Winding of simple walks on the square lattice

https://doi.org/10.1016/j.jcta.2019.105191Get rights and content

Abstract

A method is described to count simple diagonal walks on Z2 with a fixed starting point and endpoint on one of the axes and a fixed winding angle around the origin. The method involves the decomposition of such walks into smaller pieces, the generating functions of which are encoded in a commuting set of Hilbert space operators. The general enumeration problem is then solved by obtaining an explicit eigenvalue decomposition of these operators involving elliptic functions. By further restricting the intermediate winding angles of the walks to some open interval, the method can be used to count various classes of walks restricted to cones in Z2 of opening angles that are integer multiples of π/4.

We present three applications of this main result. First we find an explicit generating function for the walks in such cones that start and end at the origin. In the particular case of a cone of angle 3π/4 these walks are directly related to Gessel's walks in the quadrant, and we provide a new proof of their enumeration. Next we study the distribution of the winding angle of a simple random walk on Z2 around a point in the close vicinity of its starting point, for which we identify discrete analogues of the known hyperbolic secant laws and a probabilistic interpretation of the Jacobi elliptic functions. Finally we relate the spectrum of one of the Hilbert space operators to the enumeration of closed loops in Z2 with fixed winding number around the origin.

Introduction

Counting of lattice paths has been a major topic in combinatorics (and probability and physics) for many decades. Especially the enumeration of various types of lattice walks confined to convex cones in Z2, like the positive quadrant, has attracted much attention in recent years, due mainly to the rich algebraic structure of the generating functions involved (see e.g. [17], [7] and references therein) and the relations with other combinatorial structures (e.g. [6], [31]). The study of lattice walks in non-convex cones has received much less attention. Notable exception are walks on the slit plane [14], [18] and the three-quarter plane [16]. When describing the plane in polar coordinates, the confinement of walks to cones of different opening angles (with the tip positioned at the origin) may equally be phrased as a restriction on the angular coordinates of the sites visited by the walk. One may generalize this concept by replacing the angular coordinate by a notion of winding angle of the walk around the origin, in which case one can even make sense of cones of angles larger than 2π. It stands to reason that a fine control over the winding angle in the enumeration of lattice walks brings us a long way in the study of walks in (especially non-convex) cones.

Although the winding angle of lattice walks seems to have received little attention in the combinatorics literature, probabilistic aspects of the winding of long random walks have been studied in considerable detail [4], [5], [36], [38]. In particular, it is known that under suitable conditions on the steps of the random walk the winding angle after n steps is typically of order logn, and that the angle normalized by 2/logn converges in distribution to a standard hyperbolic secant distribution. The methods used all rely on coupling to Brownian motion, for which the winding angle problem is easily studied with the help of its conformal properties. Although quite generally applicable in the asymptotic regime, these techniques tell us little about the underlying combinatorics.

In this paper we initiate the combinatorial study of lattice walks with control on the winding angle, by looking at various classes of simple (rectilinear or diagonal) walks on Z2. As we will see, the combinatorial tools described in this paper are strong enough to bridge the gap between the combinatorial study of walks in cones and the asymptotic winding of random walks. Before describing the main results of the paper, we should start with some definitions.

We let W be the set of simple diagonal walks w in Z2 of length |w|0 avoiding the origin, i.e. w is a sequence (wi)i=0|w| in Z2{(0,0)} with wiwi1{(1,1),(1,1),(1,1),(1,1)} for 1i|w|. We define the winding angle θiwR of w up to time i to be the difference in angular coordinates of wi and w0 including a contribution 2π (resp. 2π) for each full counterclockwise (resp. clockwise) turn of w around the origin up to time i. Equivalently, (θiw)i=0|w| is the unique sequence in R such that θ0w=0 and θiwθi1w is the (counterclockwise) angle between the segments ((0,0),wi1) and ((0,0),wi) for 1i|w|. The (full) winding angle of w is then θwθ|w|w. See Fig. 1 for an example.

Main result  The Dirichlet space D is the Hilbert space of complex analytic functions f on the unit disc D={zC:|z|<1} that vanish at 0 and have finite norm fD2 with respect to the Dirichlet inner productf,gD=Df(z)g(z)dA(z)=n=1n[zn]f(z)[zn]g(z), where the measure dA(z1+iz2)1πdz1dz2 is chosen such that DdA(z)=1. See [3] for a review of its properties. We denote by (en)n=1 the standard orthogonal basis defined by en(z)zn, which is unnormalized since enD2=n. With this notation en,fD=n[zn]f(z) for any analytic function fD.

For k(0,1) we let vk:C{zR:z2k}C be the analytic function defined by the elliptic integralvk(z)14K(k)0zdx(kx2)(1kx2) along the simplest path from 0 to z, where K(k) is the complete elliptic integral of the first kind with elliptic modulus k (see Appendix A for definitions and notation). The appearance of this elliptic integral in lattice walks enumeration is a natural one since 2πK(4t) is precisely the generating function for excursions of the simple diagonal walk from the origin (see (58) in Appendix A),2πK(4t)=n=0(2nn)2t2n. The incomplete elliptic integral vk does not have a comparably simple combinatorial interpretation, but provides an important conformal mapping from a slit disk onto a rectangle in the complex plane, as detailed in Section 2.1.

For fixed k we use the conventional notationk=1k2andk1=1k1+k for the complementary modulus k and the descending Landen transformation k1 of k, which both take values in (0,1) again (see Appendix A). Using these we introduce a family (fm)m=1 of analytic functions by setting (notice the k1 in vk1(z)!)fm(z)cos(2πm(vk1(z)+1/4))cos(πm/2), which satisfies fm(0)=0. Even though vk1(z) has branch cuts at z=±k1, we will see (Lemma 6, Lemma 7, Lemma 8) that fm(z) has radius of convergence around 0 equal to 1/k1>1 and has finite norm with respect to the Dirichlet inner product, hence fmD. According to Proposition 9 the norm of fm is given explicitly byfmD2=m(qkmqkm)4, where qk(0,1) is the (elliptic) nome of modulus k (see (61) in Appendix A), which is analytic for k in the unit disk. Once properly normalized the family of functions (fˆm)m=1 provides an orthonormal basis of D, i.e.fˆm(z)fm(z)fmD=cos(2πm(vk1(z)+1/4))cos(πm/2)m4(qkmqkm),fˆn,fˆmD=1n=m.

The main technical result of this paper is the following.

Theorem 1

For ,p1 and απ2Z, let W,p(α) be the set of (possibly empty) simple diagonal walks w on Z2{(0,0)} that start at (p,0), end on one of the axes at distance ℓ from the origin, and have full winding angle θw=α.

  • (i)

    Let W,p(α)(t)wW,p(α)t|w| be the generating function of W,p(α). For k=4t(0,1) fixed, there exists a compact self-adjoint operator Yk(α) on D with eigenvectors (fm)m=1 such thatW,p(α)(t)=e,Yk(α)epD,Yk(α)fm=2K(k)π1mqkm|α|/πfm.

  • (ii)

    Let W,p(α,I)W,p(α) be the subset of the aforementioned walks that have intermediate winding angles in an interval IR, i.e. θiwI for i=1,2,,|w|1, and let W,p(α,I)(t) be the corresponding generating function. If I=(β,β+) with β±π2Z{±}, α[β,β+]π2Z and α0 or αβ±, then the generating function W,p(α,I)(t) is related to a matrix element of a compact self-adjoint operator on D with the same eigenvectors (fm)m=1, as described in the table below. The remaining cases follow from the symmetries (α,β,β+)(α,β+,β) and (α,β,β+)(α,αβ+,αβ), and the cases β±=± agree with the corresponding limits β±± (using that qk(0,1)).

  • (iii)

    The statement of (ii) remains valid for β±π4Z{±} and α[β,β+]π2Z and α0 or αβ± as long as ℓ and p are even.

See Fig. 2 for examples of walks enumerated by Theorem 1. We emphasize that the convention for the index placement in W,p(α,I), that is used also for other sets of walks throughout this work, is that the first index corresponds to the endpoint of the walk and the second index to the starting point. The reason for this choice is precisely the relation of the generating function W,p(α) to the matrix element at position (,p) of a Hilbert space operator. As we will see, composition of particular families of walks often corresponds to the composition of the respective operators (or multiplication of the respective infinite matrices). The index placement reflects the natural right-to-left ordering in the composition of these operators.

Theorem 1 is stated for fixed real values of k=4t(0,1), while from a combinatorial point of view it may be preferable to think of the generating functions W,p(α,I)(t) as formal power series in the variable t. This raises the question to what extent the eigenvalue decomposition can be understood on the level of formal power series. To this end we prove in Proposition 10 that for any m,p1 the coefficient [zp]fˆm(z)=1pep,fˆmD is an analytic function in k around 0, such that we may interpret fˆm(z)=fˆm(z;t) as taking values in the ring Rz,t of formal power series in the variables z and t, see Section 2.2 for an explicit expansion for m=1,,6. With this observation Theorem 1 can be largely recast as a set of identities on formal power series. To this end we denote the multivariate generating function of W,p(α,I) byW(α,I)(x,y;t)=,p1xypW,p(α,I)(t)Rx,y,t, and the eigenvalues of Ak(α), Jk(α,β), Bk(α,β,β+) by Am(α)(t), Jm(α,β)(t), Bm(α,β,β+)(t) respectively (this includes Yk(α)=Bk(α,,) as a special case). These eigenvalues, as given in Theorem 1(ii), are all analytic functions of t in the unit disk (see (70) and (2) for the power series representations of qk and K(k)). Provided α0, the mth eigenvalue converges to 0 as m in the formal topology of Rt, i.e. for any n0 the coefficient [tn]Am(α)(t) vanishes for all m large enough. Theorem 1 implies the convergent series identities in Rx,y,t for α0,W(α,(0,α))(x,y;t)=m=1Am(α)(t)fˆm(x;t)fˆm(y;t),W(α,(β,α))(x,y;t)=yym=1Jm(α,β)(t)fˆm(x;t)fˆm(y;t),W(α,(β,β+))(x,y;t)=xyxym=1Bm(α,β,β+)(t)fˆm(x;t)fˆm(y;t), with α, β and β+ satisfying the respective assumptions described in Theorem 1(ii) and (iii). We leave it as an open problem whether these identities and an expression for fˆm(x;t) can be derived using exclusively formal power series.

As an example of how to compute explicit generating functions with the help of Theorem 1, let us look at the set W3,3(π) of simple diagonal walks from (3,0) to (3,0) that have winding angle π around the origin. According to Theorem 1(i) its generating function satisfiesW3,3(π)(t)=e3,Yk(π)e3D=2K(k)πm=11mqkme3,fˆmD2=10t6+280t8+5661t10+, where we used the series expansions of qk and K(k) (respectively (70) and (58) in Appendix A) and that of e3,fˆmD in Section 2.2.

Application: excursions  Theorem 1 can be used to count many specialized classes of walks involving winding angles. The first quite natural counting problem we address is that of the (diagonal) excursions E from the origin, i.e. E is the set of (non-empty) simple diagonal walks starting and ending at the origin with no intermediate returns (Fig. 3a). Actually, in this case we may equally well consider simple rectilinear walks on Z2, thanks to the obvious linear mapping between the two types of walks (Fig. 3b). Even though walks wE do not completely avoid the origin, we may still naturally assign a winding angle sequence to them by imposing that the first and last step do not contribute to the winding angle, i.e. θ1w=θ0w=0 and θw=θ|w|w=θ|w|1w. In Proposition 18 we prove that the generating function for excursions with winding angle απ2Z is given (for k=4t(0,1) fixed) byF(α)(t)wEt|w|1{θw=α}=2πK(k)n=1(1qkn)21qk4nqkn(2π|α|+1). Since the summand is analytic in t around 0 and O(t2n) for any n, the relation implies an identity of formal power series in Rt for any απ2Z.

Similarly to Theorem 1(ii) one may further restrict the full winding angle sequence of w to lie in an open interval I=(β,β+) with βπ4Z<0, β+π4Z>0 and αIπ2Z. In this case it is more natural to also fix the starting direction, say w1=(1,1), and we denote the corresponding generating function by F(α,I)(t). Observe in particular that F(α,R)(t)=F(α)(t)/4. We prove in Theorem 23 that the generating function F(α,I)(t) is given by the finite sumF(α,I)(t)=π8δσ(0,δ)π2Z(cos(4σαδ)cos(4σ(2β+α)δ))F(t,4σδ),δ2(β+β), where F(t,b)απ2ZF(α)(t)eibα is the “characteristic function” associated to F(α)(t). Since F(t,b)=F(t,4b), the terms in the right-hand side of (6) corresponding to σ and δσ are actually identical, therefore leaving only δπ distinct terms. For non-integer values of b (see Proposition 18 for the full expression) F(t,b) can be expressed in closed form asF(t,b)=1cos(πb2)[1πtan(πb4)2K(k)θ1(πb4,qk)θ1(πb4,qk)],(bRZ) where θ1(z,q) is the first Jacobi theta function (see (64) for a definition). Again (6) and (7) imply the equivalent identities on the level of formal power series in Rt or Ct. In Proposition 20 we prove that the power series F(t,b) is algebraic in t for any bQZ but that is it transcendental for bZ. By inspecting the terms appearing in (6) we find that F(α,I)(t) is transcendental if and only if β+,βπ2Z+π4 or both β+,βπZ+π2 and απZ (see Theorem 23).

A special case of excursions for which the generating function is algebraic is α=0 and I=(π/4,π/2), see Fig. 4a. After removal of the first and last step of the walk and a linear transformation, these correspond precisely to walks in the quadrant starting and ending at the origin with steps in {(1,0),(1,0),(1,1),(1,1)}, also known as Gessel walks. Around 2000 Ira Gessel conjectured that the generating function for such walks is given (in our notation) by1t2F(0,(π/4,π/2))(t)=n=0t2n16n(5/6)n(1/2)n(2)n(5/3)n=1+2t2+11t4+85t6+, where (a)n=Γ(a+n)/Γ(a) is the descending Pochhammer symbol (see also Sloane's Online Encyclopedia of Integer sequences (OEIS) sequence A135404). The first computer-aided proof of this conjecture appeared in [30], and it was followed by several “human” proofs in [13], [15], [7]. Here we provide an alternative proof using Theorem 23. Indeed, we have explicitlyF(0,(π/4,π/2))(t)=18F(t,43)+18F(t,83)=14F(t,43)=12[3π2K(4t)θ1(π3,qk)θ1(π3,qk)1], where used that F(t,b)=F(t,4b). According to our discussion above this is an algebraic power series in t, a fact about F(0,(π/4,π/2))(t) that was first observed in [12]. In Corollary 24 we deduce an explicit algebraic equation for F(0,(π/4,π/2))(t), and check that it agrees with a known equation for n=0t2n+216n(5/6)n(1/2)n(2)n(5/3)n.

Application: unconstrained random walks  Let (Wi)i0 be a simple random walk on Z2 started at the origin. A natural question is to ask for the (approximate) distribution of the winding angle θjz of the random walk around some point zR2 up to time j. As mentioned before, this question has been addressed successfully in the literature in the limit j using coupling to Brownian motion. If z(Z+1/2)2, then 2θjz/logj is known [36], [4], [5], [38] to converge in distribution to the hyperbolic secant law with density 12sech(πx/2)dx (recall that sechy=1/coshy=2/(ey+ey)). If instead zZ2 and one conditions the random walk not to hit z before time j, then 2θjz/logj converges to a “hyperbolic secant squared law” with density π4sech2(πx/2)dx [36].

In Section 4 we complement these results by deriving exact statistics at finite j with the help of Theorem 1. To this end we look at the winding angles around two points in the vicinity of the starting point, namely (1/2,1/2) and the origin itself. To be precise, let θj+1/2 be the winding angle of (Wi)i around (1/2,1/2) up to time j+1/2, i.e. halfway its step from Wj to Wj+1 (see Fig. 5). Similarly, let θj+1/2 be the winding angle around the origin, ignoring the first step and with the convention that θj+1/2= if (Wi)i has returned to the origin strictly before time j+1.

If ζk is a geometric random variable with parameter k(0,1), i.e. P[ζk=j]=kj(1k) for j0, then the following “discrete hyperbolic secant laws” hold (Theorem 25):P[θζk+1/2(απ2,α+π2)]=cksech(4Tkα)+k1k1{α=0}for απ2Z,P[θζk+1/2(απ2,α+π2)]=Cksech(4Tk(απ4))sech(4Tk(α+π4))for απ2Z+π4, where ck, Ck and Tk are explicit functions of k (see Theorem 25).

As a consequence we find probabilistic interpretations of the Jacobi elliptic functions cn and dn (see Appendix A) as characteristic functions of winding angles,Eexp(ib{θζk+1/2}πZ+π2)=cn(K(k)b,k),Eexp(ib{θζk1/2}πZ)=dn(K(k)b,k). Here {}A denotes rounding to the nearest element of AR and we set θ1/2=0 by convention.

Since cn(y,1)=dn(y,1)=sech(y)=Eeiyη is the characteristic function of the aforementioned hyperbolic secant distribution, we may directly conclude the convergence in distribution as k1 of the winding angle θζk+1/2/K(k) at geometric time ζk. A more delicate singularity analysis, which is beyond the scope of this paper, would yield the same distributional limit for 2θj+1/2/log(j) as j, in accordance with previous work.

Application: loops  The last application we discuss utilizes the fact that the eigenvalues of the operators in Theorem 1 have much simpler expressions than the components ep,fmD of the eigenvectors. It is therefore worthwhile to seek combinatorial interpretations of traces of (combinations of) operators, the values of which only depend on the eigenvalues. When writing out the trace in terms of the basis (ep)p=1 it is clear that such an interpretation must involve walks that start and end at arbitrary but equal distance from the origin. If the full winding angle is taken to be a multiple of 2π then such a walk forms a loop, i.e. it starts and ends at the same point.

A natural combinatorial set-up is described in Section 5. There we consider the set Ln of rooted loops of index n, nZ{0}, which are simple diagonal walks avoiding the origin that start and end at an arbitrary but equal point in Z2 and have winding angle 2πn around the origin. The set Ln=LnevenLnodd naturally partitions into loops that visit only sites of even respectively odd parity ((x,y)Z2 with x+y even respectively odd), see Fig. 6. We introduce the generating functionsLn(t)wLnt|w||w|Lnodd(t)=wLnoddt|w||w|,Lneven(t)=wLnevent|w||w|, where we have included a factor 1/|w| for convenience, such that the generating functions of the set Ln is actually tLn(t). Observe that L1(t) is precisely the generating function of unrooted loops of index 1, i.e. rooted loops modulo rerooting (but preserving orientation), because the equivalence class of a rooted loop w of index 1 contains precisely |w| elements. This is not true for n2, since Ln contains rooted loops w that cover themselves multiple times and thus have equivalence classes with less than |w| elements.

Theorem 28 states that these generating functions for n>0 are given byLn(t)=1ntrJk(2πn,)=1nqk2n1qk2n,Lnodd(t)=1nqk2n1qk4n,Lneven(t)=1nqk4n1qk4n. Here Jk(2πn,) is the operator Jk(α,β) with α=2πn and β= appearing in Theorem 1(ii) whose mth eigenvalue qk2nm can be obtained as the β limit of the displayed expression (see the final remark of Theorem 1(ii)). Note as before that these imply the equivalent identities on the level of power series in Rt.

A simple probabilistic consequence is the following. Let (Wi)i=02 be a simple random walk on Z2 started at the origin and conditioned to return after 2 steps. For each point zR2 we let the index Iz be the signed number of times (Wi)i=02 winds around z in counterclockwise direction, i.e. 2πIz is the winding angle of (Wi)i=02 around z. If z lies on the trajectory of (Wi)i=02, then we set Iz=. We let the clusters Cn of index n be the set of connected components of {zR2:Iz=n}, and for cCn we let |c| and |c| respectively be the area and boundary length of component c. See Fig. 7 for an example. The expectation values for the total area and boundary length of the clusters of index n then satisfyE[cCn|c|]=42(2)22n[k2]qk2n1qk4n2πn2,E[cCn(|c|2)]=42(2)24n[k2]qk2n1+qk2n2π3log2, with the asymptotics as indicated. The first result should be compared to the analogous result for Brownian motion: Garban and Ferreras proved in [28] using Yor's work [40] that the expected area of the set of points with index n with respect to a unit time Brownian bridge in R2 is equal to 1/(2πn2). Perhaps surprisingly, we find that the expected boundary length all the components of index n (minus twice the expected number of such components) grows asymptotically at a rate that is independent of n, contrary to the total area.

Open question 1

Does E[cC0 finite|c|], i.e. the total area of the finite clusters of index 0, have a similarly explicit expression? Based on the results of [28] we expect it to be asymptotic to π/30 as .

Finally we mention one more potential application of the enumeration of loops in Theorem 28 in the context of random walk loop soups [33], which are certain Poisson processes of loops on Z2. A natural quantity to consider in such a system is the winding field which roughly assigns to any point zR2 the total index of all the loops in the process [24], [23]. Theorem 28 may be used to compute explicit expectation values (one-point functions of the corresponding vertex operators to be precise) in the massive version of the loop soups. We will pursue this direction elsewhere.

Discussion  The connection between the enumeration of walks and the explicitly diagonalizable operators on Dirichlet space may seem a bit magical to the reader. So perhaps some comments are in order on how we arrived at this result, which originates in the combinatorial study of planar maps.

A planar map is a connected multigraph (a graph with multiple edges and loops allowed) that is properly embedded in the 2-sphere (edges are only allowed to meet at their extremities), viewed up to orientation-preserving homeomorphisms of the sphere. The connected components of the complement of a planar map are called the faces, which have a degree equal to the number of bounding edges. There exists a relatively simple multivariate generating function for bipartite planar maps, i.e. maps with all faces of even degree, that have two distinguished faces of degree p and and a fixed number of faces of each degree (see e.g. [25]). The surprising fact is that this generating function has a form that is very similar to that of the generating function W,p(π,(π,π))(t) of diagonal walks from (p,0) to (,0) that avoid the slit {(x,0):x0} until the end. A combinatorial explanation (in the particular case of critical planar maps) appears in [21] using a peeling exploration [20], [22].

If one further decorates the planar maps by a rigid O(n) loop model [9], then the combinatorial relation extends to one between walks of fixed winding angle W,p(α) with απZ and planar maps with two distinguished faces and certain collections of non-intersecting loops separating the two faces. The combinatorics of the latter has been studied in considerable detail in [9], [8], [10], [21], which has inspired our treatment of the simple walks on Z2 in this paper. Further details on the connection and an extension to more general lattice walks with small steps (i.e. steps in {1,0,1}2) will be provided in forthcoming work.

Finally, we point out that these methods can be used to determine Green's functions (and more general resolvents) for the Laplacian on regular lattices in the presence of a conical defect or branched covering, which are relevant to the study of various two-dimensional statistical systems. As an example, in recent work [32] Kenyon and Wilson computed the Green's function on the branched double cover of the square lattice, which has applications in local statistics of the uniform spanning tree on Z2 as well as dimer systems.

The author would like to thank Kilian Raschel, Alin Bostan and Gaëtan Borot for their suggestions on how to prove Corollary 24, and Thomas Prellberg for suggesting to extend Theorem 25 to absorbing boundary conditions. The author is indebted to an anonymous referee for numerous corrections and suggestions to improve the exposition. This work was supported by a public grant as part of the Investissement d'avenir project, reference ANR-11-LABX-0056-LMH, LabEx LMH, as well as ERC-Advanced grant 291092, “Exploring the Quantum Universe” (EQU). Part of this work was done while the author was at the Niels Bohr Institute, University of Copenhagen and Institut de Physique Théorique, CEA, Université Paris-Saclay.

Section snippets

Winding angle of walks starting and ending on an axis

Our strategy towards proving Theorem 1 will be to first prove part (ii) for three special cases (see Fig. 8),J,pW,p(π2,(π2,π2)),B,pW,p(0,(π2,π2)),A,pW,p(π2,(0,π2)). We define three linear operators Jk, Bk and Ak on D by specifying their matrix elements with respect to the standard basis (ei)i1 in terms of the corresponding generating functions J,p(t), B,p(t) and A,p(t) (with k=4t) as1e,JkepD=J,p(t),e,BkepD=B,p(t),1pe,AkepD=A,p(t). The motivation to define the

Counting excursions with fixed winding angle

Recall from the introduction the set of excursions E consisting of (non-empty) simple diagonal walks starting and ending at the origin with no intermediate returns. For such an excursion wE we have a well-defined winding angle sequence (θiw)i=0|w| with θ1w=θ0w=0 and θw=θ|w|w=θ|w|1w. Our first goal is to compute the generating function of excursions with winding angle equal to απ2Z,F(α)(t)wEt|w|1{θw=α}. To this end we cannot directly apply Theorem 1(i) because the excursions start and end

Winding angle distribution of an unconstrained random walk

Let (Wi)i be the simple random walk on Z2 started at the origin. In this section we use the results of Theorem 1 to determine statistics of the winding angle θjz of (Wi)i up to time j around a lattice point zZ2 or a dual lattice point z(Z+1/2)2. In the case that (Wi)i hits the lattice point z at time i, we set θiz=θi1z and θjz= for j>i. From a physicist's point of view, we may regard the zZ2 as an obstacle with absorbing boundary conditions and z(Z+1/2)2 as one with reflecting boundary

Winding angle of loops

Another, rather interesting application of Theorem 1 is the counting of loops on Z2. To be precise, for integer n0, let the set Ln of rooted loops of index n be the set of simple diagonal walks w on Z2{(0,0)} that start and end at the same (arbitrary) point and have winding angle θw=2πn. The set Ln=LnevenLnodd naturally partitions into the even loops Lneven supported on {(x,y)Z2:x+y even, (x,y)(0,0)} and the odd loops Lnodd on {(x,y)Z2:x+y odd}.

Theorem 28

The (“inverse-size biased”) generating

References (42)

  • O. Bernardi et al.

    Counting quadrant walks via Tutte's invariant method

  • G. Borot et al.

    Loop models on random maps via nested loops: the case of domain symmetry breaking and application to the Potts model

    J. Phys. A, Math. Theor.

    (2012)
  • G. Borot et al.

    A recursive approach to the O(n) model on random maps via nested loops

    J. Phys. A, Math. Theor.

    (2012)
  • G. Borot et al.

    Nesting statistics in the O(n) loop model on random planar maps

  • J. Borwein et al.

    Pi and the AGM: A Study in Analytic Number Theory and Computational Complexity

    (1987)
  • A. Bostan et al.

    The complete generating function for Gessel walks is algebraic

    Proc. Am. Math. Soc.

    (2010)
  • A. Bostan et al.

    A human proof of Gessel's lattice path conjecture

    Trans. Am. Math. Soc.

    (2017)
  • M. Bousquet-Mélou et al.

    Walks with small steps in the quarter plane

  • M. Bousquet-Mélou et al.

    Walks on the slit plane

    Probab. Theory Relat. Fields

    (2002)
  • P.S. Bruckman

    On the evaluation of certain infinite series by elliptic functions

    Fibonacci Q.

    (1977)
  • T. Budd

    The peeling process of infinite Boltzmann planar maps

    Electron. J. Comb.

    (2016)
  • Cited by (19)

    View all citing articles on Scopus
    View full text