quenched along the critical molecular fraction (where the order parameter is ϕ = 0) from a point, T̂1 in the one phase region to a point, T̂2 below the critical temperature of demixing in the unstable region. The solid curve indicates the liquid-liquid coexistence (binodal), while the dashed curve indicates the limit of stability (spinodal). . . . . . . . . . . . . . . . . . . . 9 2.2 The global pressure-temperature phase diagram for hydrogen. (a) The full range from low to extreme pressures in logarithmic scale. The crosses indicate the experimental data for the solid-liquid melting transition presented in Diatschenko et al. [14] (blue), Datchi et al. [15] (cyan), Gregoryanz et al. [16] (pink), and Zha et al. [17] (purple). The solid black curves at low pressure (P ≤ 0.1 GPa) are the liquid-gas-solid phase transitions [18], while the solid black curve at high pressure (P > 0.1 GPa) is the Kechin equation [19] as reported in ref. [17]. The black dashed curve is the predicted continuation of the melting line based on experimental and computational evidence [17, 20, 21], while the dotted lines represent the highly-debated prediction [22–28] of the domain of solid metallic hydrogen [22,23,28–33]. The red line is the first-order fluid-fluid phase transition adopted in this section. (b) The phase diagram of hydrogen at extreme conditions, in the area of the box in (a). The open circles are experimental data presented in Zaghoo et al. [34–36] (dark brown), McWilliams et al. [37] (light brown), and Ohta et al. [38] (orange). Simulation results [21, 39–46] are spread within the grey area and shown in detail in Fig. 2.3. The fluid-fluid phase transition (solid red) and Widom line (dotted red) are represented by Eq. (2.18). The red star is the location of the fluid-fluid critical point (FFCP) as adopted in this section. . . . 14 ix 2.3 Unifying the different simulation results with experimental data of hydrogen by the generalized law of corresponding states. (a) Experimental and simulation results for the fluid-fluid phase transition (FFPT). (b) Unified representation of the FFPT by reducing pressure, P̂ = P/Pc, temperature, T̂ = T/Tc, and the critical value of the entropy, Ŝ = S/Sc. In (a) and (b), the open circles are the experimental data of Zaghoo et al. [34–36] (dark brown), McWilliams et al. [37] (light brown), and Ohta et al. [38] (orange). The computational results are indicated by the triangles: blue tints correspond to the Density Functional Theory (DFT) simulations of Bonev et al. [47] (dark blue), Morales et al. [48] (blue), Hinz et al. [44] (sky blue), and Karasiev et al. [45] (light blue). Meanwhile, green tints correspond to the Quantum Monte Carlo (QMC) simulations of Morales et al. [48] (dark sea green), Lorenzen et al. [39] (green), Perlioni et al. [41] (dark green), Mozzola et al. [42] (lime green), and Tirelli et al. [46] (yellow green). The colored stars correspond to the reported (or adopted in this section) critical points for each data set. The solid black curve is the solid-fluid phase transition line as discussed in Fig. 2.2, and the red solid line is the FFPT predicted in this section. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.4 The components of the Gibbs energy (per atom) for hydrogen in the vicinity of the fluid-fluid phase transition. (a) The Gibbs energy of reaction, GBA, as given by Eq. (2.9). The isotherms are T = 0.5Tc (orange), T = 0.75Tc (blue), T = Tc (green), T = 1.25Tc (red), and T = 1.5Tc (purple). (b) The Gibbs energy of mixing, Gmix, as given by Eq. (2.15). Gmix is shown as a function of the fraction of hydrogen atoms in the diatomic state, x, for isotherms T = 0.5Tc (blue), T = 0.75Tc (green), and T = Tc (red) at P = Pc. The solid curve corresponds to the fluid-fluid coexistence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 2.5 Equilibrium fraction of hydrogen atoms in the diatomic state, xe. (a) Equilibrium fraction-pressure diagram for T = 0.5Tc (orange), T = 0.75Tc (blue), T = Tc (green), and T = 1.25Tc (red). (b) Equilibrium fraction-temperature diagram for P = 0.75Pc (blue), P = Pc (green), P = 1.25Pc (red), P = 1.5Pc (purple). The solid and dashed black curves are, respectively, the fluid-fluid coexistence and the limit of thermodynamic stability (spinodal). . . . . . . . . . . . . . . . . . . . . 20 2.6 The pressure-density phase diagram of hydrogen based on the equation of state developed in this section. The open circles correspond to the QMC simulations of Tirelli et al. [46]. Isotherms are T̂ = 0.67 (orange), T̂ = 0.73 (red), T̂ = 0.8 (brown), T̂ = 0.87 (purple), T̂ = 0.93 (green), and T̂ = 1.0 (blue). The fluid- fluid coexistence is shown by the solid black curve. The red star is the FFCP adopted in this section. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.7 A schematic representation of the phase diagram of sulfur. The solid black curves represent the solid-liquid-gas phase transition in sulfur [49–51], while the blue curve represents the experimentally verified liquid-liquid phase transition [51]. The dashed black line is the predicted continuation of the melting line based on experimental evidence [49]. The black circles are the solid-liquid-liquid triple point (TP-SLL) and the solid-liquid-gas triple point (TP-SLG), while the red cir- cle is the liquid-gas critical point (LGCP) and the red star is the liquid-liquid critical point (LLCP). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 x 2.8 Reactions and interactions in the maximum-valence model. (a) The three types of covalent bond-forming reversible chemical reactions that may occur in the system. If two atoms without bonds (S0) collide with each other, they may form a bond and become S1 atoms. If a S0 and S1 atom collide, they may form a bond and become S1 and S2 atoms, respectively. If two S1 atoms collide with each other, they form an additional bond and become S2 atoms. (b-d) The three major interactions between atoms, in which each atom is composed of a core and shell, both with a radius σ and mass m. U(r) is the pair potential energy and r is the distance from the center of an atom. (b) The cores of each atom interact with an attractive square well of depth ε = 1 and width w = 0.4. (c) The shells may react to form covalent bonds that consist of a narrow well with depth εb = 1 and width wb = 0.02. (d) Phase segregation is coupled to polymerization via the additional attractive interactions between atoms in state S2, described by a square well of depth ε22 = 0.5 and width w22 = 0.3. . . . . . . . . . . . . . . . . . . . . . . . 27 2.9 Phase diagrams for the maximum-valence model (with ε22 = 0.5 and εb = 1.0) obtained in an NVT ensemble after t = 106 time units. (a) The isotherms in the P -ρ plane are T = 0.96 − 1.20 (red-purple) in steps ∆T = 0.02. (b) The liquid-gas and liquid-liquid critical isochores in the P -T plane are ρLG c = 0.35 and ρLL c = 0.81 as indicated by the lower and upper dashed lines, respectively. In both figures, the liquid-gas and liquid-liquid coexistence curves are calculated via the Maxwell construction and indicated by the solid curves. The liquid-gas (T LG c = 1.023, P LG c = 0.0922, ρLG c = 0.35) and liquid-liquid (T LL c = 1.187, P LL c = 2.28, ρLL c = 0.81) critical points are indicated by the red open circles, while the triple point (P TP = 0.0738, T TP = 0.995) is indicated by the black open circles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 2.10 (a) T -ρ phase diagram for the maximum-valence model (with ε22 = 0.5 and εb = 1.0) obtained in an NVT ensemble after t = 106 time units. The temperature dependence of the fraction of atoms with two bonds, x2, (b) and the average chain-length, 〈n〉, (c) in two coexisting liquid phases. The simulation data in (b) is fit to a second order polynomial. . . . . . . . . . . . . . . . . . . . . . . . . . 30 2.11 (a) The density correlation function g(r) and (b) the structure factor S(q) across the liquid-liquid transition at T = 1.00 for densities of ρ = 0.65 (blue), ρ = 0.70 (orange), ρ = 0.75 (green), ρ = 0.80 (red), ρ = 0.85 (purple), and ρ = 0.90 (black). In (a), the sharp peak, around r = 1 (in units of σ), corresponds to the length of the covalent bond, which increases upon increasing density. Simul- taneously, in (b), the maximum of the structure factor (the first peak) shifts to larger wavenumbers upon increasing density, while the second peak acquires a characteristic bump, similar to what was recently observed in sulfur [51]. The di- vergence of the structure factor at q = 0 indicates the divergence of the isothermal compressibility in the vicinity of the LLCP. The insets (dashed boxes) highlight the behavior of the maximum of the correlation function and second peak of the structure factor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 xi 2.12 Density-temperature and pressure-temperature phase diagrams for the blinking- checkers model with εAA = 1.6, εBB = 2.0, εAB = 1.08, ê = 3, and ŝ = 4. The liquid-vapor coexistence (blue curves) terminates at the liquid-vapor critical point, LVCP. The liquid-liquid coexistence (red curves) terminates at the liquid- liquid critical point, LLCP. The limits of thermodynamic stability (spinodals) are given by the dashed curves. In (a,b), the dotted line corresponds to the condition, x = 1/2, which qualitatively separates the regions enriched either by species A (at low temperatures and low densities) or species B (at high temperatures and high densities). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 2.13 Liquid-vapor (a,b) and liquid-liquid (c,d) coexistence curves for the seven sys- tems with εAA = 1.6, εBB = 2.0, e = 3, s = 4, and with various values of εBA: εBA = 1.00 (blue), εBA = 1.04 (orange), εBA = 1.08 (green), εBA = 1.12 (red), εBA = 1.16 (purple), εBA = 1.20 (pink), and εBA = 1.24 (gray). The critical points, indicated by the stars in (a,b) and open circles in (c,d), are the unique liquid-vapor (LVCP) or liquid-liquid (LLCP) critical points in the interconvert- ing system, referred to as “actual” critical points. As also indicated in Fig. 2.12, species A is enriched in the low-density, low-temperature region, while species B is enriched in the high-density, high-temperature region. For the system with εBA = 1.00, the dashed blue curves in (c,d) indicate the liquid-vapor coexistence, while in (a-d), the dotted blue line indicates the discontinuity at the triple point. . 38 2.14 Illustration of the thermodynamic path selected by the interconversion reaction for ê = 3 and ŝ = 4 at constant volume, V , and at constant number of occupied lattice sites,N1+N2, represented through (a) the liquid branch of the liquid-vapor temperature-concentration coexistence (see Fig. 2.13b), and (b) the activity, a = 1/[1+e−GBA/T̂ ], for the systems with different interaction parameters: εBA = 1.00 (blue), εBA = 1.04 (orange), εBA = 1.08 (green), εBA = 1.12 (red), εBA = 1.16 (purple), εBA = 1.20 (pink), and εBA = 1.24 (gray). For each system, the liquid- vapor critical line (LVCL) is shown by the dashed curves, while the collapsed coexistence, in (a), is illustrated by the black curve. The insets show the LV critical points for each scenario. Note that for the system with εBA = 1.00, the thermodynamic path crosses through the triple point, indicated by the dotted line in (a). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 2.15 Liquid-vapor (a,b) and liquid-liquid (c,d) diameters of the density, Eq. (2.60), and the concentration, Eq. (2.61), as a function of the distance to the virtual LVCL (or LLCL) along the thermodynamic path selected by interconversion for systems with εAA = 1.6, εBB = 2.0, ê = 3, ŝ = 4, and with various values of εBA: εBA = 1.00 (blue), εBA = 1.04 (orange), εBA = 1.08 (green), εBA = 1.12 (red), εBA = 1.16 (purple), εBA = 1.20 (pink), and εBA = 1.24 (gray). In (a-d), the dotted blue lines indicate the discontinuity for the system with εBA = 1.00 at the triple point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 xii 2.16 The liquid-vapor interfacial tension as a function of temperature (a), and also pre- sented in reduced units (b), for the system with εBA = 1.00 (blue), εBA = 1.04 (orange), εBA = 1.08 (green), εBA = 1.12 (red), εBA = 1.16 (purple), εBA = 1.20 (pink), and εBA = 1.24 (gray). In (b), the critical temperature is given by the “vir- tual” critical point (for the non-reactive binary mixture) for each concentration along the thermodynamic path selected by the interconversion reaction. The blue arrows indicate the direction of warming. (c) The reduced interfacial thickness, ζ̂ = ζ/`, and (d) the reduced relative distance between the concentration and density profiles, δ̂ = δ/`. In (a-d), the dotted lines indicate the discontinuity of the interfacial properties for the system with εBA = 1.00 at the triple point, shown by the vertical bars in (c,d). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 2.17 (a) The behavior of the liquid-vapor interfacial tension follows the power law, σ = σ0|∆T̂ |3/2, where the amplitude was found to be σ0 = 0.71, asymptotically close to the actual liquid-vapor critical temperature (see Table 2.2). (b) The behavior of the reduced liquid-vapor interfacial thickness, ζ̂ = ζ/`, follows the power law, ζ̂ = ζ̂0|∆T̂ |−0.38, where the amplitude was found to be ζ̂0 = 1.50 asymptotically close to the actual critical point. . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 2.18 Asymptotic behavior of the liquid vapor coexistence for systems with εAA = 1.6, εBB = 2.0, ê = 3, ŝ = 4, and with various values of εBA: εBA = 1.00 (blue), εBA = 1.04 (orange), εBA = 1.08 (green), εBA = 1.12 (red), εBA = 1.16 (purple), εBA = 1.20 (pink), and εBA = 1.24 (gray). (a) The temperature-density LV coexistence follows the meanfield power law, ∆T̂ ∼ |∆ρ̂|2, where ∆T̂act = 1 − T/T act c and T act c is the actual critical temperature selected by the interconverting path. Likewise, ∆ρ̂act = 1 − ρ/ρact c , where ρact c is the actual critical density. (b) The temperature-average concentration LV coexistence follows the meanfield power law, ∆T̂ ∼ |∆x̄act|2, where ∆x̄act = 1 − x̄/xact c , in which x̄ = (xL + xV)/2 and xact c is the actual critical concentration. (c,d) Illustrate, as an example, the asymptotic behavior of the system with εBA = 1.08, in which (d) shows the asymptotic behavior of each branch of the concentration coexistence. . . . . . . . 53 2.19 Comparison between the liquid-liquid (dashed curves) and liquid-vapor (solid curves) interfacial tensions as a function of temperature for the system with εBA = 1.00 (blue), εBA = 1.04 (orange), εBA = 1.08 (green), εBA = 1.12 (red). The dotted blue line indicates the discontinuity in the liquid-vapor interfacial tension. 56 2.20 Liquid-liquid interfacial properties of the systems exhibiting liquid polyamor- phism with εAA = 1.6, εBB = 2.0, ê = 3, ŝ = 4, and with various values of εBA: εBA = 1.00 (blue), εBA = 1.04 (orange), εBA = 1.08 (green), εBA = 1.12 (red). (a) the reduced thickness, ζ̂ = ζ/` of the liquid-liquid interface, and (b) the reduced shift, δ̂ = δ/`, between the density and concentration liquid-liquid profiles. In (a,b) the thickness and shift reach a finite value (marked with a blue circle) at the triple point temperature. . . . . . . . . . . . . . . . . . . . . . . . . 56 xiii 2.21 (a) The behavior of the liquid-liquid interfacial tension follows the meanfield power law, σ = σ0|∆T̂ |3/2, (dashed lines) asymptotically close to the actual liquid-liquid critical temperature (see Table 2.3). (b) The behavior of the reduced liquid-liquid interfacial thickness, ζ̂ = ζ/`, follows the meanfield power law, ζ̂ = ζ̂0|∆T̂ |−1/2, (dashed lines). In (a,b) the systems exhibiting liquid polyamorphism and a liquid-liquid critical point are shown: εBA = 1.04 (orange), εBA = 1.08 (green), εBA = 1.12 (red), and the amplitudes of the asymptotic meanfield power laws are provided in Table 2.4. . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 2.22 Normalized density and concentration liquid-vapor profiles as a function of the coordinate perpendicular to the planar interface, ẑ = z/`, given by Eqs. (2.54) and (2.55) for the system with εAA = 1.6, εBB = 2.0, εBA = 1.08, ê = 3, and ŝ = 4 at the two temperatures (a,b) that correspond to the two extrema of the liquid-vapor interfacial tension (shown in Fig. 2.16). Normalized (c) density and (d) concentration profiles for three-phase coexistence at the triple point, TTP = 0.6843, for the system with εAA = 1.6, εBB = 2.0, and εBA = 1.00. . . . . . . . . 59 2.23 Interfacial profiles of species B, ρB = ρx, in the blinking-checkers model demon- strate surface enrichment near the TP temperature, TTP = 0.68429. (a) Surface enrichment of species B for the system with εBA = 1.00. The colored curves indi- cate temperatures from T = 0.68989 to T = TTP in steps of ∆T̂ = −0.0008 in or- der of purple to red. The black curves are T = 0.68389 (dashed) and T = 0.68309 (solid). (b) Surface enrichment of species B for the system with εBA = 1.04. The curves are T = 0.6882 to T = 0.6826 in steps of ∆T = −0.0008 (blue to pink). In (a,b), the black arrows indicate the direction of decreasing temperature. Note that while the transition of a surface enriched profile (T > TTP) to a smooth pro- file (T < TTP) is discontinuous in the system with εBA = 1.00, it is continuous in the system with εBA = 1.04. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 3.1 The characteristic growth rate, Eq. (3.14), affected by the competition between diffusion, natural interconversion, and forced interconversion at ∆T̂ = −0.5. Complete phase separation (as predicted by Cahn-Hilliard’s theory for L = 0 and K = 0) is illustrated by the red curve for M = 100. Phase amplification is illustrated by the purple curves for restricted (M = 100 - solid) and unrestricted (M = 10 - dashed) cases, in which L = 10 and K = 2. Microphase separation for M = 100, L = 1, and K = 2 is illustrated by the solid green curve. When the growth rate is always negative, as illustrated by the green dashed curve (for M = 1, L = 1, and K = 2), there is no phase domain growth corresponding to a homogeneous steady state. The green circles indicate the three characteristic wavenumbers of the amplification factor: the maximum, qm, the lower cut-off, q−, and the upper cut-off, q+ . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 xiv 3.2 Steady-state phase domain morphology for different magnitudes of forced inter- conversion (after ∼ 105 time steps) numerically computed from the time evolu- tion of the order parameter, Eq. (3.7), with M = 1, L = 1/127, ∆T̂ = −0.1, ` = 64, σi = 0.1, and η = 10−5, as discussed in Section 3.3. Morphologies are shown for the middle slice of the three-dimensional system at (a) K = 0, (b) K = 5× 10−4, (c) K = 15× 10−4, and (d) K = 25× 10−4. The red regions correspond to where the value of the normalized order parameter is ϕ/ϕmax = 1, the purple regions correspond to where the value of the normalized order param- eter is ϕ/ϕmax = −1, and the other colors depict the interface between these two regions. The image in (a) depicts a metastable structure toward phase amplifica- tion [6], while the images in (b-d) are modulated steady-state structures with a characterize size, 1/q−. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 3.3 a) The amplification factor, ω(q), given by Eq. (3.15) with κ = 1, ∆T̂ = −0.1, M = 1, L = 1/127, and K = 1.3× 10−3. b) The time evolution of the structure factor, given by Eq. (3.23) in the presence of natural and forced interconversion. The black dotted line depicts the evolution of the maximum of the structure factor. Due to the external source of forced interconversion, the maximum of the struc- ture factor is interrupted at the wavenumber q−, while for complete phase separa- tion and phase amplification, the maximum of the structure factor will evolve to q = 0 for an infinite-sized system. . . . . . . . . . . . . . . . . . . . . . . . . . 80 3.4 Temporal evolution of the structure factor: a) for a system undergoing diffusion dynamics (M = 1) toward an equilibrium state in the absence of natural intercon- version (L = 0) and forced interconversion (K = 0); b) for a system undergoing a hybrid of diffusion (M = 1) and natural interconversion (L = 0.01) dynam- ics in the presence of forced interconversion (K = 1.5× 10−3) toward a steady state. The structure factor, given by Eq. (3.23), exhibits a crossover from spinodal decomposition to the nucleation regime. The dashed-black curves indicate the de- velopment of the maximum of the structure factor. The characteristic crossover time is defined in Eq. (3.24) and adopted as τ = 100. In (a) the evolution of the maximum of the structure factor moves to q = 0 for infinite-size system and satu- rates at Sm(q = 0, t→∞) = 1/(2ξ2) = 5 for ∆T̂ = −0.1. In contrast, in (b) the evolution of the maximum is interrupted at a characteristic cut-off wavenumber predicted by the characteristic phase domain growth rate, qs m(t→∞) ∝ q−, and it saturates at Sm(q−, t→∞) = 417. . . . . . . . . . . . . . . . . . . . . . . . . 82 3.5 The time evolution of the wavenumber corresponding to the maximum of the structure factor, given by Eq. (3.23), during the crossover from the early stage of spinodal decomposition, q ∝ t1/4 (green - dashed), to the nucleation regime, q ∝ t1/3 (orange - dashed) for a system undergoing diffusion dynamics in the absence of natural interconversion (L = 0) and forced interconversion (K = 0) under conditions: M = 1, ∆T̂ = −0.1, τ = 100. . . . . . . . . . . . . . . . . . 83 xv 3.6 The effect of increasing interconversion force on the phase domain growth rate for M = 100, L = 1, and ∆T̂ = −0.5. The red dashed line corresponds to the inverse maximum size of the phase domain on the length scale of the simulation box, q∗. When q− > q∗ microphase domains will form. Alternatively, when q− < q∗, the size of the simulation box will cut-off the growing phase domains. The conditions where ω(q) < 0 (dashed-dot portions of the curves) corresponds to non-growing wavenumbers. As the rate of forced interconversion increases, the growth rate is shifted down from the onset of phase separation where q− = q∗ (red, K = 1), to the microphase region (green, K = 3.75), to the termination point of domain growth (blue, K = 6.5) where q− = qm = q+ = q∗∗, and to the no growth regime for any wavenumber (orange, K = 9.25). . . . . . . . . . . . 84 3.7 Time evolution of the structure factor computed from the Fast Fourier trans- form (FFT) of Eq. (3.7) for M = 1, L = 1/127, ∆T̂ = −0.1, ` = 64, σi = 0.1, η = 10−5 depicted at times t = 6× 103 (green), t = 1.2× 104 (blue), t = 2.4× 104 (orange), t = 5× 104 (red), t = 1× 105 (pink), and t = 2× 105 (black). The open circles in (a-d) depict the computed structure factors for the four selected magnitudes of forced interconversion averaged over N = 100 real- izations with 95% confidence interval error bars, while the solid lines illustrate the behavior of the structure factors assuming a Gaussian distribution. The wavenum- ber is normalized by the size of the system, such that q = 1 corresponds to phase domains with a characteristic size of half the simulation box, `/2. . . . . . . . . . 86 3.8 The temporal evolution of the symmetry of phase separation. a) The time evolu- tion of the average order parameter, calculated by first averaging over all space and second by averaging the absolute value over N = 100 realizations, for M = 1, L = 1/127, ∆T = −0.1, σi = 0.1, η = 1.0× 10−5, and various mag- nitudes of forced interconversion, K. b) The time evolution of the N -averaged standard deviation of the averaged order parameter, calculated by first determin- ing the standard deviation of the spatially averaged order parameter and second by averaging over N = 100 realizations. This method of averaging highlights the behavioral deviation from an equal concentration of species A and B, ϕ = 0. . . . 88 3.9 Three hypothesized binary mixture systems exhibiting interconversion of species and liquid-liquid phase separation quenched from high temperature to low tem- perature (without volume change). The black dashed curve corresponds to the liquid-liquid phase coexistence in this system without interconversion and with interaction energy, ε = 2. The open circle indicates the liquid-liquid critical point (LLCP), while the crosses show the locations of T = TBA, the points correspond- ing to 50:50 interconversion for different energy change of reaction. For a system with TBA = 1.05Tc (green), no liquid-liquid phase transition will be observed upon quenching. For a system with TBA = Tc (red), the quenching process passes through the critical point. For a system with TBA = 0.95Tc (purple), there are two equilibrium solutions for the fraction of interconversion, such that upon quench- ing to the cross, phase amplification occurs with equal probability of forming an A-rich or B-rich phase. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 xvi 3.10 The phase diagram (suggested in ref. [13]) for supercooled water that exhibits a liquid-liquid phase transition (global phase diagram illustrated in Fig. 1.1). A hypothesized quenching process by compression for supercooled water is shown from the one phase region at P1 = 20 MPa (orange) to the two phase region at P2 = 120 MPa (green) along the Widom line (dashed black) which corresponds to a line of constant fraction of interconversion, lnK = 0. Two additional isobars are shown for reference at P = 40 MPa (blue) and P = 80 MPa (purple) along with the liquid-liquid coexistence (black). Phase amplification would only be possible in a system where the number of molecules changes to compensate the volume change of the interconversion reaction. . . . . . . . . . . . . . . . . . . . 92 3.11 a) Characteristic phase domain growth rate in the vicinity of the critical point (∆T̂ = −0.001) forM0 = 1, L = 0.002, andK = 2.25×10−5 calculated through Eq. (3.13), red curve, with use of the diverging molecular mobility, Eq. (3.27), and scaling inverse susceptibility in the first order epsilon expansion, χ̂−1 q=0 ∼ |∆T̂ |−γ with γ = 1 + ε/6 (ε = 4 − d). The meanfield approximation is shown by the green curve, Eq. (3.13). b) The onset, red solid curves, Eq. (3.28), and termination, red dashed curves, Eq. (3.29), of microphase separation affected by critical fluctuations for M = 1, L = 0.01, ` = 100, ν = 1/2 + ε/12. The meanfield approximation is shown by the green curves. . . . . . . . . . . . . . . 96 4.1 The spontaneous equilibrium order parameter (ϕ = ϕ0) in the lattice gas / lattice binary mixture along the liquid-vapor phase coexistence (red domain). One of the two alternative magnetizations (ϕ0 > 0 and ϕ0 < 0) in the Ising ferromagnet in zero field are shown in the red domain with blue arrows. The solid curve is the crossover from meanfield behavior (dashed) to the asymptotic scaling power law ϕ ∝ ∆T β with β = 0.326 [52, 53], while the crosses correspond to simulation results of the HL model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 4.2 Phase amplification in the HL model for systems with T = 4.0, ` = 100, and interconversion probabilities pr: (a) 100%, (b) 10%, (c) 10−4%, and (d) 10−7%. The inset in (d) shows 10−7% interconversion probability at a shorter time scale. Each curve represents one of the 100 different realizations of the MC simulations, while the solid zero line depicts the behavior of a system with 0% interconversion. The simulation snapshots in (a,d) correspond to the end of the simulation time. . . 104 4.3 The evolution of the order parameter during phase amplification. (a) The RMS of the distribution of the growth rates for different probabilities captured at the same time, t = 300. The solid curve is the crossover between σ ∝ √pr and σ ∝ pr, approximated as σ = a √ pr(1 + bpr)/(1 + √ pr). (b-d) The growth of the order parameter for different (b) probabilities at ∆T̂ = −0.11 and ` = 100, (c) system sizes at pr = 1.0 and ∆T̂ = −0.11, and (d) distances to the critical point at pr = 1.0 and ` = 100; The inset shows the power law for the initial growth of the reduced order parameter, ϕ/ϕ0 ∝ t3/4. . . . . . . . . . . . . . . . . . . . . . 106 xvii 4.4 Scaling properties of the growth of the reduced order parameter, ϕ̂ = 〈ϕ〉/ϕ0. (a) The order parameter growth with time rescaled by system size. The size depen- dence of the rescaling parameter, τ(`), is shown in the inset; in the log-log scale with a slope of 1. The colors are the same as in Fig. 4.3c. (b) The order pa- rameter growth with time rescaled by probability; the rescaling parameter τ0(pr), inversely proportional to the probability, is shown in the inset. The colors are the same as in Fig. 4.3b. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 4.5 Topological characteristics of the time dependence of phase amplification. (a) For spherical domains, the reduced deviation from the equilibrium order parameter, ∆ϕ̂ = 1 − ϕ/ϕ0 scales as t3/2. This is shown for temperatures, ∆T̂ , as: −0.11 (cyan), −0.025 (purple), −0.014 (orange), and −0.009 (brown). The inset shows the effect for a cylindrical domain at ∆T̂ = −0.11. (b) A zero curvature Schwarz- P interface is initially formed by simulating a system with Kawasaki dynamics (pr = 0) for a long time. At t − t0 = 0, the system obtains Glauber dynamics (pr = 1) and the collapse of one of the phases is shown. Amplification transitions from random-walk behavior, √ t, at short times (see inset) to exponential behavior before saturation shown by the straight line. . . . . . . . . . . . . . . . . . . . . 110 4.6 The time evolution of the concentration for the HL model with pure Ising dy- namics (pr = 1), given by Eq. (4.5) with the time-dependent susceptibility in the form of Eq. (4.4). For ∆T̂ = −0.32 (green), τ = 0.2, ϕ∞ = 0.91, a = 1.9 and b = 0.3, and for ∆T̂ = −0.11 (blue) τ = 0.2, ϕ∞ = 0.73, a = 4.6, and b = 0.4. The open circles are the computational data presented in Fig. 4.3d. . . . . . . . . 113 4.7 Effect of forced interconversion on domain size, R, normalized by the system size, `, in the HL model. (a) The time dependence of the domain growth for en- ergy source E = 5 and interconversion probability pr = 1/128 at T/Tc = 0.24 (green), T/Tc = 0.27 (blue), and T/Tc = 0.40 (red), where Tc = 4.511 [54]. The horizontal dashed line indicates the size of the simulation box, R = `. (b) Tem- perature dependence of the steady-state domain size for E = 5 and pr = 1/128. The vertical dashed line indicates the onset temperature, T ∗/Tc. (c) Depen- dence of the steady-state domain size on the energy of forced interconversion for pr = 1/256 and T/Tc = 0.31. The vertical dashed line denotes the onset source energy, E∗. In (a-c), the system is simulated on a 3-dimensional lattice of size ` = 100. The open circles are the results of MC simulations, the images are snapshots of the system for selected conditions, and the curves are the theoretical predictions. In (a-c), black denotes up-spins and white denotes down-spins. . . . 116 4.8 Domain size as a function of temperature in the HL model. (a) pr = 1/32 and (b) pr = 1/256, for a system of size ` = 100 and energies: E = 1 (orange), E = 2 (green), E = 3 (red), E = 4 (purple), E = 5 (brown), E = 6 (pink), E = 7 (gray), E = 8 (yellow), E = 9 (cyan), and E = 10 (dark blue). The solid curves are the theoretical predictions of q−, the inverse steady-state domain size, where the fit parameters are provided in Table 4.1. . . . . . . . . . . . . . . . . . . . . 118 xviii 4.9 a) Steady-state structure factors computed for the HL (open circles) and the pre- diction given by Eq. (4.7) (solid lines) for selected external energy sources (E) at ∆T̂ = −0.4, M = 1, L = 1/127, ` = 100, and averaged over N = 60 re- alizations with 95% confidence interval error bars. The insets show steady-state (t ∼ 3× 105) domain morphologies observed in the HL model for the selected energies. b) The dependence of forced interconversion on the wavenumber cor- responding to the maximum of the structure factor, qs m, in the steady-state limit. The open circles are numerical computations of structure factors determined from FFTs of the time evolution of the order parameter, given by Eq. (3.7), in the steady-state limit (t ∼ 105) for M = 1, L = 1/127, σi = 0.1, and η = 10−5, averaged over N = 100 realizations. The triangles correspond to the predic- tions of K determined from fits of Eq. (4.7) to the structure factor for the HL model, like those presented in (a). The curves illustrate the theoretical predic- tion qm(t→∞) ∝ q−, given by the full expression for q−, found from evaluating ω(q, 0) = 0 using Eq. (3.15). The colors correspond to temperatures: ∆T̂ = −0.1 (blue), ∆T̂ = −0.2 (green), ∆T̂ = −0.3 (red), and ∆T̂ = −0.4 (purple). The inset shows the relationship between K and E. . . . . . . . . . . . . . . . . . . . 121 4.10 Molecular representation and geometrical features of tetramer molecules. L- enantiomers (green), D-enantiomers (blue) and achiral transition states (cis- or trans-, red) are shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 4.11 Time dependence of the instantaneous dihedral angle of a typical tetramer in a racemic mixture at P = 0.1 for the conservative formulation of the CM model, at several values of T and kd. a) T = 0.6, b) T = 1.7, and c) T = 2.3, with kd = 5 (green), kd = 9.86 (orange) and kd = 19.86 (purple). d) Behavior of the dihedral angle at a very low value of the dihedral constant, kd = 0.001. . . . . . . . . . . 128 4.12 The phase diagram showing phase amplification in the CM model with conser- vative intermolecular forces, heterochiral bias parameter λ = 0.5, and rigidity spring constant kd = 0.001. The circles on the solid curve are the computational data for the critical temperature of equilibrium phase separation and the curve is the fit of Eq. (4.12). The snapshots depict the equilibrium states for the pres- sures P = 0.1, P = 1, P = 5, and P = 10 below the critical temperature and at P = 1 above the critical temperature. The triangles show the prediction of the critical temperature from the extrapolation of the CM model with dissipative intermolecular forces to the limit kd →∞. . . . . . . . . . . . . . . . . . . . . . 129 4.13 The change of compositional heterogeneity with chiral interconversion kinetics at T = 1.7 and P = 0.1 in the dissipative-force formulation. a) Steady-state snapshots of chiral liquid systems at various dihedral force constants (kd). b) The steady-state domain size as a function of interconversion rate, 1/τINC. The solid line is the approximation given by 1/τINC = a1/R 2 ∞ + a2/R 4 ∞, which follows from Eq. (4.15), where a1 = 4.6 × 10−3 and a2 = 3.8 × 10−4. The inset shows the linear correlation betweenR∞ and kd. The colored points highlight the results corresponding to the three dihedral force constants for which the domain growth is shown in Fig. 4.15. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 xix 4.14 Steady-state domain size, R, normalized by the size of the system, `, in the CM model. (a) The time evolution of the domain size for different interconversion rates, tuned by the rigidity parameter, kd, as kd = 3 (purple), kd = 5 (green), and kd = 9.86 (red) at the reduced pressure P = 0.1 and T/Tc = 0.35, where Tc(P = 0.1) = 2.32, as indicated on Fig. 4.12. (b) The normalized steady-state domain size as a function of temperature at P = 0.1 and kd = 5. The verti- cal dashed line indicates the onset temperature, T ∗/Tc. In (a) and (b), the open circles correspond to simulation results, the curves correspond to the theoretical predictions (see Table 4.3), and the images show snapshots of the system simu- lated at the indicated conditions. In (a-b), dark/clear spheres correspond to the L-/D-configuration of a chiral tetramer (spheres are located a tetramer’s center of mass). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 4.15 Phase domain growth for the dissipative-force formulation of the CM model. The open circles represent computational data [7], and the curves illustrates predic- tions of the time evolution of the domain size, R(t) = 1/qs m, from Eq. (3.23) for dihedral force constants: kd = 5 (green), kd = 9.86 (red), kd = 19.86 (blue), and kd → ∞ (black) for T = 1.8, Tc = 2.3, τ = 2, and M = 0.0022. The domain size is normalized by the size of the computational box, `. The dashed curves represent the predictions of the domain growth if it is not restricted by the finite size. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 4.16 The mutual diffusion coefficient affected by interconversion, given by Eq. (4.17), as a function of the normalized steady-state domain size R∞, Eq. (4.18) in the CM model. The open circles are computational data for the dissipative-force formulation of the CM model at three different dihedral force constants: kd = 5 (green), kd = 9.86 (red), and kd = 19.86 (blue) [7]. The molecular mobility in the limit kd → ∞ is approximated as M = b0T/η, where b0 = 0.94 and η is the viscosity approximated by the Arrhenius equation η ∼ eT0/T . The characteristic temperature T0 is 3.5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 4.17 Temperature dependence of the characteristic time scales in the dissipative-force formulation of the CM model for dihedral force constants: kd = 5 (green), kd = 9.86 (orange), and kd = 19.86 (purple) at P = 0.1. a) Characteristic time for liquid-liquid phase separation at the length scale of the simulation box, q = q∗. The curves are τLLPS ∝ 1/ω̃(q∗) where ω̃(q) is given by Eq. (4.16) where it was found that q∗ = 0.15 and Tc = 2.35. The condition τLLPS → ∞ corresponds to T = T ∗. b) Characteristic times for chiral interconversion. τINC ∝ 1/L, and the curves are given by Eq. (4.13). c) Characteristic self-diffusion times, where the curves are given by Eq. (4.19). . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 xx 4.18 Dihedral force constant dependence of the onset temperature of phase separation on the length scale of the simulation box in the dissipative-force formulation CM model for P = 1.0 (red circles), P = 0.5 (blue circles), and P = 0.1 (black circles). The curves are numerically calculated from the first solution of ω̃ = 0, given by Eq. (4.16), when q− = q∗ = 0.11 ≈ 1/Rmax ∼ 1/`, T = T ∗, and Tc(P = 1.0) = 3.45, Tc(P = 0.5) = 2.91, and Tc(P = 0.1) = 2.3 for different pressures (a) and in rescaled coordinates (b). The triangles, shown in (b), are obtained from the asymptotic limits of the time of liquid-liquid phase separation, τLLPS →∞, as shown in Fig. 4.17a for q∗ = 0.15 and Tc = 2.35. . . . . . . . . . 141 4.19 The conditions for microphase separation at a length scale, R∞ = 1/q−. The solid curve depicts the onset temperature of microphase separation, T = T ∗, at the scale of the computational box, Eq. (4.20), for R∞ = 1/q∗ = 7.1 and Tc = 2.4. The two dashed curves depict the lines of steady-state domain size that are smaller than the size of the computational box, R∞ = 3.6 (lower) and R∞ = 2.0 (upper), while the dotted curve depicts the growth-termination tem- perature, T = T ∗∗, Eq. (4.21). The symbols are computational data (from the observed size of the phase domain) in the dissipative-force formulation of the CM model for pressure P = 0.1. The symbols indicate the onset of phase sep- aration (open circles), the microphase region (R∞ = 1/q∗, closed circles), and the two phase region (R∞ > 1/q∗, triangle). The two simulation snapshots illus- trate the behavior of the system composed of A-rich (green), B-rich (blue), and intermediate (red) states of the CM model. [7] . . . . . . . . . . . . . . . . . . . 143 4.20 Self-diffusion (molecular mobility) of the HCS model. Simulations were con- duced for a system with b = 0 (no interconversion, either natural or forced). The open circles indicate simulation results, while the curve is the fit of M = 0.056 √ T + 0.015T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 4.21 Critical temperature of the HCS model. Internal energy, ∆U , of the HCS model as a function of temperature (blue) for the case b = 0 (no interconversion, either natural or forced). The blue circles indicate the results of simulations, while the red circles correspond to the calculated isochoric heat capacity, CV = ∂U/∂T |V . The solid red curve is an empirical fit ofCV , where the temperature corresponding to the maximum of CV is the critical temperature for the HCS model, Tc = 3.6± 0.05, in units of ε0/kB. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146 4.22 Steady-state domain size, R, normalized by the size of the system, `, in the HCS model. (a) The temperature-dependence of the normalized steady-state domain size for b = 0.005 (blue), b = 0.050 (black), and b = 0.075 (green) at ε = 10. (b) The normalized steady-state domain size as a function of interconversion probability, b, for the energy source ε = 12 and T/Tc = 0.22. The vertical dashed line indicates the onset interconversion probability, b∗. In (a) and (b), the open circles correspond to simulation of 64, 000 particles, the curves correspond to the theoretical predictions, and the images show snapshots of the system simulated at the indicated conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 xxi 4.23 Structure factor, S(q, t), in the HCS model. The formation of a steady-state bi- continuous microemulsions observed in MD simulations of the HCS model for 64,000 particles at interconversion energy source ε = 12, T/Tc = 0.1, and inter- conversion probability b = 0.05 (a) and b = 0.15 (b). The images show snapshots of the steady-state system simulated at the specified conditions. . . . . . . . . . . 149 4.24 Domain size as a function of temperature for the HCS model. (a) ε = 6 and (b) ε = 8, for system size ` = 40σ, averaged over N = 10 realizations, for interconversion probabilities, b = 0.005 (blue), b = 0.05 (black), and b = 0.075 (green). It should be noted that for low temperatures, T/Tc ≤ 0.8, the simulated systems freeze, such that the characteristic domain size is always less than the predicted size (as indicated by the triangles). . . . . . . . . . . . . . . . . . . . . 150 4.25 The onset and termination of microphase separation. (a) The steady-state domain size in the HL model for pr = 1/128 and different forced interconversion ener- gies from E = 1 (orange) to E = 10 (dark blue) in steps of ∆E = 1, just as in Fig. 4.8 (Fit parameters provided in Table 4.4). For E > E∗∗ = 7 (the ter- mination energy), the data collapse into a single line (black), indicating that the characteristic steady-state domain size remains on the order of the microscopic length scale R0(T ), which corresponds to homogeneous steady-state systems for all temperatures. For E ≤ 7, the onset of microphase separation is observed at T = T ∗(E∗), where E∗ is the onset energy. For T < T ∗, the steady-state domain size is equal to the size of the system, R = `. (Caption continues on the next page.)154 4.25 (b) The onset energy E∗ (black circles and curve) for the HL model as a function of temperature for pr = 1/128. Colored open circles and dashed curves corre- spond to steady-state domain sizes: R = 0.143 (blue), R = 0.095 (green), and R = 0.074 (red). (c) The inverse onset rigidity parameter 1/k∗d (black circles and curve) for the CM model at P = 0.1. Colored circles and dashed curves correspond to steady-state domain sizes: R = 0.32 (blue), R = 0.22 (green), and R = 0.18 (red). In (b) and (c), the blue area corresponds to phase separation on the scale of the simulation box, the white area corresponds to microphase separa- tion, and the yellow area corresponds to homogeneous steady states. The images in (b) and (c) correspond to the different final states of the systems below E∗(T ∗) and 1/k∗d(T ∗) where the size of the phase domain is on the scale of the simulation box (q∗ ∼ 1/`). In (b), phase amplification is observed since for pr = 1/128 nat- ural interconversion is relatively fast, while in (c), where natural interconversion is relatively slow for the simulated range of kd, complete two-phase separation, on the scale of the simulation box, is found. . . . . . . . . . . . . . . . . . . . . 155 A.1 Comparison between the Fisher-Wortis (FW) and an alternative symmetric ansatz (sym). obtained for the liquid-vapor coexistence, for two systems with εBA = 1.04 (a) and εBA = 1.12 (b) with εAA = 1.6, εBB = 2.0, ê = 3, and ŝ = 4. (c) The relative deviation between the symmetric and FW ansatzes. . . . . . . . . . . . . 161 A.2 Numerical calculations of the liquid-vapor surface tension for three temperatures in the system with εBA = 1.04, εAA = 1.60, εBB = 2.00, ê = 3, and ŝ = 4. The FW ansatz is given at iteration 0. The temperatures were chosen based on surface tensions with similar values, as predicted by the FW ansatz. . . . . . . . . . . . . 162 xxii A.3 Comparison of the liquid-vapor density (a) and concentration (b) interfacial pro- files, as obtained from numerical calculations after eight iterations, ρV4 and xV4, and for the FW ansatz, ρV0 and xV0, for the system with εBA = 1.04, εAA = 1.60, εBB = 2.00, ê = 3, and ŝ = 4 at temperature, T = 0.6943 (red curve in Fig. A.2). (c,d) The difference between the exact solution for the density and concentration profiles and the FW and symmetric ansatzes. In (a-d), the different profiles were aligned along their Gibbs dividing surface, such that the excess density is zero for all profiles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 xxiii List of Abbreviations and Key Notations Acronyms FFCP Fluid-fluid critical point FFPT Fluid-fluid phase transition FFT Fast Fourier transform FW Fisher-Wortis HDL High-density liquid INC Characteristic interconversion LDL Low-density liquid LLCP Liquid-liquid critical point LLPS Liquid-liquid phase separation LLPT Liquid-liquid phase transition LVCL Liquid-vapor critical line LVCP Liquid-vapor critical point MC Monte Carlo MD Molecular dynamics DMD Discrete molecular dynamics QMC Quantum Monte Carlo RG Renormalization Group RMS Root mean square SFF-TP Solid-fluid-fluid triple point SLL-TP Solid-liquid-liquid triple point SLG-TP Solid-liquid-gas triple point TP Triple point Simulated Models HL Hybrid Ising / binary lattice CM Chiral mixture of interconverting enantiomers HCS Coarse-grained hard-core-shoulder Key Notation b Interconversion probability in the HCS model ε Non-ideality interaction parameter ` Size of a system Ft Total free energy (Helmholtz energy) f = Ft/N Free energy per molecule (or per lattice site) Gt Total Gibbs energy xxiv G = Gt/N Gibbs energy per molecule (or per lattice site) Gmix Gibbs energy of mixing GBA Gibbs energy of reaction κ The square of the range of molecular interactions kB Boltzmann’s constant kd Force constant of dihedral motion in the CM model K Kinetic coefficient of forced interconversion L Kinetic coefficient of natural interconversion M Kinetic coefficient of diffusion (Mobility) µ Chemical potential µth Thermodynamic “bulk” part of the chemical potential N Number of molecules (or lattice sites) P Pressure pr Probability (rate) of interconversion in the HL model ϕ Order parameter R Phase domain size Rid Ideal gas constant ρ Density S Structure Factor T Temperature q Wavenumber qm Wavenumber corresponding to the maximum of the amplification factor qs m Wavenumber corresponding to the maximum of the structure factor q− Wavenumber corresponding to the first root of the amplification factor q∗ Wavenumber corresponding to the onset of microphase separation q∗∗ Wavenumber corresponding to the termination of microphase separation ω Amplification factor (phase domain growth rate) Ωt Total grand thermodynamic potential Ω = Ωt/N Grand thermodynamic potential per lattice site x Molecular fraction ξ Correlation length of order-parameter fluctuations xxv Chapter 1: Introduction Typically, pure substances may be found with only one gaseous or liquid state, while their solid state may exist in various polymorphic crystalline states. The existence of multiple dis- tinct liquid forms in a single component substance is more unusual since liquids lack the long- range order common to crystals. Yet, the existence of multiple amorphous liquid states in a single component substance, a phenomenon known as “liquid polyamorphism,” [9, 55, 56] has been observed or predicted in a wide variety of substances, such as superfluid helium [57, 58], high-pressure hydrogen [1, 35, 37, 38, 59], high-density sulfur [2, 51], phosphorous [60, 61], car- bon [62], silicon [63–66], silica [67,68], selenium and tellurium [69,70], and cerium [71]. Liquid polyamorphism is also highly plausible in deeply supercooled liquid water [9, 13, 55, 56, 72–80]. A substance may be found to be polyamorphic by experimentally or computationally de- tecting a liquid-liquid phase transition (LLPT), which terminates at a liquid-liquid critical point (LLCP). For example, liquid polyamorphism, via the existence of two alternative supramolec- ular structures, has been hypothesized to explain the remarkable anomalies in the thermody- namic properties of supercooled water, namely a maximum in the temperature dependence of its density and its isothermal compressibility [81, 82], a maximum in the isobaric heat capac- ity [83], and an inflection point in its surface tension [84–87]. Simulations of water-like mod- els [74–77, 79, 88–91], apparently supported by experiment [92, 93], have demonstrated the hy- 1 Figure 1.1: Hypothesized global pressure-temperature phase diagram for water. The black curves indicate the solid-solid and solid-liquid-gas phase transitions experimentally reported for wa- ter [11], while the blue curve indicates the hypothesized liquid-liquid phase transition. The dotted black curve represents the line of spontaneous ice nucleation [12]. The red circle is the liquid- vapor critical point (LVCP) and the red star is the hypothesized liquid-liquid critical point (LLCP) around 220 K and 60 MPa as suggested in Ref. [13]. pothesized LLPT in supercooled water, see Fig. 1.1. Fluid polyamorphism ultimately originates from the complex interactions between molecules or supramolecular structures. However, it may be phenomenologically modeled through the re- versible interconversion of two alternative molecular or supramolecular states [4, 9, 10]. The application of this “two-state” approach to the variety of polyamorphic substances could be just as a useful phenomenology or it may reflect the true microscopic origin of fluid polyamorphism. Indeed, there are a few substances, such as hydrogen, sulfur, phosphorous, and carbon, where the existence of alternative liquid or dense-fluid states is explicitly induced by a reversible chem- ical reaction: dimerization in hydrogen [1] or polymerization in sulfur, phosphorus, and car- bon [2]. While equilibrium phase transitions in simple fluids have been well-studied, the descrip- 2 tion of phase transformations in the presence of interconversion between alternative molecular or supramolecular states, and in fluids far from equilibrium, is much less investigated [94–96]. The interconversion of two-states significantly effects both the thermodynamics and dy- namics of fluid mixtures. In particular, molecular interconversion imposes an additional ther- modynamic constraint, the chemical-reaction equilibrium condition, which reduces the number of thermodynamic degrees of freedom. In this case, the concentration of a binary mixture is no longer an independent variable, being a function of temperature and pressure. Consequently, a bi- nary mixture with interconversion of species follows the Gibbs phase rule for a single-component substance. In the absence of interconversion, the binary mixture exhibits a critical point for each concentration. Collectively, these critical points make up a critical locus. The interconversion of species selects a single path through the planes of phase coexistence at fixed concentration, crossing the critical locus of the binary mixture at a unique point. Moreover, this path could cross the critical line, the line of triple points, or any other unique line on the phase diagram more than once. The resulting phase diagram can be viewed as the phase diagram of a polyamorphic single-component fluid with multiple fluid-fluid critical points. The thermodynamic properties, being state functions, under interconversion equilibrium conditions are the same as those in the non-reacting binary mixture at the corresponding equi- librium temperature, pressure, and overall concentration. Therefore, at any point along the ther- modynamic path selected by interconversion, the properties are affected by the proximity to the critical line of the non-reacting binary mixture. It should be noted that the evolution of the ther- modynamic properties along the path (where the equilibrium concentration changes) is generally very different from the evolution of the same property at constant composition, which would typically be studied in non-reacting binary mixtures. This line in the interconverting mixture is 3 referred to as the “virtual critical line,” which may be used to describe the anomalous behavior of the thermodynamic properties along this path. The interconversion of species also effects the dynamics of phase formation. Generally, a mixture of two liquids will phase separate if their interactions are not favorable for mutual mix- ing. The most recognizable example is almost complete separation of water and oil. Another example is a possible demixing of structural isomers, such as enantiomers with opposite molec- ular chirality [7, 97–99]. In contrast, solid ferromagnetic (such as iron) or ferroelectric (such as barium titanate) materials, in the absence of a magnetic or electric field, do not establish equilib- rium coexistence between phases with alternative magnetizations or polarizations [94, 100, 101]. However, if liquids exhibit interconversion of species, similar to the flipping of magnetic spins or electric dipoles, the thermodynamics and dynamics of phase separation will dramatically change. In this thesis, it is shown that molecular interconversion may destroy or restrict liquid-liquid phase separation. After a binary mixture, initially containing equal amounts of the alternative molecules, is quenched from the one-phase homogeneous region at a high temperature into the unstable region below the critical temperature of demixing at constant pressure, the species will separate through a process known as spinodal decomposition [102]. However, if the two species with the same density may rapidly interconvert (like in a mixture of enantiomers [7]), then to avoid the formation of an energetically unfavorable interface, the phases will compete with each other until one of them is eliminated [6]. This is the phenomenon of phase amplification, the result of the competition between the diffusive dynamics of phase separation and the “flipping” dynamics of interconversion. In magnetic and ferroelectric materials, away from the Curie tem- perature and in the absence of a magnetic or electric field, the phenomenon of phase amplification occurs naturally. In such materials, there is no restriction on the direction of magnetization or 4 polarization, meaning that there is no conservation of the number of magnetic spins or electric dipoles with a particular orientation. Similarly, in fluids with fast molecular interconversion, the conservation of the number of alternative molecules is broken and phase amplification could occur. In interconverting fluids, the growth of one phase at the expense of another stable phase occurs to avoid the formation of an energetically unfavorable interface, and consequently, the coexistence of the alternative phases is destroyed [6]. Since the interface is crucial to phase am- plification, in macroscopic systems where the interfacial energy is much smaller than the bulk energy, the system may enter a metastable state in which the energetic benefit of phase amplifi- cation is negligible [4]. Thus, the size of the system plays a crucial role in phase amplification. As discussed in Chapter 3, it is also shown that, in addition to smaller system sizes, a faster rate of the interconversion reaction and closer proximity to the LLCP of demixing are necessary for one to observe phase amplification [4, 6]. In a nonequilibrium system, if an external force causes the alternative molecules to stay in equal numbers, the striking phenomenon of steady-state, restricted (“microphase”) separation into mesoscale domains may be observed. In equilibrium, examples of mesoscale structures are present in bicontinuous or spatially modulated microemulsions [103, 104] and microphase sepa- ration of diblock or polyelectrolyte copolymers [105,106], where these mesoscale patterns are the result of the minimization of the equilibrium free energy [9,107]. As discussed in Chapter 4, it is found that the structure of the phase domains formed from nonequilibrium microphase separation resembles modulated or bicontinuous microemulsion structures [9,103–107]. However, contrary to the patterns formed in equilibrium or in metastable (“frozen”) conditions [108,109] (like those commonly observed in glasses [110,111]), these nonequilibrium structures persist in steady-state 5 due to the continuous energy supply. Thus, steady-state microphase separation is one of the sim- plest examples of dissipative structures in condensed matter. The characteristic length scale of this dissipative structure emerges as a result of the competition between forced interconversion and phase growth. If the source of forced interconversion is not sufficiently strong to overcome the natural interconversion of alternative species, then the phenomenon of phase amplification, the growth of one stable phase at the expense of another phase, is observed [4, 6, 112]. An ex- ternal racemizing energy source can be achieved in physical systems through the interactions of energy-carrying particles, such as photons, that may break intramolecular bonds [111]. It can also be seen in biological cells through a flux of energy produced by ATP [113, 114] or it could be achieved chemically through an external flux of matter or heat [115–117]. The size of the system, rate of forced interconversion, and proximity to the LLCP are also important to observe microphase separation. In this thesis, it is found that there are two key conditions for microphase separation to be observed. First, if the characteristic size of the meso- scopic steady-state microphase domains is comparable to half the size of the system, then the system will produce the same two alternative phases that would be observed without intercon- version [4, 5, 8]. As a result, the size of the system may “cut off” the system’s ability to phase separate into microdomains. Second, if the rate of forced interconversion is much faster than the natural interconversion or diffusion rate, then the external force dominates the systems’ ki- netics and no phase formation is possible. Consequently, it is shown that the size of the system and the rate of the forced interconversion reaction are crucial components to observe microphase separation [4, 5, 8]. This thesis is organized as follows: In Chapter 2, the two-state thermodynamic approach to model fluid polyamorphism is discussed. The application of this approach is considered for 6 fluid-fluid transitions in hydrogen, sulfur, and substances exhibiting water-like anomalies. In Chapter 3, the Cahn-Hilliard model of spinodal decomposition (phase separation process that occurs after a binary mixture is quenched into the unstable region), is generalized to include equilibrium (natural) and nonequilibrium (forced) interconversion of species. In Chapter 4, the generalized Cahn-Hilliard theory is applied to describe the behavior of microscopic models of mixtures, simulated through molecular dynamics (MD) or Monte Carlo (MC) methods, which exhibit both natural and forced molecular interconversion. General concluding remarks and sug- gestions for future research are presented in Chapter 5. 7 Chapter 2: Two-State Thermodynamics of Fluid Polyamorphism In this chapter, the two-state thermodynamic approach for fluid polyamorphism is intro- duced (Sec. 2.1). The approach is applied to describe the fluid-fluid transition in high-pressure hydrogen (Sec. 2.2), high-density sulfur (Sec. 2.3), and fluids exhibiting water-like anomalies (Sec. 2.4). In addition to the phase behavior of these substances, the approach is also used to describe the anomalies in the interfacial properties of fluids exhibiting interconversion between species. 2.1 Interconversion of Two States Consider a symmetric binary mixture of two species A and B with molecular fractions, xB = x and xA = 1 − x. Initially, it may be assumed that both species have the same densities (ρ = 1), viscosities, and molecular weights. This system can be described by a Landau-Ginzburg free-energy functional with a single order parameter uniquely linked to the fraction of species B as, ϕ = x/xc − 1, where xc is the critical fraction. This functional reads as F [{ϕ}] = 1 ρ ∫ V ( Ĝ(ϕ, T, P ) + 1 2 κ|∇ϕ|2 ) dV (2.1) 8 Figure 2.1: A symmetric binary mixture is quenched along the critical molecular fraction (where the order parameter is ϕ = 0) from a point, T̂1 in the one phase region to a point, T̂2 below the critical temperature of demixing in the unstable region. The solid curve indicates the liquid-liquid coexistence (binodal), while the dashed curve indicates the limit of stability (spinodal). where the first term represents the thermodynamic “bulk” free energy and the second term is included to describe the contribution to the free energy due to inhomogeneities within the system. For an isotropic system, the coefficient κ is the square of the range of intermolecular interactions, on the order of the square of the molecular size. The bulk free energy density for the system, Ĝ, is the reduced Gibbs energy, Ĝ = G/kBTc, where kB is Boltzmann’s constant and the critical temperature for liquid-liquid phase separation is Tc. The bulk free energy density may be expressed in the general form, Ĝ = ĜA + xĜBA + Ĝmix (2.2) where ĜBA = µ̂0 B− µ̂0 A is the reduced difference between the Gibbs energies (chemical potentials) of pure species A and B, referred to as the Gibbs energy change of reaction. For the symmetric 9 binary-lattice (“regular solution”) model, Ĝmix is formulated through the order parameter, ϕ, and in the meanfield approximation, it reads as Ĝmix = T̂ [( 1 + ϕ 2 ) ln ( 1 + ϕ 2 ) + ( 1− ϕ 2 ) ln ( 1− ϕ 2 )] + ε 4 (1− ϕ2) (2.3) where T̂ = T/Tc is the reduced temperature and ε is a non-ideality interaction parameter, which generally depends on temperature and pressure. The conditions for liquid-liquid phase equilib- rium is ∂Ĝmix ∂ϕ ∣∣∣∣ T,P = T̂ 2 ln ( 1 + ϕ 1− ϕ ) − ε 2 ϕ = 0 (2.4) The critical molecular fraction, xc, and critical temperature, Tc, are obtained from the thermody- namic stability conditions ∂2Ĝ ∂ϕ2 ∣∣∣∣ T,P = 0 and ∂3Ĝ ∂ϕ3 ∣∣∣∣ T,P = 0 (2.5) as xc = 0.5 and Tc = ε/(2kB) (2.6) Consider the interconversion between molecular states A and B through a reversible chem- ical reaction of the form A k1−−⇀↽−− k2 B (2.7) where k1 and k2 are the forward and reverse reaction rates, respectively. Since, for this system, the chemical reaction coordinate is the order parameter, the chemical reaction equilibrium condition 10 reads ∂Ĝ ∂ϕ ∣∣∣∣ T,P = ∂Ĝmix ∂ϕ ∣∣∣∣ T,P − ĜBA = 0 (2.8) The reaction-equilibrium condition constrains the number of thermodynamic degrees of freedom for the system. Consequently, the fraction of interconversion, given through the order parameter ϕ, is no longer an independent thermodynamic variable, but instead, becomes a function of tem- perature and pressure. The reduced Gibbs energy change of the reaction, ĜBA, can be expressed through the reaction equilibrium constant, K(T̂ , P̂ ) = k1/k2, and is given, up to second-order in temperature and pressure, as ĜBA = −T̂ lnK(T̂ , P̂ ) = ê− ŝT̂ + υ̂P̂ + γT̂ P̂ + 1 2 δT̂ 2 − 1 2 κP̂ 2 (2.9) where ê = e/(kBTc), ŝ = s/kB, and υ̂ = υ/(kBTc) are the reduced energy, entropy, and volume changes of the reaction, while γ, δ, and κ are proportional to the volumetric expansivity, isobaric heat capacity, and isothermal compressibility changes of the reaction, respectively [9, 10]. The reduced chemical potential for such a system undergoing spinodal decomposition to- wards both chemical-reaction and phase equilibrium, µ̂ = µ/kBTc, is the reduced deviation of the chemical-potential difference in solution (µ = µA − µB) from its equilibrium value, µ = 0. The reduced time-dependent chemical potential is found from the functional derivative of Eq. (2.1) as µ̂ = µ̂th − κ∇2ϕ (2.10) where the order parameter depends on space and time, ϕ = ϕ(r, t). In this form, the chemical po- tential is comprised of a thermodynamic potential as µ̂th = ∂Ĝ/∂ϕ|T,P , and a spatial-dependent 11 part, κ∇2ϕ. The thermodynamic component [118] is µ̂th = T̂ 2 ln ( 1 + ϕ 1− ϕ ) − ε 2 ϕ− ĜBA (2.11) Expanding the logarithmic term to first order in ϕ via a Taylor series around ϕ = 0 (the value of the order parameter at the initial time t = 0), gives ∂Ĝ/∂ϕ ≈ χ̂−1 q=0ϕ, where the inverse ther- modynamic susceptibility in the limit of zero wavenumber is χ̂−1 q=0 = ∂2Ĝ/∂ϕ2. In the meanfield regular-solution model, the inverse susceptibility scales as χ̂−1 q=0 ∼ ∆T̂ , where ∆T̂ = T/Tc − 1 is the reduced distance to the critical temperature. Therefore, the chemical potential defined in Eq. (2.10) in the first-order approximation becomes µ̂ ≈ ∆T̂ϕ− ĜBA − κ∇2ϕ (2.12) It should be emphasized that the reduced chemical potential, given by Eq. (2.12), when the term related to interconversion is absent, is the same as obtained in the classical Cahn-Hilliard the- ory [102]. In addition to the chemical potential, the density and entropy are obtained from derivatives of the Gibbs energy, Eq. (2.2), as ρ̂(P̂ , T̂ ) = ( ∂Ĝ ∂P̂ )−1 ϕ,T̂ = [ 1 ρ̂A + xe ∂ĜBA ∂P̂ ∣∣∣∣ ϕ,T̂ + ∂Ĝmix ∂P̂ ∣∣∣∣ ϕ,T̂ ]−1 (2.13) S(P̂ , T̂ ) = −∂Ĝ ∂T̂ ∣∣∣∣ ϕ,P̂ = SA − xe ∂ĜBA ∂T̂ ∣∣∣∣ ϕ,P̂ − ∂Ĝmix ∂T̂ ∣∣∣∣ ϕ,P̂ (2.14) where ρ̂A = ρ̂A(P, T ) and SA = SA(P, T ) are the density and entropy (per molecule) of state A 12 and xe = x(P̂ , T̂ ) is the equilibrium fraction of species. Thus, from the appropriate specification of state A and the equilibrium reaction constant, K, one may obtain all other thermodynamic properties, such as the isothermal compressibility and heat capacity, as well as the global phase diagram that includes both the vapor-liquid and the liquid-liquid transitions. In the next section, this is demonstrated by applying the two-state approach to the fluid-fluid phase transition in high- pressure hydrogen. 2.2 Fluid-Fluid Phase Transition in High-Pressure Hydrogen Experiments and simulations have discovered that at extremely high pressures, highly- dense fluid (dimeric) hydrogen dissociates into atomistic fluid hydrogen [21, 34–45, 48, 59, 119– 123]. In this section1, the two-state thermodynamic approach (Sec. 2.1) and the generalized law of corresponding states (obtained from reducing the temperature, pressure, and entropy by their critical values), in combination with the available experimental data and with the results of computations [21, 39–46, 48, 124], are utilized to predict the equation of state of hydrogen near the fluid-fluid phase transition (FFPT). 2.2.1 Phase Behavior of Hydrogen The suggested global phase diagram of hydrogen, based only on the available experimental evidence for the fluid-fluid phase transition [34–38], the solid-liquid melting transition [14–17, 125–127], and the location of solid-metallic hydrogen [22, 29–33], which is supported by the most recent computational studies [21, 39–45, 47, 48, 124, 128–130], is shown in Fig. 2.2a. It 1This section was reproduced from Nathaniel R. Fried, Thomas J. Longo, and Mikhail A. Anisimov, J. Chem. Phys., 157, 101101 (2022); https://doi.org/10.1063/5.0107043, with the permission of AIP Publishing. 13 Figure 2.2: The global pressure-temperature phase diagram for hydrogen. (a) The full range from low to extreme pressures in logarithmic scale. The crosses indicate the experimental data for the solid-liquid melting transition presented in Diatschenko et al. [14] (blue), Datchi et al. [15] (cyan), Gregoryanz et al. [16] (pink), and Zha et al. [17] (purple). The solid black curves at low pressure (P ≤ 0.1 GPa) are the liquid-gas-solid phase transitions [18], while the solid black curve at high pressure (P > 0.1 GPa) is the Kechin equation [19] as reported in ref. [17]. The black dashed curve is the predicted continuation of the melting line based on experimental and computational evidence [17,20,21], while the dotted lines represent the highly-debated prediction [22–28] of the domain of solid metallic hydrogen [22, 23, 28–33]. The red line is the first-order fluid-fluid phase transition adopted in this section. (b) The phase diagram of hydrogen at extreme conditions, in the area of the box in (a). The open circles are experimental data presented in Zaghoo et al. [34–36] (dark brown), McWilliams et al. [37] (light brown), and Ohta et al. [38] (orange). Simulation results [21, 39–46] are spread within the grey area and shown in detail in Fig. 2.3. The fluid-fluid phase transition (solid red) and Widom line (dotted red) are represented by Eq. (2.18). The red star is the location of the fluid-fluid critical point (FFCP) as adopted in this section. illustrates the fact that a huge pressure gap separates the liquid-gas [18] and fluid-fluid phase transitions in hydrogen. The adopted locations of the FFCP and the solid-fluid-fluid triple point (SFF-TP) in this section are based on the available experimental data [34–38] and on discussions present in the literature [42–44] (Table 2.1). It should be noted that the exact location of the FFCP is uncertain, as the interpretation of both the Zaghoo et al. [35] and Ohta et al. [38] experimental data have been highly debated [43, 131–133]. Most authors suggest that the experimental data of Ohta et al. [38], on the anomalies of the heating efficiency, are obtained in the supercritical region [43]. 14 Table 2.1: The suggested locations of the FFCP and the SFF-TP in hydrogen. P [GPa] T [K] ρ [g/cm3] FFCP 105 1900 0.8 SFF-TP 250 600 - The results observed by Ohta et al. [38] have been interpreted [1] as being the anomalies of the heating efficiency along the “Widom line”, the line corresponding to the maximum of the fluctuations of the order parameter, which emanates from the critical point [9, 78, 134]. The significant discrepancy between the results of different computational models makes it impossible to utilize these results for a single equation of state. However, presenting the same results in reduced variables, as suggested by the law of corresponding states, allows the com- putational results to be used along with the experimental data for thermodynamic modeling. In Fig. 2.3, all of the available computational and experimental data on the fluid-fluid phase transi- tion are presented in real units of pressure and temperature (Fig. 2.3a) and in reduced variables (Fig. 2.3b), P̂ = P/Pc and T̂ = T/Tc, where Pc and Tc are the critical pressures and temper- atures obtained (or adopted) from different works. It has been found that the simulation data based on the Quantum Monte Carlo (QMC) approach could also be collapsed into the universal phase diagram by reducing the entropy by its critical value, Ŝ = S/Sc [1]. In classical thermo- dynamics, the reference value for the entropy is arbitrary. Commonly, the value, Sc, is adopted as Sc = dP̂ /dT̂ |T=Tc [135–138], which was found to be Sc = 0.8 for all QMC simulations. 2.2.2 Phase Separation Coupled with Dimerization Thermodynamically, the fluid-fluid transition in hydrogen can be modeled through the two- state approach (Sec. 2.1), in which state A represents the free atoms of hydrogen and state B 15 represents dimerized hydrogen atoms. The total Gibbs energy per hydrogen atom (reduced by RidTc, where Rid is the ideal-gas constant) is given by Eq. (2.2), where µ̂◦A and µ̂◦B are the Gibbs energies of hydrogen in the monatomic or diatomic states, respectively, and x is the fraction of hydrogen atoms in the diatomic state. Unlike the symmetric system considered in Sec. 2.1, the mixing of free hydrogen atoms and diatomic hydrogen is not symmetric, and thus, Ĝmix may be modeled as Ĝmix(x) = T̂ [(x 2 ) ln (x 2 ) + (1− x) ln (1− x) ] + ε(T̂ , P̂ ) x(1− x) (2.15) The dimensionless non-ideality parameter, ε = ε(T̂ , P̂ ), may be approximated up to first order in ∆T̂ and ∆P̂ , as ε(T, P ) = ε0 − εT∆T̂ + εP∆P̂ (2.16) where ∆T̂ = T̂ − 1 and ∆P̂ = P̂ − 1. The FFCP parameters are determined from the thermodynamic stability criteria (Eq. 2.5), such that the critical fraction and critical temperature of hydrogen atoms is given by xc = √ 2− 1 and Tc = 2(2− √ 2)2ε0 (2.17) It should be noted that the first study to apply the two-state thermodynamic approach to high- pressure hydrogen was presented by Cheng et al. [21]. While the predictions of Cheng et al. for the FFPT are not in agreement with the results of all other simulations and experimental studies [45], their study provides a reasonable idea for how the non-ideality parameter, ε, might depend on pressure and temperature. Based on the suggested trend, εT and εP are optimized to 16 Figure 2.3: Unifying the different simulation results with experimental data of hydrogen by the generalized law of corresponding states. (a) Experimental and simulation results for the fluid- fluid phase transition (FFPT). (b) Unified representation of the FFPT by reducing pressure, P̂ = P/Pc, temperature, T̂ = T/Tc, and the critical value of the entropy, Ŝ = S/Sc. In (a) and (b), the open circles are the experimental data of Zaghoo et al. [34–36] (dark brown), McWilliams et al. [37] (light brown), and Ohta et al. [38] (orange). The computational results are indicated by the triangles: blue tints correspond to the Density Functional Theory (DFT) simulations of Bonev et al. [47] (dark blue), Morales et al. [48] (blue), Hinz et al. [44] (sky blue), and Karasiev et al. [45] (light blue). Meanwhile, green tints correspond to the Quantum Monte Carlo (QMC) simulations of Morales et al. [48] (dark sea green), Lorenzen et al. [39] (green), Perlioni et al. [41] (dark green), Mozzola et al. [42] (lime green), and Tirelli et al. [46] (yellow green). The colored stars correspond to the reported (or adopted in this section) critical points for each data set. The solid black curve is the solid-fluid phase transition line as discussed in Fig. 2.2, and the red solid line is the FFPT predicted in this section. agree with the behavior of hydrogen from the available computational data [41–44, 46, 48], and consequently, these parameters were adopted as εT = 2.062 and εP = −0.175. The asymmetric Gibbs energy of mixing is illustrated in Fig. 2.4a along with the fluid-fluid coexistence, calculated via the common tangent method, and the limit of absolute stability (spinodal), calculated via the thermodynamic stability conditions. The condition for chemical-reaction equilibrium is given by ∂Ĝ/∂x|T,P = 0, resulting in the balance of the Gibbs energy of reaction, ĜBA, and the exchange chemical potential of mixing, 17 µ̂mix = ∂Ĝmix/∂x|T̂ ,P̂ , such that thermodynamic equilibrium follows from ĜBA = −µ̂mix (2.18) The Gibbs energy of reaction, ĜBA = ĜBA(T̂ , P̂ ), may be approximated up to second order in T̂ and P̂ , as in Eq. (2.9). To balance the Gibbs energy of reaction, Eq. (2.9), with the derivative of the Gibbs energy of mixing, ĜBA is expressed as an expansion in ∆T̂ and ∆P̂ as ĜBA = u− a∆T̂ + b∆P̂ + g∆T̂∆P̂ + d 2 (∆T̂ )2 − k 2 (∆P̂ )2 (2.19) where the modified coefficients of the thermodynamic balance, Eq. (2.19), are related to the coefficients of reaction, Eq. (2.9), as: ê = u+ a− b+ g + d 2 − k 2 ŝ = a+ g + d υ̂ = b− g + k (2.20) along with γ = g, δ = d, and κ = k. If the Gibbs energy of mixing, Ĝmix, would be symmetric with respect to x, then ĜBA = −µmix = 0, could describe the conditions for both reaction equilibrium and fluid-fluid phase equilibrium [9]. However, since the monatomic and diatomic mixing is asymmetric, the condition for the balance of phase and reaction equilibrium, Eq. (2.18), is given through µ̂mix T̂ = a2 ( ε(T, P ) T̂ − ε0 )2 + a1 ( ε(T, P ) T̂ − ε0 ) + a0 (2.21) 18 Figure 2.4: The components of the Gibbs energy (per atom) for hydrogen in the vicinity of the fluid-fluid phase transition. (a) The Gibbs energy of reaction, GBA, as given by Eq. (2.9). The isotherms are T = 0.5Tc (orange), T = 0.75Tc (blue), T = Tc (green), T = 1.25Tc (red), and T = 1.5Tc (purple). (b) The Gibbs energy of mixing, Gmix, as given by Eq. (2.15). Gmix is shown as a function of the fraction of hydrogen atoms in the diatomic state, x, for isotherms T = 0.5Tc (blue), T = 0.75Tc (green), and T = Tc (red) at P = Pc. The solid curve corresponds to the fluid-fluid coexistence. where the coefficients a0 = −0.502, a1 = 0.166, and a2 = −0.071. The developed equation of state is formulated through the Gibbs energy for the system as a function of temperature and pressure. Due to the interconverting nature, the two-states of hydrogen are thermodynamically equivalent to a single component system. Consequently, this produces an equation of state in terms of the equilibrium fraction of dimerized atoms, xe = xe(T, P ), and the density of the system, ρ = ρ(T, P ). The equation of state presented in this section contains seven adjustable parameters: five from the Gibbs energy of reaction, GBA, (u, a, b, g, and k), Eq. (2.9) and two from the non-ideality parameter in the Gibbs energy of mixing (εT and εP ), Eq. (2.16). The number of adjustable parameters are reduced from the following analysis of the available computational data on hydrogen in the vicinity of the fluid-fluid critical point. From the computational heat capacity data presented by Karasiev et al. [45], the heat- 19 Figure 2.5: Equilibrium fraction of hydrogen atoms in the diatomic state, xe. (a) Equilibrium fraction-pressure diagram for T = 0.5Tc (orange), T = 0.75Tc (blue), T = Tc (green), and T = 1.25Tc (red). (b) Equilibrium fraction-temperature diagram for P = 0.75Pc (blue), P = Pc (green), P = 1.25Pc (red), P = 1.5Pc (purple). The solid and dashed black curves are, respectively, the fluid-fluid coexistence and the limit of thermodynamic stability (spinodal). capacity change of reaction is approximated to be δ ≈ 0, and from the computational isothermal- compressibility data presented in the supplemental material of Pierleoni et al. [41], it is approx- imated that κ ≈ 0.625 [mm3/GPa ·mol]. Additionally, e = −108 [kJ/mol] is adopted based on the known value of the bond dissociation energy of H2 [139]. As discussed above, εT and εP are assigned as εT = 2.062 and εP = −0.175. From these findings, the number of free parameters have been reduced to three: a, b, and g. The values of the remaining free parameters are determined as a = −4.95, b = 0.044, and g = 0.0124 from the computational and experimental data utilizing the generalized law of corresponding states (Fig. 2.3). Using the relations between these parameters and the physical parameters in Eq. (2.9), the following parameters are estimated: the entropy change of the reac- tion as s = −34.0 [J/K ·mol], the volume change of the reaction as υ = 393 [mm3/mol], and the volume-expansivity change of the reaction as γ = 0.0677 [mm3/K ·mol]. The Gibbs energy change of reaction is shown in Fig. 2.4b. It demonstrates that the pressure is the major factor in 20 the behavior of GBA. Using the Gibbs energy of mixing, Eq. (2.15), the Gibbs energy of reaction, Eq (2.9), and the variables determined from the universal phase diagram, the equilibrium fraction of hydrogen atoms in the dimerized state, xe, is determined from Eq. (2.18). The corresponding equilibrium- fraction phase diagrams are presented in Fig. 2.5(a,b). At higher temperatures and lower pres- sures, the equilibrium composition changes from the dimeric state xe = 1 to the monomeric state xe = 0. The density of species is expressed through the equilibrium fraction via Eq. (2.13), and with use of Eq. (2.15) for Ĝmix is given in the form [9] ρ̂(P̂ , T̂ ) = ( ∂Ĝ ∂P̂ )−1 T̂ = [ 1 ρ̂A + xe ∂ĜBA ∂P̂ + ∂ε ∂P̂ xe (1− xe) ]−1 (2.22) where ρ̂A = ρ̂A(P̂ , T̂ ) is the volume of the monatomic hydrogen state, and may be expressed up to second-order in ∆T̂ and ∆P̂ as ρ̂A = ρ̂c − ρ̂0∆T̂ + ρ̂1∆P̂ + ρ̂2∆T̂∆P̂ − ρ̂3 ( ∆P̂ )2 + ρ̂4 ( ∆T̂ )2 (2.23) Using the most recent QMC simulations presented in Tirelli et al. [46], ρ̂A is estimated by Eq. (2.23) with coefficients: ρ̂c = 1.01, ρ̂0 = 0.25, ρ̂1 = 0.56, ρ̂2 = 0.56, ρ̂3 = 0.21, and ρ̂4 = 0.12. The corresponding pressure-density phase diagram is presented in Fig. 2.6, and demonstrates a good agreement with the computational data in the vicinity of the FFCP. It is noted that the properties observed in experimental studies (e.g. conductivity, reflec- tivity, thermal efficiency, etc.) could be indirectly related to the order parameter for the FFPT in 21 Figure 2.6: The pressure-density phase diagram of hydrogen based on the equation of state de- veloped in this section. The open circles correspond to the QMC simulations of Tirelli et al. [46]. Isotherms are T̂ = 0.67 (orange), T̂ = 0.73 (red), T̂ = 0.8 (brown), T̂ = 0.87 (purple), T̂ = 0.93 (green), and T̂ = 1.0 (blue). The fluid-fluid coexistence is shown by the solid black curve. The red star is the FFCP adopted in this section. hydrogen, ϕ = x/xc−1. The measurable quantities (such as density or conductivity) are coupled to the order parameter. There is a remarkable analogy between the challenges in thermodynamic modeling of fluid polyamorphism in hydrogen and that in supercooled water. In both cases, there is a reasonable agreement among the scientific community on the shape and location of the first-order transition line, while the position of the corresponding FFCP and LLCP are highly uncertain and subjects of current debate in the literature [13, 34–36, 43, 78, 131–133]. This uncertainty, in both hydro- gen and water, is due to the extreme conditions of the phenomena. In supercooled water, the liquid-liquid transition is hidden below the temperature of spontaneous ice formation [75, 78], 22 while in hydrogen, the fluid-fluid transition occurs at immensely high pressures (millions of atm) [43]. Consequently, it is not surprising that the available computational or experimental data are scarce [40, 43, 44]. In this section, it has been shown that despite the uncertainty in de- termining the location of the FFCP in hydrogen, thermodynamic modeling provides a principle direction to predict the equation of state for the system. Remarkably, the law of corresponding states can be utilized to reconcile the different computational models of hydrogen and experi- ment [21, 34–45, 48, 124] into a unified equation of state. An additional parameter has been in- cluded to generalize the law of corresponding states, the entropy at the critical point (Sc), which provides the opportunity for further studies of hydrogen, both experimental and computational, to be unified under the general approach presented in this section. 2.3 Liquid-Liquid Phase Transition in High-Density Sulfur While the fluid-fluid phase transition in hydrogen was induced by the dimerization reac- tion, a fluid-fluid phase transition may also be induced by more complex chemical reactions or interactions, such as polymerization or gelation. An example of polymerization induced fluid polyamorphism was recently discovered in high-density sulfur [51]. A polymerized state of sul- fur exists above 433 K [140–144]). It has been experimentally demonstrated