Fact-checked by Grok 2 weeks ago

Cluster expansion

Cluster expansion is a fundamental mathematical technique in used to express the logarithm of the partition function as a convergent in terms of cluster or polymer interactions, enabling the rigorous computation of thermodynamic quantities such as , , and correlation functions under conditions like high , low , or strong external fields. Developed initially for analyzing the of ideal and interacting gases, it reorganizes microscopic interactions into graphical representations like Mayer graphs or tree structures, bridging microscopic models to macroscopic behavior while ensuring analyticity in relevant parameter regimes. The method originated in the 1930s with Joseph E. Mayer's work on the for classical gases, where the is expanded as a series in the activity λ, with coefficients derived from connected cluster integrals: βp(β, λ) = ∑_{n=1}^∞ C_n(β) λ^n / n!. Rigorous proofs of convergence were established in the by Penrose and , who introduced bounds such as the Penrose-Ruelle |C_n(β)| ≤ e^{2βB} (n-2) n^{n-2} [C(β)]^{n-1} / n!, guaranteeing for sufficiently small |λ| < 1 / [e^{2βB+1} C(β)], where B is the stability constant and C(β) integrates the Mayer f-function |e^{-βV(x)} - 1|. Subsequent advancements, including the polymer gas representation by Gruber and Kunz in 1970 and inductive convergence criteria by Kotecký and Preiss in 1986, extended the technique to lattice spin systems and more general interactions, with modern variants incorporating tree-graph identities and Kirkwood-Salsburg equations for improved efficiency. In practice, cluster expansions represent the system as a "gas" of polymers—irreducible geometric configurations with associated weights w(γ) encoding interactions—where the free energy emerges from the logarithm of a grand canonical partition function over compatible polymer collections, converging via conditions like ∑_{γ ≁ γ*} |w(γ)| e^{a|γ|} ≤ a - 1 for some a > 0. For spin systems like the Ising model, high-temperature expansions yield the pressure as an absolutely convergent series when βJ ≤ 0.151 on the square lattice, while low-temperature expansions apply near zero temperature. These expansions are particularly powerful for proving the existence of the thermodynamic limit, identifying phase transitions through non-analyticities in the series, and computing critical exponents in models such as hard-sphere gases (with convergence radius R_V ≥ 0.5107 in 2D) and Potts models. Beyond classical , cluster expansions have been adapted to via Glimm-Jaffe-Spencer methods from constructive and find applications in for alloy , where effective cluster interactions are fitted to energies. Recent developments emphasize abstract frameworks for tempered and stable potentials, ensuring applicability to disordered systems and long-range interactions, with radii often bounded by of Ursell functions.

Fundamentals

Definition and motivation

In statistical mechanics, the cluster expansion provides a perturbative approach to approximate thermodynamic properties of interacting particle systems by expressing the pressure as a power series in the fugacity, also known as the activity z = e^{\beta \mu}/\lambda^3, where \beta = 1/kT, \mu is the chemical potential, k is Boltzmann's constant, T is temperature, and \lambda is the thermal wavelength. This method, originally developed by Mayer and Montroll, decomposes the contributions from particle interactions into sums over "clusters" of particles, facilitating the computation of quantities like pressure and density in the grand canonical ensemble. The primary motivation for cluster expansions arises from the difficulty in exactly evaluating the partition function for systems with non-trivial interactions, such as pairwise potentials between particles, where direct summation over all configurations becomes intractable for large numbers of particles N or volumes V. By recasting the grand partition function \Xi in terms of irreducible cluster integrals, the expansion addresses these challenges, particularly for dilute gases or low-density regimes, enabling systematic approximations that reveal how interactions deviate from ideal gas behavior. This perturbative framework is especially valuable for deriving virial expansions of the equation of state, where higher-order terms capture many-body effects. Consider a classical of N in a V subject to pairwise interactions, analyzed in the grand canonical ensemble, which allows fluctuations in particle number. The grand partition function is \Xi = \sum_{N=0}^\infty \frac{z^N}{N!} \int d\mathbf{q}^N \exp\left(-\beta U(\mathbf{q}^N)\right), where U is the total and the integral is over configuration space. The grand potential is then defined as \Phi = -kT \log \Xi, which relates to the via pV = kT \log \Xi = -\Phi. The cluster expansion yields the key relation \beta p = \sum_{l=1}^\infty b_l z^l, where the coefficients b_l are the cluster coefficients, each expressed as an integral over connected clusters of l particles that encode the interaction effects (from which the virial coefficients in the density expansion can be derived). These coefficients b_l are volume-independent in the and provide a systematic way to compute thermodynamic observables order by order in z.

Historical background

The cluster expansion technique was first introduced by Joseph E. Mayer in 1937 through his derivation of the equation of state for a gas using cluster integrals, providing a systematic way to compute virial coefficients for non-ideal gases. This work was expanded in the 1940 monograph co-authored with , where the method was fully developed as a perturbation series for the partition function of interacting particle systems, emphasizing its application to real gases beyond the approximation. In the early , further contributions built on Mayer's framework by connecting the cluster integrals to f-functions—defined as f_{ij} = e^{-\beta u(r_{ij})} - 1, where u is the pair potential and \beta = 1/[kT](/page/KT)—which facilitated approximations in solving the Boltzmann transport equation for dilute gases with interactions. These developments, detailed in Mayer's subsequent papers and the 1940 book, established the expansion as a tool for handling irreducible diagrams representing correlated particle configurations, laying the groundwork for its broader use in . During the 1950s and 1960s, George E. Uhlenbeck and collaborators formalized the cluster expansion within , refining its application to both classical and quantum systems and highlighting its diagrammatic representation. A key advancement was the explicit linkage to through Ursell functions, which express connected correlations as sums over labeled graphs, providing a combinatorial interpretation that simplified computations of higher-order terms. This period saw the method gain prominence for analyzing models and fluids, with Uhlenbeck's lectures emphasizing its role in deriving thermodynamic properties from microscopic interactions. Rigorous mathematical foundations were established in the 1960s, particularly through the work of David Ruelle and , who proved of the cluster expansion for classical systems under suitable conditions on the interaction potentials. Ruelle's 1963 analysis provided the first rigorous theorem for short-range potentials, establishing analyticity of the in the for low densities and bounded interactions. Penrose extended these results to repulsive potentials, confirming in the and enabling reliable predictions for phase transitions in continuous systems. These proofs solidified the cluster expansion as a cornerstone of rigorous , distinguishing it from earlier formal series expansions.

Classical formulation

General theory

The classical cluster expansion provides a formal power series representation for thermodynamic quantities in systems of interacting particles, such as real gases, derived from the . The logarithm of the grand partition function \Xi for a classical system of particles with pairwise interactions is expressed as \log \Xi = \sum_{l=1}^\infty \frac{z^l}{l!} \int b_l(\mathbf{r}_1, \dots, \mathbf{r}_l) \, d\mathbf{r}_1 \cdots d\mathbf{r}_l, where z = e^{\beta \mu}/\Lambda^d is the (\mu the , \Lambda the thermal wavelength, d the spatial dimension, \beta = 1/kT), and b_l are the cluster coefficients that encode the irreducible contributions from l-particle configurations. This expansion, known as the Ursell-Mayer expansion, arises from decomposing the full function into sums over connected graphs, where each cluster coefficient b_l corresponds to the sum over all connected, irreducible diagrams representing the interactions among l particles. The Ursell functions, introduced earlier for the , provide the connected correlations, while Mayer's formulation extends this to the grand canonical setting by incorporating the and focusing on irreducible clusters to ensure convergence for low densities. Central to the theory are the correlation functions: the total correlation function h(\mathbf{r}) = g(\mathbf{r}) - 1, where g(\mathbf{r}) is the , and the c(\mathbf{r}), which captures irreducible direct interactions. These are related through the Ornstein-Zernike , h(\mathbf{r}) = c(\mathbf{r}) + \rho \int c(\mathbf{r}') h(|\mathbf{r} - \mathbf{r}'|) \, d\mathbf{r}', where \rho is the and the integral denotes ; this equation decomposes correlations into direct and indirect (chain-like) contributions, with the cluster expansion providing the diagrammatic basis for approximating c(\mathbf{r}). The pressure P is directly obtained from the expansion as \frac{P}{kT} = \sum_{l=1}^\infty b_l z^l, with the lowest-order coefficients given by b_1 = 1 and b_2 = -\frac{1}{2}\int f(\mathbf{r}) \, d\mathbf{r}, where f(\mathbf{r}) = e^{-\beta u(\mathbf{r})} - 1 is the Mayer f-function, representing the deviation from ideal-gas behavior due to the pairwise potential u(\mathbf{r}). Higher-order b_l involve integrals over products of f-functions summed over connected graphs.

Cluster integrals and diagrams

In the classical formulation of cluster expansions, Mayer cluster diagrams provide a graphical representation of the integrals arising in the expansion of the configuration integral. These diagrams consist of graphs where vertices correspond to particles, and bonds represent the Mayer f-functions, defined as f(\mathbf{r}_i - \mathbf{r}_j) = e^{-\beta u(\mathbf{r}_i - \mathbf{r}_j)} - 1, with u denoting the pairwise interaction potential and \beta = 1/(kT). The graphs can be labeled, where vertices are distinguished by particle indices, or unlabeled, focusing on topological structure to account for symmetries. This graphical approach facilitates the systematic enumeration of contributions to the partition function by summing over all possible bond configurations. The cluster integral b_l for l particles is given by b_l = \frac{1}{l!} \sum \int \prod_{\text{bonds } (ij)} f(\mathbf{r}_i - \mathbf{r}_j) \, d\mathbf{r}_1 \cdots d\mathbf{r}_l, where the sum is over all connected unlabeled graphs with l vertices, ensuring the integral captures only irreducible contributions from interacting clusters. Connected graphs are those that cannot be partitioned into disjoint subsets without breaking bonds, distinguishing irreducible (fully connected) clusters from reducible ones, which may factor into separate components. Rooted clusters designate one vertex as a reference point, useful for certain expansions, while unrooted clusters treat all vertices equivalently. For l=2, the integral simplifies to b_2 = -\frac{1}{2} \int f(\mathbf{r}) \, d\mathbf{r}, corresponding to a single bond diagram. For higher l, such as l=3, it includes all connected topologies like chains and triangles, summing their weighted integrals. These cluster integrals directly relate to the virial coefficients in the expansion of the , where the l-th virial coefficient satisfies B_l = -(l-1) b_l. This connection arises from the combinatorial structure of the diagrams, linking the grand canonical expansion to the virial series for the equation of state. The negative sign reflects the typical behavior of f-functions for repulsive potentials, ensuring physical consistency in low-density limits.

Configuration integral computation

The configuration integrals in classical cluster expansions, particularly the Mayer cluster integrals b_l, are evaluated using direct analytical methods for low orders when the interparticle potential is spherical. For the second-order integral b_2, which corresponds to the second virial coefficient, the computation reduces to a one-dimensional over the radial Mayer function f(r) = e^{-\beta u(r)} - 1, where u(r) is the pairwise potential and \beta = 1/[kT](/page/KT); this yields b_2 = -\frac{1}{2V} \int f_{12} \, d\mathbf{r}_1 d\mathbf{r}_2, analytically solvable for potentials like Lennard-Jones or . For the third-order b_3, the expression involves a threefold that can be simplified using spherical and symmetry reductions, allowing analytical evaluation for simple potentials, though it requires more computational effort than b_2. For higher-order integrals where direct integration becomes infeasible due to increasing complexity, methods such as Mayer sampling are employed. This technique uses and to efficiently sample configurations weighted by the product of Mayer functions, enabling accurate computation of b_l up to order 10 or higher by avoiding uniform sampling in high-dimensional spaces. These methods bridge the gap between analytical tractability and numerical precision, particularly for realistic potentials. To extend the utility of the finite series of computed cluster integrals beyond perturbation theory, resummation techniques like Padé approximants are applied to the virial expansion. These rational approximations interpolate between low- and high-density regimes, improving convergence and providing reliable equations of state from limited b_l values, as demonstrated in analyses of simple fluid expansions. A notable example is the computation of cluster integrals for hard-sphere systems, where exact low-order values are known, but higher b_l are approximated using scaled particle theory (SPT). SPT derives an equation of state by considering the work to insert a scaled hard sphere, reproducing exact b_2 and b_3 while providing estimates for higher orders that align well with numerical results up to moderate densities. Despite these advances, numerical challenges persist for high-order integrals, primarily due to the in dimensionality—the l-body spans $3l dimensions minus symmetries—leading to a computational cost that scales factorially with l. This limits practical computations to l \leq 10 for general potentials, restricting accurate virial expansions to low densities where higher terms are negligible.

Extensions and variants

High-temperature expansions

High-temperature expansions adapt cluster expansion techniques to discrete models, such as the , by expressing the partition function as a in the inverse β for small β (high temperatures). This method facilitates the computation of thermodynamic properties like the , specific heat, and through sums over lattice clusters or graphs, providing insights into phase transitions and critical behavior. Unlike continuous-space virial expansions, these series focus on bond and site configurations on lattices, using variables that capture short-range correlations at high temperatures. The formulation begins with the partition function for an Ising-like model with Hamiltonian H = -∑{} J{ij} σ_i σ_j (no external field), where σ_i = ±1 are spins on lattice sites. Each interaction term expands as exp(β J_{ij} σ_i σ_j) = cosh(β J_{ij}) [1 + σ_i σ_j tanh(β J_{ij})]. Thus, Z = ∑{σ} ∏{} cosh(β J_{ij}) [1 + σ_i σ_j v_{ij}], with v_{ij} = tanh(β J_{ij}) serving as the high-temperature bond variable for edge expansions. Expanding the product yields a sum over subgraphs G of the lattice, where each term corresponds to selecting bonds for the v_{ij} factors; the spin sum contributes 2^N only for subgraphs where every vertex has even degree (closed loops or clusters), and zero otherwise. The resulting expression is Z = 2^N ∏{} cosh(β J{ij}) ∑{G \in \mathcal{E}} ∏{(i,j) \in G} v_{ij}, where \mathcal{E} is the set of even-degree subgraphs. For uniform J, this simplifies to Z = 2^N [cosh(β J)]^{b} ∑_{G \in \mathcal{E}} [tanh(β J)]^{|E(G)|}, with b the number of bonds. To obtain the free energy, the logarithm is taken, leading to a linked-cluster expansion that isolates connected contributions: \frac{1}{N} \log Z = \log 2 + \frac{b}{N} \log \cosh(\beta J) + \sum_{k=1}^\infty u_k [\tanh(\beta J)]^k, where the coefficients u_k are high-temperature cluster sums over connected even subgraphs (irreducible clusters) of k bonds, weighted by their lattice embeddings and symmetry factors. For general small β, an equivalent series is \frac{1}{N} \log Z = \sum_{k=0}^\infty u_k \beta^k, with u_k the high-T cluster sums incorporating powers of J. These u_k are computed via linked-cluster theorems, ensuring only connected diagrams contribute to the logarithm, analogous to diagrammatic expansions in field theory. Series coefficients are generated using recursive algorithms that enumerate and weight clusters up to high orders, often exceeding 20; for example, computations reach 25th order on body-centered cubic lattices. techniques supplement for low-dimensional cases, but recursive methods dominate for higher dimensions due to their efficiency in handling symmetries. High-temperature series were first developed by van der Waerden in for the partition function via even-subgraph sums, but cluster methods were formalized in the 1950s by Domb and Sykes for extracting from the series, with Fisher extending them in the to link Ising correlations to self-avoiding walks and precise exponent estimates.

Quantum cluster expansions

Quantum cluster expansions extend the classical cluster formalism to , accounting for the wave nature of particles and quantum statistics. Developed by Kahn and Uhlenbeck in for imperfect quantum gases, this approach derives an expansion for the partition function that parallels the Mayer expansion but incorporates quantum effects through symmetrized wave functions. The method applies equally to bosons and fermions, enabling the study of phenomena like Bose-Einstein condensation in dilute gases. In the quantum case, the grand partition function \Xi is given by the trace over the Fock space: \Xi = \mathrm{Tr} \left[ \exp\left( -\beta (H - \mu N) \right) \right], where H is the many-body , \mu the , N the particle number , and \beta = 1/(kT). This can be expanded using symmetrized basis states, such as Slater determinants for fermions or permanent determinants for bosons, leading to integrals that capture irreducible contributions from groups of particles. Alternatively, representations facilitate the expansion by integrating over closed particle trajectories, which naturally incorporate quantum statistics. A primary distinction from classical cluster expansions lies in the inclusion of exchange effects arising from the indistinguishability of particles, enforced by symmetrization or antisymmetrization of the wave functions. For identical fermions, the introduces negative contributions from terms, while for bosons, positive reinforcements occur, altering the and structure of the series compared to the classical Boltzmann . These quantum corrections are essential for low-temperature or high-density regimes where de Broglie wavelengths overlap. Semiclassical approximations, such as the Wigner-Kirkwood expansion, provide systematic corrections to the classical cluster integrals by perturbing in powers of \hbar. Introduced by Wigner in 1932 and extended by Kirkwood in 1935, this method expands the quantum phase-space distribution in \hbar, yielding the quantum cluster coefficient b_l^Q as: b_l^Q = b_l^C + \hbar^2 \, c_l + O(\hbar^4), where b_l^C is the classical cluster integral, and the leading \hbar^2 correction c_l involves kinetic energy operators, such as Laplacians acting on the potential, averaged over classical configurations. Higher-order terms account for anharmonicity and non-perturbative effects but typically diverge beyond a few orders. This expansion is particularly useful for weakly quantum systems, bridging classical and full quantum treatments. Applications of quantum cluster expansions are prominent in dilute Bose and Fermi gases, where they quantify finite-size effects and virial coefficients beyond mean-field approximations, as demonstrated in trapped ultracold atomic systems.

Applications

Lattice gas and Ising models

The lattice gas model, which models hard-core particles on a with nearest-neighbor interactions and no double occupancy, is formally equivalent to the through a between variables and variables. Specifically, the occupation number n_i = 0 or $1 at site i maps to the spin \sigma_i = \pm 1 via \sigma_i = 1 - 2 n_i, transforming the grand canonical function of the lattice gas into the function of the in a field analogous to a . This equivalence, first highlighted in studies of phase transitions, implies that fluid-like density fluctuations in the lattice gas correspond directly to magnetic ordering in the , enabling shared analytical tools for both systems. Cluster expansions applied to these models yield perturbative series for observables like the magnetization m = \langle \sigma_i \rangle in the Ising formulation or the density in the lattice gas. For the Ising model in a uniform magnetic field h, the high-temperature cluster expansion expresses the magnetization as m = \tanh(\beta h) + contributions from higher-order spin clusters, where \beta = 1/(k_B T) and the cluster terms account for interaction graphs connecting multiple sites. These expansions, derived from diagrammatic representations of the partition function, systematically incorporate the effects of lattice connectivity and interaction strength, providing insights into response functions near equilibrium. High-temperature expansions, a common implementation of this approach, facilitate numerical evaluation of the series up to high orders for lattice-specific computations. In the context of critical phenomena, cluster expansions have proven essential for estimating phase transition points in lattice gas and Ising systems, where spontaneous symmetry breaking signals the onset of ordering. For the two-dimensional square-lattice Ising model without an external field, high-temperature series derived from cluster sums yield a critical temperature T_c \approx 2.269 J / k_B, where J is the coupling constant, aligning closely with Onsager's exact analytic solution T_c = 2 J / [k_B \ln(1 + \sqrt{2})] \approx 2.269 J / k_B. This agreement validates the method's accuracy for classical lattice models, allowing extrapolation to critical exponents and susceptibility divergences that characterize the transition. Cluster expansions also integrate with renormalization group theory, treating spin clusters as fundamental units that coarse-grain the lattice to reveal fixed points dictating universal critical behavior in Ising and lattice gas models. In this framework, iterative rescaling of cluster interactions flows toward the critical fixed point, unifying short-range correlations captured by expansions with long-range scaling properties observed at the transition.

Real gases and fluids

Cluster expansions provide a foundational framework for analyzing continuous systems of interacting particles, such as real gases and fluids, by expressing thermodynamic properties in terms of density expansions derived from irreducible cluster integrals. In these systems, the virial equation of state captures deviations from ideality due to pairwise and higher-order interactions, with the pressure P related to the number density \rho by \frac{P}{\rho k_B T} = 1 + B_2 \rho + B_3 \rho^2 + \cdots + B_l \rho^{l-1} + \cdots, where k_B is Boltzmann's constant, T is temperature, and the virial coefficients B_l (for l \geq 2) are computed from Mayer cluster integrals involving the configuration integrals of l particles with the Mayer f-functions f_{ij} = e^{-\beta u(r_{ij})} - 1, \beta = 1/(k_B T), and u(r_{ij}) the pair potential. These coefficients quantify the effective volume excluded or attraction induced by interactions, enabling predictions of compressibility and other properties at low to moderate densities where the series converges. A prototypical application is to hard-sphere fluids, which model the repulsive cores of simple liquids without attractive forces. The second virial coefficient B_2 is exactly B_2 = 4v, where v = \frac{\pi \sigma^3}{6} is the volume of a of \sigma. The third virial coefficient B_3 is also exactly computable via three-body diagrams, yielding B_3 = 10 v^2 in three dimensions. Higher-order coefficients B_l (for l > 3) become increasingly complex, but the Percus-Yevick approximation to the Ornstein-Zernike provides a systematic way to estimate them all, producing values like B_4 \approx 18.365 v^3 that closely match numerical results from simulations up to the fifth or sixth order. The foundational application to real gases was advanced by Joseph E. Mayer and in their 1940 analysis of imperfect gases, where cluster expansions were used to compute virial coefficients for like , incorporating realistic pair potentials such as the Lennard-Jones form to fit experimental second and third virials. This work demonstrated how cluster methods bridge microscopic interactions to macroscopic observables, such as the equation of state for nearly ideal systems with weak attractions. In contemporary models, cluster expansion principles are resummed in the Statistical Associating Fluid Theory (SAFT), which extends the hard-sphere reference via to handle chain connectivity, polar groups, and associations in complex fluids, improving predictions of phase behavior for hydrocarbons and polymers. Beyond equations of state, cluster expansions predict liquid-vapor transitions in fluids by summing virial series to identify coexistence conditions, often via the equal-area applied to partial sums or Padé that analytically continue the expansion. For instance, in square-well or Lennard-Jones potentials, truncated cluster sums locate binodals by equating chemical potentials and pressures between phases. The of the virial series, marking the density beyond which the expansion diverges, corresponds to the spinodal line where the isothermal \kappa_T = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right)_T diverges, providing an estimate of the metastable limit for superheated vapors or supercooled liquids in phase diagrams.

Convergence and rigor

Proofs of convergence

Ruelle's theorem establishes the convergence of the cluster expansion for classical particle systems with short-range, stable interactions. It proves of the Mayer series for the and functions when the activity satisfies |z| < R, where the radius R is determined by the interaction range and the inverse temperature β, specifically bounded below by an explicit function of the potential's constant and β. In the 1980s, Brydges, Fröhlich, and Spencer introduced an inductive method for proving , relying on tree-like graphs to organize the expansion and combinatorial estimates to control error terms. This approach provides rigorous bounds for the convergence radius in gas models and systems, improving upon earlier techniques by handling more general interaction structures through recursive on size. A central condition ensuring for short-range potentials is that the sum over sites r of |f(r)| < 1/e, where f(r) denotes the Mayer f-function associated with the two-body interaction. Under this condition, the integrals satisfy the bound |b_l| \leq \frac{ (c)^l }{ l! }, for some constant c depending on the model parameters, which implies exponential decay in l and guarantees the series sums to a finite value. These classical convergence proofs have been extended to quantum settings, particularly for quasi-free fields in constructive quantum field theory, where Glimm and Jaffe adapted cluster expansions to establish the existence of Euclidean field theories with polynomial interactions.

Limitations and modern developments

Cluster expansions, while powerful for perturbative regimes, exhibit slow convergence near critical points where correlation lengths diverge, limiting their accuracy for systems approaching phase transitions. This slowdown arises because higher-order clusters become increasingly important as interactions strengthen, requiring more terms for reliable approximations in low-temperature or high-density conditions. Furthermore, the series often diverges at phase transitions, where discontinuities in thermodynamic quantities like or occur, rendering the expansion inadequate for capturing the non-analytic behavior. In applications to fluids, cluster expansions break down in high-density regimes, manifesting as unphysical oscillations akin to the near discontinuities in the equation of state. For instance, virial expansions for dense gases show erratic behavior when extrapolated beyond the , necessitating alternative formulations with reducible cluster integrals to stabilize results. Modern developments address these limitations through resummation techniques, such as Padé-Borel methods, which accelerate by approximating the Borel transform of the divergent series and yield reliable estimates even near the convergence boundary. Post-2010 advancements incorporate to predict cluster expansion coefficients, enabling efficient modeling of complex multicomponent alloys by training on data to identify effective interactions. In the 2020s, approaches have enhanced expansions for two-dimensional , providing exponentially convergent representations that handle entanglement and correlations more effectively than traditional methods. Integration with has also advanced, allowing hybrid schemes where DFT supplies accurate energies for fitting, as implemented in tools like the CELL package for high-throughput materials discovery. Extensions to disordered systems and long-range interactions leverage Dobrushin-Shlosman techniques, which ensure uniqueness of Gibbs measures via graded expansions with diverging minimal scales near coexistence lines.

References

  1. [1]
    [PDF] The Cluster Expansion
    Simply stated, the cluster expansion provides a method for studying the loga- rithm of a partition function. We will use it in various situations, for instance ...
  2. [2]
  3. [3]
    Physics The Cluster Expansion in Statistical Mechanics*
    Abstract. The Glimm-Jaffe-Spencer cluster expansion from constructive quantum field theory is adapted to treat quantum statistical mechanical.
  4. [4]
    [PDF] Statistical Mechanics I - MIT OpenCourseWare
    V.B The Cluster Expansion. For short range interactions, specially with a hard core, it is much better to replace the expansion parameter V( q) by f(q ) = exp ( ...
  5. [5]
    [PDF] The influence of accidental deviations of density on the equation of ...
    ORNSTEIN and DJ', F. ZERNIKE. (Communicated by Prof. H. A. LORENTZ). (Communicated in the meetin« or June 24, 1916). In thei!' article in the Encyclopädie ...
  6. [6]
    None
    ### Historical Background on Cluster Expansions
  7. [7]
    [PDF] 100 Years of the (Critical) Ising Model on the Hypercubic Lattice
    The Ising model, introduced by Lenz in 1920, is a ferromagnetic nearest-neighbor model on a hypercubic lattice, where elementary magnets take only two opposite ...
  8. [8]
    [cond-mat/0005130] Linked-Cluster Expansion of the Ising Model
    May 8, 2000 · The linked-cluster expansion technique for the high-temperature expansion of spin model is reviewed. A new algorithm for the computation of three-point and ...
  9. [9]
    [PDF] ON THE THEORY OF CONDENSATION
    Abstract. A discussion is given of Mayer's theory of the condensation phenome- non. It is shown that the theory is not restricted to classical statistical.
  10. [10]
  11. [11]
    [PDF] 1 Unit 4-6: The Liquid-Gas Phase Transition The Lattice Gas Model ...
    Jun 4, 2021 · We can now map the lattice gas model onto the Ising model. Let si = 2ni − 1 = ±1, ni = si + 1. 2. (4.6.3). Then. H = −U X. ⟨ij⟩ si + 1. 2 sj + 1.
  12. [12]
    [PDF] IL Lattice Gas and Ising Model
    Ising model in a magnetic Geld and its relationship with the distribution of roots of the partition function. II. ISING MODEL AND LATTICE GAS. We shallin ...
  13. [13]
    Cluster Expansion for the Ising Model - AIP Publishing
    By performing a partial summation, one finds that the free energy is given in terms of connected diagrams which are composed of βJ bonds, and black circles Mn.
  14. [14]
  15. [15]
    [PDF] A SHORT COURSE ON CLUSTER EXPANSIONS* - UBC Mathematics
    These lectures are an attempt to provide some overview and review of the various kinds of expansions-Mayer expansions, high-temperature ex- pansions, expansions ...
  16. [16]
    [PDF] Constructive Quantum Field Theory - Arthur Jaffe
    The cluster expansion selects one of these vacuum states by imposing the value of the field on the boundary of a large box, before taking the infinite volume ...
  17. [17]
    [PDF] THE NATURE OF CRITICAL POINTS* Michael E. Fishert The ...
    For Ising model expansions at low-temperatures and high fields (which ... The high-temperature specific heat series for the Ising model are relatively ...
  18. [18]
    Critical point signatures in the cluster expansion in fugacities
    Jan 22, 2020 · The QCD baryon number density can formally be expanded into a Laurent series in fugacity, which is a relativistic generalization of Mayer's cluster expansion.
  19. [19]
    Divergence of activity expansions: Is it actually a problem?
    Dec 13, 2017 · Even in the divergence region (at subcritical temperatures), this behavior stays thermodynamically adequate (in contrast to the behavior of the ...
  20. [20]
    Equation of state for all regimes of a fluid: From gas to liquid
    Sep 26, 2018 · The only possibility to make the CE behavior adequate in high-density regimes is to use the equations in terms of reducible cluster integrals as ...
  21. [21]
    Revisiting Groeneveld's approach to the virial expansion |
    Feb 22, 2021 · The best-known proofs of convergence for the virial expansion are based on an inversion. First, one proves convergence of the activity ...
  22. [22]
    Machine-learning the configurational energy of multicomponent ...
    Nov 1, 2018 · The cluster expansion formalism sets up a natural mathematical framework with which to represent the properties of a crystal as a function of ...Missing: post- | Show results with:post-
  23. [23]
    CELL: a Python package for cluster expansion with a focus ... - Nature
    Aug 30, 2024 · The ab initio energies are calculated by density-functional theory (DFT) with the all-electron, full-potential electronic-structure code FHI- ...Methods · Cluster Expansion Method · Cluster Expansion Of A...