Fact-checked by Grok 2 weeks ago

Self-avoiding walk

A self-avoiding walk (SAW) is a starting at the origin in \mathbb{Z}^d, consisting of n consecutive steps to nearest-neighbor sites, such that no point is visited more than once. This restriction models the constraint, preventing self-intersections, in contrast to unrestricted random walks that allow revisits. Formally, an n-step SAW \omega = (\omega_0, \omega_1, \dots, \omega_n) satisfies \omega_0 = 0, |\omega_i - \omega_{i-1}| = 1 for each i, and \omega_i \neq \omega_j for all $0 \leq i < j \leq n. The model originated in polymer physics, where Nobel laureate Paul J. Flory introduced it in 1949 to analyze the average end-to-end distance of real polymer chains in dilute solutions. Flory argued that polymer molecules, modeled as flexible chains linked by covalent bonds, expand due to repulsive interactions between non-adjacent segments, leading to a scaling R \sim N^\nu where N is the chain length and \nu > 1/2 (unlike the Gaussian chain's \nu = 1/2). Earlier ideas trace to Werner Kuhn in the , but Flory's mean-field approximation popularized the SAW as a discrete analog for these "hindered" random walks. By the , computational simulations by Frank Rosenbluth and others began enumerating short SAWs, revealing the model's combinatorial complexity. SAWs serve as a in for studying , phase transitions, and the behavior of linear polymers under good solvent conditions. In , they approximate the swollen conformations of macromolecules, aiding predictions of , elasticity, and . Mathematically, the connective constant \mu = \lim_{n \to \infty} c_n^{1/n} (where c_n is the number of n-step SAWs) governs exponential growth, proven to exist via submultiplicativity. Exact values are known only for specific lattices; for example, on the \mu = \sqrt{2} + \sqrt{2}. Despite their simplicity, SAWs pose profound challenges: exact enumeration beyond small n is intractable, and critical exponents like \gamma (susceptibility) and \nu (end-to-end distance) require advanced techniques such as lace expansion or conformal invariance. In high dimensions (d \geq 5), SAWs asymptotically resemble simple random walks, with mean-square displacement \langle |X_n|^2 \rangle \sim n. In four dimensions, logarithmic corrections appear, \langle |X_n|^2 \rangle \sim n (\log n)^{1/2}, while in two dimensions, the scaling limit is believed to be the Schramm–Loewner evolution (SLE) curve with parameter \kappa = 8/3, rigorously proven for specific lattices. These results underscore SAWs' role in bridging combinatorics, probability, and physics, with ongoing research into their universality and extensions like interacting or directed variants.

Fundamentals

Definition

A self-avoiding walk (SAW) of length n is formally defined on a , most commonly the d-dimensional \mathbb{Z}^d for d \geq 1. It consists of a of n+1 distinct points \omega = (\omega_0, \omega_1, \dots, \omega_n), where \omega_0 is the starting point (often taken as the $0 \in \mathbb{Z}^d) and each consecutive pair satisfies \|\omega_{i+1} - \omega_i\|_1 = 1 for i = 0, 1, \dots, n-1, meaning the steps are to nearest-neighbor sites along the edges. The key constraint is the absence of self-intersections: \omega_i \neq \omega_j for all $0 \leq i < j \leq n. This setup ensures the walk traces a path without revisiting any site, modeling non-overlapping configurations in discrete space. The walks are typically undirected in the sense that steps can proceed in any of the $2d possible nearest-neighbor directions from the current site, though the sequence itself imposes a directionality to the path. In some contexts, SAWs may be restricted to directed steps (e.g., only positive coordinates), but the standard formulation allows bidirectional movement while prohibiting loops or overlaps. This no-intersection rule applies strictly to site visits, distinguishing SAWs from lattice animals or other path models that might permit edge crossings without site revisits. For illustration on the square lattice \mathbb{Z}^2 (where d=2 and nearest neighbors differ by \pm e_1 or \pm e_2, with e_1 = (1,0), e_2 = (0,1)), a length-1 SAW is simply (0,0) \to (1,0), a straight step with no opportunity for intersection. A length-2 example without overlap is the bent path (0,0) \to (1,0) \to (1,1), whereas (0,0) \to (1,0) \to (0,0) is invalid due to revisiting the origin. These short paths highlight how the self-avoidance constraint begins to limit configurations even at small scales, unlike simple random walks on the same lattice, which permit arbitrary revisits and thus explore space more freely without intersection prohibitions.

Historical development

The concept of the self-avoiding walk (SAW) originated in the context of during the mid-20th century, primarily as a model for the conformations of long-chain molecules under the , where chain segments cannot occupy the same space. In 1947, W.J.C. Orr provided the first exact enumerations of SAWs on , computing the number of walks up to length 6 to study polymer configurations at infinite dilution. This work laid early groundwork for numerical approaches, though limited by computational constraints of the time. Subsequently, in 1949, Paul J. Flory formalized the SAW as a key tool for modeling real , deriving approximate scaling relations for their end-to-end distances and emphasizing the role of self-avoidance in preventing unphysical overlaps or knots in chain statistics. The 1950s and 1960s saw the transition of SAWs from a physical model to a rigorous mathematical object, with initial focus on enumeration and bounds. Pioneering computations by and in 1959 extended enumerations to walks of length up to 9 on square and cubic lattices, enabling estimates of growth constants. In 1963, established fundamental upper and lower bounds on the number of n-step SAWs, proving that the connective constant μ (the limit of the nth root of the number of walks) exists and lies between 2^{1/d} and 2d-1 for d-dimensional lattices, where d is the dimension. These results, built on combinatorial arguments, shifted attention toward asymptotic behavior and marked the beginning of SAW as a central problem in probability and statistical mechanics. By the 1970s, SAWs gained prominence in the study of critical phenomena and phase transitions, linking polymer statistics to broader universality classes. Pierre-Gilles de Gennes, in 1972, connected SAWs to the in the limit n → 0 using framework, providing a field-theoretic perspective on critical exponents and facilitating predictions for polymer dimensions near criticality. Michael E. Fisher further advanced this by analyzing SAW scaling in relation to correlations during the same decade, highlighting shared universality with other long-range interacting systems. These developments solidified SAWs as a paradigm for non-Markovian processes in statistical physics. In the 1980s, SAWs were firmly established as a cornerstone model in statistical mechanics, with rigorous proofs of mean-field behavior above the upper critical dimension d=4 by Takao Hara and others, and lace expansion techniques introduced by Gordon Slade confirming Gaussian-like asymptotics in high dimensions. By the 1990s, the model extended to more general settings, including SAWs on arbitrary networks and graphs, as explored in Neal Madras and Gordon Slade's 1996 monograph, which analyzed random walk analogs and connectivity on non-regular structures. Connections to percolation theory also emerged, particularly through the dilute limit where SAWs model incipient infinite clusters, influencing studies of transport and exploration on percolating lattices.

Mathematical properties

Enumeration

The number c_n denotes the total number of distinct n-step self-avoiding walks starting from the origin on a given lattice, where each walk is a sequence of adjacent lattice sites without revisiting any site. This count is typically considered up to rotations and reflections only if specified, but standard enumerations fix the starting point and count all possible paths. Examples include the integer lattice \mathbb{Z}^2 (square lattice) and the honeycomb lattice, where the geometry affects the possible configurations. Exact enumeration of c_n relies on recursive algorithms and transfer-matrix methods, which systematically build walks step by step while enforcing the self-avoidance condition. Recursive approaches, such as pruning enumerators, discard invalid partial walks early to reduce computational complexity, achieving growth rates around $1.3^n for practical implementations. Transfer-matrix techniques model the enumeration on finite lattice strips, propagating connectivity states across rows to count walks of increasing length. These methods have yielded exact values for the square lattice up to n = 79, with earlier computations reaching n = 71 and n = 51 using similar algebraic and parallelized strategies. No closed-form formula for c_n exists in general, as counting self-avoiding walks of fixed length is #P-complete even on subgraphs of two-dimensional grids. The growth rate of c_n is characterized by the connective constant \mu = \lim_{n \to \infty} c_n^{1/n}, which provides a subexponential measure of branching possibilities. On the , \mu \approx 2.63815853, while on the , \mu = \sqrt{2 + \sqrt{2}} \approx 1.84775907, the only lattice where \mu is known exactly. Asymptotically, c_n \sim A \mu^n n^{\gamma - 1}, where A is a critical amplitude and \gamma is a universal critical exponent (detailed in the section on critical exponents). Generating functions, formed as \sum c_n x^n, serve as a tool to analyze these counts through series expansions and singularity analysis. The following table lists exact values of c_n for the square lattice up to n = 20, computed via recursive and transfer-matrix methods; higher values up to n = 79 are available but grow rapidly in magnitude.
nc_n
14
212
336
4100
5284
6780
72172
85916
916268
1044100
11120292
12326120
13889636
142420544
156605960
1617997620
1749142412
18134006524
19365436720
20997497644

Generating functions

The generating function for self-avoiding walks on a lattice is defined as G(x) = \sum_{n=0}^\infty c_n x^n, where c_n denotes the number of walks of length n starting from a fixed origin, and x is a fugacity variable that weights the walk length. This formal power series encodes the enumeration data c_n and facilitates the extraction of asymptotic behaviors through its analytic continuation. The radius of convergence of G(x) is $1/\mu, where \mu is the connective constant of the lattice, defined as \mu = \lim_{n \to \infty} c_n^{1/n}. The presence of a singularity at x = 1/\mu determines the dominant growth rate of c_n \sim \mu^n for large n, reflecting the exponential proliferation of self-avoiding configurations. Hammersley's theorem guarantees the existence of this limit \mu for self-avoiding walks on vertex-transitive graphs in dimensions d \geq 2, via submultiplicative arguments on the sequence c_n. For d-dimensional hypercubic lattices, bounds establish d \leq \mu \leq 2d - 1, with the upper bound arising from comparisons to spanning trees or pattern restrictions that limit branching. Approximations to G(x) employ inclusion-exclusion principles, such as those in the lace expansion, which decompose walks into non-intersecting segments to bound or compute coefficients iteratively. Pattern-avoidance methods further refine these by excluding specific intersecting motifs, yielding series expansions convergent within the radius $1/\mu. In statistical mechanics, G(x) connects to partition functions, as self-avoiding walks emerge in the zero-component limit of the O(n) model, where the fugacity x tunes the polymer density analogous to inverse temperature. Extensions include weighted generating functions for variants like bridged walks, defined as B(x) = \sum_{n=0}^\infty b_n x^n, where b_n counts bridges—self-avoiding walks conditioned to stay non-negative in a specified direction. Similarly, unrooted walk generating functions weight configurations by symmetry or closure properties, aiding in the study of cycles and polygons derived from self-avoiding paths.

Scaling and universality

Critical exponents

The critical exponents govern the asymptotic scaling behavior of self-avoiding walks (SAWs) in the limit of large step length n. The mean-square end-to-end distance scales as \langle R^2 \rangle \sim n^{2\nu}, where \nu is the Flory exponent characterizing the size of the walk. The total number of n-step SAWs, denoted c_n, obeys c_n \sim \mu^n n^{\gamma - 1}, where \mu is the connective constant and \gamma is the susceptibility exponent related to the "entropy" of configurations. These exponents are interconnected through scaling relations derived from renormalization group (RG) theory, which maps SAWs to the n \to 0 limit of the O(n) model in the continuum. A key relation is Fisher's identity, \gamma = \nu (2 - \eta), where \eta is the anomalous dimension arising from the two-point correlation function G(x) \sim |x|^{-(d-2+\eta)} for large separation |x| in d dimensions. Hyperscaling relations, valid below the upper critical dimension d=4, further link exponents via $2 - \alpha = d \nu, where \alpha describes the singularity in the free energy analog for the model. Exact values are known in low dimensions. In one dimension, SAWs are trivial straight lines, yielding \nu = 1 and \gamma = 1. In two dimensions, conformal invariance and Coulomb gas methods provide exact results: \nu = 3/4 and \gamma = 43/32, with \eta = 5/24. In three dimensions, no exact solutions exist, but high-precision numerical estimates give \nu \approx 0.587597 and \gamma \approx 1.157. Above d=4, mean-field theory applies with \nu = 1/2, \gamma = 1, and \eta = 0, matching simple random walk behavior. Rigorous bounds include Kesten's result that \nu \geq 1/2 in all dimensions d \geq 2, reflecting the stiffness induced by self-avoidance compared to unrestricted walks. In d \geq 5, the equality \nu = 1/2 is proven via lace expansion techniques.

Universality class

Self-avoiding walks (SAWs) belong to the universality class of the O(n) vector model in the limit as n approaches 0, where critical exponents describing their scaling behavior are independent of microscopic details such as the underlying lattice structure. This correspondence was first proposed by de Gennes, who mapped the excluded-volume problem of SAWs to the n=0 case of the n-component O(n) spin model, capturing the entropic repulsion that prevents self-intersections. The theoretical foundation relies on supersymmetric formulations that equate SAWs to this limiting spin model, ensuring that long-length-scale properties are governed by the same fixed point under renormalization group transformations. In high dimensions, SAWs share critical exponents with the Ising model (corresponding to n=1 in the O(n) model), such as the mean-field values, while exhibiting distinct behavior in two dimensions due to stronger fluctuations. The upper critical dimension is d_c=4, above which renormalization group flows lead to mean-field exponents like the Flory exponent ν=1/2, indicating Gaussian-like scaling similar to simple random walks; at d=4, logarithmic corrections appear. This universality is supported by rigorous proofs using lace expansion techniques, which demonstrate lattice independence for d ≥ 5. Monte Carlo simulations further confirm these mean-field exponents and their lattice independence above d=4, with simulations on hypercubic lattices yielding ν consistent with 1/2 within small corrections. Unlike self-attracting walks at the theta point, which exhibit collapsed configurations with a different tricritical scaling (ν ≈ 1/3 in three dimensions), or rigid rods characterized by orientational order and persistence length, SAWs represent flexible chains in the swollen regime driven purely by excluded volume effects. This distinction underscores the unique position of SAWs within polymer physics universality classes.

Generalizations

On networks and graphs

A self-avoiding walk on a graph G = (V, E) is defined as a path that visits each vertex at most once, generalizing the lattice case to arbitrary finite or infinite graphs without the regular structure of a lattice. On such graphs, the connective constant \mu(G), which measures the exponential growth rate of the number of n-step self-avoiding walks from a fixed origin, exists for quasi-transitive graphs and equals \lim_{n \to \infty} c_n^{1/n}, where c_n is the number of such walks of length n. Key properties of self-avoiding walks on graphs include exact computations of the connective constant for trees. For a regular tree of degree z \geq 2, \mu = z-1, reflecting the branching structure where each step after the first offers z-1 choices without revisiting vertices, as there are no cycles. On general graphs, determining \mu(G) is more challenging due to cycles and varying connectivity, often requiring bounds or approximations rather than exact values. Examples illustrate the adaptation to specific graph structures. On the complete graph K_m, self-avoiding walks correspond to Hamiltonian paths, with the number of n-step walks from a vertex given by (m-1)! for n = m-1, but the model is limited to finite graphs where walks cannot exceed m steps. Percolation clusters serve as hosts for self-avoiding walks, where walks are confined to the infinite cluster at criticality, exhibiting modified scaling behaviors influenced by the fractal dimension of the cluster. Unlike on lattices, self-avoiding walks on general graphs lack translation invariance, leading to position-dependent statistics and no uniform embedding in Euclidean space. Growth rates, captured by \mu(G), depend critically on the graph's girth (shortest cycle length) and degree distribution; for instance, graphs with large girth approximate tree-like behavior, while heterogeneous degree distributions in random graphs can yield higher connectivity constants than regular lattices. Seminal results on connective constants for general graphs are provided by Madras and Slade, who established existence and submultiplicativity properties for quasi-transitive graphs and analyzed bounds using pattern avoidance techniques. Further studies extend to expander graphs and random graphs, where self-avoiding walks achieve positive speed and \mu scales with the spectral gap or average degree, differing from lattice universality. On infinite graphs, asymptotic limits of walk lengths relate to the graph's amenability, with non-amenable graphs supporting ballistic self-avoiding walks.

Limits and asymptotics

The thermodynamic limit of self-avoiding walks (SAWs) concerns the behavior as the walk length n tends to infinity. The existence of a translation-invariant probability measure on the space of infinite SAWs is established in dimensions d \geq 3; in two dimensions, such a measure exists for half-space walks but remains an open problem in the full plane. Uniform measures on finite SAWs converge weakly in constrained settings, such as bridges and half-spaces, to limiting measures supported on infinite SAWs. Constrained variants of SAWs, such as bridges and half-space walks, exhibit analogous limiting behaviors. A bridge is an n-step SAW from the origin that stays non-negative in one coordinate and ends at a height of order n^\nu, where \nu is the scaling exponent; the uniform measure on such bridges converges to a limiting bridge measure as n \to \infty. Half-space walks, conditioned to remain in the half-plane, converge similarly, with their scaling limits related to the free SAW measure through factors involving critical exponents. Asymptotic theorems reveal that SAWs deviate from classical random walk behavior. Donsker-type invariance principles, which yield Gaussian scaling for simple random walks, fail for SAWs in dimensions d < 4, where the end-to-end distance scales super-diffusively as n^\nu with \nu > 1/2 instead of the diffusive \nu = 1/2. This super-diffusive scaling reflects the entropic repulsion from self-intersections and is governed by , such as the Flory exponent \nu, which determine the asymptotic growth. Rigorous results include Kesten's invariance principle, which demonstrates that the ratio of the number of (n+2)-step SAWs to n-step SAWs approaches \mu^2 (where \mu is the ) as n \to \infty, facilitating proofs of the infinite SAW measure's existence. For the end-to-end , invariance principles hold in high dimensions: in d \geq 5, the scaled end-to-end vector converges in distribution to a Gaussian, consistent with mean-field behavior. In two dimensions, Lawler's work, building on restriction properties, shows that if the scaling limit of planar SAWs exists and is conformally invariant, it must coincide with the SLE_ {8/3} curve. Dimensional dependence is pronounced in asymptotic behaviors. In d > 4, SAWs obey mean-field asymptotics, with \nu = 1/2, \gamma = 1, and Gaussian scaling limits dominating due to negligible self-intersection effects. At the upper critical dimension d = 4, a crossover occurs, featuring mean-field exponents perturbed by logarithmic corrections, such as in the susceptibility scaling as \mu^n n^{\gamma-1} (\log n)^{1/4}.

Applications and methods

Physical models

Self-avoiding walks (SAWs) provide a foundational model for linear chains in dilute solutions under good solvent conditions, where interactions prevent chain segments from overlapping. In Flory's seminal theory, a polymer with n monomers is represented as an SAW of length n, capturing the swelling of the chain due to these repulsive interactions, which lead to an end-to-end distance scaling as R \sim n^{3/5} in three dimensions. This model contrasts with ideal random walks by incorporating the physical constraint that monomers occupy finite volume, thus avoiding unphysical self-intersections. The Edwards model extends this discrete framework to a continuous analog, treating the polymer as a path in space governed by a measure modified by a self-avoidance potential that penalizes overlaps. This formulation, introduced in , employs path-integral techniques from to describe the of the chain, enabling analytical approximations for the swollen conformation in the presence of . The model highlights the long-range correlations induced by self-avoidance, bridging lattice-based with continuum descriptions used in . In , model the swollen phase above the theta point, where attractive interactions balance effects, leading to a from extended coils to compact globules as temperature decreases. At the theta point, behaves ideally with Gaussian statistics, but below it, attractions dominate, causing ; specifically represent the good-solvent regime where repulsion prevails. This is analogous to phenomena in other systems, with sharing universality with critical magnetic models like the Ising ferromagnet. Beyond synthetic polymers, SAWs apply to biological systems, such as modeling DNA looping where chromatin forms self-avoiding loops to regulate gene expression, and protein folding pathways that avoid steric clashes during native structure formation. Similarly, foraging patterns in animals exhibit self-avoiding trajectories to maximize resource coverage without revisiting depleted areas. Experimental validation comes from small-angle neutron scattering (SANS) studies, which confirm SAW predictions for polymer radii of gyration, showing exponents close to the precise numerical value of approximately 0.588, consistent with Flory's approximation of 0.6, for chains in good solvents like polystyrene in toluene.

Computational approaches

Computational approaches to studying self-avoiding walks (SAWs) primarily rely on numerical methods to estimate properties such as the number of walks of length n, the connective constant \mu, and , given the intractability of exact solutions for large systems. These methods include exact enumerations for short walks, simulations for longer configurations, and field-theoretic approximations that provide bounds and scaling insights. High-precision results from these techniques have validated theoretical predictions and informed applications in . Exact enumeration techniques, such as algorithms and methods, allow for the precise counting of up to moderate lengths on lattices. On the in two dimensions, these methods have enumerated all up to 71 steps using optimized approaches that reduce the state space through and pruning. In three dimensions on the simple cubic , enumerations extend to 36 steps via similar with , providing series data for to the connective constant \mu. These exact counts serve as benchmarks for approximate methods but become computationally infeasible beyond these lengths due to in the number of configurations. Monte Carlo methods enable the sampling of long SAWs by generating ensembles that approximate the . The pivot algorithm, introduced by Madras and Sokal, is a dynamic technique that starts from an initial SAW and applies random rotations (pivots) to subsections, ensuring self-avoidance while achieving rapid decorrelation; it generates effectively independent samples in time proportional to the walk length N, allowing simulations up to N \approx 10^5 steps in three dimensions. Rosenbluth sampling, originally for chains, biases the growth of walks by weighting branches proportional to their extension probability, mitigating attrition in self-avoiding growth but requiring corrections for unbiased estimates. A modern extension, the pruned-enriched Rosenbluth method (PERM) by Grassberger, addresses weight imbalances in Rosenbluth sampling by pruning low-weight partial walks and enriching (duplicating) high-weight ones, enabling efficient estimation of like the end-to-end distance scaling \nu \approx 0.587597 in three dimensions through flat-histogram variants. Field-theoretic approximations map SAWs to the n \to 0 limit of the O(n) vector model, where perturbative renormalization group methods yield estimates for exponents and \mu. In this framework, de Gennes showed that the excluded-volume problem corresponds to a \phi^4 field theory in the n=0 limit, providing non-rigorous but accurate predictions for scaling behavior via \epsilon-expansions around the upper critical dimension. Variational methods, such as those by Hammersley, derive upper and lower bounds on \mu by considering sub- and super-multiplicative sequences of walk counts, with refinements yielding \mu < 4.278 for the cubic lattice. These approaches have produced high-precision estimates, such as \mu \approx 4.68404 for the simple cubic lattice \mathbb{Z}^3, obtained via simulations with of order $10^{-7} from variance analysis in estimators. Error control in simulations involves times and finite-size scaling, ensuring convergence for exponents like \gamma \approx 1.15695 with relative errors below 0.1%.

References

  1. [1]
    [PDF] Lectures on Self-Avoiding Walks - IHES
    Despite its simple definition, the self-avoiding walk is difficult to study in a mathematically rigorous manner. Many of the important problems remain ...
  2. [2]
    [PDF] THE SELF-AVOIDING WALK - University of Regina
    This report reviews the main and most recent results about the scaling limit of self-avoiding walk (SAW) on the integer lattice in dimensions greater than or ...
  3. [3]
    The Configuration of Real Polymer Chains - AIP Publishing
    Flory; The Configuration of Real Polymer Chains. J. Chem. Phys. 1 March 1949 ... This content is only available via PDF. Open the PDF for in another window.
  4. [4]
    [PDF] Self-avoiding walks. - UBC Mathematics
    For despite its simple definition, the self-avoiding walk leads to math- ematical problems which are simple enough to state to the mathematically ...
  5. [5]
  6. [6]
    [PDF] Calculation of the connective constant for self-avoiding walks via the ...
    Feb 8, 2013 · We calculate the connective constant for self-avoiding walks on the simple cubic lattice to unprece- dented accuracy, using a novel application ...
  7. [7]
    The complexity of counting self-avoiding walks in subgraphs of two ...
    This paper offers a partial answer to the question of Welsh: it is #P-complete to compute the number of self-avoiding walks of a given length in a subgraph of a ...
  8. [8]
    [PDF] SELF-AVOIDING WALKS AND CONNECTIVE CONSTANTS
    Apr 19, 2017 · In this paper we review certain properties of the connective constant. µ(G), in particular exact values (Section 1.2), upper and lower bounds. ( ...
  9. [9]
    Counting Theorem for the Generating Function of Self‐Avoiding Walks
    Mar 1, 1971 · A generating function Gn(α, β, γ) is defined to contain the information which gives the number of self‐avoiding walks which reach the point ...
  10. [10]
    [PDF] Lectures on self-avoiding walks - UBC Math
    Dec 9, 2010 · These lecture notes focus on a number of rigorous results for self-avoiding walks on the d-dimensional integer lattice Zd. The model is defined ...
  11. [11]
    Self-avoiding walk, spin systems and renormalization - Journals
    Jan 23, 2019 · The self-avoiding walk (SAW) is a combinatorial model of lattice paths without self-intersections. In addition to its intrinsic mathematical ...Missing: seminal | Show results with:seminal
  12. [12]
    New exact exponents for two-dimensional self-avoiding walks
    These exponents give, for any p, the number of configurations of p two-dimensional self-avoiding walks of the same length l which are attached by their ends.
  13. [13]
    [PDF] The Self-Avoiding Walk: A Brief Survey∗ - UBC Math
    May 28, 2010 · Slade. Renormalisation group analysis of weakly self-avoiding walk in dimensions four and higher. To appear in Proceedings of the International.
  14. [14]
    [PDF] Monte Carlo study of four-dimensional self-avoiding walks of ... - arXiv
    Mar 30, 2017 · The logarithmic corrections arise because d = 4 is the upper critical dimension for self-avoiding walks. For d ≥ 5 self-avoiding walks are ...
  15. [15]
    The Self-Avoiding Walk | SpringerLink
    Scaling, polymers and spins. Neal Madras, Gordon Slade. Pages 35-55. Some ... "This is the first book on self-avoiding random walk and a very good one.
  16. [16]
  17. [17]
    Self-avoiding walks on percolation clusters - IOPscience
    The author studies self-avoiding walks (SAWs) on percolation clusters. A scaling function representation for R, the mean end-to-end distance, is proposed.
  18. [18]
    Self-avoiding walks and connective constants in small-world networks
    Aug 11, 2003 · Madras and G. Slade, The Self-Avoiding Walk (Birkhäuser, Boston, 1993). D.S. Gaunt and A.J. Guttmann, in Phase Transitions and Critical ...
  19. [19]
    [1510.08659] Self-avoiding walks and amenability - arXiv
    Oct 29, 2015 · The connective constant \mu(G) of an infinite transitive graph G is the exponential growth rate of the number of self-avoiding walks from a ...Missing: connectivity | Show results with:connectivity
  20. [20]
    [PDF] The self-avoiding walk in a strip
    We prove that this probability measure may be obtained by conditioning the SAW in the half plane to have a bridge at height y.
  21. [21]
    The statistical mechanics of polymers with excluded volume
    The statistical mechanics of polymers with excluded volume. S F Edwards. Published under licence by IOP Publishing Ltd Proceedings of the Physical Society ...
  22. [22]
    [PDF] Nobel Lecture, December 11, 1974 - PAUL J. FLORY
    The effect of excluded volume. The configuration on the left represents the random coil in absence of volume exclusion, the chain being equivalent to a line in.Missing: seminal paper
  23. [23]
    Modeling a Self-Avoiding Chromatin Loop: Relation to the Packing ...
    Here, we report a new, to our knowledge, method to explicitly model a DNA or chromatin loop as a semiflexible (self-avoiding) tube (Gonzalez and Maddocks, 1999, ...
  24. [24]
    Assessing Lévy walks as models of animal foraging - Journals
    Jun 1, 2011 · Reynolds [80] explored a model of individuals that avoid odour trails left by conspecifics (termed a self-avoiding walk), as occurs, for example ...