Fact-checked by Grok 2 weeks ago

Ising model

The Ising model is a foundational in used to describe and phase transitions in magnetic systems. It consists of a of discrete sites, each occupied by a that can take one of two values, typically denoted as s_i = \pm 1, representing the two possible orientations of atomic magnetic dipoles. The interactions between nearest-neighbor spins are modeled by the H = -J \sum_{\langle i,j \rangle} s_i s_j - h \sum_i s_i, where J > 0 is the ferromagnetic favoring , h is an external , and the sums run over nearest-neighbor pairs \langle i,j \rangle and all sites i, respectively. This energy function captures the competition between thermal disorder and ordered magnetic , leading to phenomena such as below a critical . The model originated from efforts to explain ferromagnetic properties theoretically. In 1920, Wilhelm Lenz proposed the idea of a lattice of interacting as a simplified representation of atomic magnetism in solids. His student, , analyzed the one-dimensional version in his 1925 doctoral thesis, deriving an exact solution for the partition function and demonstrating that no occurs at finite in one dimension—a result that initially cast doubt on the model's applicability to real three-dimensional ferromagnets. Despite this, the model's simplicity and solvability revived interest in the 1940s; in 1944, provided the exact solution for the two-dimensional Ising model without an external field, revealing a second-order at a critical T_c = \frac{2J}{k \ln(1 + \sqrt{2})}, where k is Boltzmann's constant, and computing key thermodynamic quantities like the spontaneous magnetization and specific heat. The Ising model's enduring importance lies in its role as a paradigm for studying critical phenomena and universality in phase transitions across diverse fields, including condensed matter physics, materials science, and even social sciences for modeling opinion dynamics. While exactly solvable in one and two dimensions, the three-dimensional case remains unsolved analytically, relying on numerical methods, series expansions, and approximations like mean-field theory, which predict a phase transition but overestimate the critical temperature. Extensions to frustrated lattices, quantum versions, and disordered systems have further broadened its applications, influencing breakthroughs in renormalization group theory and computational algorithms for optimization problems.

Definition and Formulation

Hamiltonian

The Hamiltonian of the Ising model specifies the total energy of a system of interacting magnetic spins arranged on a . It captures the essential physics of through pairwise interactions between spins and an optional coupling to an external . The model assumes spin variables that can point in one of two directions, typically representing atomic magnetic moments. The standard mathematical expression for the is H = -J \sum_{\langle i,j \rangle} \sigma_i \sigma_j - h \sum_i \sigma_i, where \sigma_i = \pm 1 denotes the spin at lattice site i, the sum \langle i,j \rangle runs over all pairs of nearest-neighbor sites, J is the coupling constant measuring the strength of the spin-spin interaction, and h is the external magnetic field strength. This form, with nearest-neighbor interactions, was proposed by Wilhelm Lenz in 1920 and formalized by Ernst Ising in 1925 to explain spontaneous magnetization in ferromagnets. Physically, the interaction term -J \sum_{\langle i,j \rangle} \sigma_i \sigma_j favors alignment of neighboring when J > 0 (ferromagnetic case), as parallel (\sigma_i = \sigma_j) lower the compared to antiparallel ones. The term -h \sum_i \sigma_i couples each to the external , biasing configurations where align with h. At zero and h = 0, the minimizes H with all uniformly aligned, either all up or all down, yielding -J z N / 2 where z is the coordination number and N the number of sites. If h \neq 0, the has all aligned with the . These interpretations stem directly from the model's design to mimic atomic dipole interactions in . Within statistical mechanics, the Hamiltonian enters the partition function via the Boltzmann weight, defining the equilibrium statistical properties of the system: Z = \sum_{\{\sigma\}} \exp(-\beta H), where the sum is over all $2^N possible spin configurations, and \beta = 1/(k_B T) with k_B Boltzmann's constant and T the temperature. This expression, rooted in the canonical ensemble, enables computation of averages like magnetization \langle m \rangle = (1/N) \partial \ln Z / \partial (\beta h) and free energy F = -k_B T \ln Z. Ising applied this framework in his exact solution for the one-dimensional case, demonstrating no phase transition at finite temperature. The formulation employs units where J and h have dimensions of energy, consistent with thermal energies k_B T. The nearest-neighbor restriction simplifies the model while capturing effects, and the \sigma_i = \pm 1 convention normalizes the spin magnitude to unity, focusing on directional .

Lattice and Spin Configurations

The Ising model is formulated on a spatial consisting of sites arranged in a regular structure, where each site hosts a representing a microscopic . Originally proposed by Wilhelm Lenz in 1920 as a model for involving dipoles fixed on a that can in two directions, the framework was analyzed by in 1925 for the one-dimensional case. In general, the is taken to be a d-dimensional hypercubic structure, such as a linear chain in one dimension (d=1), a square grid in two dimensions (d=2), or a cubic array in three dimensions (d=3), with spins located at the vertices of the . This hypercubic geometry ensures translational invariance and uniform nearest-neighbor , making it a choice for studying phenomena in systems. Each site i is associated with a variable \sigma_i \in \{+1, -1\}, where +1 and -1 denote the two opposing directions of the 's magnetic orientation, analogous to up or down states in a simplified . For a finite with N sites, the system admits exactly $2^N possible configurations, as each can independently adopt one of the two states, forming the configuration space of the model. Interactions in the model occur exclusively between nearest-neighbor pairs of sites, denoted by \langle i,j \rangle, which are adjacent sites connected by the 's edges—such as along the coordinate axes in the hypercubic case. Boundaries can be implemented as open (free edges) or periodic ( wrapping), though the standard setup assumes a large to approximate bulk behavior. To illustrate, consider a one-dimensional (1D) chain lattice with N=4 sites labeled sequentially as 1, 2, 3, 4. The nearest-neighbor pairs are \langle 1,2 \rangle, \langle 2,3 \rangle, and \langle 3,4 \rangle, with possible spin configurations including all-up (+1,+1,+1,+1) or alternating (+1,-1,+1,-1). In two dimensions (2D), a square lattice with, say, 2×2=4 sites forms a grid where each interior site (in larger grids) has four neighbors: horizontal and vertical adjacents, such as site (1,1) connected to (2,1) and (1,2). Configurations here might feature ferromagnetic alignment, like all spins +1 across the grid, highlighting the role of lattice connectivity in aligning spins. These examples underscore how the lattice topology dictates the interaction network, central to the model's physical interpretation.

Boundary Conditions and Variants

In the Ising model, boundary conditions specify how spins at the edges of a finite interact, influencing thermodynamic properties, particularly in low dimensions where surface effects are prominent. Periodic boundary conditions identify opposite edges of the , effectively forming a (or more generally, a higher-dimensional analogue), which minimizes finite-size artifacts and approximates infinite-volume behavior for bulk properties. This setup is standard for exact solutions, such as Onsager's derivation of the two-dimensional partition function, where the geometry ensures translational invariance. Open or free boundary conditions, in contrast, impose no interactions across lattice edges, leading to unpaired at the surface and the emergence of surface terms in the total . These terms contribute corrections of order O(1/L) to thermodynamic quantities like the specific heat or in a system of linear size L, altering finite-size scaling near criticality; for instance, in the two-dimensional critical Ising model on triangular , shape-dependent surface contributions shift the effective . Fixed boundary conditions fix to specific values, such as all +1 (positive or "+" ) or alternating signs, which is useful for studying domain walls or interfaces; this introduces an effective surface field, enhancing ordering near the and modifying lengths. Variants of the Ising model extend the standard Hamiltonian by incorporating additional interactions or disorder, often to model real materials like anisotropic magnets or alloys. The inclusion of an external magnetic field h modifies the Hamiltonian to H = -\sum_{\langle i,j \rangle} J s_i s_j - h \sum_i s_i, breaking the \mathbb{Z}_2 symmetry and suppressing spontaneous magnetization below the critical temperature; this form is central to mean-field analyses and Lee-Yang circle theorem studies of phase transitions. Anisotropic couplings, where horizontal interactions differ from vertical ones (J_x \neq J_y), yield a rectangular lattice model solvable exactly in two dimensions via transfer-matrix methods, with the critical temperature satisfying \sinh(2\beta J_x) \sinh(2\beta J_y) = 1, generalizing Onsager's isotropic result. The dilute Ising model introduces quenched through random , where each J_{ij} is J with probability p (bond occupancy) and zero otherwise, modeling diluted ferromagnets; this leads to thresholds for connectivity and modified critical behavior, with the Harris —which states that the clean fixed point is stable against weak quenched if d ν > 2, where ν is the correlation length exponent of the clean system—indicating that is relevant in three dimensions for the Ising model (since 3 × 0.63 ≈ 1.89 < 2). In finite systems with open boundaries, such variants amplify surface effects, as missing at edges exacerbate imbalance in coordination numbers, contributing additional O(L^{d-2}) surface terms to the free energy and altering crossover scaling at first-order transitions. Beyond lattices, the Ising model connects to graph theory, where spins reside on vertices of an arbitrary graph G = (V, E) with Hamiltonian H = -\sum_{(i,j) \in E} J_{ij} s_i s_j; for the antiferromagnetic case on an arbitrary graph, the Hamiltonian is often written as H = \sum_{(i,j) \in E} s_i s_j (setting J=1), whose ground-state minimization maps directly to the , with energy E = |E| - 2 \times (\text{cut size}), linking statistical mechanics to NP-hard optimization.

Historical Development

Origins and Early Work

The roots of the Ising model trace back to early 20th-century efforts to understand ferromagnetism through mean-field approximations, notably Pierre Weiss's 1907 theory, which posited an internal "molecular field" acting on atomic magnetic moments to explain the Curie law and spontaneous magnetization below a critical temperature. Weiss's approach treated the material as a collection of independent dipoles influenced by an average field from neighbors, providing a phenomenological framework but lacking a detailed microscopic description of interactions. In 1920, Wilhelm Lenz proposed a discrete lattice model for ferromagnetism, envisioning atoms as magnetic dipoles fixed at lattice sites that could orient in only two directions, up or down, with nearest-neighbor interactions favoring alignment. Lenz assigned this problem to his graduate student Ernst Ising as a thesis topic, aiming to derive ferromagnetic behavior from statistical mechanics while simplifying the complex quantum mechanical treatment of electron spins and exchange interactions prevalent in contemporary atomic physics. Ernst Ising completed his doctoral thesis in 1924 and published his results in 1925, solving the one-dimensional version of the model exactly and demonstrating that no phase transition occurs at finite temperatures, contrary to experimental observations of ferromagnetism. This early work highlighted the model's computational tractability as a classical approximation to more intricate quantum models, though its limitation to one dimension underscored the need for higher-dimensional analyses to capture real material behavior.

Key Theoretical Advances

In 1936, Rudolf Peierls provided the first rigorous proof of a phase transition in the two-dimensional using a contour argument, demonstrating that spontaneous magnetization persists at low temperatures by bounding the free energy cost of domain walls or "droplets" that disrupt ferromagnetic order. This approach highlighted the instability of disordered configurations below a critical temperature, establishing the model's relevance for understanding spontaneous symmetry breaking in lattice systems. Building on this, Hendrik Kramers and Gregory Wannier introduced a duality in 1941 that maps the high-temperature expansion of the two-dimensional square-lattice to its low-temperature counterpart, revealing a self-dual point that precisely locates the critical temperature where the two phases meet. This symmetry not only simplified the analysis of thermodynamic properties but also underscored the model's exact solvability potential. The culmination of these efforts came in 1944 with Lars Onsager's exact solution for the two-dimensional without an external field, yielding the partition function and critical temperature T_c = \frac{2J}{k \ln(1 + \sqrt{2})}, where J is the coupling constant and k is Boltzmann's constant. Onsager's transfer-matrix method computed the free energy analytically, confirming Peierls' phase transition and providing explicit expressions for spontaneous magnetization and specific heat, which diverges logarithmically at T_c. Further theoretical progress in the 1950s included the Yang-Lee theorem, proved in 1952, which shows that the zeros of the partition function for ferromagnetic lie on the unit circle in the complex fugacity plane, implying the absence of phase transitions in finite systems and analyticity in the thermodynamic limit away from criticality. This circle theorem generalized earlier insights into the distribution of partition function roots, linking them to the stability of ferromagnetic order. By 1967, Robert B. Griffiths established key correlation inequalities for Ising ferromagnets, such as the positive association of spin correlations—specifically, that the expectation of the product of spins at distinct sites is greater than or equal to the product of their individual expectations—providing bounds on fluctuations and reinforcing the model's monotonic response to temperature and field changes. These inequalities, derived from the positive definiteness of the interaction matrix, have proven foundational for proving absence of phase transitions in one dimension and analyzing higher-dimensional behaviors.

Basic Properties

Absence of Phase Transition in One Dimension

The one-dimensional , consisting of spins arranged on a linear chain with nearest-neighbor interactions, exhibits no phase transition at any finite temperature in the absence of an external magnetic field. This result follows from the exact solution of the model, which reveals that the magnetization vanishes for all temperatures above absolute zero, and thermodynamic quantities remain analytic without singularities indicative of a phase transition. The partition function for a chain of N spins with periodic boundary conditions and zero external field is given exactly by Z = \lambda_+^N + \lambda_-^N, where the eigenvalues of the transfer matrix are \lambda_+ = 2 \cosh(\beta J) and \lambda_- = 2 \sinh(\beta J), with \beta = 1/(k_B T) and J > 0 the ferromagnetic coupling strength. In the N \to \infty, the partition function is dominated by the largest eigenvalue, yielding Z \approx [2 \cosh(\beta J)]^N. The corresponding per spin is then f = -k_B T \ln[2 \cosh(\beta J)], which is smooth and differentiable for all finite T > 0, showing no non-analyticity associated with a . The zero-field magnetization per spin, m = -\frac{1}{\beta N} \frac{\partial \ln Z}{\partial h} \big|_{h=0}, vanishes identically for all T > 0 due to the symmetry of the under global spin flip, precluding . This absence of long-range order is confirmed by the approach: the ratio of eigenvalues \lambda_- / \lambda_+ = \tanh(\beta J) < 1 for finite \beta J, ensuring that contributions from the symmetric and antisymmetric sectors do not support a nonzero ordered phase at finite temperature. Spatial correlations decay exponentially, with the two-point function \langle \sigma_i \sigma_j \rangle = [\tanh(\beta J)]^{|i-j|}, implying a finite correlation length \xi = -1 / \ln[\tanh(\beta J)]. This length diverges only as T \to 0, where \tanh(\beta J) \to 1, but remains finite at any nonzero temperature, further evidencing the lack of a critical point with diverging fluctuations. At low temperatures, \xi \approx \frac{1}{2} e^{2 \beta J}, highlighting exponentially long but still finite correlations without a singularity in the thermodynamic potentials.

Phase Transitions in Higher Dimensions

In dimensions d \geq 2, the Ising model exhibits a phase transition at a finite critical temperature T_c, separating a low-temperature ordered phase with ferromagnetic alignment from a high-temperature disordered paramagnetic phase. This order-disorder transition involves spontaneous symmetry breaking of the \mathbb{Z}_2 symmetry, where the spins collectively align in the absence of an external field below T_c. Unlike the one-dimensional case, where thermal fluctuations destroy long-range order at any positive temperature, the higher-dimensional lattices allow for stable ferromagnetic configurations due to the increased number of nearest-neighbor interactions that favor alignment. Spontaneous magnetization emerges below T_c in zero external field for d \geq 2, characterized by a non-zero average magnetization m = |\langle \sigma \rangle| > 0, which vanishes continuously as T \to T_c^- according to m \sim (T_c - T)^\beta. In two dimensions, this was exactly computed, yielding \beta = 1/8, while in three dimensions, \beta \approx 0.326, and for d > 4, applies with \beta = 1/2. The phase transition belongs to the Ising universality class, where describe the singular behavior of thermodynamic quantities near T_c, such as the correlation length diverging as \xi \sim |T - T_c|^{-\nu} with \nu = 1 exactly in 2D and \nu \approx 0.63 in 3D. These exponents are dimension-dependent up to the upper critical dimension d=4, above which mean-field values dominate due to the irrelevance of fluctuations. The Mermin-Wagner theorem prohibits spontaneous magnetization at finite temperatures in d \leq 2 for models with continuous symmetries, as long-wavelength fluctuations destroy long-range order; however, the discrete \mathbb{Z}_2 symmetry of the Ising model circumvents this restriction, permitting a in 2D. In higher dimensions, the theorem does not apply, and the transition persists with robust ordering. Numerical studies of finite lattices reveal finite-size effects, where observables like magnetization scale with system size L via finite-size scaling forms, such as m(L, T) = L^{-\beta/\nu} f((T - T_c) L^{1/\nu}), allowing extraction of and T_c from simulations without requiring the . This approach is essential for probing higher-dimensional systems where exact solutions are unavailable.

Correlation Functions and Inequalities

In the Ising model, the two-point quantifies the statistical dependence between at distinct sites. It is defined as the connected correlation G(\mathbf{i}, \mathbf{j}) = \langle \sigma_{\mathbf{i}} \sigma_{\mathbf{j}} \rangle - \langle \sigma_{\mathbf{i}} \rangle \langle \sigma_{\mathbf{j}} \rangle, where the angle brackets denote the thermal average with respect to the . In the disordered (high-temperature) , this function exhibits for large separations |\mathbf{i} - \mathbf{j}|, taking the form G(\mathbf{r}) \sim e^{-r / \xi} with r = |\mathbf{r}| and \xi > 0 the correlation length, which diverges as the critical temperature is approached from above. This decay reflects the finite range of spin correlations away from criticality and has been rigorously established for the ferromagnetic Ising model on \mathbb{Z}^d (d \geq 2) in the pure phases. A cornerstone inequality for understanding these correlations is the Fortuin–Kasteleyn–Ginibre (FKG) inequality, proved in 1971 for ferromagnetic models with positive interactions J > 0. It states that for any two coordinate-wise increasing functions f and g on the spin configurations, the thermal average satisfies \langle f g \rangle \geq \langle f \rangle \langle g \rangle. In the context of the Ising model, this implies positive correlations between spins: \langle \sigma_i \sigma_j \rangle \geq \langle \sigma_i \rangle \langle \sigma_j \rangle for all sites i, j, ensuring that ferromagnetic alignments are statistically favored. The extends to more general gases and has facilitated proofs of monotonicity in diagrams. Griffiths' second inequality, derived in 1967, further constrains correlations by establishing monotonicity with respect to interaction parameters. For the ferromagnetic , it asserts that the two-point function \langle \sigma_i \sigma_j \rangle is non-decreasing in each ferromagnetic coupling J_{kl} (for k, l \neq i, j), reflecting how strengthening interactions enhances alignment. This , part of a broader class introduced by Griffiths, applies to even-moment correlations and holds under pair interactions, providing bounds that strengthen the first Griffiths on positivity. The Simon-Lieb inequality offers a quantitative bound on two-point correlations, stating that for sites \alpha and \gamma separated by a cutset B of spins, |\langle \sigma_\alpha \sigma_\gamma \rangle| \leq \sum_{b \in B} \langle |\sigma_\alpha \sigma_b| \rangle \langle |\sigma_b \sigma_\gamma| \rangle. This path-like estimate, originally due to and refined by Lieb in , controls the propagation of correlations across the and implies under suitable conditions on the interactions. Such bounds are pivotal for analyzing decay rates beyond mean-field approximations. These inequalities play a key role in establishing long-range order in the low-temperature phase, where \lim_{|\mathbf{i}-\mathbf{j}| \to \infty} \langle \sigma_{\mathbf{i}} \sigma_{\mathbf{j}} \rangle > 0 despite \langle \sigma_{\mathbf{i}} \rangle = 0 by . The FKG and Griffiths inequalities underpin extensions of the Peierls , which bounds the probability of disordering contours to show non-vanishing below the critical temperature in dimensions d \geq 2. The Simon-Lieb bound further ensures that correlations do not decay too slowly, facilitating rigorous proofs of phase coexistence in ferromagnetic settings.

Exact Solutions

One-Dimensional Solution

The one-dimensional Ising model, consisting of a linear chain of N interacting spins \sigma_i = \pm 1 with nearest-neighbor ferromagnetic coupling J > 0 and an external h, is described by the H = -J \sum_{i=1}^N \sigma_i \sigma_{i+1} - h \sum_{i=1}^N \sigma_i, assuming \sigma_{N+1} = \sigma_1. This model admits a complete exact solution via the , which systematically computes the partition function and derives all thermodynamic properties. The method exploits the Markovian structure of the interactions, allowing the Boltzmann weights to be factored into products of matrices. The \mathbf{T} is a $2 \times 2 with elements determined by the possible configurations of adjacent sites: \mathbf{T} = \begin{pmatrix} e^{\beta (J + h)} & e^{-\beta J} \\ e^{-\beta J} & e^{\beta (J - h)} \end{pmatrix}, where \beta = 1/(k_B T) and k_B is Boltzmann's . The partition function Z for the chain is then given by Z = \mathrm{Tr}(\mathbf{T}^N), where \mathrm{Tr} denotes the . For large N, Z \approx \lambda_\mathrm{max}^N, with \lambda_\mathrm{max} the largest eigenvalue of \mathbf{T}, \lambda_\mathrm{max} = e^{\beta J} \cosh(\beta h) + \sqrt{ [e^{\beta J} \sinh(\beta h)]^2 + e^{-2 \beta J} }. The Helmholtz free energy per site in the thermodynamic limit is f = -k_B T \ln \lambda_\mathrm{max}. The magnetization per site m = \langle \sigma_i \rangle follows from m = \frac{1}{\beta} \frac{\partial \ln \lambda_\mathrm{max}}{\partial h}, m = \frac{\sinh(\beta h)}{\sqrt{\sinh^2(\beta h) + e^{-4 \beta J}}}. This expression captures the response to the field at any temperature, showing saturation m \to 1 as T \to 0 or h \to \infty, and vanishing m as T \to \infty. The zero-field susceptibility \chi = \left. \frac{\partial m}{\partial h} \right|_{h=0} evaluates to \chi = \beta e^{2 \beta J}, which exhibits \chi \sim \beta e^{2 \beta J} at low temperatures T \ll J/k_B, reflecting the dominance of long-range correlations along the chain. The specific heat per site C = \frac{\partial u}{\partial T}, where u = -\frac{\partial \ln Z}{\partial \beta} is the per site, takes the form C/k_B = (\beta J)^2 \sech^2(\beta J), peaking near T \sim J/k_B and decaying exponentially as e^{-4 \beta J} at low T. In the zero-field limit h \to 0, the magnetization vanishes m = 0 for all finite T > 0, consistent with the absence of in one dimension; this result exactly reproduces Ising's original 1925 calculation, which demonstrated no at finite temperature using a recursive approach equivalent to the formalism.

Two-Dimensional Solution

In 1944, provided the exact solution for the partition function of the two-dimensional Ising model on a square with and zero external , marking a pivotal advance in . This solution was obtained by extending the to the two-dimensional case, where the partition function Z for a lattice of N sites is given by \ln Z = \frac{1}{2} \int_0^{2\pi} \frac{d\phi}{2\pi} \ln \left[ 2 \cosh 2\beta J \left(1 + \sqrt{1 - \kappa^2 \sin^2 \phi}\right) \right], with the parameter \kappa = \frac{2 \sinh 2\beta J}{\cosh^2 2\beta J}. Here, \beta = 1/(k_B T) is the inverse temperature, J is the ferromagnetic coupling constant, and the integral arises from diagonalizing the transfer matrix in momentum space, capturing the eigenvalues that determine the free energy. This expression reveals the thermodynamic properties, such as the specific heat, through differentiation with respect to temperature. The critical temperature T_c for the is determined by the condition where the integrand exhibits a logarithmic , yielding \sinh(2\beta_c J) = 1, or explicitly k_B T_c / J = \frac{2}{\ln(1 + \sqrt{2})}. Above T_c, the system is disordered with zero ; below T_c, long-range order emerges. To compute the m in the absence of an external field (h=0), Onsager collaborated with Bruria Kaufman, who in 1949 employed a method to evaluate correlation functions. This technique maps the Ising spins to a fermionic via a Jordan-Wigner-like transformation, expressing the partition function and correlations in terms of Pfaffians of skew-symmetric matrices, which reduce to determinants for practical computation. The resulting spontaneous magnetization below T_c is m = \left[1 - \left(\sinh 2\beta J\right)^{-4}\right]^{1/8}, which vanishes continuously at T_c as (1 - T/T_c)^{1/8} below T_c, with critical exponent \beta = 1/8. This formula highlights the power-law singularity in the at the critical point, underscoring the non-analyticity of the and contrasting with the mean-field exponent \beta = 1/2. The approach not only yields m but also enables exact calculations of spin-spin functions, decaying algebraically below T_c with an exponent \eta = 1/4. These results confirm the presence of a finite-temperature in two dimensions, resolving earlier uncertainties from approximate methods.

Solutions in Higher Dimensions

Unlike the two-dimensional case, the three-dimensional Ising model lacks an exact analytical solution, with progress relying on approximate methods such as high- and low-temperature series expansions. These expansions, computed to high orders, estimate the critical temperature for the simple cubic lattice at k T_c / J \approx 4.51, where J is the coupling constant and k is Boltzmann's constant. The mean-field theory, which neglects fluctuations, overestimates this value at k T_c / J = 6, highlighting the importance of beyond-mean-field corrections in three dimensions. In four dimensions and above, the Ising model reaches its upper d_c = 4, where the Gaussian fixed point dominates the flow. Above d = 4, match their mean-field values exactly, as fluctuations become irrelevant and the theory becomes perturbative around the free-field limit. The , or Cayley tree, serves as a loop-free relevant to higher dimensions, allowing an recursive via transfer-matrix methods. The is determined by the onset of in the recursion relations, satisfying \tanh(J / k_B T_c) = 1/(q-1) for coordination number q. To access dimensions below four, the ε-expansion employs perturbative renormalization group techniques around d = 4 - \varepsilon, treating ε as a small parameter to compute corrections to mean-field exponents. This approach, seminal in understanding the Wilson-Fisher fixed point, provides systematic series for valid near the upper critical dimension.

Numerical and Computational Methods

Monte Carlo Simulations

Monte Carlo simulations provide a powerful numerical approach to study the equilibrium properties of the , particularly in dimensions where exact solutions are unavailable, by generating a sequence of spin configurations distributed according to the . The foundational technique is the Metropolis algorithm, introduced in , which generates Markov chains of configurations through local updates. In this method, an initial spin configuration is selected, and at each step, a single spin is randomly chosen and proposed for flipping. The energy change \Delta E associated with this flip, computed from the H = -J \sum_{\langle i,j \rangle} s_i s_j - h \sum_i s_i, determines the acceptance probability A = \min\left(1, e^{-\beta \Delta E}\right), where \beta = 1/(k_B T). If a uniform is less than A, the flip is accepted; otherwise, the original configuration is retained. This single-spin flip proposal ensures exploration of the configuration space while biasing towards low-energy states. Near the critical temperature, the Metropolis algorithm suffers from critical slowing down, where autocorrelation times diverge due to long-range correlations, leading to inefficient sampling. To mitigate this, the Wolff algorithm, proposed in , employs cluster-based updates that flip large correlated groups of simultaneously. The procedure begins by randomly selecting a seed , then iteratively adding neighboring of the same orientation to the cluster with probability p = 1 - e^{-2\beta J}; once the cluster is formed, all in it are flipped with certainty. This rejection-free method significantly reduces autocorrelation times at criticality, improving efficiency for simulations of the Ising model in two and three dimensions by orders of magnitude compared to single-flip dynamics. From the generated configurations, equilibrium averages of observables are estimated via or time averages after discarding initial transients for equilibration. For instance, the per is m = \frac{1}{N} \left\langle \sum_i s_i \right\rangle, the average energy is \langle H \rangle, and the is \chi = \beta N \left( \langle m^2 \rangle - \langle m \rangle^2 \right), where brackets denote averages over many independent configurations. These estimators converge to exact thermodynamic quantities as the chain length increases, allowing accurate determination of phase transitions and in the Ising model. The validity of these algorithms relies on satisfying and in the underlying . requires that the transition probabilities P(\sigma \to \sigma') \pi(\sigma) = P(\sigma' \to \sigma) \pi(\sigma'), where \pi(\sigma) \propto e^{-\beta H(\sigma)} is the ; the Metropolis acceptance rule ensures this by making the ratio of forward to reverse probabilities equal to e^{-\beta \Delta E}. guarantees that from any starting , the chain can reach any other with positive probability in finite steps, implying to the unique \pi regardless of initial conditions; for the Ising model, the connected and non-zero acceptance probabilities for flips establish this property.

Transfer Matrix Approaches

The transfer matrix method offers an exact computational framework for evaluating the partition function of the Ising model on finite lattices, extending the analytical approach used in one dimension to quasi-two-dimensional geometries. By modeling the as a strip of finite width L in and infinite (or very long) extent in the perpendicular direction, the method captures essential features of phase transitions and in two dimensions while remaining computationally tractable for moderate L. This generalization builds on the one-dimensional , where the is a simple $2 \times 2 structure, but scales to higher dimensions by accounting for interactions across rows of spins. For a two-dimensional Ising model on a strip of width L and length N \gg L, the partition function Z is computed as Z = \mathrm{Tr}(T^N), where T is the of dimension $2^L \times 2^L. The matrix elements of T encode the Boltzmann weights for interactions between consecutive rows of L , with rows and columns indexed by the $2^L possible configurations (up or down for each ). In the thermodynamic limit along the length (N \to \infty), Z \approx \lambda_{\max}^N, where \lambda_{\max} is the largest eigenvalue of T, allowing direct calculation of the per as f = -\frac{1}{\beta} \log \lambda_{\max} (with \beta = 1/[kT](/page/KT)). This eigenvalue spectrum also enables the extraction of lengths via ratios of eigenvalues, such as \xi = 1 / \log(\lambda_{\max} / |\lambda_2|), where \lambda_2 is the second-largest eigenvalue. Finite-size scaling analysis leverages these transfer matrix results to extrapolate properties to the infinite-volume . and scaling functions are inferred by fitting data from strips of varying widths L, using relations like the correlation length \xi(L) \sim L / \pi at criticality or cumulants that become size-independent in the scaling regime. This approach has been instrumental in verifying exact results for the clean two-dimensional Ising model and studying perturbations, such as random fields or diluted bonds. On cylindrical geometries—strips with in the width direction—the method computes two-point correlation functions G(r) along the infinite direction from eigenvector components, providing insights into decay behaviors near criticality. Despite its power, the incurs an exponential computational cost in the strip width L, as constructing and diagonalizing the $2^L \times 2^L requires resources as O(4^L) for the full matrix or better with optimizations like symmetry exploitation. It remains feasible for widths up to L \approx 20–30 using standard numerical on modern hardware, beyond which alternatives become necessary for larger systems.

Techniques

The renormalization group (RG) techniques provide a powerful framework for analyzing the multi-scale properties and critical behavior of the Ising model by systematically integrating out short-wavelength degrees of freedom to reveal how couplings evolve under rescaling. In 1966, Leo Kadanoff introduced the block-spin transformation, a real-space coarse-graining method where groups of original spins are averaged into effective block spins, preserving the overall structure of the Hamiltonian but with rescaled lattice spacing b. This approach posits that near criticality, the effective coupling J' between block spins depends on the original coupling J and the block size b via a scaling function J' = f(J, b), enabling the study of scale-invariant physics without explicit computation of the partition function. Kenneth Wilson's formulation in 1971 advanced this idea into a continuous RG flow, treating the couplings as parameters that evolve under infinitesimal rescaling, particularly near fixed points where the system exhibits . For the Ising model, the flow equations describe how the coupling K = J / k_B T (and possibly a ) transform, with the \beta-function \beta(K) = \frac{dK}{d\ell} (where \ell = \ln b) governing the approach to the critical fixed point K_c where \beta(K_c) = 0. Linearizing around this fixed point yields the , such as the correlation length exponent \nu = 1 / y_t (with y_t the thermal eigenvalue) and the magnetic exponent y_h, which match known values for the Ising in dimensions d = 4 - \epsilon via perturbative expansions. This momentum-shell RG method, applied to a version of the Ising model (the \phi^4 theory), elucidates universality and the breakdown of below the upper critical dimension d=4. Real-space implementations, such as , offer practical approximations for models by iteratively summing over subsets of spins to obtain effective interactions. In one , exactly maps the Ising chain to a rescaled model with modified nearest-neighbor K' = \tanh K, revealing the absence of a finite-temperature fixed point and thus no , consistent with the exact solution. In two dimensions, approximate schemes, like the Migdal-Kadanoff bond-moving procedure, reduce the to a hierarchical to estimate critical couplings and exponents, though they introduce approximations that slightly deviate from exact Onsager values. On hierarchical lattices, where the geometry is self-similar by construction, real-space transformations become exact, allowing precise computation of the and exponents for the Ising model without truncation errors.

Applications

Statistical Mechanics and Magnetism

The Ising model serves as a foundational framework in statistical mechanics for understanding ferromagnetism, capturing the collective behavior of magnetic spins on a lattice interacting via nearest-neighbor couplings. In this model, spins are represented as binary variables that align parallel or antiparallel, mimicking the orientation of atomic magnetic moments in materials. The Hamiltonian, H = -J \sum_{\langle i,j \rangle} \sigma_i \sigma_j - h \sum_i \sigma_i, where J > 0 for ferromagnetic interactions, \sigma_i = \pm 1, and h is an external field, leads to a phase transition at a critical temperature T_c, below which spontaneous magnetization emerges even in zero field. This transition exemplifies a second-order phase change, where the system shifts from a disordered paramagnetic state to an ordered ferromagnetic state. In the ferromagnetic phase below T_c, are insufficient to disrupt the energetic preference for alignment, resulting in a net M = \langle \sum_i \sigma_i \rangle / N > 0. This spontaneous alignment arises from cooperative interactions that propagate order across the , analogous to the formation and observed in real ferromagnetic materials like iron, where atomic moments align below the of approximately 1043 K. The model's simplicity allows it to approximate the mean-field behavior of such metals, though real systems include additional complexities like long-range interactions and crystal anisotropy; nonetheless, the Ising framework has proven remarkably predictive for qualitative features of magnetic ordering in iron and similar alloys. Near T_c, the \chi = \partial M / \partial h diverges as \chi \sim |T - T_c|^{-\gamma}, with \gamma = 7/4 in two dimensions from solutions, signaling enhanced responsiveness to external fields as fluctuations grow critical. Similarly, the specific exhibits a singularity, particularly in two dimensions where C \sim |\ln |T - T_c||, reflecting the logarithmic divergence (corresponding to \alpha = 0 in notation) due to the sharp increase in fluctuations at the . These behaviors highlight the model's utility in elucidating aspects of phase transitions in magnetic systems. At high temperatures well above T_c, interactions become negligible compared to , and the susceptibility follows the \chi \sim 1/T, as spins behave independently like non-interacting . This high-temperature limit aligns with experimental observations in dilute magnetic systems, bridging the Ising model to classical before cooperative effects dominate near T_c.

Condensed Matter and Disordered Systems

The Ising model finds significant application in modeling phase transitions in condensed matter systems, particularly through mappings that relate spin variables to physical or composition variables. One key example is the lattice gas mapping, where Ising s_i = \pm 1 are interpreted as occupation variables for particles on a lattice site, with s_i = +1 indicating an occupied site and s_i = -1 an empty one. This equivalence transforms the ferromagnetic Ising into a model for interacting lattice gases, describing phenomena such as liquid-vapor coexistence in fluids. The critical point in this mapping occurs at half-filling, where the average is 1/2, corresponding to the ferromagnetic transition temperature of the Ising model. In the context of binary alloys, the Ising model serves as an effective description for order-disorder transitions and , where represent the local composition of two atomic , A and B, on a . Here, the order parameter, typically the staggered m = \langle s \rangle, quantifies the difference in concentrations between the two phases, driving the into A-rich and B-rich domains below the critical . This approach captures the of in systems like Cu-Zn alloys, where short-range interactions lead to a analogous to the Ising ferromagnetic . Seminal kinetic studies using methods on this model have elucidated the coarsening dynamics during , highlighting diffusive growth of domains governed by conserved order parameters. Disordered systems in are modeled by variants of the Ising framework that incorporate randomness, leading to complex landscapes of and competing interactions. The Edwards-Anderson model extends the by introducing quenched random couplings J_{ij} between spins, drawn from a Gaussian distribution, which simulates the effects of structural in real materials like dilute magnetic alloys. This randomness induces , where not all spins can align favorably, resulting in a rugged energy landscape with multiple metastable states and a spin-glass characterized by frozen at low temperatures without long-range order. The model has been pivotal in understanding the spin-glass in three dimensions, where the existence of a sharp at finite temperature remains a subject of debate, supported by numerical evidence, relying on to describe the ground-state degeneracy. Another important disordered extension is the random-field Ising model, where each spin experiences a random h_i, typically from Gaussian or bimodal distributions, mimicking impurities or defects in crystalline solids. This disorder disrupts ferromagnetic order by favoring opposite alignments in different regions. The Imry-Ma argument provides a analysis showing that in dimensions d \leq 2, random fields destroy long-range order by energetically favoring the formation of s that adapt to the field fluctuations, with the domain size as L \sim h^{-2/(d- \theta)} where \theta relates to the domain wall roughness; in higher dimensions, ordered phases can persist for weak fields. This criterion has profound implications for diluted antiferromagnets and random magnets, predicting domain destruction and algebraic decay of correlations in low dimensions.

Interdisciplinary Uses

The Ising model has found significant application in through the , a introduced in 1982 that models associative memory. In this framework, neurons are represented as Ising spins, with the coupling strengths J_{ij} learned from training patterns to minimize an energy function analogous to the Ising , enabling the network to converge to states that store and retrieve memories as stable spin configurations. In artificial neural networks, restricted Boltzmann machines (RBMs) leverage the Ising model's structure for and feature extraction. Developed in 1985, RBMs are stochastic generative models with visible and hidden units forming a , equivalent to an Ising model where the is defined by an energy function that facilitates sampling-based training algorithms like contrastive divergence, allowing efficient approximation of complex data distributions for tasks such as and pretraining deep networks. Sea ice modeling employs the to simulate the formation and evolution of melt ponds on sea ice surfaces. Here, lattice sites represent ice (+1) or open (-1), with interactions capturing and random external fields mimicking topographic variations; as effective temperature rises during summer melt, the system undergoes a second-order akin to , where connected water clusters emerge, accurately reproducing observed pond size distributions and effects critical for simulations. In the social sciences, the Ising model underpins opinion dynamics models, treating agents on a as with opinions (+1 or -1) that update via mechanisms similar to the voter model, influenced by neighboring interactions. This setup reveals phase transitions from disordered to consensus states under varying social temperatures (noise levels), providing insights into and in elections or , as demonstrated in socio-economic simulations of urban segregation and language shifts. Ecological applications of the Ising model focus on species coexistence in lattice-based metapopulations, where sites are occupied by competing or remain vacant, with spin-like states interacting through resource competition and dispersal. A 2011 lattice model shows that when resources are abundant, both species coexist in a disordered , but scarcity induces critical behavior with power-law correlations and a transition to dominance, highlighting mechanisms for biodiversity maintenance in spatially structured environments.

Extensions and Generalizations

Quantum Ising Model

The quantum Ising model extends the classical Ising model by incorporating quantum fluctuations through a transverse , which introduces non-commuting operators and leads to a quantum between ordered and disordered phases. The for the one-dimensional transverse-field quantum Ising model on a chain of s is given by H = -J \sum_{\langle i,j \rangle} \sigma^z_i \sigma^z_j - \Gamma \sum_i \sigma^x_i, where J > 0 is the ferromagnetic coupling between nearest-neighbor spins, \Gamma is the strength of the transverse field, and \sigma^{x,z}_i are the acting on site i. This model captures essential features of and serves as a for studying driven by quantum fluctuations rather than ones. In one dimension, the model exhibits a quantum at a critical transverse field \Gamma_c = J, separating a ferromagnetic (for \Gamma < J) with spontaneous magnetization along the z-direction from a paramagnetic (for \Gamma > J) aligned with the transverse field. This transition is second-order, and the model is self-dual under the \Gamma \leftrightarrow J, which maps the ordered and disordered phases onto each other, with the critical point being . The quantum Ising model is equivalent to the anisotropic quantum XY model in the limit of infinite anisotropy (\gamma = 1), where the XY Hamiltonian reduces to the Ising form after a suitable of spin axes. The one-dimensional model admits an exact solution via the Jordan-Wigner transformation, which maps the operators to non-interacting fermions, allowing in space. The resulting fermionic reveals an energy gap \Delta that closes linearly at the critical point, with \Delta \sim 2J |\Gamma/J - 1| near \Gamma_c, signaling the onset of gapless excitations and critical behavior characteristic of a (1+1)-dimensional . A powerful theoretical tool for studying the quantum Ising model is its mapping to a classical statistical mechanics problem via the path-integral formulation in Euclidean time. The partition function Z = \mathrm{Tr}(e^{-\beta H}) is represented as an imaginary-time path integral over spin configurations, where the inverse temperature \beta plays the role of the extent in the temporal direction, effectively transforming the d-dimensional quantum model into a (d+1)-dimensional classical Ising model with anisotropic couplings between spatial and temporal bonds. This quantum-to-classical correspondence facilitates the analysis of finite-temperature properties and critical exponents, particularly in higher dimensions where exact solutions are unavailable.

Anisotropic and Vector Models

The anisotropic Ising model generalizes the by allowing direction-dependent interactions, where the coupling constants differ along different axes, such as J_x \neq J_y \neq J_z. In two dimensions on a square , this model remains exactly solvable using methods, with the critical temperature T_c determined by the condition \sinh(2\beta J_x) \sinh(2\beta J_y) = 1, where \beta = 1/(k_B T) and k_B is Boltzmann's constant. This yields an anisotropic in the T-J_x/J_y plane, where T_c varies continuously with the ratio \alpha = J_y / J_x; for \alpha \to 0, the system decouples into independent one-dimensional chains with no finite-temperature transition, while for \alpha = 1, it recovers the isotropic Onsager solution with T_c \approx 2.269 J / k_B. In three dimensions, mean-field approximations or series expansions reveal similar effects, with T_c increasing along the stronger coupling direction, leading to elongated ordered regions in the . The Potts model extends the Ising framework to q discrete states per spin, \sigma_i \in \{1, 2, \dots, q\}, with the Hamiltonian H = -J \sum_{\langle i,j \rangle} \delta_{\sigma_i, \sigma_j}, reducing to the Ising model for q=2. In two dimensions, the nature of the phase transition depends critically on q: for $1 < q \leq 4, it is continuous (second-order) with power-law correlations at criticality, while for q > 4, the transition becomes first-order, characterized by a discontinuous jump in the order parameter and latent heat. This changeover at q=4 arises from the proliferation of multiple ordered states, as confirmed by exact solutions via duality and finite-lattice scaling, with the first-order regime exhibiting exponential correlation decay away from the transition. In higher dimensions, the first-order transition persists for larger q, but the critical q_c separating second- and first-order behaviors increases. Vector spin models, such as the XY and Heisenberg models, replace scalar Ising spins with continuous vector degrees of freedom, belonging to the O(n) universality class for n \geq 2. The XY model (n=2) features planar unit vectors \mathbf{S}_i = (\cos \theta_i, \sin \theta_i) with Hamiltonian H = -J \sum_{\langle i,j \rangle} \mathbf{S}_i \cdot \mathbf{S}_j, exhibiting U(1) continuous symmetry, while the Heisenberg model (n=3) uses three-dimensional vectors with SU(2) symmetry. In two dimensions, the Mermin-Wagner theorem prohibits spontaneous symmetry breaking and long-range order at any finite temperature for these models, due to Goldstone modes that destroy magnetization via infrared divergences in the spin-wave approximation; instead, the XY model shows quasi-long-range order below the Berezinskii-Kosterlitz-Thouless transition temperature, with algebraic correlations decaying as power laws. In three dimensions, both models exhibit finite-T_c second-order transitions to ferromagnetic order, with critical exponents approaching mean-field values for large n. Clock models provide a discrete approximation to vector spin systems, where each spin takes one of q equally spaced values on a circle, \theta_i = 2\pi k / q for k=0,1,\dots,q-1, with interactions H = -J \sum_{\langle i,j \rangle} \cos(q (\theta_i - \theta_j)). For q=2, it coincides with the Ising model, showing a second-order transition, while as q \to \infty, it approaches the continuous XY model. For intermediate q (e.g., q=5 to $16), the two-dimensional clock model displays a rich phase diagram with two transitions: an upper one from paramagnetic to quasi-ordered (BKT-like, with vortex unbinding) and a lower one to fully ordered phase, intermediate between Ising discreteness and XY continuity; for q > 4, the lower transition can become first-order, reflecting Potts-like multicriticality. These models are useful for studying symmetry restoration and finite-size effects in lattice simulations.

References

  1. [1]
    [PDF] An Introduction to the Ising Model Barry A. Cipra The American ...
    Apr 30, 2007 · The Ising model is concerned with the physics of phase transitions, which occur when a small change in a parameter such as temperature or ...
  2. [2]
    [2509.00632] Origins of the Ising model - arXiv
    Aug 30, 2025 · In 1925, Ernest Ising published a paper analyzing a model proposed in 1920 by Wilhelm Lenz for ferromagnetism. The model is composed of ...Missing: original English
  3. [3]
    Ernst Ising - bibliotheca Augustana
    English translation of the excerpt of the dissertation. «Beitrag zur Theorie des Ferromagnetismus»,. in: Zeitschrift für Physik, Bd. XXXI, 1925, , pp. 253-258.Missing: original paper PDF
  4. [4]
    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.
  5. [5]
    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.
  6. [6]
  7. [7]
    Shape-dependent finite-size effect of the critical two-dimensional ...
    Feb 19, 2013 · We study the finite-size behavior of the critical two-dimensional Ising model on a finite triangular lattice with free boundaries in five shapes.
  8. [8]
    [PDF] 1 The Ising model - Arizona Math
    Sep 10, 2024 · This is a third way to think about a phase transition - the correlation length diverges at the critical point. (Actually this only happens for ...
  9. [9]
    On the Onsager-Yang-Value of the Spontaneous Magnetization
    We show that the value of the spontaneous magnetization for the two- dimensional Ising model computed by Onsager is indeed, the appropriate derivative of the.<|separator|>
  10. [10]
    (PDF) Bond diluted Ising model in 2D - ResearchGate
    Aug 7, 2025 · The bond diluted Ising model is studied by Monte Carlo method. The sim-ulation is carried out on a two dimensional square lattice with ...Missing: variants | Show results with:variants
  11. [11]
    [PDF] Scaling behavior of Ising systems at first-order transitions - arXiv
    May 12, 2019 · In this paper, we study the role of the boundary conditions in the context of the simplest classical system exhibiting a magnetic FOT, i.e. the ...
  12. [12]
    [PDF] the ising antiferromagnet and max cut on random regular graphs
    The Ising antiferromagnet is an important statistical physics model with close connections to the MAX. CUT problem. Combining spatial mixing arguments with the ...
  13. [13]
    90 years of the Ising model | Nature Physics
    Dec 1, 2015 · Ernst Ising's analysis of the one-dimensional variant of his eponymous model (Z. Phys 31, 253–258; 1925) is an unusual paper in the history of early twentieth- ...Missing: thesis | Show results with:thesis
  14. [14]
    None
    Nothing is retrieved...<|control11|><|separator|>
  15. [15]
    A new look at Ernst Ising's thesis | The European Physical Journal B
    Jun 2, 2025 · The famous paper (Ising in Z Phys 31: 253, A translation to English is by Jane Ising, Tom Cummings, Ulrich Harsch is found Bibliotheca ...
  16. [16]
    [PDF] ON ISING'S MODEL OF FERROMAGNETISM
    BY MR R. PEIERLS. [Communicated by M. BORN]. [Received 16 May, read 26 October 1936]. Ising* discussed the following model of a ferromagnetic body: Assume N ...Missing: phase | Show results with:phase
  17. [17]
    [PDF] Statistics of the Two-Dimensional Ferromagnet. Part I
    approximation. The present paper is an attemptto gain sound statistical information about some model of a ferromagnet. The ...Missing: original | Show results with:original
  18. [18]
    [PDF] Crystal Statistics. I. A Two-Dimensional Model with an Order ...
    magnetic" with scalar "spins" (Ising model) is computed rigorously for the case of vanishing field. The eigenwert problem involved in the corresponding ...
  19. [19]
    [PDF] statistical-theory-equation-state-phase-transitions-II_lee_yang.pdf
    N paper I we have seen that the problem of a statistical theory of phase transitions and equations of state is closely connected with the distribution of roots ...Missing: original | Show results with:original
  20. [20]
    Correlations in Ising ferromagnets. III
    An inequality relating binary correlation functions for an Ising model with purely ferromagnetic interactions is derived by elementary arguments and used to ...Missing: original paper
  21. [21]
    One Dimensional Ising Model and Transfer Matrices
    The Ising model was first proposed in Ising's Ph.D thesis and appears in his 1925 paper based on his thesis (E. Ising, Z. Phys. 31, 253 (1925).) In his paper ...Missing: original | Show results with:original
  22. [22]
    [PDF] Lecture Notes on Statistical Mechanics - UCLA Physics & Astronomy
    Nov 3, 2016 · ... 1D Ising model! That is, the spin correlation exponentially ... with correlation length. ⇠ = 1 log. 1 tanh(βJ) . (57). 12. Page 13. As β ...
  23. [23]
    The Spontaneous Magnetization of a Two-Dimensional Ising Model
    The spontaneous magnetization of a two-dimensional Ising model is calculated exactly. The result also gives the long-range order in the lattice.
  24. [24]
    Exponential Decay of Truncated Correlations for the Ising Model in ...
    Nov 28, 2019 · The truncated two-point function of the ferromagnetic Ising model on ( ) in its pure phases is proven to decay exponentially fast throughout the ordered regime.
  25. [25]
    [PDF] Exponential decay of truncated correlations for the Ising model in ...
    Oct 7, 2019 · Abstract The truncated two-point function of the ferromagnetic Ising model on Zd (d ≥ 3) in its pure phases is proven to decay exponentially ...
  26. [26]
    The Ginibre inequality | Communications in Mathematical Physics
    Essentially, the Ginibre inequality states that certain functions on the cycle group of a graph are positive definite. This has been proved for arbitrary graphs ...Missing: second | Show results with:second
  27. [27]
    General Formulation of Griffiths' Inequalities - Project Euclid
    We present a general framework in which Griffiths inequalities on the correlations of ferromagnetic spin systems appear as natural consequences of general.Missing: original | Show results with:original
  28. [28]
  29. [29]
    [PDF] Long range order for random field Ising model - Research Collection
    May 8, 2024 · We define the classical Ising model and use the so-called Griffiths–Peierls argument to prove existence of long range order for low temperatures ...
  30. [30]
    Physics A Refinement of Simon's Correlation Inequality*
    A general formulation is given of Simon's Ising model inequality: <<Vy) = Σ (σaσb) <\σbσyy where Bis any set of spins separating α from γ. We. beB show that ...
  31. [31]
    [cond-mat/0201180] 25th-order high-temperature expansion results ...
    Jan 11, 2002 · 25th-order high-temperature series are computed for a general nearest-neighbor three-dimensional Ising model with arbitrary potential on the simple cubic ...
  32. [32]
    Critical Exponents in 3.99 Dimensions | Phys. Rev. Lett.
    Jan 24, 1972 · Critical Exponents in 3.99 Dimensions. Kenneth G. Wilson and Michael E. Fisher.
  33. [33]
    [PDF] THE RENORMALIZATION GROUP AND THE ~EXPANSION
    The e expansion about 4 dimensions was developed by Fisher and Wilson [40]. ... The Gaussian model can be obtained by modifying the Ising model. First one ...
  34. [34]
    [PDF] Equation of State Calculations by Fast Computing Machines
    A general method, suitable for fast computing machines, for investigatiflg such properties as equations of state for substances consisting of interacting ...Missing: algorithm | Show results with:algorithm
  35. [35]
    [PDF] Introduction to Cluster Monte Carlo Algorithms
    2.4 Ising Model​​ The Metropolis algorithm can be applied directly to this system. Local trial moves amount to the inversion of a single spin and are accepted or ...
  36. [36]
    [PDF] Introduction to Monte Carlo methods for an Ising Model of a ... - arXiv
    Mar 3, 2008 · This discussion serves as an introduction to the use of Monte Carlo simulations as a useful way to evaluate the observables of a ferromagnet.
  37. [37]
    [PDF] Chapter 8 Markov chain Monte Carlo - Arizona Math
    We prove that π(x) is the stationary distribution by showing it satisfies detailed balance. ... Example (Ising model): Fix a finite subset Λ of the lattice Zd.<|control11|><|separator|>
  38. [38]
    [PDF] Scaling laws for ising models near Tc*
    A model for describing the behavior of Ising models very near Tc is introduced. The description is based upon dividing the Ising model into cells which are ...
  39. [39]
    Renormalization Group and Critical Phenomena. I. Renormalization ...
    Nov 1, 1971 · The Kadanoff theory of scaling near the critical point for an Ising ferromagnet is cast in differential form.
  40. [40]
    Real-space renormalization in statistical mechanics | Rev. Mod. Phys.
    May 16, 2014 · Block spin renormalization of Ising models. We list here different calculations using the block spin renormalization method as applied to the ...
  41. [41]
    Exactly soluble Ising models on hierarchical lattices | Phys. Rev. B
    Jul 1, 1981 · Certain approximate renormalization-group recursion relations are exact for Ising models on special hierarchical lattices, as noted by Berker ...Missing: real- space
  42. [42]
    [PDF] Ising Model for Ferromagnetism
    Under an applied field, there is a smooth crossover from T > TC to T < TC. Figure 2.6b shows the equilibrium value of M as a function of T for a small but ...
  43. [43]
    [PDF] One- and two-dimensional Ising model
    Nov 8, 2021 · The Ising model is one of the most important models in statistical physics. It is analytically exactly solvable in one and two dimensions.Missing: original | Show results with:original
  44. [44]
    [PDF] Chapter 3 - Ferromagnetism alla Ising
    Above Tc the slope will be smaller than one and there will only be a trivial solution m = 0 [Fig. 3.12(c)] and below Tc the slope will be greater than 1 and two ...
  45. [45]
    [PDF] The Ising Model and Real Magnetic Materials - SciELO
    The results show that the Ising model and real magnetic materials have provided an unusually rich and productive field for the interaction between theory and ...
  46. [46]
    [PDF] Chapter 5 Ising model and phase transitions
    Figure 5.5: Exact (per spin) thermodynamic functions for the 1d Ising model. ... first derived by Yang. The correlation length is the inverse of the lattice mass.
  47. [47]
    The Ising model in two dimensions: summary of results
    ... free, cylindrical or periodic (toroidal) boundary conditions may be imposed. ... 2) The spins in the upper boundary row interact with a row of spins all fixed to ...
  48. [48]
    4.4: Ising model - Weiss molecular-field theory - Physics LibreTexts
    Sep 19, 2022 · This is the famous Curie-Weiss law, which shows that the susceptibility diverges at the approach to the Curie temperature T c . In the ...
  49. [49]
    Random-Field Instability of the Ordered State of Continuous Symmetry
    Nov 24, 1975 · We consider phase transitions in systems where the field conjugate to the order parameter is static and random. It is demonstrated that when ...
  50. [50]
    Neural networks and physical systems with emergent collective ...
    Apr 15, 1982 · Computational properties of use of biological organisms or to the construction of computers can emerge as collective properties of systems.
  51. [51]
    [PDF] A Learning Algorithm for Boltzmann Machines* - Computer Science
    ACKLEY,. HINTON. AND SEJNOWSKI work, in minimizing G, is finding the set of ... Fahlman, Hinton, and Sejnowski (1983) compare Boltzmann machines with.
  52. [52]
    Social applications of two-dimensional Ising models - AIP Publishing
    Apr 1, 2008 · I review three socio-economic models of economic opinions, urban segregation, and language change and show that the well-known two-dimensional Ising model ...Missing: seminal | Show results with:seminal
  53. [53]
    [1010.2538] Coexistence and critical behaviour in a lattice model of ...
    Oct 12, 2010 · Coexistence and critical behaviour in a lattice model of competing species. Authors:Jacek Wendykier, Adam Lipowski, Antonio Luis Ferreira.Missing: 2003 | Show results with:2003
  54. [54]
    [PDF] The One-Dimensional king Model with a Transverse Field The Ising ...
    The Ising model with a transverse field is of some physical interest, because it corresponds to the pseudospin formulation of several phase transition problems.
  55. [55]
    [PDF] The quantum Ising chain for beginners - Condensed Matter Theory
    Jan 9, 2024 · We present here various techniques to work with clean and disordered quan- tum Ising chains, for the benefit of students and non-experts.
  56. [56]
    [PDF] 14 Quantum Field Theory and Statistical Mechanics - Eduardo Fradkin
    We will see below that at g = 1 the energy gap of the spectrum of the Ising model vanishes, and the system is at its phase transition point. using the results ...<|control11|><|separator|>
  57. [57]
    [PDF] Euclidean path integrals for quantum Ising and XY models
    Quantum - classical mapping (Euclidean path integral or imaginary time path integral) d-dimensional quantum systern at T=0. → (d+1)-dimensional statistical ...
  58. [58]
    Critical temperature in 2D square Ising model with anisotropic ...
    Abstract. To solve the 2D square Ising model incorporating anisotropy, we discuss a graphical representation of Kramers-Wannier duality.
  59. [59]
    First- and Second-Order Phase Transitions in Potts Models
    Sep 10, 1979 · Finite-lattice approximations yield the changeover from continuous to first-order phase transitions predicted by Baxter but not observed in ...Missing: 2D | Show results with:2D
  60. [60]
    Phase transitions in the -state clock model | Phys. Rev. E
    The two-dimensional (2D) Ising model shows a spontaneous symmetry-breaking (SSB) phase transition at a finite temperature [1, 2] . The 2D X Y model does not ...