Fact-checked by Grok 2 weeks ago

Percolation threshold

In , a branch of and physics that models the formation of connected s in random media, the percolation threshold is the critical occupation probability p_c at which an infinite spanning emerges, enabling long-range connectivity or "percolation" across the system. This threshold marks a from disconnected finite s to a connected , analogous to critical points in , and was first formalized in the context of fluid flow through porous materials with randomly blocked paths. Introduced by Broadbent and Hammersley in 1957 as a model for random processes in media like soils or networks, the concept applies to site percolation (random occupation of lattice sites) and bond percolation (random occupation of edges between sites). The value of p_c depends on the dimensionality and geometry; for example, in , p_c = [1](/page/1), meaning full occupation is required for , while in two-dimensional square , percolation has an p_c = 1/2 and percolation p_c \approx 0.5927. In three dimensions, such as the simple cubic , p_c \approx 0.2488 and p_c \approx 0.3116. Above p_c, a giant dominates, exhibiting properties and scaling behaviors described by that vary with . Percolation thresholds have broad applications beyond pure theory, including modeling electrical in composite materials, where the threshold predicts the onset of bulk as conductive particles reach sufficient ; in porous rocks for ; and resilience, such as the fraction of nodes that must fail before a communication fragments. In , it informs spreading thresholds in random contact s, and in , it guides the design of disordered systems like polymers or gels. Exact solutions remain rare, limited to specific low-dimensional cases, with higher-dimensional values often obtained via numerical simulations or series expansions.

Fundamentals of Percolation Theory

Definition and Core Concepts

The percolation threshold represents a fundamental concept in , introduced by Broadbent and Hammersley in to model the random spread of a through a , such as the flow of liquid in a lattice-like structure where permeability arises from stochastic processes. This framework captures the transition from localized to extended connectivity in disordered systems, analogous to phase transitions in statistical mechanics. In percolation, the threshold p_c^{\text{site}} is the critical occupation probability at which an infinite of occupied sites emerges in an infinite , where each site is independently occupied with probability p, and clusters form via nearest-neighbor connections among occupied sites. Similarly, in bond percolation, the threshold p_c^{\text{bond}} is the critical probability for bonds (edges between sites) to be present, enabling a spanning or infinite through connected bonds. These thresholds mark the point where the system undergoes a from disconnected components to a macroscopic connected structure. The percolation probability P(p), also denoted \theta(p), quantifies the likelihood that a given site belongs to the infinite , serving as the order parameter for the transition; it satisfies P(p) = 0 for p < p_c and P(p) > 0 for p > p_c. In finite systems of size N, this is approximated by the relative size of the largest , with the order parameter formally defined as \lim_{N \to \infty} \frac{S_{\max}}{N} = P(p), where S_{\max} is the size of the largest . The subcritical regime (p < p_c) features only finite with exponentially decaying tail probabilities for large cluster sizes, ensuring no spanning . In contrast, the supercritical regime (p > p_c) exhibits a unique infinite with positive density P(p), enabling long-range across the . These regimes highlight the sharp nature of the transition, akin to in other physical .

Critical Phenomena and Universality

At the percolation threshold p = p_c, the system exhibits critical characterized by the divergence of the correlation length \xi, which quantifies the spatial extent of connected s and scales as \xi \sim |p - p_c|^{-\nu}, where \nu is the correlation length . Near this critical point, key observables display power-law behaviors governed by . For p > p_c, the percolation strength P(p), defined as the probability that a site belongs to the infinite , scales as P(p) \sim (p - p_c)^\beta, with \beta the exponent. The , typically the size, diverges as S(p) \sim |p - p_c|^{-\gamma} on both sides of p_c, where \gamma is the exponent. These exponents capture the singular at the transition, analogous to phase transitions in other statistical systems. The universality hypothesis posits that critical exponents depend only on the dimensionality d and the range of interactions, not on microscopic details, placing systems into universality classes. In two dimensions, exact values for percolation exponents have been derived using mapping to the Potts model and conformal field theory techniques, yielding \beta = 5/36 and \nu = 4/3. Percolation corresponds to the q \to 1 limit of the q-state Potts model via the Fortuin-Kasteleyn representation, which unifies Ising-like models with geometric cluster descriptions. In this framework, exact solutions in two dimensions leverage conformal invariance to compute exponents and scaling functions precisely. In finite systems of linear size L, finite-size scaling bridges simulations to infinite-volume criticality, where quantities like P(L, p) scale as P(L, p) \sim L^{-\beta/\nu} f((p - p_c) L^{1/\nu}), with f a universal scaling function. The threshold p_c is estimated by identifying where scaling-invariant ratios, such as crossing probabilities in rectangular geometries or the cumulant U = 1 - \frac{\langle S^4 \rangle}{3 \langle S^2 \rangle^2} ( cluster size moments), become independent of L. Hyperscaling relations connect exponents to dimensionality, such as d \nu = 2 - \alpha, where \alpha is the exponent for the singular part of the "specific heat" analog (mean-squared cluster size fluctuations). This relation holds below the upper critical dimension d_c = 6 for , validating scaling assumptions in low dimensions while failing above d_c due to mean-field dominance.

Types of Percolation Models

Lattice-Based Models

Lattice-based models in consider discrete, regular grids where sites or bonds are randomly occupied to study connectivity transitions. These models, foundational to the field, abstract physical systems like porous media or lattices into graphs where occurs through nearest-neighbor connections. Introduced by Broadbent and Hammersley in their seminal work on stochastic processes modeling fluid flow through random media, lattice models emphasize the role of structural homogeneity in determining percolation behavior. In site percolation, each (site) of the is independently occupied with probability p, and unoccupied otherwise. Connectivity forms between occupied sites that are nearest neighbors, creating clusters of linked sites; the percolation threshold marks the point where an infinite cluster emerges, spanning the . This model captures scenarios where blockages occur at particle positions rather than connections between them. Bond percolation, in contrast, involves the edges (bonds) between nearest-neighbor sites on the , each independently open with probability p to allow passage, or closed otherwise. Paths connect sites through sequences of open bonds, with the threshold indicating the emergence of an infinite connected component via these bond-linked routes. This variant is particularly suited to modeling flow through channels or pipes in a . A is the mixed site-bond percolation model, where sites are occupied independently with probability p_s and bonds are open independently with probability p_b. Here, requires both an occupied site and an open bond, leading to correlated occupations that influence the effective threshold; the model interpolates between pure site and bond cases, enabling analysis of crossover behaviors through the joint parameter space. The z, defined as the number of nearest neighbors per site in the , plays a central role in approximations for thresholds. In the , applicable to high-dimensional s or tree-like Bethe lattices where loops are negligible, the critical probability approximates p_c \approx 1/(z-1) for bond , reflecting the needed for infinite connectivity. This approximation captures the scaling behavior as dimensionality increases, providing an upper bound for finite-dimensional systems. Simple lattices illustrate these concepts: the has coordination number z = 4, with each site connected to four orthogonal neighbors, forming a planar suitable for studying two-dimensional transitions. The triangular lattice, with z = 6, connects sites to six equidistant neighbors in a hexagonal arrangement, offering denser connectivity and often serving as a dual to the in studies. Dimer coverings represent special cases of , where the lattice is tiled with dimers (pairs of adjacent or ) either fully or partially. In full coverings, every is paired via a , restricting subsequent percolation to the remaining and effectively raising the frustration for connectivity; thresholds for such perfect matchings emerge as limiting behaviors in or models on bipartite lattices like the square grid. Partial coverings, achieved through random sequential addition, similarly constrain cluster formation, linking dimer statistics to percolation universality.

Continuum and Overlapping Models

Continuum percolation extends the concepts of to continuous space, where geometric objects such as disks or spheres are placed randomly according to a , allowing for overlaps. The coverage fraction η, defined as the expected number of objects covering any given point, is given by η = ρ v, with ρ the of object centers and v the volume (or area in ) of a single object. The percolation threshold η_c is the critical value of η at which a spanning emerges from the union of overlapping objects, marking the transition from isolated clusters to long-range connectivity. A canonical example is the Boolean model of overlapping disks in two dimensions, where disks of fixed radius are centered at Poisson points. For unit radius disks, high-precision simulations yield η_c ≈ 1.12808737(6), corresponding to a critical covered area fraction φ_c = 1 - e^{-η_c} ≈ 0.6763. This threshold has been determined through efficient Monte Carlo methods that track cluster formation near criticality. In general, for continuum systems, the percolation threshold relates to the excluded volume or area between objects. For spheres in three dimensions, η_c = \frac{4}{3} \pi r^3 \rho_c, where ρ_c is the critical density at which overlaps form a percolating network; numerical estimates place η_c ≈ 0.2895 for unit radius spheres, though focus remains on the structural analogy across dimensions. Void describes the of empty space amid packed objects, such as in hard-core packings where overlaps are forbidden, contrasting with the occupied percolation in overlapping models. This process is to the overlapping case, where the for void spanning corresponds to a critical occupied of 1 - φ_c ≈ 0.3237 in , beyond which isolated void pockets form without long-range ; this duality arises because the blocking structures in packed systems mirror the percolating clusters in the overlapping model. Random sequential adsorption (RSA) provides another continuum framework, involving the irreversible deposition of non-overlapping objects onto a substrate until occurs, with assessed for the adsorbed phase prior to saturation. In for disks, spanning clusters form at a reduced coverage φ_p ≈ 0.36, well below the limit φ_j ≈ 0.547, as determined by simulations tracking cluster growth during sequential addition. Polymers in space can be modeled as percolating paths, such as self-avoiding walks or random walks that connect via overlaps or proximity, reaching when the of chain segments enables a spanning . For random walks in , the threshold occurs at a critical segment where the effective mimics overlapping objects, with η_c similarly to disk models but adjusted for path dimensionality and self-avoidance.

Network and Graph Models

In network and graph models of , the focus shifts from regular to arbitrary or structures, where connectivity emerges through edges linking nodes without imposed geometric regularity. These models capture the behavior of complex systems like social networks, communication infrastructures, and biological webs, where the is often irregular and heterogeneous. here typically involves randomly occupying or removing edges (bond percolation) or nodes ( percolation), with the threshold defined as the probability p_c at which a giant spanning a finite fraction of the system appears. Unlike models, the absence of spatial allows for exact analytical treatments in many cases, particularly for systems or large s. Bond percolation on graphs proceeds by retaining each edge independently with probability p, effectively removing edges with probability $1-p, until a emerges at the p_c. This process models scenarios such as random link failures in communication networks, where the graph's structure determines the onset of global connectivity. In random graphs, the marks the transition from fragmented small to a macroscopic connected , analogous to the of long-range in physical systems. Seminal analyses show that for sparse random graphs, this threshold aligns with the point where the expected number of connections supports unbounded growth. Site percolation on networks, in contrast, involves removing nodes (and their incident edges) with probability $1-p, leading to the threshold p_c where the surviving subgraph develops a giant component. For the Erdős–Rényi random graph G(n,p') with n nodes and edge probability p', the mean degree is \langle k \rangle = (n-1)p' \approx np', and the site percolation threshold is p_c = 1/\langle k \rangle, meaning the giant component forms when the average surviving degree exceeds 1. This result arises from the branching process approximation, where clusters grow like a Galton-Watson process with offspring distribution Poisson(\langle k \rangle p). For tree-like graphs, such as infinite regular trees with no cycles, the site percolation threshold is exactly p_c = 1, as any p < 1 disconnects the structure into finite branches. A general criterion for the emergence of the giant component in random graphs with arbitrary degree distributions is provided by the Molloy-Reed condition, which states that a giant component exists if \langle k^2 \rangle / \langle k \rangle > 2, or equivalently, the site percolation threshold is p_c = \langle k \rangle / (\langle k^2 \rangle - \langle k \rangle), where \langle k \rangle and \langle k^2 \rangle are the first and second moments of the degree distribution. This criterion, derived from analysis of the , applies to graphs and highlights how heterogeneity in degrees influences robustness. For Erdős–Rényi graphs, where degrees are Poisson-distributed, it recovers p_c = 1/\langle k \rangle, but for broader distributions, higher variance in degrees lowers p_c by facilitating easier cluster coalescence. Scale-free networks, characterized by degree distributions P(k) \sim k^{-\gamma} with $2 < \gamma < 3, exhibit exceptional robustness to random node or edge failures, with p_c = 0 due to the diverging second moment \langle k^2 \rangle, ensuring a giant component persists even for infinitesimal p. This implies that random percolation nearly always yields global connectivity, as rare high-degree hubs anchor the structure. However, targeted attacks removing high-degree nodes first raise p_c to a finite value, exposing fragility; for \gamma > 3, p_c becomes positive even under random failure, resembling mean-field behavior. These insights underscore the dual nature of scale-free topologies in real-world systems like the . Percolation on interdependent , where in one network depend on specific in another (e.g., power grids relying on communication lines), introduces coupling that amplifies failures through cascades, drastically lowering the overall threshold compared to isolated networks. In mutually connected pairs of Erdős–Rényi networks, random failure of a fraction $1-p of triggers iterative collapses: a fails if disconnected in its primary network or if its interdependent partner fails, leading to a first-order percolation transition at p_c \approx 0.58 for equal-sized networks with \langle k \rangle = 3, far above the single-network value of \approx 0.33. This cascading mechanism explains vulnerabilities in coupled infrastructures, where even minor initial damage propagates system-wide. Explosive percolation refers to modified growth processes, such as the Achlioptas process, where edges are added selectively to suppress large clusters—e.g., by choosing the edge connecting the smallest pair of clusters from randomly sampled options—resulting in an apparently abrupt transition to the . Initially observed in random graphs, this yields a sharper-than-usual crossover, mimicking a discontinuous , though rigorous analysis confirms it remains continuous but with suppressed critical window and anomalous scaling exponents. Such processes highlight how non-random rules can alter the of percolation, with applications to controlled network design.

Percolation Thresholds in Low Dimensions

One-Dimensional Systems

In one-dimensional systems, the percolation threshold for bond percolation on a linear chain is exactly p_c^{\text{bond}} = 1, meaning an infinite spanning cluster forms only if every bond is occupied, as any unoccupied bond creates a disconnection that prevents long-range connectivity. Similarly, for site percolation, the threshold is p_c^{\text{site}} = 1, since even a single unoccupied site acts as a gap that isolates clusters and blocks spanning across the chain. Below the (p < 1), clusters in one-dimensional percolation exhibit an exponential size distribution, with the probability of a cluster of size S following a geometric form w_S = (1 - p)^2 p^{S-1}, leading to finite clusters and no infinite component due to the absence of alternative paths around gaps. This trivial threshold and rapid decay contrast with higher dimensions, where mean-field approximations begin to apply for large d. In long-range one-dimensional percolation, bonds connect sites with probability P(r) \sim r^{-\sigma}, where r is the distance; for \sigma > 2, the threshold remains p_c = 1 with no percolation possible, but for \sigma < 2, long-range links enable at p_c < 1, allowing infinite clusters through power-law connections that bypass local gaps. Numerical studies confirm this regime shift, showing p_c = 0 for small \sigma and finite p_c < 1 approaching 1 as \sigma nears 2. For directed one-dimensional percolation, the threshold stays at p_c = 1 in equilibrium models, as directed bonds along the chain require full occupancy for spanning paths, though nonequilibrium variants introduce temporal dynamics where activity can propagate below unity probability in processes like the contact model. One-dimensional percolation models serve as simple baselines for applications in linear polymers, where chain connectivity mimics site or bond occupation to predict gelation or mechanical reinforcement thresholds, and in one-dimensional transport, such as electron conduction in nanowires, where gaps model scattering events limiting current flow.

Two-Dimensional Lattices and Variants

In two-dimensional percolation models, the square lattice serves as a fundamental example for both site and bond percolation. For bond percolation on the square lattice, the critical threshold is exactly p_c = 0.5, derived from the self-duality of the lattice under the star-triangle transformation. For site percolation, high-precision numerical simulations yield p_c \approx 0.592746, obtained through hull-gradient methods that refine estimates by analyzing cluster boundaries near criticality. The triangular lattice exhibits exact thresholds due to its symmetry and duality relations. Site percolation has p_c = 0.5, reflecting the lattice's equivalence to its dual under occupation duality. Bond percolation achieves criticality at p_c = 2 \sin(\pi/18) \approx 0.3473, also exact via the star-triangle approximation, which maps the problem to solvable polynomial equations. The honeycomb lattice, dual to the triangular lattice, shows complementary behavior. Its site percolation threshold is approximately p_c \approx 0.6970, determined by gradient-percolation simulations that track spanning clusters across occupation gradients. Bond percolation occurs at p_c = 1 - 2 \sin(\pi/18) \approx 0.6527, exactly following from the duality with the triangular lattice's bond threshold. Archimedean lattices, the 11 uniform tilings of the plane by regular polygons, extend these results to more complex coordination numbers z. Site percolation thresholds vary systematically with lattice geometry, as quantified by hull-walk simulations on finite systems extrapolated to infinity. For example, the kagome lattice (3.6.3.6) has p_c \approx 0.6527, while the snub square (3^4.4^2) reaches higher values around 0.586. Bond thresholds follow an approximate relation p_c \approx 1/(z - 1 + \sqrt{2(z-2)}), which captures the scaling with average degree for many cases, though exact values require numerical refinement. The following table summarizes representative site and bond thresholds for select Archimedean lattices, based on high-precision computations:
Lattice (Schläfli symbol)Coordination zSite p_c (approx.)Bond p_c (approx.)
Triangular (3^6)60.5000000.3473
Square (4^4)40.59270.5000
Honeycomb (6^3)30.69700.6527
Kagome (3.6.3.6)40.65270.5244
Elongated triangular (3^3.4^2)60.54910.4194
Snub square (3^4.4^2)50.58640.4755
These values highlight how increasing z generally lowers thresholds, reflecting enhanced connectivity. Distorted lattices introduce geometric modifications, such as bond-bending models where angles between bonds deviate from ideality, altering local connectivity. In bond-bending frameworks on square lattices, thresholds shift upward with increasing distortion angle, as Monte Carlo simulations show reduced spanning probability due to anisotropic clustering; for example, distortions beyond 10° can raise the bond threshold by up to 5% from the undistorted value. Similar effects occur in triangular lattices, where shear distortions elevate site thresholds, emphasizing the role of lattice isotropy in criticality. Two-uniform and covering lattices, such as the (4,8)^2 tiling (also known as the ), combine squares and octagons for non-Archimedean uniformity. Site percolation on (4,8)^2 yields p_c \approx 0.417, lower than the square lattice due to higher average coordination, estimated via mean Euler characteristic methods on finite approximations. Medial lattices, as duals to these coverings, exhibit reciprocal thresholds, facilitating bounds through duality arguments. Inhomogeneous two-dimensional lattices incorporate spatial gradients in occupation probability, p(x), leading to a critical manifold where local thresholds adjust to global spanning. For a linear gradient, the effective local threshold follows p_c(x) \approx p_c^{\hom} / (1 + g x), with g the gradient strength and p_c^{\hom} the homogeneous value; this form arises from renormalization group analysis of interface roughness in gradient percolation, ensuring percolation along the gradient direction at adjusted probabilities. Correlated occupations, where site states exhibit spatial dependencies (e.g., positive correlations promoting clustering), raise thresholds above uncorrelated cases. In two-dimensional square lattices with short-range correlations, positive dependencies increase p_c by up to 10% for correlation lengths comparable to the lattice spacing, as simulations reveal suppressed long-range connectivity from local aggregates; this contrasts with negative correlations, which lower thresholds by enhancing dispersion.

Two-Dimensional Continuum Systems

In two-dimensional continuum percolation, systems involving overlapping geometric shapes provide a fundamental framework for understanding connectivity transitions without underlying lattice structures. For overlapping disks of unit radius, the critical reduced density at which percolation occurs is φ_c ≈ 0.676, corresponding to an area coverage parameter η_c ≈ 1.128, determined through efficient Monte Carlo simulations using union-find algorithms with periodic boundary conditions. Similarly, for overlapping squares, the threshold varies with orientation: aligned squares yield η_c ≈ 1.099 and φ_c ≈ 0.667, while randomly rotated squares exhibit η_c ≈ 0.982 and φ_c ≈ 0.626, highlighting the influence of shape alignment on connectivity. These values underscore how isotropic overlaps in continuous media lead to percolation thresholds distinct from discrete lattices, with universality class matching two-dimensional percolation (correlation length exponent ν = 4/3). Random sequential adsorption (RSA) in two dimensions introduces irreversibility, where disks are sequentially placed without overlap until jamming occurs, affecting percolation of the adsorbed phase. The jamming coverage for monodisperse disks reaches θ_j ≈ 0.547, beyond which no further adsorption is possible. Percolation of the adsorbed phase (defined by disk centers within distance 2r) emerges before jamming, though spanning clusters remain finite at saturation in standard models due to the irreversible nature limiting large cluster growth. In RSA variants involving self-avoiding walks, extended objects formed by k-step walks on a continuous plane or quasi-continuous embedding exhibit percolation thresholds that decrease with increasing k, reflecting enhanced connectivity from elongated shapes. For linear k-mers modeled as self-avoiding chains, numerical studies show the critical coverage θ_p^* diminishing as chain length grows, approaching lower values for larger k due to improved bridging across voids before full jamming. This behavior emphasizes the role of particle anisotropy in lowering the onset of global connectivity in irreversible deposition processes. Random quasi-lattices, such as those derived from or of , bridge continuum and lattice models, with percolation thresholds approaching those of square lattices but exhibiting slight deviations due to irregular cell geometries. Site percolation on random in 2D yields p_c ≈ 0.714, close to the square lattice value of ≈0.593, as computed via that account for the network's topological randomness. These structures model disordered media like cellular materials, where thresholds reflect averaged local coordinations akin to lattices but adapted to continuous spatial distributions. Percolation in slab geometries, which confine continuum systems to finite thickness while allowing infinite extent in one direction, behaves as quasi-one-dimensional, with thresholds increasing toward the 1D limit of 1 as thickness decreases. In thin slabs of overlapping or voids, the critical density rises monotonically with reducing height, eventually mimicking exact 1D blockage where any finite occupation prevents spanning, as analyzed in scaling studies of confined random media. This crossover illustrates dimensional reduction effects in continuum settings. Colored or AB percolation in two-dimensional continua involves two species (A and B) with mutual exclusion, where connectivity forms via same-species overlaps, and thresholds are informed by duality relations analogous to lattice models. In antagonistic RSA variants for disks, where A and B cannot overlap, the single-species percolation threshold aligns with standard RSA values, but dual exclusion leads to symmetric critical points where A and B clusters emerge complementarily, leveraging Monte Carlo estimates for exclusion-driven duality. This setup models competitive resource allocation in random media, with thresholds tied to balanced occupation fractions.

Percolation Thresholds in Three Dimensions

Three-Dimensional Lattices

In three-dimensional lattices, unlike their two-dimensional counterparts, exact percolation thresholds remain unknown and are determined through high-precision numerical simulations or rigorous bounds. These lattices exhibit coordination numbers typically ranging from 6 to 12, influencing the critical occupation probabilities required for spanning clusters. Site percolation involves random occupation of vertices, while bond percolation concerns random occupation of edges connecting nearest neighbors. Seminal Monte Carlo studies have provided precise estimates for common Bravais lattices, revealing that thresholds decrease with increasing coordination number due to enhanced connectivity opportunities. The simple cubic (SC) lattice, with a coordination number of 6, serves as a benchmark for isotropic 3D percolation. Numerical simulations yield a site percolation threshold of approximately 0.3116 and a bond percolation threshold of approximately 0.2488. For the body-centered cubic (BCC) lattice (coordination number 8), the site threshold is about 0.246 and the bond threshold about 0.1803. The face-centered cubic (FCC) lattice (coordination number 12) has the lowest thresholds among these, with site percolation at roughly 0.199 and bond percolation at about 0.1202. These values highlight the inverse relationship between coordination number and threshold, as higher connectivity facilitates percolation at lower occupation fractions. Recent high-precision simulations (as of 2022) refine the SC site threshold to ≈0.3116077.
Lattice TypeCoordination NumberSite Threshold (p_c^{\text{site}})Bond Threshold (p_c^{\text{bond}})
Simple Cubic (SC)6≈ 0.3116≈ 0.2488
Body-Centered Cubic (BCC)8≈ 0.246≈ 0.1803
Face-Centered Cubic (FCC)12≈ 0.199≈ 0.1202
Distortions introducing anisotropy, such as in where interlayer spacing varies, shift the percolation threshold relative to the isotropic case. For instance, increasing distortion—measured by the deviation in bond lengths or angles—generally raises the threshold, as reduced effective connectivity in certain directions hinders cluster formation. In a distorted , simulations show the site percolation threshold increasing monotonically with distortion amplitude when the connection range equals or exceeds the base lattice constant. This effect is pronounced in quasi-two-dimensional , where extreme anisotropy can elevate the threshold toward two-dimensional values. Dimer percolation on three-dimensional lattices extends the monomer (single-site) model by occupying pairs of adjacent sites with rigid dimers, imposing geometric constraints that alter connectivity. The threshold for spanning clusters in dimer coverings on the simple cubic lattice is approximately 0.2555 in terms of the fraction of covered sites, but the effective probability required for dimer placement is higher than for uncorrelated monomers due to exclusion and pairing restrictions, which limit the density of viable connections. This variant models systems like molecular assemblies or matching problems, where the constraint of paired occupations raises the critical density needed for percolation compared to independent site occupation. Subnet lattices, formed by embedding subgraphs within a host three-dimensional lattice, effectively reduce the average coordination number (z) by selecting sparse connections, which in turn lowers the percolation threshold relative to the full lattice. For example, subnets with reduced z (e.g., from 6 to 4 in cubic embeddings) exhibit thresholds decreased by up to 10–15%, as the sparser topology mimics lower-dimensional behavior while retaining volumetric scaling. This reduction arises because fewer edges per vertex necessitate higher occupation fractions for percolation in the full lattice, but targeted subgraphs optimize paths, lowering the effective p_c. Such models are relevant for analyzing diluted or hierarchical systems.

Three-Dimensional Continuum and Void Models

In three-dimensional continuum percolation models, the overlapping spheres model serves as a fundamental prototype for understanding connectivity in random media without underlying lattice structures. In this model, spheres of equal radius are placed randomly in space, allowing overlaps, and percolation occurs when the union of the spheres forms a connected cluster spanning the system. The critical covered volume fraction at which this happens, denoted φ_c, is φ_c ≈ 0.2896, corresponding to a critical reduced density η_c ≈ 0.341 (where η = (4/3)π r^3 n and n is the number density). This value was determined through large-scale Monte Carlo simulations using efficient cluster identification algorithms, revealing critical exponents consistent with mean-field behavior above the upper critical dimension. Void percolation models complement this by examining connectivity in the unoccupied space surrounding packed or overlapping particles, relevant for transport in porous materials. In random close packings of hard spheres, which achieve a packing fraction of approximately 0.64, the void space exhibits percolation for probing particles or flow paths up to a critical relative size p_c ≈ 0.03 times the sphere radius, reflecting the narrow and tortuous nature of pores in dense configurations. This low threshold arises from the limited porosity (about 0.36) and the geometric constraints that block large-scale void connectivity at higher probe sizes, as quantified by dilation-based simulations of the pore network. For overlapping spheres, the void phase percolates at a covered volume fraction of approximately 0.958, marking the point where the interstitial space loses connectivity. Random sequential adsorption (RSA) provides another continuum framework, where hard spheres are sequentially added without overlap until jamming occurs, mimicking irreversible deposition processes. In 3D, the jamming volume fraction θ_j ≈ 0.38, below the random close packing limit due to the kinetic arrest of the deposition process. Percolation of the adsorbed spheres emerges prior to jamming at a reduced density η_c ≈ 0.3–0.4, depending on the specific RSA protocol and correlation length, as determined by monitoring cluster growth during the adsorption kinetics. This range highlights how the irreversible nature of RSA leads to looser structures compared to equilibrium packings, with percolation driven by local ordering in the early stages of deposition. Extensions to full dimer coverings in continuum settings adapt lattice-based dimer models to continuous media, where pairs of overlapping or adjacent objects represent bonds in a matching. In 3D, thresholds for such models, analogous to perfect matchings on random graphs embedded continuously, occur at occupation probabilities around 0.5 for the dimer phase to percolate, though exact values remain simulation-dependent due to the complexity of continuous pairings. These models are useful for studying matching percolation in disordered continua, bridging discrete coverings to fluid-like systems. In random 3D media modeled by Gaussian random fields, percolation of level sets—regions above a critical height u_c—occurs at a volume fraction p_c ≈ 0.16 for the excursion sets, capturing connectivity in correlated noise typical of natural heterogeneous materials. This threshold, derived from extreme value theory and Monte Carlo sampling of the field, underscores the role of correlation length in lowering p_c below lattice values, with applications to cosmic web structures and porous geomaterials. Other continuum and void models, such as those for foams and granular media, rely heavily on simulations to estimate thresholds. In 3D foams modeled as or soap bubble networks, the liquid fraction at percolation is approximately 0.36, aligning with random close packing porosity, where void channels connect via . For granular media, simulations of polydisperse sphere assemblies yield site percolation thresholds around 0.31 for the solid phase. These models highlight the interplay of geometry and disorder in real-world continua, often validated through finite-size scaling in numerical packings.

Percolation Thresholds in Higher Dimensions

Hypercubic and Standard Lattices

In high-dimensional hypercubic lattices, the percolation thresholds for both site and bond variants decrease monotonically with dimension d, reflecting reduced geometric constraints and a transition toward mean-field behavior above the upper critical dimension of 6. For site percolation, numerical estimates yield p_c ≈ 0.197 in 4D and ≈ 0.141 in 5D, while for bond percolation, p_c ≈ 0.160 in 4D and ≈ 0.118 in 5D. These values illustrate the dimensional crossover, where critical exponents deviate from low-dimensional universality but approach mean-field values (such as β = 1, γ = 1, ν = 1/2) precisely for d ≥ 6, with logarithmic corrections appearing at d = 6. In the large-d limit, the bond threshold satisfies p_c^bond ∼ 1/(2d) asymptotically, derived from lace expansion techniques that confirm mean-field criticality. For site percolation, a leading approximation is p_c^site ≈ 1/(2d - 1), capturing the effective coordination in the tree-like structure of high-dimensional clusters. Continuum percolation models in dimensions d > 3, such as systems of overlapping hyperspheres, similarly exhibit thresholds that approach mean-field limits as d increases. The critical reduced density η_c scales as ∼ 2^{-d}, ensuring the average number of overlaps per particle N_c = 2^d η_c converges to 1, consistent with the approximation where criticality occurs when the effective branching factor equals . This behavior underscores the loss of spatial correlations in high dimensions, aligning continuum results with discrete lattice mean-field predictions. Slab geometries, featuring finite extent in perpendicular directions (d_⊥ < ∞) and infinite extent in parallel directions, provide a framework to study this dimensional interpolation quantitatively. in such anisotropic systems smoothly bridge low-d (non-mean-field) and high-d (mean-field) regimes, with critical probabilities increasing toward low-d values as slab thickness decreases, enabling rigorous bounds on universality classes via lace expansion on finite-width lattices. The infinite-dimensional limit of standard lattices is exactly captured by tree-like structures such as the Bethe lattice (Cayley tree), where loops are absent and mean-field theory applies precisely. For bond percolation on a Bethe lattice with coordination number z, the threshold is p_c = 1/(z - 1); in the hypercubic correspondence, z = 2d yields p_c = 1/(2d - 1), reproducing the large-d site approximation and serving as a benchmark for high-d universality.

Long-Range, Directed, and Hierarchical Models

Long-range percolation models extend traditional short-range variants by incorporating connections that decay as a power law with distance, typically with probability proportional to $1/r^{d+\sigma} where r is the site separation, d is the embedding dimension, and \sigma > 0 controls the range. In one dimension (d=1), the behavior depends critically on \sigma: for \sigma > 1, the model has threshold p_c = 1, as long-range bonds do not sufficiently enhance to enable percolation below full occupancy. For \sigma < 1, long-range effects dominate, yielding a lower threshold p_c < 1, with numerical studies confirming non-mean-field exponents. This regime bridges one- and higher-dimensional behaviors. Directed percolation introduces anisotropy, modeling nonequilibrium processes like epidemic spreading or fluid invasion along a preferred direction, often on lattices such as the square grid. Unlike isotropic undirected percolation, it lacks time-reversal symmetry and belongs to a distinct universality class characterized by absorbing states. For bond-directed percolation in two dimensions (1+1 spacetime dimensions), the critical threshold is p_c^\text{bond} \approx 0.6445, above which a spanning cluster propagates indefinitely in the directed sense. The order parameter exponent \beta \approx 0.276 governs the density of the active phase near criticality, differing from the undirected value \beta \approx 0.139 and reflecting slower cluster growth due to directionality. This model captures real-world phenomena like forest fire propagation, where fire spreads only downwind. Variants with multiple outgoing neighbors, such as directed bootstrap percolation, lower the threshold by requiring fewer initial seeds for activation while allowing higher out-degrees. In these processes, a site activates if it receives input from at least r active predecessors, but with multiple possible outputs per site, percolation occurs at smaller occupation probabilities compared to single-output cases; for instance, on random directed graphs, increasing out-degree reduces p_c by enhancing spreading efficiency. Bootstrap-like rules, often analyzed in directed settings, exhibit phase transitions where higher connectivity mimics lower-dimensional thresholds, with critical probabilities scaling inversely with neighbor count in mean-field approximations. Hierarchical lattices, constructed recursively with branching ratio b (e.g., diamond-like structures where each level replaces bonds with b parallel paths), admit exact renormalization-group analysis for percolation thresholds. The fixed point of the decimation map yields p_c as a function of b; for b=2, typical diamond lattices give p_c \approx 0.618, increasing with b due to enhanced redundancy but always below one-dimensional values. This approach reveals exact critical exponents, such as \nu = \ln b / \ln \lambda where \lambda is the rescaling eigenvalue, providing insights into universality without simulations. Such models approximate fractal or scale-invariant systems, with p_c decreasing as branching promotes local connectivity. Hyperbolic lattices, embedding graphs in negatively curved spaces (e.g., {7,3} tessellation with exponential volume growth), exhibit percolation thresholds lower than their two-dimensional Euclidean counterparts due to the abundance of short paths and tree-like structure. Numerical invasion percolation yields bond thresholds p_c \approx 0.18 for common hyperbolic tilings, well below the square lattice's p_c \approx 0.5, as negative curvature facilitates unique infinite clusters at lower densities. Unlike planar lattices with a single threshold, hyperbolic ones feature two: a lower p_c for infinite cluster emergence and a higher p_u for uniqueness, driven by the geometry's suppression of loops. This makes them relevant for modeling networks with hierarchical or scale-free properties. In site-bond directed percolation, sites and bonds occupy independently with probabilities p_s and p_b, forming a critical manifold in the (p_s, p_b) plane where the transition occurs along a curve separating percolating and non-percolating phases. For the square lattice in 1+1 dimensions, the manifold is a smooth line with endpoints at the pure site (p_s^c \approx 0.7055) and bond (p_b^c \approx 0.6447) thresholds, exhibiting directed universality across the interior. This combined model highlights how mixed occupations shift the effective dimensionality, with the critical surface computable via series expansions or finite-size scaling.

Computational Methods and Approximations

Exact Solutions and Series Expansions

Exact solutions for the percolation threshold exist in two dimensions for specific lattice models, leveraging symmetries such as duality and transformation equivalences. For bond percolation on the self-dual square lattice, duality arguments establish that the critical probability satisfies p_c = 1 - p_c, yielding the exact value p_c = 0.5. Similarly, for site percolation on the triangular lattice, a matching duality relation implies p_c = 0.5. The star-triangle transformation provides exact solutions for site percolation on certain lattices, such as the honeycomb, by equating the partition functions of equivalent configurations and solving the resulting algebraic equations for the critical point. Series expansions offer a powerful analytical approach to estimate percolation thresholds in higher dimensions where exact solutions are unavailable. These expansions enumerate the number of finite clusters of size n, generating power series for quantities like the mean cluster size S(p) = \sum_n n s_n p^n / \sum_n s_n p^n, where s_n is the number of n-site clusters and p is the occupation probability. High-order series, computed up to order 40 or more through systematic enumeration on supercomputers, allow estimation of p_c via the ratio method, which analyzes the convergence of ratios r_n = s_{n+1}/s_n to identify the singularity at p_c. For example, in three-dimensional bond percolation on the cubic lattice, such series expansions for the cluster statistics \sum s_n p^n provide precise approximations to p_c by fitting or differential approximants to the series data. The transfer-matrix method enables exact computations for finite systems, particularly in two dimensions, by representing the partition function as a product of matrices along one direction. For percolation on strips or cylinders of finite width L, the method calculates the exact finite-size critical probability p_c(L), which approaches the infinite-system threshold as L \to \infty via finite-size scaling analysis. This approach exploits the Markovian structure of cluster growth across slices, yielding eigenvalues that determine connectivity probabilities and allowing extrapolation to bulk properties. In hierarchical lattices, the renormalization group (RG) framework admits exact fixed points due to the self-similar structure, enabling precise determination of critical behavior without approximations. By decimating bonds or sites over scale factors, the RG flow maps the occupation probability to an effective parameter, with the unstable fixed point corresponding to p_c. For standard lattices, RG provides approximate but insightful fixed points through real-space decimation or block-spin transformations. For inhomogeneous percolation, exact critical manifolds—hypersurfaces in parameter space separating percolating and non-percolating phases—can be derived using duality extensions. In certain two-dimensional lattices, these manifolds are given by algebraic relations equivalent to the product of local bond probabilities equaling a constant, as obtained from triangle or star-duality conditions. Dimer models relate to percolation through perfect matchings on bipartite graphs, where the Kasteleyn matrix provides an exact solution in two dimensions via Pfaffian orientation. The determinant of this signed adjacency matrix yields the exact partition function for dimer coverings, which in the scaling limit connects to percolation interfaces or spanning clusters on planar domains.

Numerical Simulations and Bounds

Numerical simulations play a crucial role in estimating percolation thresholds for systems where exact analytical solutions are unavailable, particularly in dimensions greater than two. Monte Carlo methods, such as the efficient algorithm developed by , enable rapid generation and analysis of cluster statistics by tracking the addition of sites or bonds without redundant computations. This approach has been widely adopted to compute quantities like the spanning probability P(p, L) on finite lattices of size L, where finite-size scaling techniques identify the threshold p_c through the crossing point of curves for different system sizes. Theoretical bounds provide rigorous constraints on p_c, aiding in the validation of simulation results. For lattices with coordination number z, the bond percolation threshold satisfies p_c^\text{bond} \leq 1/(z-1), derived from the branching process approximation on the , which serves as a lower bound for regular lattices. Additionally, site percolation thresholds obey p_c^\text{site} \geq p_c^\text{bond}, as site occupation implies potential bond connections in the dual model. In two dimensions, duality relations impose tighter bounds; for self-dual lattices like the square, the bond threshold is exactly p_c^\text{bond} = 1/2, while for non-self-dual cases, duality between primal and dual graphs yields intervals such as p_c^\text{site} \in [0.407, 0.592] for the before refinements. High-precision simulations have refined estimates for key systems, such as the site percolation threshold on the three-dimensional simple cubic lattice, determined to be p_c^\text{site} = 0.3116077 \pm 0.0000002 using multimethod approaches combining with finite-size scaling across lattices up to L=1024. For continuum percolation on random lattices or void models, simulations often employ event-driven techniques to efficiently detect overlaps and connectivity in disk or sphere packings, avoiding exhaustive pairwise checks by processing collision events sequentially. These methods have estimated thresholds like \eta_c \approx 0.341 for overlapping spheres in 3D, where \eta is the reduced density. Error analysis in these simulations accounts for finite-size biases, where the apparent threshold shifts as p_c(L) = p_c + a L^{-1/\nu} + b L^{-\omega}, with \nu the correlation length exponent (\nu \approx 0.876 in 3D) and \omega a correction-to-scaling exponent. Extrapolations from multiple sizes mitigate systematic errors, achieving precisions below $10^{-7} in well-studied cases. Recent advances since 2020 incorporate machine learning for accelerated cluster identification; for instance, unsupervised neural networks trained on configuration images detect spanning clusters with near-perfect accuracy, reducing computational demands for large-scale 3D simulations by orders of magnitude.

Applications and Extensions

Disordered and Inhomogeneous Systems

In percolation systems, quenched randomness in site occupation probabilities p_i or coordination numbers z introduces variability that can perturb the critical behavior relative to homogeneous cases. The Harris criterion provides a framework to assess the relevance of such short-range correlated : if the correlation length exponent \nu satisfies d\nu > 2, where d is the dimension, the is irrelevant, preserving the and percolation threshold p_c of the clean system; for standard bond or site on , this holds in d \geq 2, with uncorrelated random p_i yielding an effective p_c approximately equal to the mean occupation probability for weak distributions. Stronger or marginal , as in one dimension where \nu = 1 = 2/d, can shift p_c and alter exponents, though numerical studies confirm minimal deviation for typical models above the lower . Inhomogeneous systems feature spatially varying occupation probabilities p(\mathbf{r}), such as linear gradients or localized random p_i, leading to effective thresholds that depend on the variation's strength. For weak inhomogeneities, like small gradients in p(\mathbf{r}), the global percolation threshold approximates the spatial average \bar{p}, as local fluctuations do not significantly disrupt the overall connectivity, akin to in transport properties. Stronger gradients introduce dual thresholds: a lower local p_c for short-range clusters and a higher global p_c for spanning paths across the system, observed in simulations of two-phase media where macroscopic inhomogeneity doubles the transition points. In random p_i landscapes with bounded variance, the effective p_c remains close to the homogeneous value if variations are subextensive, but sharp interfaces can elevate it by impeding cross-region bonds. Correlated disorder, particularly with power-law spatial correlations in occupation probabilities decaying as r^{-\alpha} with \alpha < 2, renders the perturbation relevant even in higher dimensions, increasing p_c beyond the uncorrelated case due to enhanced clustering of occupied sites. This long-range correlation modifies , as shown in analyses, where the effective shifts and rare regions contribute to slower near p_c. Griffiths phases emerge in these regimes, characterized by power-law relaxation times and non-universal exponents arising from locally supercritical rare regions in subcritical bulk, leading to singular behavior without a true . Such phases, first identified in diluted Ising models mapping to , manifest in correlated site-diluted lattices with algebraic correlations, where the density of infinite clusters exhibits logarithmic corrections. Multitype or AB percolation extends the model to multiple particle types (e.g., A and B) on a , where sites are randomly assigned types with probability q for A and $1-q for B, and bonds form only between like types, introducing inter-type without . The threshold for each type is determined by to the q-state via the random-cluster representation, where the joint equates to the Fortuin-Kasteleyn clusters in the Potts ensemble at q=1 for pure but adjusted for type fractions. In the AB case, the critical p_c^A and p_c^B differ if type probabilities are unequal, with the overall spanning governed by the dominant type's , as derived from duality relations in two dimensions and generalized to higher dimensions via series expansions. Interactions via shared exclusion raise the effective p_c compared to independent overlays, with phase diagrams showing coexistence regions analogous to Potts ordering. Bootstrap percolation introduces deterministic growth from an initial seed set, where a site becomes occupied if at least r neighbors are occupied, contrasting models by focusing on the for complete occupation rather than infinite clusters. The critical seed probability p_c for full decreases with , but in high dimensions (d \gg r), the transition becomes discontinuous, featuring a jump in the final occupied fraction due to supercritical avalanches triggered by critical droplets. For r=2 on hypercubic lattices, p_c \sim \pi^2 / (3 \ln d) as d \to \infty, with the discontinuity arising from the suppression of subcritical configurations, leading to abrupt closure above a secondary . Recent advances in the 2020s have applied to amorphous materials, using three-dimensional to extract thresholds from real microstructures. nano- reveals bicontinuous pore networks in dealloyed amorphous alloys, where the percolation threshold for transport corresponds to volume fractions lower than in crystalline counterparts due to disordered . In Fe-based amorphous soft magnets, electrical percolation thresholds around 0.15 are tuned via microstructural disorder, with tomographic reconstructions quantifying cluster size distributions to validate effective medium predictions. These techniques enable direct measurement of p_c in polydisperse amorphous powders, mapping caking kinetics to percolation transitions with thresholds influenced by correlations.

Real-World and Multidimensional Applications

In , percolation thresholds play a crucial role in designing conductive composites, where the onset of electrical occurs when filler particles form a spanning network. For instance, in ()-polymer composites, such as those with high-density polyethylene matrices, the percolation threshold for achieving is approximately 0.15 wt%, enabling applications in and sensors by minimizing filler content while maximizing performance. In , percolation theory models the spread of infectious diseases on contact networks, particularly in the susceptible-infected-recovered () framework, where the critical transmission probability p_c = \frac{1}{\langle k \rangle} marks the onset of a large-scale , with \langle k \rangle denoting the average degree of the network. This threshold determines whether interventions like can prevent outbreaks by keeping below the percolation point in random networks. Hydrology applies percolation thresholds to understand fluid flow through porous rocks and aquifers, where the void fraction must exceed a critical value to form connected pathways for movement. In continuum models of porous media, such as those simulating or formations, the percolation threshold for void governs permeability, with values around 0.71 for the void fraction (corresponding to a volume fraction of approximately 0.29) in the overlapping spheres model representing spaces. Biological systems leverage percolation concepts in neural networks, where connectivity thresholds influence emergent properties; for example, optimal identifies critical hubs whose targeted modulation affects global integration, potentially relevant to states of awareness, as demonstrated in animal models and analyses of human connectomes. In protein folding, thresholds dictate the transition from disordered to structured states, as seen in denatured proteins where a threshold of approximately 0.9 enables diffusive exploration of conformational space before aggregation. Multidimensional extensions of percolation address anisotropic media, where the effective percolation threshold manifests as a tensor reflecting directional dependencies in , such as in layered materials where bond occupations vary by . At the threshold, clusters exhibit fractal dimensions that quantify , with values like d_f \approx 2.53 in aiding analysis of irregular structures in composites or geological formations. Recent applications as of 2025 include quantum percolation in topological insulators, where disorder induces percolation transitions that preserve robust edge states, as in quantum anomalous Hall systems tolerant to uncorrelated noise up to thresholds near classical limits. In , percolation thresholds in fusion-based networks, such as photonic setups, set the photon loss tolerance at around $1/(n-1) for degree-n graphs, enabling scalable fault-tolerant computing.

References

  1. [1]
    [PDF] Percolation Theory - MIT
    Oct 9, 2002 · Definition 5 The percolation threshold pc is the concentration (occupation probability) p at which an infinite cluster appears for the first ...
  2. [2]
    [PDF] Percolation processes
    Oct 24, 2008 · S. R. Broadbent and J. M. Hammersley (1957). Percolation processes ... The present paper is a preliminary exploration of percolation.
  3. [3]
    Percolation Threshold -- from Wolfram MathWorld
    In this context, the percolation threshold is the fraction of lattice points that must be filled to create a continuous path of nearest neighbors from one side ...Missing: physics | Show results with:physics
  4. [4]
    Percolation Threshold - an overview | ScienceDirect Topics
    Percolation Threshold refers to the critical value of the occupation probability in percolation, at which a sudden transition occurs from a disconnected state ...V Modifications And... · 3.2. 3 Requirements For... · Semiconducting Chalcogenide...
  5. [5]
    [PDF] Percolation theory - arXiv
    Jul 11, 2005 · For a much more in-depth review of percolation on lattices and the mathematical methods involved in its study, and for the proofs of most of ...
  6. [6]
    Renormalization-Group Approach to Percolation Problems
    Aug 11, 1975 · This collection marks the 30th anniversary of the discovery of high-temperature superconductors. The papers selected highlight some of the ...Missing: Halperin Ma 1971
  7. [7]
    Universality of the Crossing Probability for the Potts Model for q=1,2 ...
    Apr 18, 2002 · We check the percolation through Fortuin-Kasteleyn clusters near the critical point on the square lattice by using representation of the Potts ...
  8. [8]
    Finite-size scaling of directed percolation in the steady state
    Oct 22, 2007 · We introduce a ratio U , Eq. (5.6) , which in a certain sense takes on the role in critical dynamics that the famous Binder cumulant plays in ...
  9. [9]
    [PDF] 2 Percolation - Arizona Math
    2.1 Definition of the model. Percolation can be defined on any lattice in any number of dimensions. It also comes in two flavors - bond percolation and site ...
  10. [10]
    Mixed percolation as a bridge between site and bond percolation
    By using mixed percolation as a bridge between site and bond percolation, we derive a new inequality between the critical points of these processes that is ...
  11. [11]
    [PDF] arXiv:2111.10975v1 [cond-mat.stat-mech] 22 Nov 2021
    Nov 22, 2021 · For bond percolation, Bethe-lattice behavior pc = 1/(z − 1) is expected to hold for large z, and the finite-z correction is confirmed to satisfy.
  12. [12]
    Site percolation on square and simple cubic lattices with extended ...
    Feb 12, 2021 · The percolation thresholds decrease monotonically with the coordination number, and linear behavior implies that the dependence of p c on z ...
  13. [13]
    [PDF] arXiv:1507.04411v2 [cond-mat.stat-mech] 25 Sep 2015
    Sep 25, 2015 · Here, we propose a novel procedure for generating random dimer coverings of a given lattice. We then compute the bond percolation threshold on ...
  14. [14]
    [PDF] Continuum Percolation Thresholds - arXiv
    Dec 8, 2012 · We apply this approach to percolation in continuum models, finding overlaps between objects with real-valued positions and orientations. In ...Missing: Boolean | Show results with:Boolean
  15. [15]
    [PDF] arXiv:1910.05072v2 [cond-mat.dis-nn] 8 Dec 2019
    Dec 8, 2019 · known value of the percolation threshold of discs, i.e., ellipses with ε = 1, is η◦ c = 1.128 087 37(6), respectively n◦ c = η◦ c / πr2 ...<|control11|><|separator|>
  16. [16]
    Effect of dimensionality on the continuum percolation of overlapping ...
    Aug 7, 2025 · These results lead to lower bounds on the percolation threshold density η(c), which become progressively tighter as d increases and exact ...Missing: dual | Show results with:dual
  17. [17]
    Precise determination of the void percolation threshold for two ...
    Jul 1, 2000 · The void percolation threshold is calculated for a distribution of overlapping spheres with equal radii, and for a binary-sized distribution ...Missing: disks 2D numerical
  18. [18]
    [PDF] Percolation of Binary Disk Systems: Modeling and Theory - OSTI
    Jun 28, 2016 · This work utilizes Monte Carlo models to predict percolation thresholds for a two dimensional systems containing disks of two different radii.Missing: Boolean | Show results with:Boolean
  19. [19]
    Continuum percolation of overlapping disks with a distribution of ...
    Aug 26, 2013 · In this paper we consider a continuum percolation model of overlapping disks in two dimensions where distribution of the radii of the disks has ...Missing: Boolean η_c
  20. [20]
    [PDF] PERCOLATION AND RANDOM GRAPHS - of Remco van der Hofstad
    dν = 2 − α. (1.80). More precisely, it is predicted that the scaling relations in (1.28) hold for all d, while the hyperscaling relations in (1.80) only ...
  21. [21]
    [PDF] Random Graphs - Stanford Network Analysis Project
    Percolation threshold: how many edges need to be added before the giant component appears? As the average degree increases to z = 1, a giant component suddenly ...
  22. [22]
    [PDF] A CRITICAL POINT FOR RANDOM GRAPHS WITH A GIVEN ...
    Aug 14, 2000 · In this paper we consider two parameters of certain random graphs: the num- ber of vertices and the number of cycles in the largest ...
  23. [23]
    Explosive Percolation in Random Networks - Science
    Whether nonrandom selection rules can delay (or accelerate) percolation in such models, which have become known as Achlioptas processes, has received much ...
  24. [24]
  25. [25]
    [PDF] Discontinuity of the Percolation Density in One Dimensional l/\x—y\2 ...
    For long range models with an asymptotic power law falloff, Kx&μ/\x\~s, the above criterion shows the absence of percolation if s > 2 (regardless ofμand the.
  26. [26]
    Directed paths on percolation clusters | Journal of Statistical Physics
    Nov 21, 1991 · Above the percolation threshold, the scaling behavior is governed by the standard “random energy” exponents (ω=1/3 and ζ=2/3 in 1+1 dimensions).
  27. [27]
    Applications of dynamic bond percolation theory to the dielectric ...
    Our recently-developed dynamic bond percolation model is extended and applied to polymer electrolytes by assuming an approximate form for the relaxation of the ...
  28. [28]
    [PDF] Percolation, statistical topography, and transport in random media
    For example, an infinite percolation cluster at the percolation threshold is statistically self- similar in the scaling range of [A,o, oo ], where A,o is ...
  29. [29]
  30. [30]
    [PDF] Exact percolation probabilities for a square lattice - arXiv
    Mar 31, 2022 · Thus, currently, the most accurate value of the percolation threshold for site percolation on a square lattice p𝑐 = 0.592 746 050 792 10(2) has ...
  31. [31]
    Site percolation thresholds for Archimedean lattices | Phys. Rev. E
    Jul 1, 1999 · Precise thresholds for site percolation on eight Archimedean lattices are determined by the hull-walk gradient-percolation simulation method ...
  32. [32]
    [1910.12376] Bond percolation thresholds on Archimedean lattices ...
    Oct 27, 2019 · Here we report the result of large parallel calculations to produce what we believe may become the reference values of bond percolation ...
  33. [33]
    Determination of the bond percolation threshold for the Kagome lattice
    Jul 10, 1997 · The hull-gradient method is used to determine the critical threshold for bond percolation on the two-dimensional Kagome lattice (and its dual, ...
  34. [34]
    Percolation in a distorted square lattice | Phys. Rev. E
    Jan 11, 2019 · This paper presents a Monte Carlo study of percolation in a distorted square lattice, in which the adjacent sites are not equidistant.Missing: bending | Show results with:bending
  35. [35]
    Bond percolation in distorted square and triangular lattices - arXiv
    Sep 12, 2025 · This article presents a Monte Carlo study on bond percolation in distorted square and triangular lattices. The distorted lattices are generated ...Missing: bending models
  36. [36]
    [PDF] arXiv:0708.3250v3 [cond-mat.stat-mech] 13 Jan 2008
    Jan 13, 2008 · Global physical properties of random media change qualitatively at a percolation threshold, where isolated clusters merge to form one infinite.
  37. [37]
    [PDF] Critical exponents of planar gradient percolation - arXiv
    We study some aspects of “gradient percolation.” This is a model of inhomogeneous site percolation where the probability for each site to be occupied varies ...
  38. [38]
    [PDF] arXiv:1305.6941v2 [cond-mat.stat-mech] 11 Feb 2014
    Feb 11, 2014 · We study the effect of positive correlations on the critical threshold of site and bond percolation in a square lattice with d = 2.
  39. [39]
    Correlation-space description of the percolation transition in ...
    Oct 4, 2007 · We explore the percolation threshold shift as short-range correlations are introduced and systematically varied in binary composites.Missing: uncorrelated | Show results with:uncorrelated
  40. [40]
    Precise determination of the bond percolation thresholds and finite ...
    Jan 1, 1998 · Extensive Monte Carlo simulations were performed to study bond percolation on the simple cubic (sc), face-centered-cubic (fcc), and body-centered-cubic (bcc) ...
  41. [41]
    Percolation in a simple cubic lattice with distortion | Phys. Rev. E
    It is observed that the percolation threshold always increases with distortion if the connection threshold is equal to or greater than the lattice constant of ...
  42. [42]
    [0709.3626] Dimer percolation and jamming on simple cubic lattice
    Sep 23, 2007 · The percolation threshold is estimated as p_c^\text{perc} \approx 0.2555 \pm 0.0001. The jamming threshold is estimated as p_c^\text{jamm} ...Missing: 3D | Show results with:3D
  43. [43]
    Site and bond percolation thresholds on regular lattices with ...
    The critical number (per site) of these objects for percolation between neighboring occupied sites defines the threshold p c .
  44. [44]
    Precise determination of the critical threshold and exponents in a ...
    Precise determination of the critical threshold and exponents in a three-dimensional continuum percolation model. M D Rintoul and S Torquato.
  45. [45]
    Percolation analysis for estimating the maximum size of particles ...
    Feb 28, 2019 · The percolation threshold was found to be ∼ 0.03 for all random overlapping sphere simulations, consistent with the work of Priour [6] and ...
  46. [46]
    Random sequential addition of hard spheres in high Euclidean ...
    Dec 20, 2006 · We investigate the structural characteristics of random sequential addition (RSA) of congruent spheres in d-dimensional Euclidean space R d in the infinite- ...Missing: 0.547 | Show results with:0.547
  47. [47]
    [1208.0328] Percolation through Voids around Overlapping Spheres ...
    Aug 1, 2012 · Abstract:The percolation threshold for flow or conduction through voids surrounding randomly placed spheres is rigorously calculated.Missing: 3D seminal
  48. [48]
    Reionization through the lens of percolation theory - Oxford Academic
    They found, however, that the initial Gaussian random field in that model has β ∼ 0.5–0.6. We have also seen that the power-law slope τ of the size ...
  49. [49]
    Effect of porosity and pore size distribution on elastic modulus of foams
    Jan 1, 2024 · For void ratio less than the percolation threshold Φ c , all the curves collapse onto one single master curve independent of the parameters ...Effect Of Porosity And Pore... · 2. Method: Numerical Model · 3. Results And Discussion
  50. [50]
    Site percolation in randomly packed spheres | Phys. Rev. B
    Nov 15, 1979 · The site-percolation threshold for randomly packed spheres is 0.310, with a critical volume fraction of 0.183. The mean coordination is 6.Missing: void close 3D p_c 0.03
  51. [51]
    Mean-field critical behaviour for percolation in high dimensions
    Jul 6, 1989 · The triangle condition is known to imply that various critical exponents take their mean-field (Bethe lattice) values ( γ = β = 1 , δ = Δ t = 2 ...
  52. [52]
    [PDF] Asymptotics in High Dimensions for Percolation
    Abstract. We prove that the critical probability for bond or site percolation on Zd is asymptotically equal to 1/(2d) as d → ∞. If the probability of a bond.
  53. [53]
    On the percolation threshold for a d-dimensional simple hypercubic ...
    The technique provides convincing evidence that none of the conjectures are correct and gives a quantitative understanding of the quality of the approximations.Missing: 3D | Show results with:3D
  54. [54]
    [1208.3720] Effect of Dimensionality on the Continuum Percolation ...
    Aug 18, 2012 · We find that the bounds converge to one another as the space dimension increases, but the lower bound provides an excellent estimate of \eta_c ...Missing: field 1.436
  55. [55]
    [PDF] Approximation on slabs and uniqueness for Bernoulli percolation ...
    The present work addresses two fundamental problems in percolation theory which have not yet been considered for the model described above. The first one is to ...
  56. [56]
    Mean-Field Critical Behaviour for Percolation in High Dimensions
    At this point we need to make the following definitions. Definition 13. Given a bond configuration {nb} and a set A of sites, we define. (a) the connected ...
  57. [57]
    One-dimensional long-range percolation: a numerical study - arXiv
    Oct 1, 2016 · In this paper we study bond percolation on a one-dimensional chain with power-law bond probability C/ r^{1+\sigma}, where r is the distance length between ...
  58. [58]
    Oriented Bond Percolation and Phase Transitions: an Analytic ...
    Aug 7, 2025 · We use it to give pc < 0.647 with a 99.999967% confidence. As Monte Carlo simulations suggest that pc ≈ 0.6445, this bound is fairly tight.Missing: 2D | Show results with:2D
  59. [59]
    Directed percolation to sustained turbulence in Couette flow
    Feb 15, 2016 · These values are close to the universal DP exponent (0.276) and fits with a fixed exponent of 0.276 are shown in red. In the numerical ...
  60. [60]
    Bootstrap percolation in directed and inhomogeneous random graphs
    Nov 25, 2015 · We perform a thorough analysis of bootstrap percolation on a novel model of directed and inhomogeneous random graphs, where the distribution of ...Missing: multiple outputs
  61. [61]
    Renormalization Group Approach to Percolation in Hierarchical ...
    Feb 18, 2022 · In this paper we study the percolation problem on a hierarchical lattice, where exact results for the critical exponents can be obtained from a decimation ...Missing: diamond | Show results with:diamond
  62. [62]
    [1708.05876] Percolation Thresholds in Hyperbolic Lattices - arXiv
    Aug 19, 2017 · Abstract:We use invasion percolation to compute numerical values for bond and site percolation thresholds p_c (existence of an infinite cluster) ...
  63. [63]
    Some Exact Critical Percolation Probabilities for Bond and Site ...
    Some Exact Critical Percolation Probabilities for Bond and Site Problems in Two Dimensions. M. F. Sykes · J. W. Essam. Wheatstone Physics Laboratory, King's ...
  64. [64]
    Series Expansion of the Percolation Threshold on Hypercubic Lattices
    May 7, 2018 · En passant, we present new perimeter polynomials for site and bond percolation and numerical values for the growth rate of bond animals.Missing: 5d 0.111
  65. [65]
    Percolation transitions in two dimensions | Phys. Rev. E
    Sep 30, 2008 · The bond percolation threshold for the diced lattice follows from a duality (d) transformation of the kagome lattice model, and therefore ...
  66. [66]
    A real-space renormalization group study of correlated percolation
    The critical behaviour of a quenched site-correlated percolating system in two dimensions has been studied using the real-space renormalization group.
  67. [67]
    The critical manifolds of inhomogeneous bond percolation on bow ...
    Oct 24, 2012 · We give a conditional derivation of the inhomogeneous critical percolation manifold of the bow-tie lattice with five different probabilities, a ...Missing: product local thresholds
  68. [68]
    Exact solution of the 2d dimer model: Corner free energy, correlation ...
    The complete and detailed fermionic solution of the dimer model on the square lattice with an arbitrary number of monomers is presented.Missing: threshold | Show results with:threshold<|control11|><|separator|>
  69. [69]
    Percolation thresholds for photonic quantum computing - Nature
    Mar 6, 2019 · Of all infinite graphs of maximum degree n, the minimum bond-percolation threshold is that of the degree-n Bethe lattice, and equals 1/(n−1).<|separator|>
  70. [70]
    [PDF] Introduction to Bernoulli percolation - IHES
    Oct 7, 2018 · We refer to [20, 6, 13] for books on the subject. We will often define Bernoulli percolation on the infinite lattice Zd. We therefore consider.
  71. [71]
    The effects of disorder on Harris-criterion violating percolation - arXiv
    Apr 8, 2021 · The Harris criterion states whether the critical behavior at a phase transition from a disordered state to an ordered state will be altered by ...
  72. [72]
    Absorbing state phase transitions with quenched disorder
    Jun 23, 2004 · Quenched disorder—in the sense of the Harris criterion—is generally a relevant perturbation at an absorbing state phase transition point.Missing: threshold | Show results with:threshold
  73. [73]
    [PDF] Advanced Statistical Physics: Quenched random systems - LPTHE
    Nov 23, 2022 · ... Harris criterium ... probability. 4. Page 5. 1.2 Properties. 1 QUENCHED RANDOM SYSTEMS. 1.2.2 Frustration.
  74. [74]
    Double-threshold percolation behavior of effective kinetic coefficients
    Aug 11, 2008 · We investigate thresholds of percolation in two-phase macroscopically inhomogeneous media. It is shown that, unlike standard percolation ...
  75. [75]
    Inhomogeneous percolation models for spreading phenomena in ...
    The generating functions approach is exploited in order to get a generalization of the Molloy–Reed criterion for inhomogeneous joint site–bond percolation in ...<|separator|>
  76. [76]
    How Inhomogeneous Site Percolation Works on Bethe Lattices
    Mar 1, 2016 · In this paper, we first investigate inhomogeneous site percolation on Bethe Lattices with two occupation probabilities, and then extend the result to ...Missing: 2D | Show results with:2D
  77. [77]
    [PDF] arXiv:2001.06184v1 [cond-mat.stat-mech] 17 Jan 2020
    Jan 17, 2020 · The region with generic power-laws is known as Griffiths phase. It is known to emerge from rare region effects due to the presence of quenched ...
  78. [78]
    Dynamic phase transition in the contact process with spatial disorder
    Feb 24, 2020 · The regime in which the power-law decay of ρ ( t ) with changing exponents is observed is known as the Griffiths phase. The relaxation behavior ...
  79. [79]
    Percolation and Connectivity in AB Random Geometric Graphs - arXiv
    Apr 1, 2009 · This is a generalization of the AB percolation model on discrete lattices. We show the existence of percolation for all d > 1 and derive bounds ...
  80. [80]
    [1003.5583] Bootstrap Percolation on Complex Networks - arXiv
    Mar 29, 2010 · We consider bootstrap percolation on uncorrelated complex networks. We obtain the phase diagram for this process with respect to two parameters.Missing: dimensions | Show results with:dimensions
  81. [81]
    Bootstrap percolation on complex networks | Phys. Rev. E
    Jul 1, 2010 · We give the phase diagram showing the thresholds with respect to both the extent of damage to the network, and to the size of the initial seed ...
  82. [82]
    Formation of three-dimensional bicontinuous structures via molten ...
    Jun 9, 2021 · This work studies a method of molten salt dealloying via real-time in situ synchrotron three-dimensional X-ray nano-tomography.
  83. [83]
    Tailoring microstructure in a soft-magnetic Fe-based amorphous ...
    A novel Fe-based amorphous alloy is tailored by electrical percolation threshold. Quantitative correlations between microstructure and properties were ...
  84. [84]
    Percolation-based simulation to predict caking kinetics of ...
    We perform particle-based simulations and show that, if caking is mapped into a percolation transition, the role of spatial heterogeneities is well captured.Missing: tomography | Show results with:tomography
  85. [85]
    Composites with carbon nanotubes and graphene: An outlook
    Nov 2, 2018 · The percolation threshold of CNTs in high-density polyethylene matrix is 0.15 wt %, compared to 1 wt % for graphene in the same polymer (24).
  86. [86]
    [PDF] Epidemic dynamics on complex networks - arXiv
    As mentioned in section one, the network SIR model is equal to the bond percolation; and clearly, vaccination for individual can be considered as the site ...Missing: p_c | Show results with:p_c
  87. [87]
    The upper percolation threshold and porosity–permeability ... - Nature
    Jul 4, 2022 · The percolation threshold is defined as the minimum fraction of nodes required to exhibit connectivity in order for a pathway to form between ...
  88. [88]
    Finding influential nodes for integration in brain networks using ...
    Jun 11, 2018 · Thus, optimal percolation theory predicts essential nodes in brain networks. This could be used to identify targets of interventions to modulate ...
  89. [89]
    [PDF] A Monte Carlo simulation of a protein (CoVE) in a matrix of random ...
    Nov 11, 2019 · The threshold porosity (cp = 1- c) for the percolation of the CoVE protein in denatured phase is found to be around 0.9. Note that cp is much ...
  90. [90]
    Emergence of superconductivity in the cuprates via a universal ...
    Oct 18, 2018 · The cuprate superconductors are known to be strongly anisotropic; in the site-percolation model, this translates to anisotropy within the ...
  91. [91]