Fact-checked by Grok 2 weeks ago

Percolation theory

Percolation theory is a mathematical framework in statistical physics and probability that models the emergence of large-scale connectivity in random media, such as the formation of spanning clusters in lattices where sites or bonds are occupied with independent probability p. Introduced by Simon R. Broadbent and John M. Hammersley in 1957, it originated as a model for infiltration through porous structures like crystals or mazes, contrasting with by emphasizing randomness in the medium rather than the propagating agent. The theory examines Bernoulli percolation processes, where the probability p determines whether an infinite connected cluster forms, marking a transition at the critical p_c. Central to percolation theory are two primary models: bond percolation, where edges of a (e.g., the \mathbb{Z}^d) are present independently with probability p, and site percolation, where vertices are occupied with probability p. Below p_c, clusters are finite and localized; at p_c, a unique infinite emerges with properties, exhibiting scaling behaviors characterized by such as the correlation length exponent \nu (approximately 4/3 in two dimensions). Exact values of p_c are known for certain cases, including p_c = 1/2 for bond percolation on the two-dimensional and p_c = 1 in one dimension, where no infinite forms for p < 1. These thresholds connect to broader critical phenomena, analogous to Ising model phase transitions, and have been analyzed using renormalization group methods since the 1970s. Percolation theory finds applications across disciplines, including modeling fluid flow in porous media for oil recovery, where cluster sizes predict extraction efficiency, and electrical conductivity in disordered networks, where the effective conductivity vanishes as (p - p_c)^\mu near the threshold (\mu \approx 1.3 in 2D). In materials science, it describes rigidity percolation in elastic composites, with thresholds linked to mechanical stability, as in the central-force model where p_c \approx 0.66 in 2D for floppy-to-rigid transitions on the bond-diluted triangular lattice. Epidemiological models use it for disease spread on contact networks, while in computer science, it informs fault-tolerant designs in integrated circuits. Extensions include continuum percolation for overlapping particles and bootstrap percolation for dependent occupations, broadening its relevance to complex systems like forest fires or social networks.

Fundamentals

Definition and Scope

Percolation theory is a fundamental framework in statistical physics that examines the connectivity properties of random media, specifically the formation and evolution of connected clusters as elements of the medium become occupied according to a probabilistic rule. In its core setup, consider a lattice or graph where each site or bond is independently occupied (termed "open") with probability p \in [0,1] through Bernoulli trials, or unoccupied ("closed") with probability $1-p. A cluster is defined as a connected component consisting of open sites or bonds linked via nearest-neighbor adjacency, and the theory investigates how these clusters grow and interconnect as p varies, particularly the emergence of a spanning cluster that either traverses the entire system or extends infinitely in the thermodynamic limit. The scope of percolation theory encompasses the study of phase transitions in disordered systems, where randomness in the medium leads to abrupt changes in global connectivity properties, such as the percolation threshold beyond which a giant connected component dominates. Unlike deterministic connectivity problems in , which assume fixed structures, percolation emphasizes stochastic processes in heterogeneous environments, providing insights into emergent phenomena in complex systems. This probabilistic approach distinguishes it as a tool for modeling real-world scenarios involving uncertainty, including the flow of fluids through porous rocks—where open pores allow permeation—and the spread of diseases across a population grid, analogous to infection propagating through connected individuals. By focusing on these connectivity thresholds and cluster statistics, percolation theory bridges mathematics, physics, and applied sciences, offering a versatile model for understanding how local randomness yields macroscopic behavior without relying on specific microscopic interactions.

Historical Development

Early ideas in percolation theory were anticipated in the 1940s by 's theory of polymer gelation, which modeled the formation of infinite networks in branching polymerization processes, establishing a critical threshold for gel formation. In the late 1940s and early 1950s, explored related problems such as the coverage of lines by random intervals, laying groundwork for continuum percolation models. These investigations addressed stochastic distributions in one-dimensional settings, such as the probability of complete coverage or gaps in random placements, which anticipated later applications to fluid flow through disordered media. In the 1950s, the theory gained formal structure through the work of Simon R. Broadbent and , who introduced the in 1957 to simulate the flow of fluids or gases through porous materials, such as charcoal in gas masks. This model also served as an analogy for the spread of diseases in lattice-based structures, like plant blight in orchards arranged on a square grid. The framework formalized random occupation of bonds or sites with probability p, marking the birth of percolation as a probabilistic tool for connectivity in random graphs. The 1960s saw significant formalization, with Hammersley, Harry Kesten, and Robert T. Smythe advancing discrete site and bond percolation on lattices. A pivotal contribution was the introduction of the critical probability p_c, the threshold above which an infinite connected cluster emerges with positive probability, rigorously established through bounds and existence proofs. Theodore E. Harris provided a key lower bound for p_c in 1960 for the square lattice, demonstrating the phase transition from finite to infinite clusters. Kesten's early work further refined these concepts, proving finiteness of moments and connectivity properties. During the 1970s and 1980s, percolation theory deepened its ties to statistical physics, particularly through connections to the via the q-state in the limit q \to 1. Researchers like M. P. M. den Nijs applied renormalization group techniques to analyze critical exponents, revealing scaling behaviors akin to magnetic phase transitions. The universality hypothesis, positing that critical phenomena in percolation depend only on dimensionality and not lattice specifics, was explored by and others, aligning percolation with broader classes of phase transitions. A landmark event was 's 1980 proof that p_c = 1/2 exactly for bond percolation on the two-dimensional square lattice, providing the first precise solution in a non-trivial case. Studies in random media around this era, influenced by advances in probability at international conferences, elevated percolation's mathematical rigor. In the modern era, percolation integrated with fractal geometry following Benoit Mandelbrot's 1982 analysis of cluster structures as self-similar fractals, highlighting their scale-invariant properties at criticality. Post-1990s developments emphasized computational methods, with simulations enabling exploration of higher dimensions and scaling limits, alongside exact results like uniqueness of the infinite cluster above p_c.

Core Models

Discrete Percolation Models

Discrete percolation models form the foundational framework of percolation theory, focusing on lattice structures where connectivity is determined by random occupations of sites or bonds. These models abstract the flow of fluids through porous media or the spread of information in networks by assigning probabilistic openness to elements of a regular grid. Introduced in the seminal work by , they emphasize independent Bernoulli trials for occupation, leading to cluster formations that reveal connectivity patterns. In bond percolation, each edge (bond) of a lattice graph is independently occupied (open) with probability p \in [0,1], and closed otherwise; two vertices are connected if there exists a path between them consisting entirely of open bonds. This model captures scenarios where connections between sites are probabilistic, such as cracks in a material or links in a communication network. Clusters are defined as the connected components under this openness rule, with percolation occurring when an infinite cluster emerges. Site percolation, in contrast, involves independently occupying each vertex (site) of the lattice with probability p, where clusters form from adjacent open sites sharing an edge. Adjacency is typically defined by a neighborhood structure, such as the (four orthogonal neighbors on a square lattice) or the (eight neighbors, including diagonals). This variant models occupations of positions themselves, like occupied pores in a filter or active nodes in a grid-based system. These models are studied on various regular lattices, with the square, triangular, and honeycomb lattices serving as canonical examples due to their symmetry and analyzable properties. On the square lattice, bond percolation exhibits self-duality, implying a critical occupation probability of exactly p_c = 1/2. The triangular and honeycomb lattices form a dual pair, where the bond percolation threshold on one relates inversely to the other via duality relations established by Sykes and Essam. Mixed models extend the basic frameworks by introducing dependencies or alternative rules, such as correlated percolation where occupations are not independent, or bootstrap percolation where sites become occupied if they have a sufficient number of already open neighbors, leading to growth dynamics from an initial seed configuration. , introduced by Chalupa, Leath, and Reich, models irreversible activation processes on lattices, with clusters expanding monotonically under threshold rules. Mathematically, these models are formulated using a probability measure \mathbb{P}_p on the configuration space \Omega, where \Omega = \{0,1\}^E for bond percolation (with E the edge set) or \Omega = \{0,1\}^V for site percolation (with V the vertex set); each element is 1 (open) with probability p independently. The percolation event is captured by \Theta(p) = \mathbb{P}_p(\text{there exists a spanning cluster}), quantifying the probability of global connectivity in finite or infinite lattices.

Continuum and Other Variants

Continuum percolation extends the concepts of discrete percolation to continuous spaces, such as \mathbb{R}^d, where connectivity arises from the overlap of randomly placed shapes rather than fixed lattice sites. In this framework, the occupied set is typically formed by the union of grains—such as disks in 2D or spheres in 3D—centered at points drawn from a Poisson point process with intensity \lambda. The analog to the occupation probability p in discrete models is the reduced density \eta = \lambda \times V, where V is the volume of a typical grain; percolation occurs above a critical \eta_c, with simulations indicating \eta_c \approx 1.13 in 2D for equal-radius disks (corresponding to a covered area fraction of approximately 0.68). This model captures phenomena like the formation of infinite clusters in random media without underlying discreteness. The Boolean model is a foundational instance of continuum percolation, where germs (Poisson points) are each associated with an identical grain, such as a ball of fixed radius r, and the percolating set is their union; the covered volume fraction at criticality is \phi_c = 1 - e^{-\eta_c}, with \phi_c \approx 0.68 in 2D and \approx 0.29 in 3D for spheres. More generally, germ-grain models allow arbitrary grains (possibly random and non-spherical) attached to germs from a point process, encompassing both overlapping and non-overlapping configurations; these generalize the Boolean model by permitting complex grain shapes while preserving the Poisson-driven randomness. Interpretations include the occupancy view (union of grains as occupied) and the vacancy view (complement as unoccupied voids), both yielding phase transitions analogous to discrete cases but in infinite domains. Directed percolation introduces anisotropy by orienting connections, such as arrows on bonds that permit flow only in specified directions (e.g., forward in time-like dimensions), leading to distinct critical behavior from isotropic models. This variant models processes like epidemic spreading or fluid invasion with preferred directions and belongs to its own universality class, characterized by exponents differing from standard percolation, as established through field-theoretic analyses and simulations. Seminal work by Grassberger and Janssen highlighted its role in nonequilibrium phase transitions with absorbing states. Other variants include percolation on trees, such as the Bethe lattice with coordination number z, where exact solvability yields a critical probability p_c = 1/(z-1) via recursive branching arguments, serving as a mean-field benchmark without loops. Long-range percolation modifies connectivity by allowing bonds between sites at distance r with probability decaying as $1/r^{d+\sigma} (\sigma > 0), altering effective dimensionality: for \sigma < 2, it exhibits mean-field-like behavior with faster cluster growth, while \sigma > 2 recovers short-range universality. Invasion percolation, a dynamic greedy process, simulates slow displacement by iteratively occupying the lowest-threshold pore among accessible ones, naturally reaching criticality without tuning parameters and producing clusters akin to critical percolation.

Phase Transitions

Subcritical and Supercritical Regimes

In the subcritical regime, where the occupation probability p is less than the p_c, all clusters are finite , and the probability of an infinite open cluster is zero, denoted \theta(p) = 0. The distribution of cluster sizes exhibits , with the probability that the cluster containing a given site has size at least n bounded by P_p[|C(v)| \geq n] \leq C_1 e^{-C_2 n} for positive constants C_1, C_2 > 0 depending on p and the . In finite systems of size N, the largest cluster scales as O(\log N) in , reflecting the absence of spanning clusters and the dominance of small, localized components. The mean cluster size \chi(p) = \mathbb{E}_p[|C(v)|] remains finite but diverges as p \to p_c^-, signaling the approach to the . In the supercritical regime, for p > p_c, an infinite open cluster emerges with positive probability \theta(p) > 0, defined as the limiting probability that a given site belongs to the infinite cluster. This infinite cluster is unique almost surely in \mathbb{Z}^d for d \geq 2, and the density \theta(p) provides a measure of the fraction of sites connected to it. Finite clusters coexist with the infinite one, but their sizes decay exponentially, similar to the subcritical case. Near the critical point, \theta(p) \approx (p - p_c)^\beta for p > p_c, where \beta > 0 is a critical exponent characterizing the order parameter's onset (detailed scaling laws are discussed in the section). The correlation length \xi(p), which governs the typical extent of connectivity, scales as \xi(p) \sim |p - p_c|^{-\nu} with \nu > 0, diverging on both sides of p_c. Key theoretical insights include the Russo-Seymour-Welsh (RSW) theory in two dimensions, which establishes bounds on crossing probabilities in rectangular regions, ensuring stability of the supercritical phase and uniform percolation probabilities away from boundaries.90037-0) This framework underpins proofs of \theta(p_c) = 0 in and facilitates understanding of arm events and interface properties in the supercritical regime.

Critical Phenomena and Exponents

At the percolation threshold p = p_c, the order parameter \theta(p_c) = 0, and there is no infinite cluster (proven for d = 2 and d \geq 19, conjectured otherwise as of 2025), yet the system exhibits critical fluctuations with diverging and power-law cluster size distributions, marking the onset of long-range correlations. Simultaneously, the \chi(p_c), defined as the mean size of finite s, diverges to infinity, signaling critical fluctuations. These behaviors exemplify the singular central to percolation theory, where geometric and statistical properties exhibit scale-invariant, fractal-like structures. Percolation exhibits universality, wherein critical behaviors depend solely on the spatial dimension d and the range of interactions, rather than microscopic details like lattice type. Distinct universality classes arise, such as the two-dimensional (2D) class characterized by exact solvability via conformal invariance, contrasting with the three-dimensional (3D) class, which relies on numerical estimates. Hyperscaling, valid below the upper critical dimension d = 6, relates exponents through $2 - \alpha = d \nu, where \alpha governs the specific heat analog and \nu the correlation length; this fails in mean-field regimes for d \geq 6. Key scaling exponents quantify these singularities. The order parameter exponent \beta describes \theta(p) \sim (p - p_c)^\beta for p > p_c, with exact 2D value \beta = 5/36 and 3D estimate \beta \approx 0.41. The correlation length exponent \nu captures \xi \sim |p - p_c|^{-\nu}, yielding \nu = 4/3 exactly in 2D and \nu \approx 0.88 in 3D. Susceptibility diverges as \chi \sim |p - p_c|^{-\gamma}, with \gamma = 43/18 in 2D and \gamma \approx 1.8 in 3D. For cluster size distribution at criticality, P(s) \sim s^{-\tau}, where \tau = 187/91 \approx 2.055 in 2D and \tau \approx 2.18 in 3D. Scaling forms encapsulate these relations, such as the cluster size probability P(s,p) \sim s^{-\tau} f((p - p_c) s^\sigma), with \sigma = 1/(\beta + \gamma). Fisher relations link exponents, including \tau = 2 + 1/(\delta + 1), where \delta = \gamma / \beta describes the magnetization analog at criticality.90060-7) Renormalization group theory provides insights by identifying the critical fixed point, enabling derivation of universality and hyperscaling through iterative coarse-graining. Critical clusters display fractal geometry, with the incipient infinite cluster having fractal dimension d_f = 91/48 \approx 1.896 in 2D. The hull, or external perimeter, exhibits dimension D_h = 7/4 = 1.75 in 2D, reflecting self-similar boundary roughness. These properties underscore the scale-free nature of percolation at p_c, with renormalization group flows confirming the irrelevance of short-range details in determining long-wavelength behavior.

Analytical and Computational Approaches

Threshold Determination

Determining the p_c, the critical occupation probability at which an infinite cluster emerges, is a central challenge in percolation theory. Analytical methods provide exact values for select models, bounds for general cases, and approximations via expansions or mean-field treatments. These approaches exploit symmetries, inequalities, and diagrammatic conditions to pinpoint or constrain p_c without relying on simulations. Exact results are available for specific low-dimensional lattices due to duality and matching properties. For bond percolation on the two-dimensional , duality between the lattice and its implies that the threshold occurs where the probability of connection equals the probability of blockage, yielding p_c = [1](/page/1)/2. Similarly, for site percolation on the two-dimensional triangular lattice, a star-triangle and matching lattice arguments establish p_c = [1](/page/1)/2. On the , an infinite tree with z, the absence of loops allows an exact recursive solution for the probability of finite clusters, giving p_c = [1](/page/1)/([z](/page/1)-[1](/page/1)). In higher dimensions or more complex lattices, exact solutions are rare, so bounds and approximations are employed. The Aizenman-Newman , derived from tree-graph approximations to the , provides a lower bound for percolation on the d-dimensional hypercubic : p_c \geq 1/(2d-1). Self-dual approximations extend duality ideas to non-self-dual lattices by constructing equivalent self-dual hypergraphs, yielding estimates like p_c \approx 1/2 for certain quasi-planar models where exact duality does not apply directly. Series expansions offer a perturbative to locate p_c by analyzing the radius of convergence of for quantities like the mean size. High-temperature expansions (analogous to low-p regimes) generate coefficients for the \chi(p) = \sum_s s^2 n_s(p), where n_s(p) is the density of s-, and singularities indicate p_c. Low-temperature expansions (high-p side) similarly probe the order parameter via connectivity probabilities. These methods, rooted in enumerations, have been applied to lattices like the square and to refine estimates. Mean-field theory simplifies the problem by neglecting spatial correlations, treating clusters as branching processes on the , which predicts p_c = 1/(z-1) for coordination number z. This approximation becomes exact above the upper d_c = 6, where fluctuations are suppressed, and critical behavior matches mean-field values. The triangle condition, a diagrammatic from lace , \sum_{x,y} \tau(x)\tau(y)\tau(|x-y|) < \infty near criticality (where \tau(r) is the two-point connectivity), confirms mean-field validity when it holds, as in d > 6. Analogies to self-avoiding walks () aid in bounding p_c, as the percolation relates to the SAW , providing inequalities like those from connective constant estimates to constrain thresholds in dimensions where duality fails.

Simulation Techniques

Monte Carlo methods form the cornerstone of numerical simulations in percolation theory, enabling the estimation of key quantities such as the percolation strength \theta(p), which represents the probability that a site belongs to the infinite cluster, and the \chi(p), the mean size of finite clusters, through repeated fixed-occupation probability p runs on finite lattices. These simulations generate random configurations of occupied sites or bonds and identify connected clusters using union-find data structures, averaging observables over many independent realizations to reduce statistical noise. To determine the critical occupation probability p_c in the , finite-size scaling analysis is applied, leveraging the divergence of the correlation length \xi \sim |p - p_c|^{-\nu} near criticality; for finite linear system size L, quantities like the effective p_c(L) shift as p_c(L) - p_c \sim L^{-1/\nu}, allowing extrapolation via fits to data from multiple L values. This approach, pioneered in early studies of lattice percolation, provides high-precision estimates of p_c and by analyzing how observables scale with L at fixed p near p_c. Efficient cluster enumeration is crucial for large-scale simulations, and the Newman-Ziff algorithm achieves this by incrementally occupying sites in random order while maintaining a dynamic record of cluster sizes and mergers through a union-find structure with path compression, computing spanning probabilities and moments for all p from 0 to 1 in linear time O(N). For dynamical variants, invasion percolation simulates slow fluid invasion by repeatedly selecting the lowest-threshold pore at the cluster boundary, modeling gradient-driven processes without fixed p, and revealing invasion clusters with the same universality as static percolation. Renormalization simulations approximate the flow numerically via block spinning, where coarse-graining transforms a fine of b into an effective coarse by defining a site as occupied if a (or other rule) of its b^d sub-sites are connected, yielding relations for effective p'(p) and fixed points to estimate p_c and exponents. Large-cell Monte Carlo variants enhance accuracy by averaging over many realizations within each block, mitigating finite-size effects and confirming universality across models. High-performance computing adaptations include parallel labeling algorithms that partition the into subdomains assigned to processors, performing local Hoshen-Kopelman labeling followed by inter-processor of identities via relaxation iterations, enabling simulations on lattices up to $10^4 sites in high dimensions. Error analysis in these simulations often employs , where subsets of configurations are omitted to compute variance in estimates like p_c, providing unbiased for finite-sample statistics. Recent advances post-2010 include GPU-accelerated implementations that parallelize , site occupation, and cluster identification across thousands of threads, achieving speedups of 10-100x for 2D studies and enabling analysis of larger systems for estimation. Extensions of the Newman-Ziff algorithm to bootstrap variants, where clusters grow by adding sites meeting a neighborhood , further optimize simulations on complex lattices like Archimedean tilings by tracking activation cascades efficiently.

Applications

In Physics and Materials Science

In physics and materials science, percolation theory provides a framework for understanding transport and structural properties in disordered systems, particularly near the percolation threshold where connectivity emerges. One key application is in modeling electrical conductivity through random resistor networks, where bonds or sites are randomly occupied with conducting or insulating elements. In these networks, the effective conductivity vanishes below the percolation threshold p_c and follows a power-law behavior above it, \sigma \sim (p - p_c)^\mu, capturing the onset of long-range connectivity. Effective medium theory approximates this conductivity by treating the network as a homogeneous medium with an averaged conductivity, offering reasonable predictions away from the threshold but underestimating the critical behavior near p_c. In two dimensions, numerical studies confirm the critical exponent \mu \approx 1.3, highlighting the universality of this scaling in lattice models. Percolation also informs magnetic properties in diluted ferromagnets, where magnetic ions are randomly placed on a . The percolation threshold determines the onset of an infinite magnetic , enabling long-range order, but this geometric transition precedes the thermal Ising transition that governs below a critical . In such systems, the diluted exhibits a percolation-driven loss of at p_c, distinct from the Ising model's cooperative alignment, as dilution disrupts connectivity before thermal dominates. This distinction is crucial for understanding phase diagrams in materials like halides, where site or bond dilution shifts the magnetic response. In through porous media, percolation theory models under , \mathbf{q} = -\kappa \nabla h, where flow emerges only above p_c due to connected pathways. The effective scales as \kappa \sim (p - p_c)^t, with the exponent t \approx 1.3 in and t \approx 2 in 3D, reflecting and effects in heterogeneous rocks or soils. This scaling aids upscaling from microscopic structures to macroscopic flow properties, validated by critical path analysis. Fracture mechanics employs invasion percolation to describe crack propagation in brittle materials, where fluid invasion or stress selects the path of least resistance, mimicking unstable growth. Introduced as a variant of standard percolation, this model simulates dendritic fracture patterns by iteratively filling pores or bonds with the lowest threshold, leading to ramified clusters without loops. In rocks, it captures subcritical crack advance under hydraulic pressure, relevant to enhanced oil recovery or geothermal systems. Representative examples include (CNT) composites, where governs electrical conductivity in matrices; aligned CNTs lower the threshold to below 1 wt%, enabling lightweight conductors for . In geological reservoirs, quantifies fluid flow connectivity in heterogeneous sandstones, predicting breakthrough times and permeability from distributions, as in fluvial or formations. Recent applications in the 2020s extend to electrode design, where optimizes ion and pathways in composite cathodes; for instance, single-walled CNTs as fillers in NCM electrodes enhance conductivity near the threshold, improving and rate performance in lithium-ion systems.

In Biological and Ecological Systems

Percolation theory has been applied to by mapping susceptible-infected-recovered () models on lattices to bond percolation processes, where the occupation probability corresponds to the probability of between neighboring sites. In this framework, the threshold aligns with the p_c, below which isolated infection clusters form without widespread outbreaks, and above which a giant emerges, representing a large-scale . This analogy allows for the analysis of spatial spread in structured populations, such as on square lattices, where from describe the size distribution of outbreak clusters near the threshold. In ecological systems, percolation models by treating landscapes as random lattices where occupied sites represent suitable patches, and determines dispersal viability. As approaches the , a spanning cluster forms, enabling long-range dispersal and persistence; below this threshold, fragmentation isolates small clusters, increasing risk for reliant on . These models predict critical abundance levels for distributions, where fragmented patches coalesce into viable networks only above specific fractions, informing strategies against loss. Percolation theory elucidates virus assembly by analyzing protein interaction graphs in capsid formation, where subunits connect via bonds to form stable shells analogous to percolating clusters. In this view, the geometric layout and interaction of viral proteins exhibit percolation transitions, with the critical probability governing the emergence of a complete capsid structure from incomplete assemblies. Removing subunits disrupts , leading to fragmentation below the , which mirrors observed in virus-like particles and highlights biophysical constraints on assembly efficiency. Biochemical processes in crowded cellular environments leverage to model and reaction networks, where macromolecules form disordered lattices that hinder or facilitate connectivity. In such settings, clusters enhance catalytic efficiency by creating extended networks for substrate access, particularly near the where shredded structures maximize interaction surfaces. Crowding alters diffusion-limited reaction rates, with -based simulations showing reduced enzymatic activity below critical densities due to isolated clusters, underscoring the role of in cellular . Representative applications include forest fire spread models, where trees on a ignite with probability p, and fire propagates through connected occupied sites until the percolation threshold determines whether the blaze spans the entire forest. In coral reef systems, percolation analogs assess for larval dispersal, treating reef patches as sites in a marine to identify fragmentation thresholds that isolate populations and reduce genetic exchange. Studies on COVID-19 spatial transmission in the 2020s have used percolation to map epidemic progression across regions, revealing directed percolation waves where case clusters expand critically until containment measures shift occupancy below p_c, as observed in Chile's outbreak dynamics.

In Networks and Social Sciences

In network percolation, the emergence of a in random graphs serves as a foundational example of phase transitions in abstract structures. In the , where edges are present independently with probability p, a spanning a finite fraction of vertices arises when the average degree \langle k \rangle = np = 1, marking the ; below this value, components remain small and tree-like, while above it, a macroscopic cluster dominates the graph's connectivity. This threshold highlights the abrupt shift from fragmentation to cohesion, analogous to fluidity in but applied to relational ties in networks. k-core percolation extends this by iteratively removing nodes with degree less than k, revealing the network's resilient backbone; for , the k-core threshold occurs at a critical average degree \langle k \rangle_c satisfying the self-consistent equation u = \exp(\langle k \rangle (u^{k-1} - 1)) having a solution u < 1, with \langle k \rangle_c \sim k \ln k for large k, and the process yielding a discontinuous transition in heterogeneous networks. Social contagion processes leverage to model the spread of information, behaviors, or diseases across human networks, where adoption depends on local influences rather than random bonds. In Watts' , individuals adopt a if a fraction r of their neighbors have it, leading to global cascades from a small when the network's distribution and threshold variance align near the point; simulations on random graphs show cascades spanning up to 75% of nodes for r \approx 0.1 and broad degree heterogeneity. This framework captures phenomena like viral trends or opinion shifts, distinct from independent by incorporating endogenous activation rules that amplify clustering. In scale-free networks with power-law degree distributions P(k) \sim k^{-\gamma}, robustness against random failures is enhanced for \gamma < 3, with the bond p_c \approx 1/(\langle k^2 \rangle / \langle k \rangle - 1), explaining the Internet's resilience to breakdowns but vulnerability to targeted attacks on hubs. Applications in social sciences underscore percolation's role in dissecting systemic risks and dynamics. Financial contagion during the 2008 crisis has been analogized to percolation cascades in interbank networks, where liquidity shocks propagate through overlapping exposures, leading to widespread defaults if the fraction of failed links exceeds a connectivity threshold; models reveal that dense core structures amplify contagion, with up to 20% node removal triggering systemic collapse in empirical datasets. On social media, echo chambers emerge as densely connected clusters in user interaction graphs, fostering polarized information flow akin to subcritical percolation regimes where low bridge probabilities isolate communities. Studies of misinformation dynamics show false narratives spreading faster than facts in polarized networks, forming giant components more readily due to community structures. Variants like percolation on scale-free networks further adapt these insights, while hybrid discrete-continuum models integrate graph edges with spatial influences for urban social ties, blending bond occupation with density gradients to predict community fragmentation.

References

  1. [1]
    [PDF] 1 What is Percolation?
    The origins of the mathematical theory of percolation may be found in the work of Flory (1941), Broadbent (1954), and Broadbent and Hammersley. (1957).
  2. [2]
    [PDF] Percolation processes
    Oct 24, 2008 · S. R. Broadbent and J. M. Hammersley (1957). Percolation processes. Mathematical. Proceedings of the Cambridge Philosophical Society, 53, pp ...
  3. [3]
    [PDF] Percolation Theory - MIT
    Oct 9, 2002 · Percolation theory is the simplest model displaying a phase transition. The analytic solutions to 1d and mean-field percolation are presented.
  4. [4]
    [PDF] Introduction to Percolation N Giordano -- Purdue University
    What is percolation? • The percolation threshold - connection with phase transitions and critical phenomena. • Fractals and fractal scaling. ➤ upscaling from ...
  5. [5]
    Percolation processes | Mathematical Proceedings of the ...
    Percolation studies how a medium's random properties influence a fluid's percolation, such as solute diffusing through solvent.
  6. [6]
    Percolation, statistical topography, and transport in random media
    Oct 1, 1992 · A review of classical percolation theory is presented, with an emphasis on novel applications to statistical topography, turbulent diffusion, and heterogeneous ...
  7. [7]
    Percolation and Epidemic Thresholds in Clustered Networks
    Aug 23, 2006 · Indeed, the susceptible-infected-removed (SIR) model of epidemic spreading can be mapped into a bond percolation problem [5–8] .
  8. [8]
    [PDF] On Hammersley's Method for One-Dimensional Covering Problems
    1940's and early 1950's. I will relate these problems to one dimensional continuum percolation, a subject which has attracted interest and attention.
  9. [9]
  10. [10]
  11. [11]
    [PDF] Percolation theory - arXiv
    Jul 11, 2005 · Percolation as a mathematical theory was intro- duced by Broadbent and Hammersley [4], as a stochastic way of modeling the flow of a fluid or ...
  12. [12]
    Moore neighborhood (eight neighbors) and Von Neumann ...
    A model where we open and close vertices rather than edges, is called site percolation (figure 2 a) while the model described above is more properly called bond ...
  13. [13]
    The critical probability of bond percolation on the square lattice ...
    1980 The critical probability of bond percolation on the square lattice equals 12 1 2. Harry Kesten · DOWNLOAD PDF + SAVE TO MY LIBRARY. Comm. Math. Phys.
  14. [14]
    Exact Critical Percolation Probabilities for Site and Bond Problems ...
    An exact method for determining the critical percolation probability, pc, for a number of two‐dimensional site and bond problems is described.
  15. [15]
    Germ-Grain Model -- from Wolfram MathWorld
    In continuum percolation theory, the so-called germ-grain model is an obvious generalization of both the Boolean and Boolean-Poisson models which is driven ...
  16. [16]
    Universality classes in nonequilibrium lattice systems
    Aug 17, 2004 · This article reviews our present knowledge of universality classes in nonequilibrium systems defined on regular lattices.Missing: seminal paper
  17. [17]
    Scaling behavior of the directed percolation universality class - arXiv
    May 10, 2005 · Abstract: In this work we consider five different lattice models which exhibit continuous phase transitions into absorbing states.Missing: seminal | Show results with:seminal
  18. [18]
    Long-Range Percolation Model -- from Wolfram MathWorld
    Intuitively, a d -dimensional discrete percolation model is said to be long-range if direct flow is possible between pairs of graph vertices or graph edges ...
  19. [19]
    Invasion percolation: a new form of percolation theory - IOPscience
    A new kind of percolation problem is described which differs from ordinary percolation theory in that it automatically finds the critical points of the system.
  20. [20]
  21. [21]
  22. [22]
    Uniqueness of the infinite cluster and continuity of connectivity ...
    Feb 19, 1987 · For independent translation-invariant irreducible percolation models, it is proved that the infinite cluster, when it exists, must be unique ...
  23. [23]
  24. [24]
  25. [25]
  26. [26]
    Fast Monte Carlo algorithm for site or bond percolation | Phys. Rev. E
    Jun 27, 2001 · We describe in detail an efficient algorithm for studying site or bond percolation on any lattice. The algorithm can measure an observable quantity in a ...Missing: original paper
  27. [27]
    Parallel cluster labeling for large-scale Monte Carlo simulations
    The algorithm divides physical space into rectangular cells which are assigned to processors and combines a serial local labeling procedure with a relaxation ...
  28. [28]
    Newman-Ziff algorithm for the bootstrap percolation
    Jun 1, 2019 · We propose very efficient algorithms for the bootstrap percolation and the diffusion percolation models by extending the Newman-Ziff algorithm of the classical ...Missing: enumeration | Show results with:enumeration
  29. [29]
    Percolation and Conduction | Rev. Mod. Phys.
    Oct 1, 1973 · A "Green's function" derivation of the effective medium theory, which makes contact with similar treatments of disordered alloys, is presented.<|separator|>
  30. [30]
  31. [31]
    Phase transitions in diluted magnets: Critical behavior, percolation ...
    Transition metal halides provide realizations of Ising,XY, and Heisenberg antiferromagnets in one, two, and three dimensions.
  32. [32]
    [PDF] Invasion-percolation fracturing during primary migration - arXiv
    [2010] introduced a fracture mechanics model of subcritical. 66 crack propagation and coalescence, based on the assumption of linear elastic behavior of the.
  33. [33]
    Effect of nanotube alignment on percolation conductivity in carbon ...
    Sep 19, 2005 · Percolation conductivity of a stick network depends on alignment as well as concentration. We show that both dependences exhibit critical (power-law) behavior.
  34. [34]
    Percolation Theory for Flow in Porous Media - SpringerLink
    Nov 15, 2005 · "Allan Hunt's book Percolation Theory for Flow in Porous Media provides an overview of how percolation theory … can be used to calculate ...
  35. [35]
    Improving High-Energy Lithium-Ion Batteries with Carbon Filler
    Nov 10, 2020 · The SWCNTs were found to be the best conductive filler for NCM electrodes. “The measured conductivity is consistent with percolation theory …
  36. [36]
    Epidemics, the Ising-model and percolation theory - PubMed Central
    In the frame of elementary percolation theory, we show that: (i) the percolation probability can be associated with the probability of a person being infected ...
  37. [37]
    Epidemic Percolation Networks, Epidemic Outcomes, and ...
    Epidemic percolation networks (EPNs) are directed random networks that can be used to analyze stochastic “Susceptible-Infectious-Removed” (SIR) and ...
  38. [38]
    Detecting Critical Scales in Fragmented Landscapes
    Percolation theory is the study of connectivity in stochastically generated structures (Stauffer and Aharony 1985). Thus, connectivity measures developed in ...
  39. [39]
    Habitat fragmentation, percolation theory and the conservation of a ...
    In this paper, we use percolation theory to consider how an environment may fragment. We then develop a stochastic, spatially explicit, individual–based model ...
  40. [40]
    Percolation Theory for the Distribution and Abundance of Species
    The models predict that species distributions will exhibit percolation critical thresholds, i.e., critical population abundances at which the fragmented patches ...
  41. [41]
    Percolation Theory Reveals Biophysical Properties of Virus-like ...
    Jul 23, 2021 · Percolation theory shows that a viral capsid's geometric characteristics, including subunit layout and interaction network, are key for its ...
  42. [42]
    The Percolation Phase Transition (Collapse) in Virus Capsids - NIH
    Here, we apply a graph- and percolation theory to examine the effect of removing capsid subunits on capsid stability and fragmentation. Based on the structure ...Missing: formation | Show results with:formation
  43. [43]
    Function of metabolic and organelle networks in crowded and ... - NIH
    It was further conjectured that as a highly shredded object, a percolation cluster may, in principle, bestow more catalytic power to cytoplasmic enzymes.
  44. [44]
    [PDF] Effect of macromolecular crowding on the rate of diffusion-limited ...
    The effect of macromolecular crowding on the enzymatic reaction rates has been modelled by an MC algorithm based on diffusion and percolation phenomena. The ...
  45. [45]
    Effect of macromolecular crowding on the rate of diffusion-limited ...
    Feb 13, 2009 · A simple numerical model is proposed here based on percolation and diffusion in disordered systems to study the effect of macromolecular ...
  46. [46]
    Percolation Theory and Fire Spread - Taylor & Francis Online
    Abstract. Percolation theory deals with the statistics of random arrays. Two-dimensional percolation theory has been used as a model of forest fires.
  47. [47]
    Spatiotemporal Analysis of the Spread of the COVID-19 Epidemic in ...
    Mar 12, 2025 · Percolation describes the critical behaviour of spatial cells that progressively change their state until they compromise an entire given ...Missing: 2020s | Show results with:2020s
  48. [48]
    [PDF] ON THE EVOLUTION OF RANDOM GRAPHS - SNAP: Stanford
    The evolution of random graphs may be considered as a (rather simplified) model of the evolution of certain real communication-nets,. e. g. the railway-, road- ...
  49. [49]
    k-core (bootstrap) percolation on complex networks - cond-mat - arXiv
    Feb 28, 2006 · Abstract: We develop the theory of the k-core (bootstrap) percolation on uncorrelated random networks with arbitrary degree distributions.
  50. [50]
    A simple model of global cascades on random networks - PNAS
    (i) Unlike epidemiological models, where contagion events between pairs of individuals are independent, the threshold rule effectively introduces local ...
  51. [51]
    [PDF] Financial Networks and Contagion
    We study cascades of failures in a network of interdependent finan- cial organizations: how discontinuous changes in asset values.<|control11|><|separator|>
  52. [52]
    The spreading of misinformation online - PNAS
    Finally, we introduce a data-driven percolation model mimicking rumor spreading and we show that homogeneity and polarization are the main determinants for ...
  53. [53]
    Percolation on complex networks: Theory and application - arXiv
    Jan 28, 2021 · The intention of this paper is to offer an overview of these applications, as well as the basic theory of percolation transition on network ...