Percolation theory

Last updated

In statistical physics and mathematics, percolation theory describes the behavior of a network when nodes or links are added. This is a geometric type of phase transition, since at a critical fraction of addition the network of small, disconnected clusters merge into significantly larger connected, so-called spanning clusters. The applications of percolation theory to materials science and in many other disciplines are discussed here and in the articles Network theory and Percolation (cognitive psychology).

Contents

Introduction

A three-dimensional site percolation graph Perc-wiki.png
A three-dimensional site percolation graph
Bond percolation in a square lattice from p=0.3 to p=0.52 Transition de percolation 2.gif
Bond percolation in a square lattice from p=0.3 to p=0.52

A representative question (and the source of the name) is as follows. Assume that some liquid is poured on top of some porous material. Will the liquid be able to make its way from hole to hole and reach the bottom? This physical question is modelled mathematically as a three-dimensional network of n × n × n vertices, usually called "sites", in which the edge or "bonds" between each two neighbors may be open (allowing the liquid through) with probability p, or closed with probability 1 – p, and they are assumed to be independent. Therefore, for a given p, what is the probability that an open path (meaning a path, each of whose links is an "open" bond) exists from the top to the bottom? The behavior for large n is of primary interest. This problem, called now bond percolation, was introduced in the mathematics literature by Broadbent & Hammersley (1957), [1] and has been studied intensively by mathematicians and physicists since then.

In a slightly different mathematical model for obtaining a random graph, a site is "occupied" with probability p or "empty" (in which case its edges are removed) with probability 1 – p; the corresponding problem is called site percolation. The question is the same: for a given p, what is the probability that a path exists between top and bottom? Similarly, one can ask, given a connected graph at what fraction 1 – p of failures the graph will become disconnected (no large component).

A 3D tube network percolation determination Tube Network Percolation.gif
A 3D tube network percolation determination

The same questions can be asked for any lattice dimension. As is quite typical, it is actually easier to examine infinite networks than just large ones. In this case the corresponding question is: does an infinite open cluster exist? That is, is there a path of connected points of infinite length "through" the network? By Kolmogorov's zero–one law, for any given p, the probability that an infinite cluster exists is either zero or one. Since this probability is an increasing function of p (proof via coupling argument), there must be a criticalp (denoted by pc) below which the probability is always 0 and above which the probability is always 1. In practice, this criticality is very easy to observe. Even for n as small as 100, the probability of an open path from the top to the bottom increases sharply from very close to zero to very close to one in a short span of values of p.

Detail of a bond percolation on the square lattice in two dimensions with percolation probability p = 0.51 Bond percolation p 51.png
Detail of a bond percolation on the square lattice in two dimensions with percolation probability p = 0.51

History

The Flory–Stockmayer theory was the first theory investigating percolation processes. [2]

The history of the percolation model as we know it has its root in the coal industry. Since the industrial revolution, the economical importance of this source of energy fostered many scientific studies to understand its composition and optimize its use. During the 30' and 40'[ when? ], the qualitative analysis by organic chemistry left more and more room to more quantitative studies. [3]

In this context, the British Coal Utilisation Research Association (BCURA) was created in 1938. It is a research association funded by the coal mines owners. In 1942, Rosalind Franklin, who then recently graduated in chemistry from the university of Cambridge, joined the BCURA. She started research on the density and porosity of coal. During the Second World War, coal was an important strategic resource. It was used as a source of energy, but also was the main constituent of gas masks.

Coal is a porous medium. To measure its 'real' density, one was to sink it in a liquid or a gas whose molecules are small enough to fill its microscopic pores. While trying to measure the density of coal using several gases (helium, methanol, hexane, benzene), and as she found different values depending on the gas used, Rosalind Franklin showed that the pores of coal are made of microstructures of various lengths that act as a microscopic sieve to discriminate the gases. She also discovered that the size of these structures depends on the temperature of carbonation during the coal production. With this research, she obtained a PhD degree and left the BCURA in 1946. [4]

In the mid fifties, Simon Broadbent worked in the BCURA as a statistician. Among other interests, he studied the use of coal in gas masks. One question is to understand how a fluid can diffuse in the coal pores, modeled as a random maze of open or closed tunnels. In 1954, during a symposium on Monte Carlo methods, he asks questions to John Hammersley on the use of numerical methods to analyze this model. [5]

Broadbent and Hammersley introduced in their article of 1957 a mathematical model to model this phenomenon, that is percolation.

Computation of the critical parameter

For most infinite lattice graphs, pc cannot be calculated exactly, though in some cases pc there is an exact value. For example:

Percolation front Front de percolation.png
Percolation front

[11]

This indicates that for a given degree distribution, the clustering leads to a larger percolation threshold, mainly because for a fixed number of links, the clustering structure reinforces the core of the network with the price of diluting the global connections. For networks with high clustering, strong clustering could induce the core–periphery structure, in which the core and periphery might percolate at different critical points, and the above approximate treatment is not applicable. [12]

Universality

The universality principle states that the numerical value of pc is determined by the local structure of the graph, whereas the behavior near the critical threshold, pc, is characterized by universal critical exponents. For example the distribution of the size of clusters at criticality decays as a power law with the same exponent for all 2d lattices. This universality means that for a given dimension, the various critical exponents, the fractal dimension of the clusters at pc is independent of the lattice type and percolation type (e.g., bond or site). However, recently percolation has been performed on a weighted planar stochastic lattice (WPSL) and found that although the dimension of the WPSL coincides with the dimension of the space where it is embedded, its universality class is different from that of all the known planar lattices. [13] [14]

Phases

Subcritical and supercritical

The main fact in the subcritical phase is "exponential decay". That is, when p < pc, the probability that a specific point (for example, the origin) is contained in an open cluster (meaning a maximal connected set of "open" edges of the graph) of size r decays to zero exponentially in r. This was proved for percolation in three and more dimensions by Menshikov (1986) and independently by Aizenman & Barsky (1987). In two dimensions, it formed part of Kesten's proof that pc = 1/2. [15]

The dual graph of the square lattice 2 is also the square lattice. It follows that, in two dimensions, the supercritical phase is dual to a subcritical percolation process. This provides essentially full information about the supercritical model with d = 2. The main result for the supercritical phase in three and more dimensions is that, for sufficiently large N, there is[ clarification needed ] an infinite open cluster in the two-dimensional slab 2 × [0, N]d − 2. This was proved by Grimmett & Marstrand (1990). [16]

In two dimensions with p < 1/2, there is with probability one a unique infinite closed cluster (a closed cluster is a maximal connected set of "closed" edges of the graph). Thus the subcritical phase may be described as finite open islands in an infinite closed ocean. When p > 1/2 just the opposite occurs, with finite closed islands in an infinite open ocean. The picture is more complicated when d ≥ 3 since pc < 1/2, and there is coexistence of infinite open and closed clusters for p between pc and 1 − pc.

Criticality

Zoom in a critical percolation cluster (Click to animate) Percolation zoom.gif
Zoom in a critical percolation cluster (Click to animate)

Percolation has a singularity at the critical point p = pc and many properties behave as of a power-law with , near . Scaling theory predicts the existence of critical exponents, depending on the number d of dimensions, that determine the class of the singularity. When d = 2 these predictions are backed up by arguments from conformal field theory and Schramm–Loewner evolution, and include predicted numerical values for the exponents. Most of these predictions are conjectural except when the number d of dimensions satisfies either d = 2 or d ≥ 6. They include:

See Grimmett (1999). [17] In 11 or more dimensions, these facts are largely proved using a technique known as the lace expansion. It is believed that a version of the lace expansion should be valid for 7 or more dimensions, perhaps with implications also for the threshold case of 6 dimensions. The connection of percolation to the lace expansion is found in Hara & Slade (1990). [18]

In two dimensions, the first fact ("no percolation in the critical phase") is proved for many lattices, using duality. Substantial progress has been made on two-dimensional percolation through the conjecture of Oded Schramm that the scaling limit of a large cluster may be described in terms of a Schramm–Loewner evolution. This conjecture was proved by Smirnov (2001) [19] in the special case of site percolation on the triangular lattice.

Different models

Applications

In biology, biochemistry, and physical virology

Percolation theory has been used to successfully predict the fragmentation of biological virus shells (capsids), [21] [22] with the fragmentation threshold of Hepatitis B virus capsid predicted and detected experimentally. [23] When a critical number of subunits has been randomly removed from the nanoscopic shell, it fragments and this fragmentation may be detected using Charge Detection Mass Spectroscopy (CDMS) among other single-particle techniques. This is a molecular analog to the common board game Jenga, and has relevance to the broader study of virus disassembly. Interestingly, more stable viral particles (tilings with greater fragmentation thresholds) are found in greater abundance in nature. [21]

In ecology

Percolation theory has been applied to studies of how environment fragmentation impacts animal habitats [24] and models of how the plague bacterium Yersinia pestis spreads. [25]

See also

Related Research Articles

<span class="mw-page-title-main">Random walk</span> Mathematical formalization of a path that consists of a succession of random steps

In mathematics, a random walk, sometimes known as a drunkard's walk, is a random process that describes a path that consists of a succession of random steps on some mathematical space.

<span class="mw-page-title-main">Percolation</span> Filtration of fluids through porous materials

In physics, chemistry, and materials science, percolation refers to the movement and filtering of fluids through porous materials. It is described by Darcy's law. Broader applications have since been developed that cover connectivity of many systems modeled as lattices or graphs, analogous to connectivity of lattice components in the filtration problem that modulates capacity for percolation.

In statistical mechanics, a universality class is a collection of mathematical models which share a single scale-invariant limit under the process of renormalization group flow. While the models within a class may differ dramatically at finite scales, their behavior will become increasingly similar as the limit scale is approached. In particular, asymptotic phenomena such as critical exponents will be the same for all models in the class.

<span class="mw-page-title-main">Random graph</span> Graph generated by a random process

In mathematics, random graph is the general term to refer to probability distributions over graphs. Random graphs may be described simply by a probability distribution, or by a random process which generates them. The theory of random graphs lies at the intersection between graph theory and probability theory. From a mathematical perspective, random graphs are used to answer questions about the properties of typical graphs. Its practical applications are found in all areas in which complex networks need to be modeled – many random graph models are thus known, mirroring the diverse types of complex networks encountered in different areas. In a mathematical context, random graph refers almost exclusively to the Erdős–Rényi random graph model. In other contexts, any graph model may be referred to as a random graph.

Critical exponents describe the behavior of physical quantities near continuous phase transitions. It is believed, though not proven, that they are universal, i.e. they do not depend on the details of the physical system, but only on some of its general features. For instance, for ferromagnetic systems, the critical exponents depend only on:

<span class="mw-page-title-main">Giant component</span> Large connected component of a random graph

In network theory, a giant component is a connected component of a given random graph that contains a significant fraction of the entire graph's vertices.

In statistical physics, directed percolation (DP) refers to a class of models that mimic filtering of fluids through porous materials along a given direction, due to the effect of gravity. Varying the microscopic connectivity of the pores, these models display a phase transition from a macroscopically permeable (percolating) to an impermeable (non-percolating) state. Directed percolation is also used as a simple model for epidemic spreading with a transition between survival and extinction of the disease depending on the infection rate.

<span class="mw-page-title-main">Erdős–Rényi model</span> Two closely related models for generating random graphs

In the mathematical field of graph theory, the Erdős–Rényi model refers to one of two closely related models for generating random graphs or the evolution of a random network. These models are named after Hungarian mathematicians Paul Erdős and Alfréd Rényi, who introduced one of the models in 1959. Edgar Gilbert introduced the other model contemporaneously with and independently of Erdős and Rényi. In the model of Erdős and Rényi, all graphs on a fixed vertex set with a fixed number of edges are equally likely. In the model introduced by Gilbert, also called the Erdős–Rényi–Gilbert model, each edge has a fixed probability of being present or absent, independently of the other edges. These models can be used in the probabilistic method to prove the existence of graphs satisfying various properties, or to provide a rigorous definition of what it means for a property to hold for almost all graphs.

<span class="mw-page-title-main">Percolation threshold</span> Threshold of percolation theory models

The percolation threshold is a mathematical concept in percolation theory that describes the formation of long-range connectivity in random systems. Below the threshold a giant connected component does not exist; while above it, there exists a giant component of the order of system size. In engineering and coffee making, percolation represents the flow of fluids through porous media, but in the mathematics and physics worlds it generally refers to simplified lattice models of random systems or networks (graphs), and the nature of the connectivity in them. The percolation threshold is the critical value of the occupation probability p, or more generally a critical surface for a group of parameters p1, p2, ..., such that infinite connectivity (percolation) first occurs.

An important question in statistical mechanics is the dependence of model behaviour on the dimension of the system. The shortcut model was introduced in the course of studying this dependence. The model interpolates between discrete regular lattices of integer dimension.

<span class="mw-page-title-main">Harry Kesten</span> American mathematician (1931–2019)

Harry Kesten was a Jewish American mathematician best known for his work in probability, most notably on random walks on groups and graphs, random matrices, branching processes, and percolation theory.

In the context of the physical and mathematical theory of percolation, a percolation transition is characterized by a set of universal critical exponents, which describe the fractal properties of the percolating medium at large scales and sufficiently close to the transition. The exponents are universal in the sense that they only depend on the type of percolation model and on the space dimension. They are expected to not depend on microscopic details such as the lattice structure, or whether site or bond percolation is considered. This article deals with the critical exponents of random percolation.

Conductivity near the percolation threshold in physics, occurs in a mixture between a dielectric and a metallic component. The conductivity and the dielectric constant of this mixture show a critical behavior if the fraction of the metallic component reaches the percolation threshold.

<span class="mw-page-title-main">Water retention on random surfaces</span> Study of water distribution

Water retention on random surfaces is the simulation of catching of water in ponds on a surface of cells of various heights on a regular array such as a square lattice, where water is rained down on every cell in the system. The boundaries of the system are open and allow water to flow out. Water will be trapped in ponds, and eventually all ponds will fill to their maximum height, with any additional water flowing over spillways and out the boundaries of the system. The problem is to find the amount of water trapped or retained for a given surface. This has been studied extensively for random surfaces.

In mathematics and probability theory, continuum percolation theory is a branch of mathematics that extends discrete percolation theory to continuous space. More specifically, the underlying points of discrete percolation form types of lattices whereas the underlying points of continuum percolation are often randomly positioned in some continuous space and form a type of point process. For each point, a random shape is frequently placed on it and the shapes overlap each with other to form clumps or components. As in discrete percolation, a common research focus of continuum percolation is studying the conditions of occurrence for infinite or giant components. Other shared concepts and analysis techniques exist in these two types of percolation theory as well as the study of random graphs and random geometric graphs.

First passage percolation is a mathematical method used to describe the paths reachable in a random medium within a given amount of time.

Robustness, the ability to withstand failures and perturbations, is a critical attribute of many complex systems including complex networks.

In statistical mechanics, bootstrap percolation is a percolation process in which a random initial configuration of active cells is selected from a lattice or other space, and then cells with few active neighbors are successively removed from the active set until the system stabilizes. The order in which this removal occurs makes no difference to the final stable state.

Percolation surface critical behavior concerns the influence of surfaces on the critical behavior of percolation.

In probability theory, the van den Berg–Kesten (BK) inequality or van den Berg–Kesten–Reimer (BKR) inequality states that the probability for two random events to both happen, and at the same time one can find "disjoint certificates" to show that they both happen, is at most the product of their individual probabilities. The special case for two monotone events was first proved by van den Berg and Kesten in 1985, who also conjectured that the inequality holds in general, not requiring monotonicity. Reimer later proved this conjecture. The inequality is applied to probability spaces with a product structure, such as in percolation problems.

References

  1. 1 2 Broadbent, Simon; Hammersley, John (1957). "Percolation processes I. Crystals and mazes". Mathematical Proceedings of the Cambridge Philosophical Society. 53 (3): 629–641. Bibcode:1957PCPS...53..629B. doi:10.1017/S0305004100032680. ISSN   0305-0041. S2CID   84176793.
  2. Sahini, M.; Sahimi, M. (2003-07-13). Applications Of Percolation Theory. CRC Press. ISBN   978-0-203-22153-2. Archived from the original on 2023-02-04. Retrieved 2020-10-27.
  3. van Krevelen, Dirk W (1982). "Development of coal research—a review". Fuel. 61 (9): 786–790. doi:10.1016/0016-2361(82)90304-0.
  4. The rosalind franklin papers - the holes in coal: Research at BCURA and in Paris, 1942-1951. https://profiles.nlm.nih.gov/spotlight/kr/feature/coal Archived 2022-07-07 at the Wayback Machine . Accessed: 2022-01-17.
  5. Hammersley, JM; Welsh, DJA (1980). "Percolation theory and its ramifications". Contemporary Physics. 21 (6): 593–605. Bibcode:1980ConPh..21..593H. doi:10.1080/00107518008210661.
  6. Bollobás, Béla; Riordan, Oliver (2006). "Sharp thresholds and percolation in the plane". Random Structures and Algorithms. 29 (4): 524–548. arXiv: math/0412510 . doi:10.1002/rsa.20134. ISSN   1042-9832. S2CID   7342807.
  7. MEJ Newman; RM Ziff (2000). "Efficient Monte Carlo algorithm and high-precision results for percolation". Physical Review Letters. 85 (19): 4104–4107. arXiv: cond-mat/0005264 . Bibcode:2000PhRvL..85.4104N. doi:10.1103/physrevlett.85.4104. PMID   11056635. S2CID   747665.
  8. Erdős, P. & Rényi, A. (1959). "On random graphs I.". Publ. Math. (6): 290–297.
  9. Erdős, P. & Rényi, A. (1960). "The evolution of random graphs". Publ. Math. Inst. Hung. Acad. Sci. (5): 17–61.
  10. Bolloba's, B. (1985). "Random Graphs". Academic.
  11. Berchenko, Yakir; Artzy-Randrup, Yael; Teicher, Mina; Stone, Lewi (2009-03-30). "Emergence and Size of the Giant Component in Clustered Random Graphs with a Given Degree Distribution". Physical Review Letters. 102 (13): 138701. Bibcode:2009PhRvL.102m8701B. doi:10.1103/PhysRevLett.102.138701. ISSN   0031-9007. PMID   19392410. Archived from the original on 2023-02-04. Retrieved 2022-02-24.
  12. Li, Ming; Liu, Run-Ran; Lü, Linyuan; Hu, Mao-Bin; Xu, Shuqi; Zhang, Yi-Cheng (2021-04-25). "Percolation on complex networks: Theory and application". Physics Reports. Percolation on complex networks: Theory and application. 907: 1–68. arXiv: 2101.11761 . Bibcode:2021PhR...907....1L. doi:10.1016/j.physrep.2020.12.003. ISSN   0370-1573. S2CID   231719831.
  13. Hassan, M. K.; Rahman, M. M. (2015). "Percolation on a multifractal scale-free planar stochastic lattice and its universality class". Phys. Rev. E. 92 (4): 040101. arXiv: 1504.06389 . Bibcode:2015PhRvE..92d0101H. doi:10.1103/PhysRevE.92.040101. PMID   26565145. S2CID   119112286.
  14. Hassan, M. K.; Rahman, M. M. (2016). "Universality class of site and bond percolation on multi-multifractal scale-free planar stochastic lattice". Phys. Rev. E. 94 (4): 042109. arXiv: 1604.08699 . Bibcode:2016PhRvE..94d2109H. doi:10.1103/PhysRevE.94.042109. PMID   27841467. S2CID   22593028.
  15. Kesten, Harry (1982). Percolation Theory for Mathematicians. Birkhauser. doi:10.1007/978-1-4899-2730-9. ISBN   978-0-8176-3107-9.
  16. Grimmett, Geoffrey; Marstrand, John (1990). "The Supercritical Phase of Percolation is Well Behaved". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 430 (1879): 439–457. Bibcode:1990RSPSA.430..439G. doi:10.1098/rspa.1990.0100. ISSN   1364-5021. S2CID   122534964.
  17. Grimmett, Geoffrey (1999). Percolation. Grundlehren der mathematischen Wissenschaften. Vol. 321. Berlin: Springer. doi:10.1007/978-3-662-03981-6. ISBN   978-3-642-08442-3. ISSN   0072-7830. Archived from the original on 2020-02-23. Retrieved 2009-04-18.
  18. Hara, Takashi; Slade, Gordon (1990). "Mean-field critical behaviour for percolation in high dimensions". Communications in Mathematical Physics. 128 (2): 333–391. Bibcode:1990CMaPh.128..333H. doi:10.1007/BF02108785. ISSN   0010-3616. S2CID   119875060. Archived from the original on 2021-02-24. Retrieved 2022-10-30.
  19. Smirnov, Stanislav (2001). "Critical percolation in the plane: conformal invariance, Cardy's formula, scaling limits". Comptes Rendus de l'Académie des Sciences. I. 333 (3): 239–244. arXiv: 0909.4499 . Bibcode:2001CRASM.333..239S. CiteSeerX   10.1.1.246.2739 . doi:10.1016/S0764-4442(01)01991-7. ISSN   0764-4442.
  20. Adler, Joan (1991), "Bootstrap percolation", Physica A: Statistical Mechanics and Its Applications, 171 (3): 453–470, Bibcode:1991PhyA..171..453A, doi:10.1016/0378-4371(91)90295-n .
  21. 1 2 Brunk, Nicholas E.; Twarock, Reidun (2021). "Percolation Theory Reveals Biophysical Properties of Virus-like Particles". ACS Nano. 15 (8). American Chemical Society (ACS): 12988–12995. doi: 10.1021/acsnano.1c01882 . ISSN   1936-0851. PMC   8397427 . PMID   34296852.
  22. Brunk, N. E.; Lee, L. S.; Glazier, J. A.; Butske, W.; Zlotnick, A. (2018). "Molecular Jenga: the percolation phase transition (collapse) in virus capsids". Physical Biology. 15 (5): 056005. Bibcode:2018PhBio..15e6005B. doi:10.1088/1478-3975/aac194. PMC   6004236 . PMID   29714713.
  23. Lee, L. S.; Brunk, N.; Haywood, D. G.; Keifer, D.; Pierson, E.; Kondylis, P.; Zlotnick, A. (2017). "A molecular breadboard: Removal and replacement of subunits in a hepatitis B virus capsid". Protein Science. 26 (11): 2170–2180. doi:10.1002/pro.3265. PMC   5654856 . PMID   28795465.
  24. Boswell, G. P.; Britton, N. F.; Franks, N. R. (1998-10-22). "Habitat fragmentation, percolation theory and the conservation of a keystone species". Proceedings of the Royal Society of London B: Biological Sciences. 265 (1409): 1921–1925. doi:10.1098/rspb.1998.0521. ISSN   0962-8452. PMC   1689475 .
  25. Davis, S.; Trapman, P.; Leirs, H.; Begon, M.; Heesterbeek, J. a. P. (2008-07-31). "The abundance threshold for plague as a critical percolation phenomenon". Nature. 454 (7204): 634–637. Bibcode:2008Natur.454..634D. doi:10.1038/nature07053. hdl: 1874/29683 . ISSN   1476-4687. PMID   18668107. S2CID   4425203.

Further reading