Fact-checked by Grok 2 weeks ago

Phase transition

A phase transition is a physical process in which a undergoes a qualitative change in its state, transitioning between distinct phases such as , , or gas, typically triggered by variations in , , or other external parameters, and marked by singularities or discontinuities in the or its derivatives. These transitions are ubiquitous in nature, manifesting in everyday phenomena like the of or of , as well as in complex systems such as ferromagnets developing below the . Phase transitions are classified by according to Ehrenfest's scheme, with transitions featuring discontinuities in the first derivatives of the (e.g., or ) and involving absorption or release, as seen in the liquid-gas transition where coexisting phases separate via a discontinuous jump. In contrast, second-order transitions exhibit continuous first derivatives but discontinuities in higher-order ones, lacking and occurring at critical points where phases become indistinguishable, exemplified by the superconducting transition in certain materials. Near second-order transitions, emerge, characterized by divergences in response functions like and correlation length, governed by universal scaling laws and that transcend microscopic details, revealing deep connections across diverse systems from fluids to quantum magnets. These behaviors, first systematically studied in the context of the , underpin modern understandings in and have profound implications for materials design, including high-temperature superconductors and phase-change memory devices.

Historical Development

Early Empirical Observations

Joseph Black's experiments in the 1760s provided the first systematic empirical evidence distinguishing heat absorbed during phase changes from that causing temperature rise in single phases. Observing that equal masses of ice and water, when heated, required significantly more thermal input to convert ice to water at 0°C than to elevate the temperature of already liquid water, Black quantified the latent heat of fusion for ice as approximately 144 times the heat capacity of water per degree Fahrenheit. This demonstrated that during melting, temperature remained constant at the transition point despite continued heat application, challenging prevailing caloric theories and highlighting the energy barrier inherent to solid-liquid phase shifts. Black extended these findings to , noting analogous during , where water at 100°C absorbed substantial without increase until fully converted to . His 1762 lecture at the formalized these observations, establishing as a tool for probing phase boundaries and revealing that phase transitions involve discrete energy quanta tied to molecular rearrangements rather than continuous . These results, derived from precise use post-1700s improvements, underscored the reproducibility of transition temperatures under constant , laying empirical groundwork for later thermodynamic models. Preceding , informal observations of phenomena—like the sharp freezing of bodies or irregular heating in —date to , but lacked quantification until reliable thermometry enabled controlled replication. 's work thus marked the onset of rigorous , confirming transitions as objective, measurable discontinuities in material properties driven by thresholds.

Emergence of Theoretical Frameworks

The foundational thermodynamic framework for phase transitions was established by through his phase rule, derived in the papers "On the Equilibrium of Heterogeneous Substances" published in 1876 and 1878. This rule expresses the F in a multiphase system as F = C - P + 2, where C denotes the number of independent chemical components and P the number of phases, with the +2 accounting for and as variables under conditions. Gibbs' formulation enabled quantitative predictions of phase coexistence and stability, shifting analysis from purely empirical observations to rigorous thermodynamic constraints, though it remained phenomenological without microscopic underpinnings. In 1933, advanced this framework by classifying phase transitions according to the order of discontinuities in thermodynamic derivatives. transitions exhibit jumps in first-order derivatives of potentials like or volume (manifesting as ), while second-order transitions show discontinuities in second-order derivatives such as specific heat or , with continuous first derivatives. This scheme highlighted the need to distinguish transition types based on thermodynamic singularities, influencing subsequent theoretical developments despite limitations in handling where higher derivatives diverge. Lev Landau's 1937 theory marked a pivotal phenomenological advance for second-order , introducing an order parameter \eta to quantify between disordered and ordered phases. Near the temperature T_c, Landau expanded the G(\eta, T) as G = G_0 + a(T - T_c)\eta^2 + b\eta^4 + \cdots, where the term drives the and higher even powers ensure stability; minimization yields \eta = 0 above T_c and \eta \propto \sqrt{T_c - T} below, predicting mean-field exponents like \beta = 1/2. This symmetry-based approach, rooted in group theory, explained diverse (e.g., ferromagnetic ordering) via universal forms, bridging to microscopic order while approximating fluctuations inadequately near criticality.

Key Milestones in 20th Century

In 1933, introduced a thermodynamic classification of phase transitions based on the order of discontinuities in derivatives of the , distinguishing transitions (discontinuous first derivatives like or ) from higher-order ones where lower derivatives remain continuous. This framework, rooted in empirical observations of singularities in equations of state, provided an initial systematic categorization but later proved insufficient for capturing microscopic behaviors in continuous transitions. Lev Landau advanced the field in 1937 with a general phenomenological theory for second-order phase transitions, employing the concept of an order parameter to describe symmetry breaking and expanding the Gibbs free energy in powers of this parameter near the critical point. Landau's approach explained the emergence of new phases through minimization of the free energy functional, incorporating fluctuations via coupling to external fields, and applied successfully to phenomena like superfluidity in helium-4. However, as a mean-field theory, it overestimated critical exponents by neglecting long-range correlations. A pivotal exact result came in 1944 when solved the two-dimensional for ferromagnetic order-disorder transitions on a square lattice with zero external field, deriving the partition function and demonstrating a logarithmic divergence in specific heat without , alongside finite below the critical temperature. This solution exposed limitations in mean-field approximations like Landau's, as the deviated from classical predictions (e.g., specific heat exponent α=0 with logarithmic singularity rather than mean-field jump), and underscored the role of dimensionality in transition behavior. The 1960s saw preparatory scaling hypotheses from researchers like and Benjamin Widom, positing universality in across systems with similar symmetries and dimensions, but microscopic justification awaited Kenneth Wilson's 1971 formulation of the transformation. Wilson's iteratively coarse-grains the system's , revealing fixed points that govern behavior and enabling computation of non-mean-field exponents via epsilon expansions near upper critical dimensions, thus resolving long-standing discrepancies in and earning him the 1982 . This development shifted focus from phenomenological models to scale-invariant microscopic theories, profoundly influencing understanding of continuous phase transitions.

Fundamental Concepts

Definition and Thermodynamic Basis

A phase transition is a change in the of a from one to another, where a represents a homogeneous and mechanically stable configuration of matter with uniform physical properties throughout. These transitions manifest as discontinuities or singularities in thermodynamic variables such as volume, entropy, or specific heat capacity, distinguishing them from smooth variations within a single . Empirically observed examples include the melting of ice at 0°C and 1 atm, where solid and liquid coexist, or the boiling of at 100°C under the same conditions. The thermodynamic foundation of phase transitions rests on the minimization of the system's appropriate potential, which dictates stability. For processes at constant and volume, the F = U - TS (with U as and S as ) is minimized; at constant and pressure, the G = F + PV (where P is pressure and V volume) governs stability./23:_Phase_Equilibria/23.02:_Gibbs_Energies_and_Phase_Diagrams) Stable phases correspond to global minima of these potentials, and a transition occurs when the free energies of competing phases become equal, enabling coexistence along a phase boundary in the . This equality implies that the chemical potentials \mu of the phases match, as G = \mu N for a single-component with N particles. Along coexistence curves, the Clapeyron equation \frac{dP}{dT} = \frac{\Delta H}{T \Delta V} relates the slope of the boundary to the change \Delta H and change \Delta V of the transition, derived from the condition dG = 0 for both phases at . Phase transitions introduce non-analyticities in the , reflecting the emergence of collective order or structural reorganization driven by and interparticle interactions, as opposed to analytic continuations within phases. This framework, rooted in classical , provides a causal explanation for why systems spontaneously shift phases: the drive toward minimization favors the configuration with the lowest potential under prevailing conditions.

Order Parameters and Symmetry

In continuous phase transitions, the order parameter serves as a measurable quantity that is zero in the symmetric, disordered phase and acquires a nonzero expectation value in the ordered phase, quantifying the degree of ordering and distinguishing the phases thermodynamically. This parameter must transform irreducibly under the system's symmetry group, ensuring that its expansion in the Landau free energy respects the underlying symmetries. The appearance of a nonzero order parameter below the critical temperature signals , where the or state selects a configuration that lacks the full symmetry of the or governing the system, even though all states collectively restore the symmetry. In , this is captured by expanding the density as f(\phi) = f_0 + r(T - T_c) \phi^2 + u \phi^4 + \cdots, where \phi is the order parameter, r > 0, and u > 0; above T_c, the minimum is at \phi = 0 (symmetric phase), while below T_c, minima occur at finite \phi = \pm \sqrt{-r(T - T_c)/(2u)}, selecting a broken-symmetry direction. Higher-order invariants, such as cubic terms (v \phi^3), can induce transitions if present, but SSB remains tied to the stabilization of ordered states with reduced symmetry. Specific examples illustrate the interplay: in ferromagnetic transitions, the magnetization \mathbf{M} acts as the order parameter, breaking continuous in spin space (SO(3)) as \mathbf{M} aligns spontaneously along a , with M \propto (T_c - T)^{1/2} near T_c in mean-field . For the superfluid transition in helium-4, the complex scalar order parameter \psi = |\psi| e^{i\theta} breaks U(1) phase symmetry, enabling off-diagonal long-range and superflow. In nematic liquid , the tensorial order Q_{ij} breaks isotropic (O(3)) down to uniaxial D_{\infty h}, reflecting molecular without preferred reversal. These cases highlight how the order 's representation under the dictates the possible broken phases and associated Goldstone modes, which emerge as massless excitations restoring continuous symmetries in the low-temperature phase. For first-order transitions, such as liquid-gas coexistence, an order parameter like the density difference \Delta \rho = \rho_\ell - \rho_g jumps discontinuously, but SSB is absent in the strict sense, as both phases share the same symmetry group, with the transition driven by free-energy minimization rather than continuous symmetry reduction. In contrast, SSB in continuous transitions underpins universality classes, where critical exponents depend on the dimensionality, range of interactions, and symmetry of the order parameter, as formalized in renormalization group theory beyond mean-field approximations.

States of Matter Involved

Phase transitions primarily involve transformations between the classical states of matter: solid, liquid, gas, and plasma. In the solid state, atoms or molecules are arranged in a fixed, ordered lattice with definite shape and volume, resisting deformation under moderate forces. The liquid state features particles in close proximity but with sufficient kinetic energy to flow and conform to container shapes while maintaining volume. Gases consist of widely spaced particles moving freely, expanding to fill containers and exhibiting neither fixed shape nor volume. Plasma, often regarded as the fourth classical state, comprises ionized particles—free electrons and positive ions—prevalent in high-temperature environments like stars or lightning, where thermal energy overcomes atomic binding. These states are distinguished by macroscopic properties such as density, compressibility, and response to external fields, with transitions driven by changes in temperature, pressure, or composition. Common transitions include melting (solid to liquid), occurring at the where vibrational energy disrupts lattice order, as in to at 0°C under standard pressure; freezing (liquid to solid), the reverse process; (liquid to gas), such as at 100°C; and (gas to liquid). transforms solids directly to gas, exemplified by (solid CO₂) at -78.5°C, while deposition reverses this, as in formation. converts gas to via high energy input, and recombination yields gas from . Within solids, phase transitions can shift between polymorphic forms, like to under extreme pressure, without altering the overall solid state. Beyond classical states, phase transitions access non-classical or exotic states under specialized conditions, such as Bose-Einstein condensates formed by cooling bosons to near (achieved experimentally in 1995 with rubidium-87 atoms at 170 nK), where quantum coherence dominates. Superfluids, like liquid helium-4 below 2.17 K, exhibit frictionless flow via transitions involving Cooper pairs. These transitions highlight how varying thermodynamic parameters reveals diverse macroscopic behaviors, though classical solid-liquid-gas-plasma interconversions remain foundational to most observed phenomena.

Classifications

Ehrenfest Classification

The Ehrenfest classification, introduced by physicist in 1933, categorizes phase transitions based on the continuity of derivatives of the G(T, P) with respect to T and P. The order of a transition is defined as the lowest n such that the nth-order derivative of G is discontinuous at the transition point, while lower-order derivatives remain continuous. This thermodynamic approach aimed to generalize the distinction between transitions like (discontinuous volume) and hypothetical continuous ones, without relying on microscopic details. In transitions, the first derivatives— S = -\left(\frac{\partial G}{\partial T}\right)_P and V = \left(\frac{\partial G}{\partial P}\right)_T—exhibit discontinuities, implying L = T \Delta S and a region of coexistence where both phases are . Examples include the solid-liquid transition in at 0°C and 1 , where jumps by about 9% upon , and the of liquids. These transitions involve and or effects due to the energy barrier between phases. Second-order transitions feature continuous first derivatives but discontinuous second derivatives, such as the specific heat C_P = -T \left(\frac{\partial^2 G}{\partial T^2}\right)_P or the coefficient \alpha = \frac{1}{V} \left(\frac{\partial V}{\partial T}\right)_P. Ehrenfest initially applied this to the superconducting transition in mercury below 4.15 at zero field, where resistivity drops discontinuously but appears continuous (though later measurements refined this). No occurs, and the transition is reversible without . Higher-order transitions (n > 2) are defined similarly, with discontinuities in even higher derivatives, but empirical examples are scarce and often reclassified under modern schemes due to subtler singularities near critical points. The Ehrenfest scheme provided a foundational phenomenological framework but overlooks divergences (rather than mere discontinuities) in derivatives at , as revealed by later ; for instance, the liquid-gas critical point at 31°C and 73.8 atm for CO₂ shows infinite , not fitting neatly into finite-order discontinuities. Despite these limitations, it remains a reference for distinguishing transitions by thermodynamic response functions.

First-Order and Continuous Transitions

First-order phase transitions feature a discontinuous change in the first derivatives of the , such as the G with respect to ( S = -∂G/∂T) or pressure (volume V = ∂G/∂P), leading to Q = T ΔS and coexistence of separated by a finite barrier. This discontinuity manifests as a jump in the order parameter, enabling and , where the system can persist in a higher-free-energy until overcomes the barrier./04%3A_Phase_Transitions/4.01%3A_First_order_phase_transitions) The Clapeyron dP/dT = ΔH / (T ΔV) governs the slope of the coexistence curve, with ΔH denoting the of transition. Prominent examples include the solid-liquid transition in at the (0.01°C, 611.657 ), where and coexist with a volume contraction ΔV ≈ -1.6 × 10^{-6} m³/mol and of 6.01 kJ/mol, and the liquid-vapor transition along the boiling curve up to the critical point (373.946°C, 22.064 MPa)./04%3A_Phase_Transitions/4.01%3A_First_order_phase_transitions) Solid-solid transformations, such as the α-to-γ phase change in iron at 912°C under , also qualify, involving atomic rearrangements with associated s around 0.9 kJ/mol. Continuous phase transitions, termed second-order in the Ehrenfest scheme, maintain continuity in first derivatives of G but exhibit discontinuities or divergences in second derivatives, such as specific heat C = -T ∂²G/∂T², without or phase coexistence./04%3A_Phase_Transitions/4.02%3A_Continuous_phase_transitions) The order parameter η evolves continuously from zero, often following mean-field power laws near the critical T_c, with susceptibilities diverging as |T - T_c|^{-γ} where γ ≈ 1 in classical theory. These transitions lack a barrier, proceeding via correlated fluctuations over diverging scales ξ ~ |T - T_c|^{-ν}, underpinning universality classes beyond Ehrenfest's thermodynamic criteria. Key instances encompass the ferromagnetic transition in iron at T_c = 1043 K, where M vanishes continuously above the Curie point amid diverging , and the superconducting-normal transition in mercury at 4.15 K under zero field, marked by zero-resistance onset without jump./04%3A_Phase_Transitions/4.02%3A_Continuous_phase_transitions) The liquid-gas critical point in at 31.0°C and 7.38 MPa exemplifies endpoint termination of lines, yielding isotropic fluid with vanishing distinctions in and divergence. Unlike cases, continuous transitions evade , driven instead by restoration through thermal agitation.

Quantum and Topological Classifications

Quantum phase transitions differ from thermal phase transitions by occurring at temperature, where the absence of means that changes in the are driven by quantum fluctuations tuned via a non-temperature control parameter, such as strength, , or . These transitions emerge as the system approaches a , where the energy landscape undergoes qualitative changes, often leading to enhanced quantum fluctuations that can influence finite-temperature properties over a fan-shaped quantum critical region in the . Quantum phase transitions are classified as continuous or based on whether the order parameter changes discontinuously or through divergent correlation lengths, with continuous ones exhibiting universality classes analogous to but distinct from classical critical points due to the role of in effective theories. Prominent examples include the superconductor-to-normal metal transition under suppression of pairs, observable in high-temperature superconductors like YBa₂Cu₃O₇₋δ at fields exceeding 100 T, and the metal-insulator transition in materials such as vanadium dioxide (VO₂) tuned by doping, where quantum fluctuations dictate the or mechanisms. In theoretical models, the Bose-Hubbard model at integer filling demonstrates a quantum phase transition from superfluid to at a critical hopping-to-interaction ratio U/t ≈ 5.8–16.7 depending on dimensionality, marking the onset of incompressible behavior without in the strict T=0 limit but with precursors at low temperatures. Experimental signatures include non-Fermi liquid behavior, such as linear resistivity versus temperature in heavy-fermion compounds like CeCu₆₋ₓAuₓ near x=0.1, attributed to proximity to antiferromagnetic quantum critical points. Topological phase transitions delineate phases of matter distinguished not by or local order parameters, but by global topological invariants that remain robust against smooth deformations, provided underlying like time-reversal or particle-hole are preserved. These transitions typically involve the closing and reopening of an energy gap at high-symmetry points in momentum space, driven by tuning parameters that alter band topology, and evade conventional Landau-Ginzburg descriptions due to the absence of a dual scalar order parameter pairing trivial and nontrivial sectors. Classification schemes for topological phases and their transitions follow the Altland-Zirnbauer () tenfold way, categorizing systems into 10 classes (A, AI, AII, AIII, BDI, C, CI, CII, D, DIII) based on combinations of time-reversal (TRS), particle-hole (PHS), and chiral (S) symmetries, with topological invariants computed via or groups that predict the number and nature of gapless boundary modes. In two dimensions, the integer quantum Hall effect exemplifies a topological transition where plateaus in Hall conductance σ_xy = n e²/h (n integer) separate Chern insulator phases, with transitions occurring via dissipationless edge state reconfiguration under varying or filling factor, as realized in GaAs heterostructures at cryogenic temperatures below 1 K. Three-dimensional topological insulators, such as Bi₂Se₃, undergo transitions to trivial insulators by breaking TRS with magnetic doping, closing the bulk gap while preserving helical protected by Z₂ invariants. Recent extensions include hybrid classifications for interacting systems, where fractional topological order in fractional quantum Hall states at filling ν=1/3 introduces excitations, with phase boundaries mapped via entanglement in cold-atom realizations. These classifications underscore causal distinctions from symmetry-broken phases, as topological protection arises from band geometry rather than energetic minimization alone.

Types of Phase Transitions

Structural and Crystallographic Transitions

Structural phase transitions refer to changes in the arrangement of atoms within a crystalline solid that alter the crystal symmetry or lattice structure, typically induced by variations in temperature, pressure, or composition, without changing the material's chemical identity. These transitions occur between distinct solid phases and are distinguished from liquid-solid or gas-solid changes by the preservation of long-range order, though the specific symmetry and topology of that order evolve. Empirical observations, such as shifts in X-ray diffraction patterns, confirm these alterations, reflecting causal mechanisms rooted in minimizing free energy through atomic rearrangements. Crystallographic transitions specifically involve modifications to the unit cell parameters, symmetry, or coordination environments, often manifesting as distortions like tilting of polyhedra or shear deformations. They can proceed via two primary mechanisms: displacive, where atoms undergo collective, diffusionless shifts with minimal bond breaking, leading to continuous or nearly continuous changes; or reconstructive, involving bond rupture, atomic , and nucleation-growth processes that disrupt and rebuild the topology. Displacive mechanisms predominate in transitions preserving structural similarity, such as martensitic transformations, while reconstructive ones require thermal activation to overcome energy barriers associated with , as evidenced by kinetic studies showing and release. Order-disorder subtypes, a variant of displacive transitions, arise from randomizing positional or orientational , like cation site occupancy in alloys. In metals, prominent examples include the allotropic transformation in tin from white (tetragonal) to gray () at 13.2°C, a reconstructive transition driven by density changes and volume expansion of 27%, which proceeds via and growth due to the incompatibility of lattices. Iron exhibits multiple structural shifts, such as body-centered cubic (α) to face-centered cubic (γ) at 912°C, involving reconstructive diffusion to accommodate packing efficiency under . Martensitic transitions in steels, by contrast, are displacive, featuring rapid, shear-dominated austenite-to-martensite conversion below 727°C, with variants oriented by habit planes to minimize , as quantified by invariant line strain analysis. Ceramics display analogous transitions, such as the displacive ferroelectric shift in (BaTiO3) from paraelectric cubic to tetragonal at 120°C, where off-center Ti displacements break inversion , enabling ; this is continuous near the point, with soft modes signaling instability. Zirconia (ZrO2) undergoes a reconstructive tetragonal-to-monoclinic transition upon cooling below 1170°C, generating 3-5% volume expansion that induces cracking unless stabilized, as in yttria-partially stabilized variants; the mechanism involves oxygen coordination changes from 7 to 8, confirmed by high-resolution electron . In silicates like , the α-β inversion at 573°C is displacive, rotating SiO4 tetrahedra to alter trigonal , with no required, highlighting how vibrations couple to macroscopic strain. These transitions underpin materials functionality, influencing mechanical toughness via toughening in ceramics or enabling shape-memory effects in alloys through reversible displacive paths. Pressure-induced variants, such as isosymmetric second-order shifts increasing coordination (e.g., in silicates at gigapascal ranges), demonstrate how alters bonding preferences without loss, as revealed by diamond-anvil cell experiments. Source credibility in this domain favors experimental from peer-reviewed journals over theoretical models alone, given occasional discrepancies between simulations and observed in reconstructive cases.

Magnetic and Superconducting Transitions

Magnetic phase transitions occur when the magnetic ordering of spins in a material changes with temperature or external fields, often exhibiting critical behavior near the transition point. A prominent example is the ferromagnetic-to-paramagnetic transition at the T_c, above which vanishes and the material behaves as a paramagnet. For pure iron, T_c = 1043 K; has T_c = 1388 K; and T_c = 627 K. These transitions are typically second-order, characterized by a continuous order parameter— the M —that follows a power-law decay M \propto (T_c - T)^\beta below T_c, with \beta \approx 0.325 in three dimensions from Ising simulations, deviating from mean-field \beta = 0.5. Antiferromagnetic transitions occur at the Néel temperature T_N, where staggered orders antiparallel spins; for instance, in MnO, T_N = 116 K. Ferrimagnetic materials like (Fe_3O_4) show transitions at T_c = 858 K, involving unequal antiparallel sublattices. These magnetic transitions involve in spin orientations, with susceptibility diverging as \chi \propto |T - T_c|^{-\gamma} near T_c, where \gamma \approx 1.24 experimentally for ferromagnets. Fluctuations become long-range correlated, leading to observable in neutron scattering, revealing spin waves below T_c that soften at the transition. In applied fields, first-order transitions can emerge, as in manganites where accompanies metal-insulator changes. Superconducting phase transitions mark the onset of zero electrical resistance and the —expulsion of magnetic fields—below a critical T_c. Discovered in mercury at T_c = 4.15 in 1911, conventional superconductors follow Bardeen-Cooper-Schrieffer (, where electrons form pairs via phonon-mediated attraction, opening an energy gap \Delta \propto T_c. The transition is second-order in BCS, with specific heat showing a discontinuity \Delta C / C_n \approx 1.43 at T_c, and exponential tail C_s - C_n \propto e^{-\Delta / kT} below. High-temperature superconductors, such as YBa_2Cu_3O_7 with T_c = 93 and HgBa_2Ca_2Cu_3O_8 reaching 134 under pressure, deviate from BCS pairing mechanisms, possibly involving magnetic fluctuations. In superconductors, the order parameter is a complex scalar \psi representing the density of Cooper pairs, with |\psi|^2 \propto (T_c - T) near T_c in Ginzburg-Landau phenomenology. Quantum phase transitions in superconductors occur at T=0 under doping or pressure, separating superconducting from insulating states, as observed in cuprates where T_c domes with carrier concentration. Critical fields H_{c1}, H_{c2} bound the phase, with type-I showing abrupt Meissner expulsion and type-II forming vortices. Recent studies confirm field-induced transitions within superconducting states, like in CeRh_2As_2 at T_c = 0.26 K with H_{c2} > 14 T.

Transitions in Mixtures and Fluids

In mixtures of two or more components, phase transitions exhibit greater complexity than in pure substances due to compositional variations across phases, governed by the , which states that the F = C - P + 2, where C is the number of components and P is the number of phases, assuming and as intensive variables. For a binary (C=2), a two-phase (P=2) is univariant (F=2), manifesting as tie lines in -composition phase diagrams at fixed pressure, where the overall composition determines phase fractions via the . Vapor-liquid transitions in binary fluid mixtures typically form lens-shaped coexistence regions in phase diagrams, bounded by saturated liquid and vapor curves that meet at a critical point, beyond which the phases become indistinguishable. These diagrams reveal phenomena such as azeotropic behavior, where mixtures or condense at constant composition, complicating processes; for instance, the ethanol-water system exhibits a minimum-boiling at 78.2°C and 95.6 wt% at . Critical curves in pressure-temperature-composition space for binary mixtures often extend from the critical points of pure components, with possible upper or lower critical endpoints marking the termination of three-phase lines. Liquid-liquid phase separations occur in partially miscible fluid mixtures when thermodynamic instability drives demixing into compositionally distinct phases, often visualized as curves enclosing a two-phase region that pinches off at a consolute (critical) point. In solutions or organic-aqueous mixtures, upper consolute points arise from entropy-driven mixing at low temperatures and enthalpy-favored separation at higher temperatures, while lower consolute points reflect the inverse; within the unstable region accelerates via infinitesimal fluctuations, contrasting metastable outside it. For multicomponent fluids, random-matrix approaches predict emergent critical behavior even with many interacting species, enabling tunable phase diagrams for applications like programmable emulsions. In supercritical fluid mixtures, phase transitions blur as crossing the critical locus yields a single homogeneous phase without latent heat, yet density fluctuations near the mixture critical point mimic pure-fluid criticality, with universal exponents describing compressibility divergence. These transitions underpin industrial processes such as enhanced oil recovery, where CO₂-hydrocarbon mixtures exploit miscibility pressure thresholds around 10-30 MPa depending on temperature and composition. Experimental phase diagrams for such systems, derived from equations of state like Peng-Robinson, confirm that deviations from ideal mixing amplify critical shifts, with non-ideal interactions quantified by second virial coefficients influencing coexistence curves.

Exotic and Recent Types

The Berezinskii–Kosterlitz–Thouless (BKT) transition exemplifies an exotic infinite-order phase transition in two-dimensional systems with continuous rotational symmetry, such as the classical XY model, where thermal fluctuations lead to the unbinding of vortex-antivortex pairs at a critical temperature T_{BKT}. Below T_{BKT}, the system exhibits quasi-long-range order with power-law decay of correlations, circumventing the Mermin-Wagner theorem's prohibition on true long-range order in 2D; above T_{BKT}, correlations decay exponentially due to free vortices. This transition, theoretically predicted between 1972 and 1974, manifests in diverse systems including thin superconducting films, 2D superfluids, and Josephson junction arrays, with experimental confirmation in ultrathin disordered NbN films showing sharpness consistent with BKT scaling. Unlike conventional transitions, it lacks divergent correlation length at the critical point but features essential singularities in specific heat and superfluid density, jumping discontinuously to zero at T_{BKT}. The in amorphous solids, such as polymers or metallic glasses, involves a kinetic slowdown where molecular rearrangements freeze upon cooling, shifting the material from a viscous liquid-like to a rigid, non-equilibrium glassy at the temperature T_g. This phenomenon, observable over a range of temperatures rather than sharply, does not qualify as a thermodynamic phase transition due to the absence of , discontinuities in or volume, or singularities in the ; instead, it reflects a dynamical crossover dependent on cooling rate, with T_g shifting by tens of degrees for rates varying from 1 /min to 10^5 /s. Theoretical debates persist, with some models interpreting it as an underlying topological transition in the network of structural excitations or defects, though underscores its non-equilibrium nature without broken or phase coexistence. In polymers, T_g correlates with flexibility and intermolecular forces, typically ranging from 140°C to 370°C depending on composition and processing. Time crystal phases, proposed in 2012, constitute a recent exotic class where systems spontaneously break continuous or discrete time-translation symmetry, manifesting persistent oscillations in time without net energy input, distinct from spatial crystals. In equilibrium contexts, continuous time crystals remain theoretically challenging due to thermodynamic constraints, but discrete time crystals—realized in periodically driven (Floquet) quantum many-body systems—have been experimentally observed since 2016 in trapped ions, diamonds, and spin chains, exhibiting subharmonic response and robustness against perturbations. Phase transitions to these states often occur via nonequilibrium mechanisms, such as crossing an exceptional point where Floquet modes coalesce, separating dissipative time crystal orders; a 2024 experiment demonstrated a transition from continuous to discrete time crystals in driven oscillators, marked by frequency locking at \omega / 2. Recent 2025 observations in spin maser systems reveal a first-order transition to a time crystal phase when feedback strength surpasses a threshold, stabilizing oscillations amid dissipation. These transitions highlight nonequilibrium universality, with applications in quantum sensing and simulation, though stability requires isolation from decoherence.

Characteristic Properties

Phase Coexistence and Latent Heat

In phase transitions, phase coexistence occurs at the transition temperature and pressure where two thermodynamically distinct phases, such as and , maintain with equal chemical potentials, enabling arbitrary proportions of each phase to exist without net driving force for change. This arises because the densities of the phases are identical, balancing the tendency for one phase to convert into the other. The coexistence region manifests as a flat plateau in temperature- or pressure- diagrams, reflecting the discontinuous jump in or at the transition. Latent heat accompanies this coexistence in first-order transitions, representing the change \Delta H absorbed or released per unit (or ) to convert between phases at constant , without altering the system's . Quantitatively, the L = T \Delta S, where \Delta S is the discontinuity between phases, derived from and the definition of as over . For endothermic processes like or , heat is absorbed to overcome intermolecular forces; exothermic processes like release it. In second-order transitions, by contrast, no exists, as and volume remain continuous, with changes occurring via higher-order derivatives of the . The Clapeyron equation governs the geometry of the coexistence curve in phase diagrams: \frac{dP}{dT} = \frac{L}{T \Delta V}, linking the L to the slope of the boundary, where \Delta V is the volume change across phases. This relation, applicable to transitions like solid-liquid or liquid-gas, predicts how pressure alters transition temperatures; for instance, increased pressure favors the denser phase, steepening the curve for \Delta V < 0. Experimentally, is measured via calorimetry, tracking heat input during isothermal phase conversion, with values scaling with molecular interactions—e.g., higher for hydrogen bonding in water than in noble gases. Deviations from ideality in real systems, such as supercooling or nucleation barriers, can delay observable coexistence but do not alter the underlying thermodynamic equality.

Critical Points and Exponents

In continuous phase transitions, the critical point denotes the thermodynamic conditions—typically a critical temperature T_c and pressure P_c—where the first-order coexistence boundary terminates, and the two phases become indistinguishable, with properties such as density or magnetization exhibiting no jump but rather singular divergences in derivatives like compressibility or susceptibility./Physical_Properties_of_Matter/States_of_Matter/Supercritical_Fluids/Critical_Point) This occurs because fluctuations grow to macroscopic scales, eliminating latent heat while response functions diverge as the system approaches T_c along paths where the reduced temperature t = |T - T_c|/T_c \to 0. For fluids, the liquid-vapor critical point exemplifies this, with T_c = 647.096 \, \mathrm{K} and P_c = 22.064 \, \mathrm{MPa} for water, beyond which supercritical states exist without phase boundaries./Physical_Properties_of_Matter/States_of_Matter/Supercritical_Fluids/Critical_Point) The singular behaviors near T_c are universally described by power-law dependencies governed by critical exponents, which capture divergences independent of microscopic details within the same universality class defined by dimensionality, symmetry, and range of interactions. These exponents arise from the scaling hypothesis, where the free energy's singular part f_s(t, h) \sim |t|^{2 - \alpha} \tilde{f}(h / |t|^{\beta \delta}), with h the conjugate field to the order parameter. Key exponents include \alpha for the specific heat C \sim |t|^{-\alpha}, where \alpha > 0 implies divergence and \alpha < 0 a cusp; \beta for the order parameter \psi \sim (-t)^\beta below T_c; \gamma for the susceptibility \chi \sim |t|^{-\gamma}; \delta from the critical isotherm \psi \sim h^{1/\delta} at T_c; \nu for the correlation length \xi \sim |t|^{-\nu}; and \eta from the spatial correlation function G(r) \sim 1/r^{d-2+\eta} at criticality, with d the dimension. Scaling relations interconnect these exponents, such as the Rushbrooke equality \alpha + 2\beta + \gamma = 2, which holds under the scaling hypothesis and hyperscaling, validated numerically for models like the 3D Ising universality class relevant to uniaxial magnets and binary fluids. In mean-field theory, valid above the upper critical dimension d=4, the exponents are \alpha=0 (discontinuity), \beta=1/2, \gamma=1, \delta=3, \nu=1/2, and \eta=0, but fluctuations reduce them below d=4; for the , high-precision simulations yield \beta \approx 0.3265, \gamma \approx 1.2371, \nu \approx 0.6299, and \alpha \approx 0.110, satisfying scaling to within 0.1%.
ExponentQuantityScaling Form3D Ising Value (approx.)Mean-Field Value
\alphaSpecific heat$C \simt^{-\alpha}$
\betaOrder parameter\psi \sim (-t)^\beta0.3260.5
\gammaSusceptibility$\chi \simt^{-\gamma}$
\deltaCritical isotherm\psi \sim h^{1/\delta}4.793
\nuCorrelation length$\xi \simt^{-\nu}$
\etaCorrelation functionG(r) \sim r^{-(d-2+\eta)}0.0360
These values, derived from Monte Carlo simulations and series expansions, confirm universality, as fluids like CO_2 (T_c = 304.2 \, \mathrm{K}) exhibit matching exponents despite differing Hamiltonians. Deviations occur in low dimensions or with long-range interactions, but the exponents robustly predict phenomena like diverging fluctuations over lengths \xi \sim 10^{-9} \, \mathrm{m} near T_c.

Universality and Scaling Laws

In continuous phase transitions, universality refers to the observation that the singular behavior near the critical point, characterized by critical exponents, depends only on the dimensionality of the system, the symmetry of the order parameter, and the range of interactions, rather than on microscopic details. Systems sharing these features belong to the same universality class and exhibit identical values of critical exponents, such as the specific heat exponent α, susceptibility exponent γ, and correlation length exponent ν. For instance, the three-dimensional encompasses uniaxial ferromagnets like the uniaxial antiferromagnet, binary fluid mixtures, and the liquid-gas transition in simple fluids, all displaying the same critical exponents despite differing Hamiltonians. This class is marked by Z₂ symmetry and short-range interactions, with measured exponents including α ≈ 0.110, β ≈ 0.326, and γ ≈ 1.237 from high-precision simulations and experiments on materials like nickel. Scaling laws emerge from the hypothesis that the singular part of the free energy density scales as f_s(t, h) = |t|^{2-α} Φ(h / |t|^{β+γ}), where t is the reduced temperature and h the ordering field, leading to power-law divergences in response functions. This ansatz implies relations among exponents, reducing the independent ones to two for Ising-like transitions; examples include the Rushbrooke scaling relation α + 2β + γ = 2, the Josephson hyperscaling relation 2 - α = dν (valid below the upper critical dimension), and Widom scaling βδ = β + γ. These laws have been verified empirically, for example, in the specific heat C ∝ |t|^{-α} for the 3D Ising class, where α > 0 indicates a cusp rather than divergence due to weak first-order effects in some realizations, but consistent scaling holds across experiments on superfluid transitions and ferromagnetic alloys. Universality and scaling underpin the classification of continuous transitions, with deviations signaling different classes, such as the XY model for superfluids (O(2) symmetry) or Heisenberg model for isotropic magnets (O(3) symmetry), each with distinct exponents like ν ≈ 0.671 for 3D XY versus ν ≈ 0.711 for 3D Ising. Finite-size scaling extends these laws to systems of limited extent L, predicting shifts in pseudocritical temperatures as t_L ~ L^{-1/ν} and rounded singularities, enabling extraction of exponents from simulations of lattice models like the Ising model on finite grids. Experimental confirmation includes neutron scattering data on ferromagnets showing correlation functions obeying scaling forms g(r) ~ r^{-(d-2+η)} f(r/ξ), with η ≈ 0.036 for 3D Ising.

Critical Phenomena and Fluctuations

Critical phenomena encompass the distinctive singular behaviors in thermodynamic and transport properties that emerge near the critical points of continuous (second-order) phase transitions, where distinct phases become indistinguishable and the correlation length diverges. These singularities manifest as power-law dependencies on the reduced t = |T - T_c|/T_c, quantified by universal critical exponents that classify systems into universality classes based on spatial dimensionality, symmetry of the order parameter, and interaction range. For instance, in the three-dimensional , representative of uniaxial ferromagnets and fluid liquid-vapor transitions, the exponents include \alpha \approx 0.110, \beta \approx 0.326, \gamma \approx 1.237, \delta \approx 4.79, \eta \approx 0.0364, and \nu \approx 0.6299. The specific heat C diverges as C \propto t^{-\alpha} above and below T_c, reflecting enhanced energy fluctuations; the order parameter m vanishes as m \propto (-t)^{\beta} for T < T_c; magnetic susceptibility \chi (or compressibility) diverges as \chi \propto t^{-\gamma}; and at T_c, the critical isotherm follows m \propto h^{1/\delta} under conjugate field h. Scaling relations interconnect these exponents, such as Rushbrooke's inequality \alpha + 2\beta + \gamma \geq 2 (equality holding in hyperscaling regimes) and Josephson's hyperscaling $2 - \alpha = d\nu, valid below the upper critical dimension d=4. Universality implies identical exponents for disparate systems sharing the same class, as confirmed experimentally in fluids and magnets, underscoring that long-wavelength fluctuations, not microscopic specifics, dictate criticality. Fluctuations near the critical point amplify dramatically due to the diverging correlation length \xi \propto t^{-\nu}, enabling cooperative effects over mesoscopic scales and invalidating mean-field approximations that neglect them. The variance of the order parameter scales with susceptibility via fluctuation-dissipation relations, \langle (\Delta m)^2 \rangle \propto \chi / V , but near criticality, \chi's divergence yields system-spanning fluctuations observable in scattering experiments. Spatial correlations decay via the Ornstein-Zernike form G(r) \sim e^{-r/\xi}/r^{(d-1)/2} for r \gg \xi, transitioning at criticality to a power law G(r) \sim 1/r^{d-2+\eta}, capturing anomalous short-distance behavior. These critical fluctuations underpin nonclassical exponents, drive renormalization group flows to fixed points governing universality, and explain phenomena like critical opalescence in fluids, where density fluctuations scatter light intensely. In dynamical contexts, slowing relaxation times \tau \propto \xi^z with dynamic exponent z \approx 2 further highlight fluctuation dominance.

Theoretical Approaches

Phenomenological Theories (Landau Theory)

Lev Landau formulated a phenomenological theory for second-order phase transitions in 1937, aiming to provide a general framework applicable to diverse systems exhibiting continuous symmetry breaking, such as ferromagnets, superconductors, and superfluids. The approach relies on thermodynamic principles rather than microscopic details, introducing an order parameter—a scalar or vector quantity, denoted typically as \eta, that remains zero in the high-temperature disordered phase and acquires a nonzero value in the low-temperature ordered phase, quantifying the degree of order. Central to the theory is the expansion of the Gibbs free energy G(T, \eta) near the critical temperature T_c as a power series in \eta, constrained by symmetry requirements of the system: G = G_0 + \frac{1}{2} r \eta^2 + \frac{1}{4} u \eta^4 + \cdots, where r = a(T - T_c) with a > 0, and u > 0 ensures . For systems under \eta \to -\eta (e.g., Ising-like ferromagnets), odd-powered terms vanish, preserving the expansion's form. is found by minimizing G with respect to \eta: above T_c, r > 0 yields \eta = 0; below T_c, r < 0 gives \eta^2 = -r / u = -a(T_c - T)/u, implying the order parameter vanishes continuously as T \to T_c^- with exponent \beta = 1/2. This mean-field approximation neglects fluctuations in \eta, treating it as uniform and predicting classical critical exponents, such as susceptibility \chi \propto |T - T_c|^{-1} (\gamma = 1) and specific heat discontinuity (\alpha = 0). For first-order transitions, the theory accommodates them via a negative quartic coefficient (u < 0), requiring a stabilizing sixth-order term, or an explicit cubic term \frac{1}{3} v \eta^3 when symmetry allows (e.g., in liquid crystals), leading to a discontinuous jump in \eta and latent heat. The theory's validity holds sufficiently far from T_c, where fluctuation effects are small, as quantified by the Ginzburg criterion, which delineates a regime where mean-field predictions break down due to long-range correlations. Extensions, such as the Ginzburg-Landau functional incorporating spatial gradients \nabla \eta, enable descriptions of interfaces and vortex structures but remain within the phenomenological paradigm. Despite limitations near criticality—where renormalization group methods reveal non-mean-field exponents—Landau theory provides qualitative insights into symmetry breaking and has influenced applications in materials design and quantum phase transitions.

Statistical Mechanics Models (Ising Model)

The Ising model, introduced by Wilhelm Lenz in 1920 and analyzed by Ernst Ising in his 1925 doctoral thesis, represents a foundational lattice-based approach to modeling cooperative phenomena such as ferromagnetism in statistical mechanics. In this model, a regular lattice of sites hosts classical variables s_i = \pm 1, interacting via nearest-neighbor couplings that favor alignment, with the system's behavior governed by temperature and an optional external magnetic field. The model Hamiltonian is \mathcal{H} = -J \sum_{\langle i,j \rangle} s_i s_j - h \sum_i s_i, where J > 0 denotes the ferromagnetic coupling strength, the sum over \langle i,j \rangle runs over adjacent pairs, and h is the field strength; this form captures the energy minimization through alignment without quantum effects. In one dimension, Ising exactly solved the model in 1925, demonstrating the absence of a finite-temperature phase transition: the magnetization vanishes for any T > 0, as thermal fluctuations disrupt long-range order, with the partition function yielding Z = [2 \cosh(\beta J)]^N for zero field and periodic boundaries, where \beta = 1/(k_B T). This result aligns with the Mermin-Wagner theorem's prohibition of continuous symmetry breaking in low dimensions due to infrared divergences in fluctuations. Extending to two dimensions on a square lattice, Onsager provided the exact solution in 1944 via methods, revealing a second-order phase transition at critical k_B T_c / J = 2 / \ln(1 + \sqrt{2}) \approx 2.269, below which emerges as M \sim (1 - T/T_c)^{1/8} for h = 0. The and functions follow from the largest eigenvalue of the , confirming logarithmic specific heat divergence (\alpha = 0) and power-law decay of correlations above T_c. Higher dimensions resist exact solutions, prompting approximations like , which replaces interactions with an effective field h_{\text{eff}} = h + z J m, where z is and m = \langle s_i \rangle; this yields a Curie-Weiss transition at T_c = z J / k_B with classical exponents (\beta = 1/2, \gamma = 1), overestimating T_c by about 50% in but capturing qualitative to ordered states. The model's , sharing exponents with short-range symmetric \phi^4 theories, links it to real ferromagnetic transitions in materials like iron, where vibrations and anisotropies introduce deviations but preserve core near criticality. simulations and series expansions refine 3D exponents (\beta \approx 0.326, \gamma \approx 1.237), validated against neutron scattering data. Extensions incorporate quenched or long-range interactions, altering exponents; for instance, infinite-range limits recover mean-field exactly. The Ising model's solvability and to or dimer coverings underpin its role in elucidating , from fluid-vapor coexistence to dynamics, emphasizing emergent order from local rules without invoking unverified global mechanisms.

Renormalization Group and Modern Methods

The renormalization group (RG) framework, developed by Kenneth G. Wilson in 1971, revolutionized the understanding of phase transitions by elucidating the emergence of scale invariance and universality at critical points through iterative coarse-graining of microscopic degrees of freedom. In this approach, short-wavelength fluctuations are integrated out, effectively rescaling the system to larger length scales while tracking the evolution of effective Hamiltonians or coupling constants via a flow equation; fixed points of this flow dictate the critical behavior, classifying operators as relevant, irrelevant, or marginal based on their scaling dimensions, which determine whether perturbations grow or decay under rescaling. This resolves the failure of mean-field theories below the upper critical dimension d_c = 4 for models like the Ising ferromagnet, where fluctuations become dominant, leading to non-classical exponents. Wilson's momentum-shell RG technique, applied to continuum field theories, enabled perturbative calculations via the \epsilon-expansion around d = 4 - \epsilon, yielding series for critical exponents such as \eta \approx \epsilon^2 / 54 + O(\epsilon^3) for the Ising model, which match experimental data when resummed or combined with higher-order terms. Universality arises because systems with identical symmetries, range of interactions, and dimensionality map to the same infrared fixed point under RG flow, explaining why diverse materials exhibit shared critical exponents despite differing microscopics. Real-space implementations, such as block-spin transformations on lattice models, provide non-perturbative insights, particularly for low dimensions, by recursively averaging spins over blocks and deriving recursion relations for couplings. Modern extensions include the functional RG formalism, introduced by Christof Wetterich in 1993, which employs a continuous flow equation for the effective average action \Gamma_k as the cutoff k is lowered from to scales, allowing treatment of strongly correlated systems beyond weak-coupling assumptions. This method has been applied to compute diagrams and exponents in frustrated magnets and fermionic systems, often incorporating vertex expansions or truncated schemes for tractability. Numerical RG variants, like the density-matrix RG (DMRG) algorithm devised by Steven White in 1992, efficiently handle quantum transitions in one by optimizing low-entanglement product states, revealing entanglement and critical points in models such as the quantum Ising with exponents \nu \approx 1 matching exact solutions. While DMRG draws conceptual parallels to RG through successive truncation of , it prioritizes entanglement minimization over strict coarse-graining, enabling ground-state computations for systems up to thousands of sites with accuracies rivaling exact . Further advancements encompass RG methods, generalizing DMRG to higher dimensions via (PEPS), which approximate ground states near criticality and extract dimensions from transfer spectra, though challenged by entanglement growth in d > 1. These techniques, combined with simulations incorporating RG-inspired finite-size , have refined exponent estimates, such as \gamma / \nu \approx 2.02 for the 3D Ising universality class, aligning with high-precision series expansions. Overall, and its evolutions underscore causal hierarchies in transitions, where ultraviolet details decouple from infrared physics at fixed points, providing a foundational tool for predictive modeling across and .

Experimental Investigation

Classical Experimental Techniques

Calorimetry has been a foundational for detecting phase transitions through measurements of in transitions and specific anomalies in continuous transitions. In transitions, such as or , the absorption or release of manifests as a plateau in during heating or cooling at constant , quantified by integrating input over time using adiabatic or isothermal calorimeters. For second-order transitions, like the point in ferromagnets, specific exhibits a logarithmic or power-law singularity near the critical , as observed in early experiments on where the peak sharpened with sample purity. Modern implementations, such as (), scan samples at controlled rates to resolve transition enthalpies with precisions below 1 J/g, though classical setups relied on ice-bath comparisons for accuracy. Dilatometry complements by tracking volume or linear expansion changes, which reveal discontinuities in transitions due to density jumps and anomalies in coefficients α near critical points. For instance, in structural transitions like the martensitic transformation in steels, dilatometers detect abrupt changes of up to 5% at transition temperatures, often coupled with thermal . Push-rod or optical dilatometry measures relative expansions δL/L to 10^{-6}, enabling mapping of boundaries in alloys. In , α diverges as |T - T_c|^{-α} with exponent α ≈ 0.11 for 3D Ising universality, as verified in fluids via capillary tube observations of blurring. Magnetometry probes magnetic phase transitions by quantifying M or χ as functions of T and applied H. In ferromagnets, the vanishes continuously at T_c in second-order transitions, measured via superconducting quantum interference device () magnetometers or classical vibrating sample magnetometers (VSM) with sensitivities to 10^{-6} emu. peaks diverge as |T - T_c|^{-γ} with γ ≈ 1.24, distinguishing second-order from transitions where in M-H loops indicates . These techniques confirmed Ehrenfest discontinuities in early studies of order-disorder transitions in alloys like Cu-Zn. Pressure-volume-temperature (PVT) measurements delineate liquid-gas critical points through isotherms showing vanishing and meniscus disappearance, as first demonstrated by Cagniard de la Tour in 1822 using sealed glass tubes heated under pressure. Classical mercury piston-cylinder apparatus mapped van der Waals loops, resolving the critical isotherm's inflection at (∂P/∂V)_T = 0 and (∂²P/∂V²)_T = 0. For solids, high-pressure cells combined with dilatometry detect pressure-induced transitions, such as in phases where volume reductions signal denser polymorphs.

Advanced Probes and Spectroscopy

Neutron scattering techniques provide detailed insights into the dynamical properties of materials near phase transitions, particularly for magnetic and structural changes. Inelastic neutron scattering has been instrumental in identifying soft modes and critical fluctuations, as demonstrated in studies of SnSe where it confirmed the role of softening in the transition mechanism. Elastic neutron scattering maps out phase diagrams by resolving reflections associated with ordering transitions, such as in CsPbCl3, revealing successive transformations at 47°C, 42°C, and 37°C driven by rotational instabilities. These methods leverage neutrons' sensitivity to magnetic moments and isotopic contrasts, enabling bulk-sensitive probes unavailable to light-based techniques, though sample requirements limit applicability to larger crystals or powders. Synchrotron-based probes offer high spatial and for of structural rearrangements. Time-resolved and have captured ultrafast in light-induced transitions, such as the insulator-to-metal switch in VO2 on picosecond timescales, highlighting metastable intermediate states. In pressure-induced scenarios, analysis detects coordination changes during transitions in GaAs and , correlating spectral edge shifts with density increases up to 20-30%. These techniques, often combined with anvil cells, resolve atomic displacements with precision but require intense sources to overcome signal-to-noise challenges near critical points. Vibrational spectroscopies, including Raman and infrared, track symmetry-breaking via mode softening or splitting at transition temperatures. Raman spectroscopy has quantified phonon frequency shifts in Weyl semimetal WTe2 during lithium intercalation-induced transitions, linking spectral changes to electronic band restructuring. Fourier-transform infrared spectroscopy monitors chain ordering in lipid bilayers, detecting gel-to-liquid crystalline transitions through methylene stretching band positions around 2850 cm⁻¹. Nuclear magnetic resonance, particularly magic-angle spinning variants, complements these by probing local environments in porous materials, as in zeolite phase transformations where chemical shift variations indicate framework reconstructions under thermal stress. Ultrafast enable nonequilibrium studies of , revealing pathways inaccessible to equilibrium methods. Pump-probe schemes with pulses have simulated topological transitions by exciting , producing absorption signatures of band inversion on scales. In supercritical fluids, time-domain THz spectroscopy observes formation and critical slowing down near the consolute point, with relaxation times diverging as |T - T_c|^{-zν} where zν ≈ 2-3. These approaches, powered by free-electron lasers, quantify energy dissipation and order parameter evolution but demand sophisticated modeling to disentangle coherent and incoherent contributions.

Recent Observations (2023-2025)

In April 2025, researchers experimentally observed the magnonic Dicke superradiant phase transition in a hybrid - system, where in spheres coupled ultrastrongly to , resulting in a macroscopic of the cavity field beyond the critical coupling strength predicted by the Dicke model. This observation confirmed the emergence of a superradiant phase in a non-equilibrium setting, with the transition marked by a discontinuity in the photon number and magnon polarization, aligning with expectations for finite systems. In August 2025, evidence of driven-dissipative phase transitions was reported in a one-dimensional of 21 superconducting nonlinear resonators, where multimode occurred as pump power exceeded a critical , leading to abrupt jumps in resonator amplitudes and phases. The experiment utilized to probe and , providing direct visualization of the phase boundary via and demonstrating scalability to larger arrays for studying collective quantum effects. October 2025 marked the experimental realization of a time rondeau crystal in a non-equilibrium quantum system, characterized by periodic breaking of time-translation symmetry with a complex temporal structure analogous to a rondeau pattern, achieved through Floquet driving in ensembles of spins or oscillators. This extended prior time crystal observations by incorporating higher-order temporal correlations, with the phase transition tuned via drive amplitude and observed through time-resolved spectroscopy revealing stabilized subharmonic responses. In November 2024, parametric control of quantum transitions was demonstrated in ultracold KRb + KRb reactions near Feshbach resonances, where tuning induced crossings of universal critical points, manifesting as non-monotonic variations in reactive loss rates consistent with Efimov physics scaling. The observations, conducted in optical dipole traps at nanokelvin temperatures, highlighted how detuning from resonance probes the transition's , with reaction rates deviating from Wigner threshold laws by factors up to 10 near criticality.

Applications and Implications

Materials Science and Engineering

Phase transitions underpin the design and processing of engineering materials by enabling precise control over microstructure, which directly influences mechanical, thermal, and electrical properties. In ferrous alloys, the allotropic transformations in pure iron—such as the diffusion-mediated shift from body-centered cubic (BCC) α-ferrite to face-centered cubic (FCC) γ-austenite at 912°C, followed by reversion to BCC δ-ferrite at 1394°C—form the basis for understanding more complex alloy behaviors. These transitions dictate the solubility of interstitial elements like carbon, enabling the formation of austenite as a precursor for subsequent hardening phases. In production and , time-temperature-transformation (TTT) diagrams map the kinetics of diffusional changes from , such as the eutectoid decomposition into at approximately 727°C and 0.76 wt% carbon, or bainite formation at intermediate temperatures. Non-diffusional martensitic transformations, triggered by rapid below the martensite start temperature (typically 200–400°C depending on carbon content), produce a supersaturated, body-centered tetragonal with exceeding 60 HRC due to distortion from trapped carbon atoms. This shear-dominated process, occurring at speeds up to 1000 m/s, minimizes atomic diffusion and preserves non-equilibrium compositions, critical for applications requiring high strength-to-weight ratios in automotive and components. Phase diagrams serve as predictive tools in alloy engineering, delineating phase fields to optimize compositions for targeted properties; for example, in nickel-titanium alloys, the austenite to monoclinic B19' transition at around 50–55°C enables shape memory effects exploited in stents and actuators. in aluminum-copper alloys involves and growth of θ'' precipitates from a supersaturated during aging at 120–200°C, increasing yield strength from 100 to over 400 via coherent strain fields that impede dislocation motion. In processes, solidification phase transitions control formation and microsegregation, where undercooling by 100–200 K can extend solid solubility limits, yielding metastable phases with improved resistance or magnetic properties. Advanced applications leverage controlled phase transitions for functional materials, such as in superalloys where γ' precipitate coherence with the FCC matrix enhances creep resistance at temperatures above 1000°C, vital for blades. in alloys, occurring via continuous composition modulation without barriers below the spinodal line, produces nanoscale lamellae that strengthen materials like Cu-Ni-Fe without coarsening. These phenomena, informed by thermodynamic modeling and methods, allow engineers to simulate multicomponent equilibria, reducing empirical trial-and-error in developing high-performance alloys for extreme environments.

Cosmology and Early Universe

In the hot Big Bang model, the early underwent a series of phase transitions as it expanded and cooled, analogous to thermodynamic phase changes in condensed matter systems but driven by the restoration and spontaneous breaking of gauge in quantum field theories. These transitions occurred when the thermal energy scale matched the vacuum expectation values of Higgs-like fields, leading to changes in the effective and the equation of state. Key examples include the (QCD) transition, separating the quark-gluon plasma phase from the confined hadronic phase, and the electroweak transition, where the SU(2)_L × U(1)_Y breaks to U(1)_EM, unifying weak and electromagnetic forces at high temperatures. Earlier (GUT) transitions, posited around 10^{15}-10^{16} GeV at times ~10^{-36} seconds post-, would have broken a larger unified but are mitigated by cosmic , which dilutes potential relics like magnetic monopoles. The QCD phase transition, occurring at temperatures of approximately 150-170 MeV and cosmic times around 10-20 microseconds after the Big Bang, marks the confinement of quarks into hadrons and the emergence of light pion degrees of freedom. In the Standard Model, lattice QCD simulations indicate this is a smooth crossover rather than a sharp first- or second-order transition, with no latent heat release but a change in the speed of sound from relativistic to non-relativistic values, impacting big bang nucleosynthesis (BBN) precursor dynamics like neutrino decoupling at ~1 MeV. Extensions beyond the Standard Model, such as those with axions or strong dynamics, could render it first-order, potentially generating gravitational waves (GWs) via bubble nucleation or sound waves in the plasma, though the peak frequency would be in the kHz range, beyond current pulsar timing array sensitivities. This transition influences the entropy release and baryon-to-photon ratio, constraining cosmological parameters through cosmic microwave background (CMB) anisotropies and light element abundances. The electroweak phase transition (EWPT), at temperatures near 100-160 GeV and times ~10^{-12} seconds, is crucial for electroweak baryogenesis, requiring a strongly first-order transition to satisfy Sakharov's conditions via out-of-equilibrium bubble expansion and CP violation. In the minimal Standard Model, perturbative calculations and lattice studies show it as a crossover, insufficient for generating the observed baryon asymmetry, necessitating extensions like supersymmetry or singlet scalars that enhance the Higgs potential barrier. First-order scenarios predict stochastic GW backgrounds from colliding bubbles, turbulent fluid motion, and magnetic field generation, with peak strains detectable by future space-based interferometers like LISA in the millihertz band, providing indirect probes of beyond-Standard-Model physics. Topological defects, such as Z-strings or domain walls, could form if the transition involves discrete symmetries, though stable networks risk overclosing the universe unless biased annihilation mechanisms operate. These phase transitions collectively shape the universe's thermal history, relic densities, and large-scale structure seeds. First-order transitions generally produce non-thermal relics and GW signals, while crossovers align with smoother evolution consistent with CMB precision data from Planck, which favor an adiabatic equation-of-state transition without excessive defects. preceding GUT-scale events resolves the overproduction problem by rapid expansion, reducing defect density below observational limits from searches and CMB distortions. Ongoing lattice and effective field theory computations refine transition orders, with implications for production via misaligned scalars or freeze-in during reheating.

Analogues in Complex Systems

In complex systems beyond equilibrium , analogues of phase transitions arise as abrupt qualitative shifts in macroscopic properties triggered by gradual changes in parameters, often driven by nonlinear interactions, loops, and heterogeneity rather than . These phenomena exhibit hallmarks like critical thresholds, behaviors, and , though they frequently occur in non-equilibrium settings without conjugate fields or , leading to deviations from classical universality classes. For instance, transitions in random networks model the sudden of global connectivity, where the fraction of occupied bonds serves as an order parameter that vanishes continuously below a critical probability p_c \approx 1/\langle k \rangle in mean-field approximations for networks with average degree \langle k \rangle. In heterogeneous scale-free networks, this transition hybridizes into a discontinuous followed by a critical-like region, reflecting robustness to random failures but vulnerability to targeted attacks on hubs. Social and economic systems display similar analogues through and formation, where agent-based models reveal phase transitions from fragmented states to synchronized behaviors as strength or information flow increases. In socio-economic contexts, emerges when agents prioritize mimicking over independent signals, yielding a second-order transition characterized by power-law distributions of cascade sizes near criticality, akin to avalanche exponents in . Empirical analyses of financial markets, for example, identify crash precursors via variance spikes and long-range correlations, signaling proximity to a where evaporates, though debates persist on whether these are true discontinuities or amplified fluctuations due to external shocks. Biological and ecological systems further exemplify these analogues, with tipping points marking shifts between stable states via and positive feedbacks. In , gradual stressors like nutrient loading can precipitate first-order-like transitions, such as shallow lake , where clear-water equilibria flip to turbid states irreversibly without hysteresis recovery, quantified by fold bifurcations in minimal models. in groups or neuronal networks shows continuous transitions to coherent motion or synchronized firing as or crosses thresholds, with order parameters like alignment velocity following mean-field exponents \beta \approx 1/2, though finite-size effects and broaden the critical region in real . These non-physical analogues underscore causal roles of local rules in generating emergent discontinuities, but their prediction remains challenged by sparse and model , contrasting the tunable of laboratory physical transitions.

Controversies and Open Questions

Debates on Quantum Phase Transitions

Quantum phase transitions (QPTs) at temperature are theoretically described as continuous changes in the driven by quantum fluctuations, yet significant debates surround their realization in specific systems, particularly regarding adherence to the Landau-Ginzburg-Wilson paradigm of order parameter fluctuations. One central controversy concerns deconfined quantum critical points (DQCPs), proposed in to explain transitions between magnetically ordered phases and valence bond solids in two-dimensional quantum antiferromagnets, where fractionalized excitations emerge without confinement, challenging the conventional separation of phases by a single order parameter. Critics argue that such points may not exist as stable continuous transitions, potentially collapsing to weakly due to dangerous irrelevant operators or events that destabilize the fixed point, as evidenced by numerical studies showing first-order signatures in related models. Recent experimental and theoretical work has intensified this debate, with some analyses indicating that DQCPs fail general standards expected for continuous transitions, as observed in pressurized SrCu₂(BO₃)₂ where correlations do not align with deconfined criticality predictions. Conversely, other studies report anomalous logarithmic persisting near putative DQCPs in SU(N) models, suggesting robustness against certain perturbations and potential hidden orders beyond standard descriptions. These conflicting findings highlight unresolved questions on the stability of DQCPs under anisotropies, long-range interactions, or doping, with proposals for dualities enhancing (e.g., SO(5)) offering pathways to reconciliation but requiring further verification. A parallel debate focuses on QPTs in itinerant systems, where the Hertz-Millis-Moriya () theory predicts Gaussian critical behavior damped by Fermi liquid quasiparticles, leading to mean-field exponents above the upper . However, experimental observations in heavy-fermion compounds and cuprates reveal non-Fermi liquid transport and singular susceptibilities inconsistent with predictions, attributed to the theory's neglect of non-analytic bosonic self-energies and vertex corrections that generate long-range interactions. For antiferromagnetic quantum critical points in metals, scaling analyses often show deviations from exponents, such as enhanced dynamical , necessitating theories incorporating critical fluctuations or SYK-like models for strange metal phases. Sachdev's classifications of metallic QPTs into categories involving reconstruction underscore these issues, emphasizing that conventional fails for low-dimensional or strongly correlated cases, prompting ongoing searches for universal non-quasiparticle descriptions. These debates underscore broader open questions, including the role of in smearing QPTs via Griffiths phases and the extent to which finite-temperature crossovers mimic true T=0 criticality, with experimental probes like neutron scattering and quantum oscillations providing indirect evidence but lacking consensus on interpretation. Resolution may require advanced numerics, quantum simulations, or new materials exhibiting tunable parameters near proposed critical points.

Dimensionality and Relativistic Paradoxes

In statistical mechanics, the occurrence and nature of phase transitions depend critically on spatial dimensionality, with lower dimensions suppressing long-range order due to enhanced fluctuations. In one dimension, continuous phase transitions at finite temperature are generally absent for systems with short-range interactions, as thermal fluctuations destroy any incipient order; this is exemplified by the exactly solvable 1D Ising model, which exhibits no spontaneous magnetization at any temperature above absolute zero. In two dimensions, discrete symmetries like the Ising model permit a finite-temperature transition, but continuous symmetries face prohibition from the Mermin-Wagner theorem, which demonstrates that Goldstone modes lead to infrared divergences, preventing spontaneous symmetry breaking at any finite temperature. Controversies persist in quasi-two-dimensional systems, such as layered materials, where finite-size effects or anisotropic interactions can induce effective transitions via dimensional crossover, challenging strict application of these theorems. Relativistic effects introduce additional paradoxes when considering phase transitions in boosted or high-energy regimes. A notable apparent contradiction arises from : a system at rest in one , say a fluid below its freezing , appears denser in a boosted due to Lorentz along the motion , suggesting a phase change (e.g., to ) that contradicts the rest- . This observer-dependent challenges the -invariance of thermodynamic phases, but resolution lies in the covariant transformation of the full ; thermodynamic variables like , temperature, and transform such that Lorentz-invariant scalars (e.g., proper or ) determine the phase consistently across . In practice, relativistic hydrodynamics with phase transitions, as simulated in heavy-ion collisions, confirms that transitions proceed via bubble nucleation without violating , provided dissipative effects and expansion are accounted for. Intersections of dimensionality and amplify these issues in quantum field theories. In low-dimensional relativistic models, such as (2+1)-dimensional systems, critical points exhibit non-trivial fixed points, but analogously to Mermin-Wagner can suppress Bose-Einstein condensation or , consistent with no long-range order at finite temperature. Debates center on whether relativistic relations (linear vs. quadratic) alter the lower ; while non-relativistic cases have clear cutoffs, relativistic theories in effective low dimensions (e.g., via dimensional reduction) may evade strict prohibitions through topological defects or long-range correlations, though empirical verification remains elusive in condensed-matter analogues like . These unresolved tensions highlight the need for frame-covariant analyses to reconcile fluctuations across dimensions and velocities.

Definitional and Interpretive Disputes

The Ehrenfest classification of phase transitions, proposed in 1933, defined them by the lowest-order derivative of the Gibbs free energy exhibiting a discontinuity: first-order for jumps in first derivatives like entropy or volume (associated with latent heat), and higher-order for discontinuities in subsequent derivatives, such as specific heat in second-order transitions. This thermodynamic approach assumed continuity in lower derivatives and aligned with observable jumps in macroscopic properties. However, the Ehrenfest scheme proved inadequate for many real systems, particularly continuous transitions where thermodynamic derivatives remain continuous but diverge at the critical point, as exemplified by the vanishing of correlation length in the two-dimensional below the critical temperature. Such divergences reflect singularities in the partition function rather than simple jumps, prompting a shift to a statistical mechanics-based definition: phase transitions occur where the becomes non-analytic in the (infinite volume or particle number), often tied to the emergence or loss of long-range order or phase coexistence satisfying Gibbs conditions. This modern framework, developed through works by Landau, Peierls, and in –1950s, emphasizes and universality classes over derivative orders, rendering the Ehrenfest labels descriptive but not fundamental. Definitional disputes persist in borderline cases lacking clear non-analyticity or equilibrium phase separation. The glass transition, observed in supercooled liquids around temperatures like 200–300 K for silica-based glasses, involves a kinetic arrest into a non-ergodic amorphous state without latent heat, hysteresis, or two coexisting equilibrium phases, leading most theorists to classify it as a dynamical crossover rather than a thermodynamic phase transition. Proponents of viewing it as a phase transition invoke topological changes in energy landscapes or ergodicity breaking, but empirical evidence from calorimetry and spectroscopy shows no singularity in equilibrium response functions, only apparent specific heat steps from frozen configurational entropy. In quantum phase transitions, tuned by non-thermal parameters like or magnetic fields at , definitions extend the classical non-analyticity criterion to ground-state properties, with quantum fluctuations replacing thermal ones to drive order-disorder changes, as in the superfluid-Mott insulator transition in optical lattices observed around 2000s experiments with Bose-Einstein condensates. Interpretive challenges arise in finite-size systems, where true singularities are absent due to the lack of a , complicating experimental identification; some argue for operational definitions based on scaling of gaps or entanglement divergences instead. These disputes highlight tensions between rigorous infinite-system ideals and practical finite-sample observations, with no on excluding quantum cases lacking macroscopic .

References

  1. [1]
    [PDF] Lecture 9: Phase Transitions
    Connecting phase to smoothness properties allows to shift focus from phases themselves to the transformations between phases called phase transitions. Phase ...
  2. [2]
    6. Phase Transitions — Introduction to Statistical Mechanics
    Phase Transitions¶. As you change the macroscopic variables of a system, sometimes its properties will abruptly change, often in a dramatic way.
  3. [3]
    Phases and phase transitions in disordered quantum systems
    Formally, we can define a phase transition as a singularity in the free energy as function of the external parameters such as temperature, pressure, magnetic ...
  4. [4]
    [PDF] 5. Phase Transitions - DAMTP
    In practice, we nearly always deal with first, second and (very rarely) third order transitions. The liquid-gas transition releases latent heat, which means ...<|separator|>
  5. [5]
    [PDF] Handout 13 - Phase transitions
    First order transitions are therefore discontinuous. Second order transitions have discontinuities in the second derivatives of G: (. ∂2G. ∂T2.
  6. [6]
    [PDF] Chapter 6 Phase transitions
    critical point : The first-order phase boundary between gas and liquid becomes second order right at the critical point. The two phases have then equal ...
  7. [7]
    [PDF] INTRODUCTION to PHASE TRANSITIONS - UBC Physics
    A 1st order transition has a kink, & its derivative is discontinuous. A 2nd order transition has a kink in the derivative; and so on. One then gets the ...
  8. [8]
    [PDF] Phase Transitions and Critical Phenomena - Heidelberg University
    Oct 18, 2021 · Phase transitions are property changes induced by modifying thermodynamic variables. First-order transitions involve discontinuities, while ...
  9. [9]
    [PDF] Phase Transitions and Critical Phenomena
    Jun 5, 2020 · Phase transitions are changes between states of matter, characterized by rapid changes in properties and critical behavior, induced by external ...
  10. [10]
    [PDF] Critical Phenomena and the Renormalization Group - UBC Physics
    Critical phenomena are second-order phase transitions, occurring at the critical point, where there is no distinction between phases. They are studied using  ...
  11. [11]
    [PDF] Phase Transitions and Collective Phenomena - TCM
    This course covers phase transitions, critical phenomena, Ginzburg-Landau theory, scaling, topological and quantum phase transitions, and collective modes.
  12. [12]
    Phase transitions, critical phenomena, and instabilities - PubMed
    Transformations among many of the diverse states of matter arise from microscopic interactions involving very many (approximately 10(23)) constituent particles.
  13. [13]
    Joseph Black and Latent Heat - American Physical Society
    The latent heat that Black discovered greatly slows the melting of snow and ice. He gave the first account of this work on April 23, 1762 at the University of ...Missing: 1761 | Show results with:1761
  14. [14]
    Joseph Black's Pioneering Discoveries about Heat | Encyclopedia.com
    In 1761, Joseph Black (1728-1799), an English chemist, discovered that ice, while it was in the process of melting, did not warm up until it was completely ...
  15. [15]
    Joseph Black, carbon dioxide, latent heat, and the beginnings of the ...
    Black's discovery of carbon dioxide was the first major advance in the discovery of the respiratory gases. At the time the topic was known as “pneumatic ...
  16. [16]
    [PDF] The discovery of latent heat 250 years ago
    In this note we will look at the work of two natural scientists, Jean André DELUC and Joseph BLACK who, shortly after the invention of reliable thermometers and.
  17. [17]
    How Gibbs derived the Phase Rule | carnotcycle - WordPress.com
    Sep 1, 2015 · The Phase Rule formula was first stated by the American mathematical physicist Josiah Willard Gibbs in his monumental masterwork On the ...<|separator|>
  18. [18]
    Historic Note No. 1: Gibbs' Phase Rule - Thermo-Calc Software
    This article is the first in a three-part series tracing the evolution of CALPHAD from the Gibbs Phase Rule to the first lattice stabilities.
  19. [19]
    The Ehrenfest Classification of Phase Transitions: Introduction and ...
    The first classification of general types of transition between phases of matter, introduced by Paul Ehrenfest in 1933, lies at a crossroads in the thermod.
  20. [20]
    [1612.03062] A Look Back at the Ehrenfest Classification ... - arXiv
    Dec 9, 2016 · Translation and Commentary of Ehrenfest's 1933 paper introducing the notion of phase transitions of different order. Authors:Tilman Sauer.
  21. [21]
    On the theory of phase transitions - Inspire HEP
    The question about continuous phase transitions (without latent heat) is investigated from the general thermodynamical point of view.
  22. [22]
    [PDF] ON THE THEORY OF PHASE TRANSITIONS 25
    Landau), he constructed a general theory of phase transitions of the second order (1936–. 1937). The Ginzburg-Landau equation is of great importance in the ...
  23. [23]
    A look back at the Ehrenfest classification
    Apr 5, 2017 · A translation of Paul Ehrenfest's 1933 paper, entitled Phase transitions in the usual and generalized sense, classified according to the singularities of the ...
  24. [24]
    (PDF) Landau and Theory of Phase Transitions - ResearchGate
    Dec 27, 2016 · This article describes the history of the creation and further development of Landau's famous works on phase transi- tions, diamagnetism of electron gas ( ...
  25. [25]
    Crystal Statistics. I. A Two-Dimensional Model with an Order ...
    The partition function of a two-dimensional ferromagnetic with scalar spins (Ising model) is computed rigorously for the case of vanishing field.
  26. [26]
    Fifty Years of the Exact Solution of the Two-Dimensional Ising Model ...
    Nov 1, 1995 · Abstract: The exact solution of the two-dimensional Ising model by Onsager in 1944 represents one of the landmarks in theoretical physics.
  27. [27]
    [PDF] Kenneth G. Wilson - Nobel Lecture
    The renormalization group approach that was defined in 1971 embraces both practical approximations leading to actual computations and a formalism.17 The full ...
  28. [28]
    Renormalization Group and Critical Phenomena. II. Phase-Space ...
    Nov 1, 1971 · Renormalization Group and Critical Phenomena. II. Phase-Space Cell Analysis of Critical Behavior. Kenneth G. Wilson.
  29. [29]
    [PDF] 05. Thermodynamics of Phase Transitions I - DigitalCommons@URI
    Dec 16, 2015 · Consider a 1-component fluid system with phases I and II. The Gibbs free energy has a different functional dependence on its natural variables ...Missing: "peer | Show results with:"peer
  30. [30]
    [PDF] BASIC THERMODYNAMICS AND KINETICS OF PHASE ...
    Thermo- dynamic phase transitions in materials and condensed matter, the subject of this book, occur when there is a singularity in the free energy function of ...<|separator|>
  31. [31]
    [PDF] SIO 224 Thermodynamics of phase transitions
    When discussing phase transformations, we usually work in the P, T domain in which G, the Gibbs free energy is the most useful function.
  32. [32]
    7.5: Landau Theory of Phase Transitions - Physics LibreTexts
    May 24, 2020 · Landau's theory of phase transitions is based on an expansion of the free energy of a thermodynamic system in terms of an order parameter.
  33. [33]
    [PDF] Lectures on Landau Theory of Phase Transitions
    Jul 9, 2015 · This transition is predicted by Landau (mean field) theory to be continuous, but it is believed to be first order in physical situations, due to ...
  34. [34]
    [PDF] Landau Theory of Second Order Phase Transitions
    Second order phase transitions occur when a new state of reduced symmetry develops continuously from the disordered (high temperature) phase.
  35. [35]
    [PDF] 1 Lecture 28 Physics 404 The Landau theory of phase transitions is ...
    There is a latent heat associated with this phase transition. An example order parameter is the difference in density of the liquid and. Page 3. 3 vapor phases, ...
  36. [36]
    Phases of Matter - NASA Glenn Research Center
    The normal phases of matter are solid, liquid, and gas. Plasma is considered a fourth phase.
  37. [37]
    Fundamentals of Phase Transitions - Chemistry LibreTexts
    Jan 29, 2023 · Phase transition is when a substance changes from a solid, liquid, or gas state to a different state. Every element and substance can transition from one phase ...Phase Changes · How Phase Transition works · Brief Explanation of a Phase...
  38. [38]
    List of Phase Changes Between States of Matter - ThoughtCo
    Jun 9, 2025 · There are eight phase changes: melting, freezing, vaporization, condensation, deposition, sublimation, ionization, and recombination.
  39. [39]
    States of matter & phase transitions – scienceinschool.org
    Jan 15, 2021 · States of matter & phase transitions Teach article · Vaporization – liquid to gas · Melting – solid to liquid · Condensation – gas to liquid.
  40. [40]
    States of Matter & Phase Transitions - S'Cool LAB
    States of Matter & Phase Transitions · Vaporisation - liquid to gas · Melting - solid to liquid · Condensation - gas to liquid.
  41. [41]
    [PDF] PDF - School of Physics and Astronomy
    Ehrenfest classification of the “order” of a Phase Transition. First order transitions have a discontinuous change of first derivative of the free energy ...
  42. [42]
    Lecture 20: Phase transition - ESM Intranet Site
    Ehrenfest proposed to classify phase transitions by their order. A first-order phase transition is categorized by a discontinuity in the first derivatives of ...
  43. [43]
    [PDF] Phase Transitions
    A second order phase transition does not have any latent heat associated with it; the entropy is continuous at TC. In a second order phase transition the order ...
  44. [44]
    Quantum Phase Transitions - ScienceDirect.com
    Quantum phase transition (QPT) occurs at zero temperature where thermal fluctuations are absent and instead the transition is driven by quantum fluctuations.
  45. [45]
    Dynamical quantum phase transitions (Review Article)
    Nov 1, 2016 · Dynamical quantum phase transitions reveal universality, scaling, connection to the topology, and many other interesting features. Here we ...
  46. [46]
    Visualizing quantum phases and identifying quantum phase ...
    Feb 2, 2021 · Identifying quantum phases and phase transitions is key to understanding complex phenomena in statistical physics.
  47. [47]
    [1911.10196] Geometry of quantum phase transitions - arXiv
    Nov 19, 2019 · In this article we provide a review of geometrical methods employed in the analysis of quantum phase transitions and non-equilibrium dissipative phase ...
  48. [48]
    Geometry of quantum phase transitions - ScienceDirect.com
    Jan 18, 2020 · This review focuses on the role of the geometric phases in the investigation of equilibrium and non-equilibrium quantum phase transitions.
  49. [49]
    Topological phase transitions - ScienceDirect.com
    Topological phase transitions are a new class of quantum critical phenomena, driven by a dual order parameter, and cannot be described by Landau theory.Topological Phase... · Topological Phases And Their... · Acknowledgements
  50. [50]
    [PDF] Chapter 5 Topological Phase Transitions - TCM
    The aim of this section is to discuss the phenomenology of this type of transition which lies outside the usual classification scheme. In classifying states of ...
  51. [51]
    Topological Quantum Phase Transitions (Chapter 13)
    May 4, 2017 · We will refer to topological quantum phase transitions (TQPT) as those that separate two phases where at least one of them is topologically non-trivial.
  52. [52]
    [PDF] TOPOLOGICAL PHASE TRANSITIONS AND ... - Nobel Prize
    Oct 4, 2016 · Haldane developed theoretical methods to describe phases of matter that cannot be identified by their pattern of symmetry breaking. In a 1982 ...Missing: history | Show results with:history
  53. [53]
    [1901.11468] Universality classes of topological phase transitions ...
    Jan 31, 2019 · A unified framework is employed to address the quantum criticality of the topological phase transitions in one to three spatial dimensions.
  54. [54]
    Classification of Topological Phase Transitions and van Hove ...
    We study quantum phase transitions in graphene superlattices in external magnetic fields, where a framework is presented to classify multiflavor Dirac fermion ...
  55. [55]
    [PDF] Introduction to Structural Phase Transitions - staff.uni-mainz.de
    Structural phase transitions occur when distortions in a crystalline solid cause a change of symmetry at discrete values of temperature, pressure, or chemical ...
  56. [56]
    Phase Transitions
    Phase transitions occur due to changes in temperature, pressure, or chemical synthesis. Solid-solid transitions are classified as first-order (reconstructive) ...
  57. [57]
    [PDF] Phase Transitions in Crystals - AIC Cristallografia
    Sep 9, 2016 · Phase transitions on crystallographic interest are those involving solid phases: with the change of the external conditions (temperature ...
  58. [58]
    The mechanism driving a solid–solid phase transition in a ... - NIH
    This study thus reveals the interplay between the conformational changes and the crystal phases – the underlying mechanism that drives the phase transition.
  59. [59]
    [PDF] Phase transitions in layered crystals - arXiv
    Oct 7, 2011 · Thermodynamics allows only two mechanisms of crystal phase transitions: (a) by nucleation-and-growth and (b) by instant change at a critical ...<|separator|>
  60. [60]
    [PDF] Some interesting phase transitions in solids
    Interesting phase transitions include low-dimensional solids, polytypes, organic solids, plastic crystals, glasses, silver selenide, spin-state transitions, ...
  61. [61]
    The Effects of Crystal Structure on Metal Properties
    Structural Transitions: Temperature-Induced Transitions: Metals like iron exhibit different crystal structures at different temperatures. For instance, iron ...
  62. [62]
    [PDF] Phase Transformations in Materials - IAM, KIT
    Phase Transformations in Materials. Displacive vs. Reconstructive. The reconstructive phase transformation can be suppressed in case diffusion processes.Missing: transitions | Show results with:transitions
  63. [63]
    4.6 Structural phase transitions - Condensed Matter Physics - Fiveable
    Structural phase transitions play a crucial role in condensed matter physics by altering material properties · Understanding different types of transitions ...
  64. [64]
    Transform to deform | Nature Materials
    Nov 30, 2022 · Phase-transformation-induced plasticity has also been reported in ionically bonded zirconia, through a displacive structural transition.Missing: examples | Show results with:examples
  65. [65]
    [PDF] Phase Transformations in Materials - KIT - IAM
    The transformation by nucleation and growth following super cooling below the equilibrium transformation temperature is a reconstructive transformation. L α β.
  66. [66]
    Structural mechanisms of pressure-induced isosymmetric second ...
    Jul 1, 2025 · This study provides new evidence for isosymmetric second-order phase transitions triggered by pressure and an increase in coordination number.
  67. [67]
    Study captures crystal phase changes in unprecedented detail
    Jul 23, 2025 · Using custom-built computer simulations, Cornell researchers have visualized solid-solid phase transitions in unprecedented detail, ...
  68. [68]
    Curie Temperature - NETZSCH Analyzing & Testing
    Discover the Curie Temperature's impact on ferromagnetic materials. Learn about key temperatures for elements like iron, cobalt, and nickel.
  69. [69]
    Tutorial: a beginner's guide to interpreting magnetic susceptibility ...
    Apr 19, 2022 · In this tutorial, we provide a guide to the interpretation of magnetic susceptibility data with a special emphasis on the Curie–Weiss law, a simple but ...
  70. [70]
    [PDF] Phase Transition of Fe3O4 Magnetic Material Based on Observation ...
    Mar 23, 2021 · Curie temperature is the critical temperature for the ferromagnetic phase transition of a solid material to become paramagnetic due to heating ...
  71. [71]
    A First-Order Phase Transition Induced by a Magnetic Field - Science
    An electronic (metal-to-insulator) phase transition of the first order, which can be caused by an external magnetic field, was discovered in Nd1/2Sr1/2MnO3.Missing: sources | Show results with:sources
  72. [72]
    BCS Theory of Superconductivity - HyperPhysics
    The exponential increase in heat capacity near the critical temperature suggests an energy bandgap for the superconducting material. This evidence for a bandgap ...<|separator|>
  73. [73]
    [PDF] 14 Superconductivity and BCS theory - Rutgers Physics
    Today, superconductivity has been observed in a wide vari- ety of materials (see Table 14.1), with transition temperatures reaching up as high as high as 134 K.
  74. [74]
    Quantum phase transition from superconducting to insulating-like ...
    Feb 17, 2022 · Here we report the first observation of a quantum phase transition from a superconducting state to an insulating-like state as a function of pressure.
  75. [75]
    Field-induced transition within the superconducting state of CeRh2As2
    Aug 26, 2021 · We show that CeRh 2 As 2 has extremely high superconducting critical fields of up to 14 T despite a superconducting transition temperature T c of only 0.26 K.
  76. [76]
    12.2: Gibbs Phase Rule - Chemistry LibreTexts
    Nov 24, 2022 · The Gibbs phase rule is obtained considering that the number of independent variables is given by the total number of variables minus the constraints.
  77. [77]
    The Gibbs phase rule - DoITPoMS
    The Gibbs phase rule links the number of phases, number of components and number of degrees of freedom in a system. A phase (p) can be defined as the homogenous ...
  78. [78]
    13.2: Phase Diagrams- Binary Systems - Chemistry LibreTexts
    Apr 12, 2022 · The open circles are critical points; the dashed curve is the critical curve. The dashed line a–b illustrates retrograde condensation at . ...
  79. [79]
    8.6: Phase Diagrams for Binary Mixtures - Chemistry LibreTexts
    Jul 19, 2021 · As suggested by the Gibbs Phase Rule, the most important variables describing a mixture are pressure, temperature and composition.
  80. [80]
    [PDF] Fluid-phase behavior of binary mixtures in which one component ...
    A, B, C, D, and E 共in inset兲 are binary mixture critical lines. Cw and Cs are the vapor-liquid critical points for water and solute 1, respectively. DCP1 is a ...
  81. [81]
    [PDF] Phase Separation in Mixtures
    Phase separation occurs when the average Gibbs energy is lowered, and the equilibrium state is phase separated. This can happen in liquid-gas or solid-liquid ...
  82. [82]
    Phase-Separating Binary Polymer Mixtures: The Degeneracy ... - NIH
    The unstable region can be found above the spinodal (blue dashed line). The binodal and spinodal intersect at the critical point (solid circle), where they ...
  83. [83]
    Phase separation in fluids with many interacting components - PMC
    Here, we leverage approaches from random-matrix theory and statistical physics to describe the emergent phase behavior of fluid mixtures with many species whose ...Missing: "peer | Show results with:"peer
  84. [84]
    Binary Systems | PNG 520: Phase Behavior of Natural Gas and ...
    Critical Point (Pc,Tc): The temperature and pressure for which liquid and vapor are indistinguishable. Again, this definition is applicable both for mixtures ...
  85. [85]
    Computing the phase diagram of binary mixtures: A patchy particle ...
    The black dashed line is the line of critical points of the mixture. Tie lines are not vertical lines in this representation. FIG. 7. FIG. 7. Projection of ...INTRODUCTION · Computational methods · Critical points · Coexistence region
  86. [86]
    Meta-analysis of critical points to determine second virial coefficients ...
    Previous work demonstrated that the critical point of a phase diagram (two parameters) of a binary polymer mixture provides too little information to allow a ...
  87. [87]
    [2207.13748] The Kosterlitz-Thouless phase transition - arXiv
    Jul 27, 2022 · This is a set of notes recalling some of the most important results on the XY model from the ground up. They are meant for a junior researcher wanting to get ...
  88. [88]
    Sharpness of the Berezinskii-Kosterlitz-Thouless Transition in ...
    We present a comprehensive investigation of the Berezinskii-Kosterlitz-Thouless transition in ultrathin strongly disordered NbN films.
  89. [89]
    [PDF] The Berezinskii-Kosterlitz-Thouless Transition - UBC Physics
    The occurrence of the Berezinskii-Kosterlitz-Thouless transition is marked by the transition of bound vortex-antivortex pairs at low temperatures to isolated ...
  90. [90]
    Perspective: The glass transition | The Journal of Chemical Physics
    In its most general sense, the glass transition refers to the generic change in a many-body system from an equilibrium fluid state to a non-equilibrium ...INTRODUCTION · II. STYLISED FACTS OF THE... · VI. RECENT THEORETICAL...
  91. [91]
    Glass transition as a topological phase transition | Phys. Rev. E
    The glass transition is described as a phase transition in the system of topologically protected excitations in matter structure.
  92. [92]
    Glass Transition Temperature - an overview | ScienceDirect Topics
    Glass transition temperatures are very variable ranging from 140 °C to 370 °C according to the grades, curing process and the moisture content.
  93. [93]
    Observation of time crystal in a spin maser system - Nature
    May 5, 2025 · This time crystal phase emerges only when the feedback strength exceeds a critical value, at which the system experiences a first order phase ...<|control11|><|separator|>
  94. [94]
    Nonequilibrium Transition between Dissipative Time Crystals
    A new unifying scenario is presented in which two different time-crystalline orders are separated by a nonequilibrium phase transition.Abstract · Article Text · INTRODUCTION · EXCEPTIONAL POINT AS...
  95. [95]
    [2402.12378] Observation of a phase transition from a continuous to ...
    Feb 19, 2024 · Here, we explore a phase transition from a continuous time crystal to a discrete time crystal. A CTC with a characteristic oscillation ...
  96. [96]
    [PDF] Lecture 9: Phase Transitions - Matthew D. Schwartz
    (7) This equation, called the Clapeyron equation, determines the shape of the phase boundary. is called the latent heat. we have simply that Q = H: the heat ...
  97. [97]
    11.3 Phase Change and Latent Heat - Physics | OpenStax
    Mar 26, 2020 · Because this energy enters or leaves a system during a phase change without causing a temperature change in the system, it is known as latent ...<|separator|>
  98. [98]
    23.4: The Clausius-Clapeyron Equation - Chemistry LibreTexts
    Mar 4, 2025 · The Clapeyron equation, which tells us that the slope of the coexistence curve is related to the ratio of the molar enthalpy between the phases to the change ...
  99. [99]
    [PDF] Critical Exponents and the Renormalization Group - UBC Physics
    At this temperature, our system passes through a second-order phase transition: the magnetization m changes continuously (figure 2), but is derivative changes ...
  100. [100]
    [2411.02604] Computing critical exponents in 3D Ising model via ...
    Nov 4, 2024 · In this study, we computed three critical exponents (\alpha, \beta, \gamma) for the 3D Ising model with Metropolis Algorithm using Finite-Size Scaling Analysis.
  101. [101]
    (PDF) Critical Exponents of the 3-D Ising Model - ResearchGate
    Aug 6, 2025 · PDF | We present a status report on the ongoing analysis of the 3D Ising model with nearest-neighbor interactions using the Monte Carlo
  102. [102]
    Critical exponents of the Ising model in three dimensions with long ...
    Jun 27, 2022 · The critical behavior of the Ising model in three dimensions on a lattice with site disorder is studied by applying Monte Carlo simulation techniques.
  103. [103]
    [PDF] Universality Explained - PhilSci-Archive
    Apr 16, 2016 · Universality is the multiple realizability where systems near critical points have identical critical behavior, like liquid-gas and ...
  104. [104]
    Universality, Phase Transitions and Statistical Mechanics
    Universality means that singularities associated with second order phase transitions are universal, not depending on model details.
  105. [105]
    Universal scaling behavior of non-equilibrium phase transitions - arXiv
    Jan 11, 2005 · One of the most impressive features of continuous phase transitions is the concept of universality, that allows to group the great variety of ...
  106. [106]
    [PDF] Chapter 3 The Scaling Hypothesis - TCM
    1A function f(x) is said to be homogeneous of degree k if it satisfies the relation f(x) = bkf(bx). Phase Transitions and Collective Phenomena. Page 2. 38.
  107. [107]
    [PDF] 7. Widom's scaling ansatz and scaling laws In the past few Sections ...
    In other words, for an Ising-like (second order) phase transitions, there are only two independent critical exponents. Once two of the critical indices are ...
  108. [108]
    Scaling laws - Scholarpedia
    Oct 19, 2009 · Scaling laws are the expression of physical principles in the mathematical language of homogeneous functions.Introduction · Scaling laws · Critical exponents · Exponents and space dimension
  109. [109]
    [PDF] 9. Phase transitions and finite size scaling
    Second order transitions are classified by their critical exponents, which characterize the behaviour at the critical point. The most important here are.
  110. [110]
    5 Phase Transitions‣ Statistical Physics by David Tong - DAMTP
    A phase transition is an abrupt, discontinuous change in the properties of a system, like steam condensing to water or water freezing to ice.
  111. [111]
    [PDF] Phase transitions and critical phenomena 1 Introduction
    Apr 29, 2024 · This report provides an overview of phase transitions and critical phenomena, focusing on their theoretical foundations.
  112. [112]
    [hep-lat/9806012] Critical Exponents of the 3D Ising Universality ...
    Jun 11, 1998 · Our best estimates for the critical exponents are nu= 0.6298(5) and eta= 0.0366(8), where the given error estimates take into account the ...
  113. [113]
    Structure of fluctuations near mean-field critical points and spinodals ...
    Mar 22, 2007 · We analyze the structure of fluctuations near critical points and spinodals in mean-field and near-mean-field systems.
  114. [114]
    Critical Correlations in the Ising Model | Phys. Rev.
    The correlation function is not of the Ornstein-Zernike type, just above the critical point, but rather 𝐶 ⁡ ( 𝑅 ) ∝ 𝑅 − 1 − 𝜂 for 𝑅 ≪ 1 𝜅 , and 𝐶 ⁡ ( 𝑅 ) ∝ 𝜅 𝜂 ...
  115. [115]
    Critical fluctuations and slowing down of chaos - Nature
    May 14, 2019 · Fluids cooled to the liquid–vapor critical point develop system-spanning fluctuations in density that transform their visual appearance.
  116. [116]
    The dynamic critical exponent z for 2d and 3d Ising models from five ...
    Feb 16, 2022 · We calculate the dynamical critical exponent z for 2d and 3d Ising universality classes by means of minimally subtracted five-loop ε expansion ...
  117. [117]
    [PDF] Advanced information on the Nobel Prize in Physics 2003
    introduced by Lev Landau in connection with his 1937 theory of second order ... In accordance with Landau's general theory of second-order phase transitions, the.
  118. [118]
    [PDF] Thermodynamical Phase transitions, the mean-field theories ... - arXiv
    Feb 27, 2014 · For the second order phase transitions, first order derivative of Gibbs free ... Landau. (1908–1968). In a landmark paper[2] of 1937 Lev Landau ...
  119. [119]
    [PDF] Chapter 3: Landau Theory for Phase Transitions
    In our treatment of mean-fielf theory of phase transitions, we have seen that the central point is the behavior of the free energy close to the phase ...
  120. [120]
    [PDF] Ginzburg-Landau Expansion
    Ginzburg-Landau Theory focuses on the Free Energy Difference between two phases. It assumes that one phase is related to another by a small parameter which ...
  121. [121]
    Nobel Prize in Physics Honors Theoretical Work on ... - AIP Publishing
    Dec 1, 2003 · Their work followed Landau's earlier approach to second-order phase transitions. To characterize the change of a system from a disordered to an ...
  122. [122]
    [PDF] 1 Unit 4-4: Critical Exponents within the Mean-Field Approximation ...
    Apr 4, 2021 · For the Ising model at h = 0, the Hamiltonian has an inversion symmetry with respect to si ↔ −si, so the free energy must satisfy f(m, T) = f(− ...
  123. [123]
    [PDF] arXiv:cond-mat/9511003v2 10 Nov 1995
    Ising studied the model and found its exact solution in one dimension and showed that there is no phase transition from para to ferromagnetism [6]. Since then ...<|separator|>
  124. [124]
    Computing critical exponents in 3D Ising model via pattern ... - arXiv
    Nov 4, 2024 · Two ways to improve our critical exponent values would be to use larger system sizes, spanning a range of 1-2 orders of magnitude (10-100x) ...
  125. [125]
    [PDF] the infinite range ising model, landau theory and critical exponents
    In this Hamiltonian each spin interacts with the sum of all of the rest of the spins. This the latter are large in number, the fluctuations of the expectation ...
  126. [126]
    [PDF] THE RENORMALIZATION GROUP AND THE ~EXPANSION
    The modern formulation of the renormalization group is explained for both critical phenomena in classical statistical mechanics and quantum field theory.
  127. [127]
    [PDF] Concepts of Renormalization in Physics - arXiv
    Aug 24, 2005 · Renormalization was introduced as a tool to predict physical properties in phase transitions, as will be described in this article, and Kenneth ...
  128. [128]
    [PDF] New applications of the renormalization group method in physics
    Feb 28, 2011 · Abstract. The renormalization group method developed by Ken Wilson more than four decades ago has revolutionized the way we think.
  129. [129]
    [quant-ph/0109024] Entanglement, quantum phase transitions, and ...
    Sep 5, 2001 · We investigate the role of entanglement in quantum phase transitions, and show that the success of the density matrix renormalization group (DMRG)
  130. [130]
    [PDF] arXiv:2309.02484v1 [cond-mat.stat-mech] 5 Sep 2023
    Sep 5, 2023 · By systematizing the underpinnings of this universality, Kenneth Wilson formulated a far-reaching renor- malization group principle [1] and in ...<|separator|>
  131. [131]
    The renormalization group: Critical phenomena and the Kondo ...
    Oct 1, 1975 · This review covers several topics involving renormalization group ideas. The solution of the s -wave Kondo Hamiltonian, describing a single magnetic impurity ...
  132. [132]
    Always relevant | Nature Physics
    Nov 9, 2023 · The modern formulation of the renormalization group relies heavily on work by Kenneth Wilson in the 1970s, for which he was awarded the Nobel ...Missing: review | Show results with:review
  133. [133]
    Analysis of structural and magnetic phase transition behaviors of La ...
    In our previous study [7], we have analyzed the structural and magnetic phase transition of La1−xSrxCrO3 by differential scanning calorimetry (DSC), dilatometry ...
  134. [134]
    Phase transitions with dilatometer - Linseis
    A dilatometer can be used to determine phase transitions in materials, as a phase change is also accompanied by a change in the coefficient of thermal ...
  135. [135]
    A quantitative criterion for determining the order of magnetic phase ...
    Jul 11, 2018 · In the magnetocalorics field, techniques for determining the order of phase transitions are distinguished as calorimetric or magnetic. The ...
  136. [136]
    The curious case of the structural phase transition in SnSe insights ...
    Jun 3, 2023 · These phenomena are entirely consistent with the soft phonon view of the phase transition observed with inelastic neutron scattering and ...
  137. [137]
    Neutron-scattering study on phase transitions of CsPb | Phys. Rev. B
    May 15, 1974 · Neutron-scattering experiments have elucidated the mechanism of the successive phase transitions of CsPb C l 3 at 47, 42, and 37°C.
  138. [138]
    Neutron scattering studies of structural phase transitions at ...
    Jul 1, 1974 · The neutron scattering technique has been particularly useful in obtaining detailed knowledge of dynamical properties near the phase transition.
  139. [139]
    Ultrafast X-ray imaging of the light-induced phase transition in VO 2
    Dec 22, 2022 · Here we use time- and spectrally resolved coherent X-ray imaging to track the prototypical light-induced insulator-to-metal phase transition in vanadium ...
  140. [140]
    X ray absorption spectroscopy investigation of phase transition in Ge ...
    Dec 1, 2006 · The phase transitions of GaAs, GaP and Ge under pressure have been investigated by x-ray absorption spectroscopy (XAS).
  141. [141]
    Phase Transformations in Porous Materials Studied by In Situ Solid ...
    Aug 4, 2020 · The combination of both in situ MAS NMR and in situ powder X-ray diffraction allowed us to obtain new insights into two phase transformation mechanisms in ...
  142. [142]
    Phase transitions in low dimensional materials characterized by ...
    We report how Raman spectroscopy is used to characterize a lithium intercalation induced phase transition in the Weyl semimetal WTe2.<|separator|>
  143. [143]
    Membrane lipid phase transitions and phase organization studied ...
    In this review, we discuss how FTIR spectroscopy can be utilized to study lipid phase structure and organization, and lipid phase transitions, in lipid model ...
  144. [144]
  145. [145]
    Direct observation of ultrafast cluster dynamics in supercritical ...
    Dec 3, 2024 · Near the critical point, the stability of large clusters leads to a reduction in molecular motions, a phenomenon known as critical slowing down.
  146. [146]
    Pressure‐Controlled Electronic and Structural Phase Transitions in ...
    Feb 14, 2025 · In situ, high-pressure ultrafast spectroscopy is used to study the nonequilibrium dynamics of Fermi and Bose particles in SnSe.
  147. [147]
    Observation of the magnonic Dicke superradiant phase transition
    Apr 4, 2025 · Two-level atoms ultrastrongly coupled with single-mode cavity photons are predicted to exhibit a quantum phase transition, entering a phase ...<|separator|>
  148. [148]
    Experimental observation of a time rondeau crystal | Nature Physics
    Oct 14, 2025 · Recently, breaking of time-translation symmetry was observed in non-equilibrium systems, producing so-called time crystals. Here we investigate ...
  149. [149]
    Parametric tuning of quantum phase transitions in ultracold reactions
    Nov 26, 2024 · One of the landmark experiments in the field is the creation of Bose-Einstein condensates (BECs) using dilute gases of alkali metal atoms in the ...
  150. [150]
    Phase Transitions in Materials
    This book covers basic thermodynamics, kinetics, temperature-composition diagrams, diffusion, nucleation, melting, solidification, spinodal decomposition, and ...
  151. [151]
    [PDF] CHAPTER 10: PHASE TRANSFORMATIONS
    How does the rate of transformation depend on time and T? 1. • How can we slow down the transformation so that we can engineering non-equilibrium structures?<|separator|>
  152. [152]
    Martensitic Transformation - an overview | ScienceDirect Topics
    Martensitic transformation is defined as a diffusionless phase change in certain alloys, characterized by the formation of martensite from austenite upon ...
  153. [153]
    the formation of martensite - Total Materia
    Abstract. Martensitic transformation represents a key microstructural change in steels, occurring through a diffusionless shear mechanism rather than atomic ...
  154. [154]
    [PDF] Martensite in Steels
    The transformation goes to all the trouble of ensuring that the shape deformation is macroscopically an invariant–plane strain because this reduces the strain ...
  155. [155]
    Thermodynamic Modeling of Multicomponent Phase Equilibria
    Phase diagrams are visual representations of the state of a material as a function of temperature, pressure and concentrations of the constituent components ...
  156. [156]
    Phase Transitions in Materials
    10 - Melting · 11 - Transformations involving precipitates and interfaces · 12 - Spinodal decomposition · 13 - Phase field theory · 14 - Method of concentration ...
  157. [157]
    (PDF) Phase Diagram Applications - Academia.edu
    Phase diagrams are essential for alloy design, processing, performance, and troubleshooting metallurgical issues. Understanding phase diagrams enhances alloy ...<|separator|>
  158. [158]
    [2008.09136] Phase transitions in the early universe - arXiv
    Aug 20, 2020 · The aim is to provide the necessary basics to understand first-order phase transitions in the early universe, to outline how they leave imprints in ...
  159. [159]
    [PDF] Phase Transitions in the Early Universe - webspace.science.uu.nl
    The following report offers a review of two cosmological phase transitions: the elec- troweak and the QCD phase transitions in the early universe.
  160. [160]
    [2212.00156] Signatures of a High Temperature QCD Transition in ...
    Nov 30, 2022 · We show that a first order QCD phase transition at \sim7 TeV is consistent with the Subaru Hyper-Suprime Cam candidate event, while a \sim 70 ...
  161. [161]
    A first-order deconfinement phase transition in the early universe ...
    May 1, 2024 · This paper discusses a first-order deconfinement phase transition in the early universe, its potential, and the gravitational wave spectrum, ...
  162. [162]
    Cosmological implications of the QCD phase transition in the Early ...
    Aug 30, 2022 · A QCD phase transition is predicted to occur during the early stages of the dynamical evolution of the Universe, with the order and the ...
  163. [163]
    Gravitational waves from a first-order electroweak phase transition
    Jan 22, 2018 · We review the production of gravitational waves by an electroweak first-order phase transition. The resulting signal is a good candidate for ...
  164. [164]
    [PDF] Lecture 2B: Phase Transitions - CERN Indico
    ◦ Fast (slow) phase transition at high (low) temperature → high (low) frequency. ◦ GWs from electroweak phase transition at T ∼ 100 GeV in the LISA (mHz) band.
  165. [165]
    [PDF] Topological Defects in Cosmology - McGill Physics
    Topological defects are produced during phase transitions in the very early Universe. ... topological defects in field theory and cosmology is relativistic.
  166. [166]
    [PDF] Topological Defects in Cosmology
    In the context of the standard Big Bang theory, cosmological phase transitions are produced by the spontaneous breaking of a fundamental symmetry, such as the ...
  167. [167]
    SciPost Phys. Lect. Notes 24 (2021) - Phase transitions in the early ...
    Feb 15, 2021 · The aim is to provide the necessary basics to understand first-order phase transitions in the early universe, to outline how they leave imprints in ...Missing: review | Show results with:review
  168. [168]
    Percolation on complex networks: Theory and application - arXiv
    Jan 28, 2021 · The intention of this paper is to offer an overview of these applications, as well as the basic theory of percolation transition on network ...
  169. [169]
    Inducing effect on the percolation transition in complex networks
    Sep 9, 2013 · Percolation transition on complex networks occurs in a wide range of natural, technological and socioeconomic systems. The emergence of ...
  170. [170]
    Information and phase transitions in socio-economic systems
    In this paper we explore two closely related phase transitions which appear in social economic systems. One is herding behaviour, where individuals behave more ...
  171. [171]
    Phase transitions in evolutionary dynamics | Chaos - AIP Publishing
    Dec 29, 2022 · Sharp changes in state, such as transitions from survival to extinction, are hallmarks of evolutionary dynamics in biological systems.
  172. [172]
    Tipping elements in the Earth's climate system - PNAS
    The tipping point is the corresponding critical point—in forcing and a feature of the system—at which the future state of the system is qualitatively altered.
  173. [173]
    Phase transitions in biology: from bird flocks to population dynamics
    Oct 20, 2021 · We provide a brief review of phase transitions and their role in explaining biological processes ranging from collective behaviour in animal flocks to neuronal ...Abstract · Introduction · Phase transitions in physics · Phase transitions in biology
  174. [174]
    Universal early warning signals of phase transitions in climate systems
    Apr 5, 2023 · This paper provides a proof of concept for this approach as applied to lattice phase transitions: a deep neural network trained exclusively on two-dimensional ...
  175. [175]
    Deconfined Quantum Criticality, Scaling Violations, and Classical ...
    Dec 23, 2015 · There is ongoing debate about whether the proposed “deconfined critical point” exists in such magnets or whether there is instead a very weak ...<|separator|>
  176. [176]
    [2307.02547] Deconfined quantum criticality lost - arXiv
    Jul 5, 2023 · Over the past two decades, the enigma of the deconfined quantum critical point (DQCP) attracted broad attention across condensed matter and ...
  177. [177]
    Deconfined quantum critical point lost in pressurized SrCu 2 (BO 3 ) 2
    Feb 22, 2025 · The controversy is more clearly revealed very recently that the DQCP “fails” a series of general standards from a quantum entanglement ...
  178. [178]
    Evolution of entanglement entropy at SU(N) deconfined quantum ...
    Feb 7, 2025 · In our work, we demonstrate that an anomalous logarithmic behavior in the entanglement entropy (EE) persists in a class of models analogous to the DQCP.
  179. [179]
    Deconfined Quantum Critical Points: Symmetries and Dualities
    Sep 22, 2017 · We propose several dualities for the deconfined QCP with SU(2) spin symmetry which together make natural the emergence of a previously suggested SO(5) symmetry.
  180. [180]
    Fermi-liquid instabilities at magnetic quantum phase transitions - arXiv
    Jun 13, 2006 · The Hertz-Millis-Moriya theory of quantum phase transitions is described in detail. The breakdown of the latter is analyzed in several examples.Missing: criticisms | Show results with:criticisms
  181. [181]
    Non-Hertz-Millis scaling of the antiferromagnetic quantum critical ...
    May 3, 2023 · The most common of such phase transitions in Nature is the onset of antiferromagnetic (AF) spin-density wave (SDW) order in a metal. This ...
  182. [182]
    Fermi-liquid instabilities at magnetic quantum phase transitions
    Aug 17, 2007 · This review discusses instabilities of the Fermi-liquid state of conduction electrons in metals with particular emphasis on magnetic quantum critical points.Article Text · Introduction · Landau Fermi-Liquid Theory · ConclusionsMissing: criticisms | Show results with:criticisms
  183. [183]
    [2407.15919] Lectures on the quantum phase transitions of metals
    Jul 22, 2024 · Abstract:Quantum phase transitions of metals involve changes in the Fermi surface, and can be divided into three categories.
  184. [184]
    Quantum phase transitions - IOPscience
    Nov 3, 2003 · It will give an overview on a variety of different quantum transitions, critically discuss open theoretical questions, and frequently make ...<|control11|><|separator|>
  185. [185]
    Phase transitions and dimensional cross-over in layered confined ...
    Apr 21, 2025 · Spatial dimension is one of the main factors influencing the existence and nature of phase transitions, and this issue has inspired innumerable ...Missing: controversies | Show results with:controversies
  186. [186]
    Paradox regarding phase transitions in relativistic systems
    Sep 16, 2019 · A phase transition should occur when the attracting forces between particles becomes high relative to their relative motion (broadly speaking).Why is Einstein solution to the twin paradox different from the one on ...A problem with proper acceleration in the Twin Paradox [closed]More results from physics.stackexchange.comMissing: dimensionality | Show results with:dimensionality
  187. [187]
    Relativistic hydrodynamics with phase transition
    Aug 19, 2024 · Our findings suggest that 'hydrodynamic series breakdown near transition points” is valid exclusively for second-order phase transitions, not for crossovers or ...
  188. [188]
    FRG analysis for a relativistic BEC in arbitrary spatial dimensions
    Apr 24, 2025 · Our numerical results are consistent with the Mermin-Wagner theorem, and this consistency is also analytically confirmed from the flow equation.<|separator|>
  189. [189]
    [PDF] Relativistic Phases in Condensed Matter
    Oct 22, 2018 · ... phase transition . . . 17. 1.4 ... Figure 1.7: The Klein paradox: because of their relativistic nature, electrons in graphene can propagate.
  190. [190]
    [PDF] On the classification of phase transitions
    Ehrenfest's classification involves sidewise derivatives; therefore, to define Ehren- fest type second-order phase connection, we need the following notion.<|separator|>
  191. [191]
    [PDF] Phase Transitions and Critical Phenomena - Heidelberg University
    May 25, 2023 · As science evolved, the first real theoretical contact with phase transitions was possible with the development of a notion of tem- perature.
  192. [192]
    Theoretical perspective on the glass transition and amorphous ...
    Jun 20, 2011 · The glass transition is when a liquid becomes too viscous to flow, below Tg, and is not a thermodynamic transition. It's a state of matter with ...
  193. [193]
    [PDF] Quantum Phase Transitions in periodically quenched systems
    Jun 11, 2024 · Quantum phase transitions encompass a variety of phenomena that occur in quan- tum systems exhibiting several possible symmetries.