-
PDF
- Split View
-
Views
-
Cite
Cite
Andreas E Kyprianou, Ana Osojnik, Tony Shardlow, Unbiased ‘walk-on-spheres’ Monte Carlo methods for the fractional Laplacian, IMA Journal of Numerical Analysis, Volume 38, Issue 3, July 2018, Pages 1550–1578, https://doi.org/10.1093/imanum/drx042
- Share Icon Share
Abstract
We consider Monte Carlo methods for simulating solutions to the analogue of the Dirichlet boundary value problem in which the Laplacian is replaced by the fractional Laplacian and boundary conditions are replaced by conditions on the exterior of the domain. Specifically, we consider the analogue of the so-called ‘walk-on-spheres’ algorithm. In the diffusive setting, this entails sampling the path of Brownian motion as it uniformly exits a sequence of spheres maximally inscribed in the domain. As this algorithm would otherwise never end, it is truncated when the ‘walk-on-spheres’ comes within |$\varepsilon>0$|of the boundary. In the setting of the fractional Laplacian, the role of Brownian motion is replaced by an isotropic |$\alpha$|-stable process with |$\alpha\in(0,2)$|. A significant difference to the Brownian setting is that the stable processes will exit spheres by a jump rather than hitting their boundary. This difference ensures that disconnected domains may be considered and that, unlike the diffusive setting, the algorithm ends after an almost surely finite number of steps.
1. Introduction
For practical purposes, because it is impossible to take the limit, one truncates the series of estimates for large |$n$|, and the central limit theorem gives |$\mathcal{O}(1/n)$| upper bounds on the variance of the |$n$|-term sum, which serves as a numerical error estimate.
Although forming the fundamental basis of most Monte Carlo methods for diffusive Dirichlet-type problems, (1.3) is an inefficient numerical approach. Least of all, this is because the Monte Carlo simulation of |$u(x)$| is independent for each |$x\in D$|. Moreover, it is unclear how exactly to simulate the path of a Brownian motion on its first exit from |$D$|, that is to say, the quantity |$W_{\tau_D}$|. This is because of the fractal properties of Brownian motion, making its path difficult to simulate. This introduces additional numerical errors over and above that of Monte Carlo simulation.
A method proposed by Muller (1956), for the case that |$D$| is convex, subsamples special points along the path of the Brownian motion to the boundary of the domain |$D$|. The method does not require a complete simulation of its path and takes advantage of the distributional symmetry of the Brownian motion. To describe the so-called ‘walk-on-spheres’, we need to first introduce some notation. We may thus set |$\rho_0 = x$| for |$x\in D$| and define |$r_1$| to be the radius of the largest sphere inscribed in |$D$| that is centred at |$x$|. This sphere we will call |$S_1 = \{y\in\mathbb{R}^d\colon |y-\rho_0| = r_1\}$|. To avoid special cases, we, henceforth, assume that the surface area of |$S_1\cap \partial D$| is zero (this excludes, for example, the case that |$x = 0$| and |$D$| is a sphere centred at the origin).
Now set |$\rho_1\in D$| to be a point uniformly distributed on |$S_1$| and note that, given the assumption in the previous sentence, |$\mathbb{P}_x(\rho_1\in \partial D) = 0$|. Construct the remainder of the sequence |$(\rho_n,\,n \geq 1)$| inductively. Given |$\rho_{n-1}$|, we define the radius, |$r_n$|, of the largest sphere inscribed in |$D$| that is centred at |$\rho_{n-1}$|. Calling this sphere |$S_n$|, we have that |$S_n = \{y\in \mathbb{R}^d \colon |y-\rho_{n-1}| =r_n\}$|. We now select |$\rho_n$| to be a point that is uniformly positioned on |$S_n$|. Once again, we note that if |$\rho_{n-1}\in D$| almost surely, then the uniform distribution of both |$\rho_{n-1}$| and |$\rho_{n}$| ensures that |$\mathbb{P}(\rho_{n}\in \partial D) = 0$|. Consequently, the sequence |$\rho_n$| continues for all |$n\geq1$|. In the case that |$\rho_n$| approaches the boundary, the sequence of spheres |$S_n$| become arbitrarily small in size.
Thanks to the strong Markov property and the stationary and independent increments of Brownian motion, it is straightforward to prove the following result.
Fix |$x\in D$| and define |$\rho'_1 = W_{\tau_{S'_1}}$|, where |$\tau_{S'_1}=\inf\{t>0 \colon W_t \in S'_1 \}$| and |$S'_1$| is the largest sphere, centred at |$x$|, inscribed in |$D$|. For |$n\geq 2$|, given |$\rho'_{n-1}\in D$|, let |$\rho'_n = W_{\tau_{S'_n}}$|, where |$\tau_{S'_n}=\inf\{t>0 \colon W_t \in S'_n \}$| and |$S'_n$| is the largest sphere, centred at |$\rho'_{n-1}$|. Then, the sequences |$(\rho_n, n\geq 0)$| and |$(\rho'_n, n\geq 0)$| have the same law.
As an immediate consequence, |$\lim_{n\to\infty}\rho_n$| almost surely exists and, moreover, it is equal in distribution to |$W_{\tau_D}$|. The sequence |$\rho{:=} (\rho_n, n\geq 0)$| may now replace the role of |$(W_t, t\leq \tau_D)$| in (1.2), and hence in (1.3), albeit that one must stop the sequence |$\rho$| at some finite |$N$|. By picking a threshold |$\varepsilon>0$|, we can choose |$N(\varepsilon)$| as a cut-off for the sequence |$\rho$|, such that |$N(\varepsilon) = \min\{n\geq 0: \inf_{z\in\partial D}|\rho_n -z|\leq \varepsilon\}$|. Intuitively, one is inwardly ‘thickening’ the boundary |$\partial D$| with an ‘|$\varepsilon$|-skin’ and stopping once the walk-on-spheres hits the |$\varepsilon$|-skin. As the sequence |$\rho$| is random, |$N(\varepsilon)$| is also random. Starting with Theorem 6.6 of Muller (1956) and the classical computations in Motoo (1959), it is known that |$ \mathbb{E}_x[N(\varepsilon)] = \mathcal{O}(|\!\log\varepsilon|). $| To be more precise, we have the following result.
Suppose that |$D$| is a convex domain. There exist constants |$c_1,c_2>0$|, such that |$\mathbb{E}_x[N(\varepsilon)] \leq c_1\,{\left\vert{\log \varepsilon}\right\vert} + c_2$|, |$\varepsilon\in(0,1)$|.
Just as with the classical Dirichlet setting, the solution to (1.7) has a Feynman–Kac representation, expressed as an expectation at first exit from |$D$| of the associated stable process. The theorem below is proved in this article in a probabilistic way. Similar statements and proofs we found in the existing literature take a more analytical perspective. See for example, the review in Bucur (2016), as well as the monographs Bliedtner & Hansen (1986), Bucur & Valdinoci (2016) and Bogdan et al. (2009); the articles Bogdan & Byczkowski (1999), Ros-Oton & Serra (2014) and Ros-Oton (2016) and references therein.
The case that |$D$| is a ball can be found, for example, in Theorem 2.10 of Bucur (2016). We exclude the case |$d=1$|, because convex domains are intervals for which exact solutions are known; see again Bucur (2016) or the forthcoming Theorem 3.1 lifted from Blumenthal et al. (1961). Theorem 1.3 follows in fact as a corollary of a more general result stated later in Theorem 6.1, which is proved in the appendix.
In this article, our objective is to demonstrate that the walk-on-spheres method may also be extended to the setting of the Dirichlet problem with fractional Laplacian. In particular, we will show that, thanks to various distributional and path properties of stable processes, notably spatial homogeneity, isotropy, self-similarity and that it exits |$D$| by a jump, simulations can be made unbiased, without the need to truncate the algorithm at an |$\varepsilon$| tolerance. Although there exist many methods for numerically examining the fractional Dirichlet problem (1.7), which mostly appeal to classical methodology for diffusive operators (see, for example, Zoia et al., 2007; Huang & Oberman, 2014; Dybiec & Szczepaniec, 2015; Szczepaniec & Dybiec, 2015; Acosta et al., 2016; D’Elia & Gunzburger, 2016; Nochetto et al., 2016), to name a few, but not all of the existing literature, we believe that no other work appeals to the walk-on-spheres algorithm in this context.
The remainder of this article is structured as follows. In the next section, we give a brief historical review of Theorem 1.2 and its proofs as well as providing a new, short proof. In Section 3, we show how an old result of Blumenthal et al. (1961) can be used to give an exact simulation of the paths of stable processes. In Section 4, we introduce the walk-on-spheres algorithm for the fractional Laplacian Dirichlet problem. We start with domains |$D$| that are convex, but not necessarily bounded. Our main result shows that the walk-on-spheres algorithm ends in an a.s. finite number of steps (without the need for approximation), which can be stochastically bounded by a geometric distribution. Moreover, the parameter of this distribution does not depend on the starting point of the walk-on-spheres algorithm. Section 5 looks at the extensions to nonconvex domains. In Section 6, we consider a fractional Poisson equation, where an inhomogeneous term is introduced on the right-hand side of the fractional Laplacian Dirichlet problem (1.7). Appealing to the related results concerning the resolvent of stable processes until first exit from the unit ball, we are able to develop the walk-on-spheres algorithm further. Finally, in Section 7, we discuss some numerical experiments to illustrate the methods developed as well as their implementation.
2. The classical setting
As promised above, we give a brief historical review of the classical walk-on-spheres algorithm and, below, for completeness, we provide a proof of Theorem 1.2, which, to the authors’ knowledge, is new. The walk-on-spheres algorithm was first derived by Muller (1956). In Theorem 6.1 of his article, Muller claims that one can compare |$\mathbb{E}_x[N(\varepsilon)]$| with the mean number of steps of a walk-on-spheres process that is stopped when it reaches an |$\varepsilon$|-skin of the tangent hyperplane that passes through a point on |$\partial D$| that is closest to |$x$|. Although the claim is correct (indeed the proof that we give for our main result Theorem 4.1 below provides the basis for an alternative justification of this fact), it is not entirely clear from Muller’s reasoning. Motoo (1959) uses Muller’s comparison of the mean number of steps to prove Theorem 1.2. He considers the total expected occupation of an appropriately time changed version of Brownian motion when crossing each sphere of the walk until touching the aforementioned |$\varepsilon$|-skin of the tangent hyperplane. Using the self-similarity of Brownian motion, Motoo argues that the time change during passage to the boundary of each sphere is such that the expected occupation across each step is uniformly bounded below. It follows that the sum of these weighted expected occupations can be bounded below by |$\mathbb{E}_x[N(\varepsilon)]$|. On the other hand, the aforesaid sum can also be bounded above by the total expected time-changed occupation until exiting the half-space (as defined by the tangent plane), which can be computed explicitly, thereby providing the |$|\!\log\varepsilon|$| comparison.
Following the foundational work of Muller and Motoo, there have been many reproofs and generalizations of the original algorithm to different processes and domain types. Notable in this respect are the works of Mikhailov (1979) and Binder & Braverman (2012), who consider nonconvex domains, and Sabelfeld (1991), who appeals to renewal theory to analyse the growth in |$\varepsilon$| of the mean number of steps to completion of the walk-on-spheres algorithm. His method also allows for variants of the algorithm in which the sphere sizes do not need to be optimally inscribed in |$D$|. Later, Sabelfeld & Talay (1995) gave an elementary proof of the |$|\!\log \varepsilon|$| bound. Mascagni and colleagues have extensively developed the walk-on-spheres algorithm in applications (see, for example, Given et al., 2001, 2002; Hwang & Mascagni, 2001; Hwang et al., 2003; Mackoy et al., 2013.
We break the proof into two parts. In the first part, we analyse the walk-on-spheres process over one step, by considering the distance of the next point in the algorithm from the orthogonal tangent hyperplane of the first point. (Note the existence of a tangent hyperplane requires convexity of the domain.) In the second part of the proof, we use this analysis to build a supermartingale, from which the desired result follows via optional stopping.
This provides an implicit expression for |$\theta_1$| in terms of the orthogonal distance |$\rho_0$| from the nearest tangent hyperplane. See Fig. 1.
One easily verifies that there is a constant |$\lambda\in(0,1)$|, such that |$\sup_{u\in [0,1]}{\it{\Lambda}}(u)<\lambda$|.
Next, we move to the second part of the proof. At each step of the walk-on-spheres, we can construct the quantities |$\zeta_{n+1}$|, the orthogonal distance of |$\rho_{n+1}$| to the tangential hyperplane that passes through the closest point on |$\partial D$| to |$\rho_n$|; and |$\theta_n$|, the angle that is subtended at |$\rho_n$| between the aforesaid point and |$\rho_{n+1}$|. Note that |$\varepsilon$| is an absorbing state for the sequence |$(\zeta_n, n\geq 0)$| in the sense that, if |$\zeta_n = \varepsilon$|, then |$\zeta_{n+k} = \varepsilon$| for all |$k\geq 0$|. We may thus write |$N(\varepsilon) \leq N'(\varepsilon){:=}\min\{n\geq 0: \zeta_n = \varepsilon\}$|.
The result now follows by taking logarithms. □
3. Exact simulation of stable paths
The key ingredient to the walk-on-spheres in the Brownian setting is the knowledge that spheres are exited continuously and uniformly on the boundary of spheres. In the stable setting, the inclusion of path discontinuities means that the process will exit a sphere by a jump. The analogous key observation that makes our analysis possible is the following result, which gives the distribution of a stable process issued from the origin, when it first exits a unit sphere.
This result provides a method for constructing precise sample paths of stable processes in phase space (i.e., exploring sample paths as ordered subsets of |$\mathbb{R}^d$| rather than as functions |$[0,\infty]\to \mathbb{R}^d$|). Choose a tolerance |$\epsilon$| and initial point |$X_0 =x$|. Denote by |$E_1$| a sampling from the distribution given in Theorem 3.1. This gives the exit from a ball of radius one when |$X$| is issued from the origin. By the scaling property (1.6) and the stationary and independent increments, |$x +\epsilon\, E_1$| is distributed as the exit position from a ball of radius |$\epsilon$| centred at |$x$| when the process is issued from |$x$|. Hence, we define |$X_1=x + \epsilon\, E_1$| and then, inductively for |$n\geq 1$|, generate |$X_{n+1}$| as the exit point of the ball centred on |$X_n$| with radius |$\epsilon$| by noting this is equal in distribution to |$X_n + \epsilon \, E_{n+1}$|, where |$E_{n+1}$| is an i.i.d. copy of |$E_1$|. It is important to remark for later that the value of |$\epsilon$| in this algorithm does not need to be fixed and may vary with each step. Note, however, the method does not generate the corresponding time to exit from each ball. Therefore, the sample paths that are produced, while being exact in the distribution of points that the stable process will pass through, cannot be represented graphically in time, as there is only an equal mean duration to exiting each sphere. If the tolerance |$\epsilon$| is altered on each step, then even this mean duration feature is lost. The method is used to generate Fig. 2.

Example sample paths for the |$\alpha$|-stable Levy process generated using the exit distribution in Theorem 3.1 for spheres of radius |$10^{-6}$|. Rows show sample paths in two and three dimensions for |$\alpha=0.9$| (left) and |$\alpha=1.8$| (right). The yellow lines indicate jumps of the process and blue dots show where the process has been.
On account of classical Feynman–Kac representation, simulation of solutions to parabolic and elliptic equations involving the fractional Laplacian and, more generally, the infinitesimal generator of a Lévy process are synonymous with the simulation of the paths of the associated stochastic process. On account of the fact that such equations occur naturally in mathematical finance in connection with (exotic) option pricing, there are already many numerical and stochastic methods in existence for the general Lévy setting. The reader is referred, for example, to the books Cont & Tankov (2004), Boyarchenko & Levendorskiĭ (2002) and the references therein. Other sources offering simulation techniques can be found (e.g., Janicki & Weron, 1994; Asmussen & Rosiński, 2001; Cohen & Rosinski, 2007; Cohen et al., 2010). Similarly to works in mathematical finance, they are mostly focused on the approximation of the stable process (and indeed the general Lévy process) by a compound Poisson process or a power series representation of the path, with a diffusive component to mimic the effect of small jumps. To our knowledge, however, the walk-on-spheres approach to path simulation has not been used in the context of simulating stable processes to date, nor, as alluded to above, to the end of numerically solving Dirichlet-type problems for the fractional Laplacian.
4. Walk-on-spheres for the fractional Laplacian
We start by describing the walk-on-spheres for the fractional Laplacian Dirichlet problem (1.7) on a convex domain |$D$|. The domain |$D$| may be unbounded, as long as |$D^{\textrm{c}}$| has nonzero measure (even though Theorem 1.3 requires boundedness). Fix |$x\in D$|. The walk-on-spheres |$(\rho_n$|, |$n\geq 0)$|, with |$\rho_0 = x$| is defined in a similar way to the Brownian setting in the sense that, given |$\rho_{n-1}$|, the distribution of |$\rho_{n}$| is selected according to an independent copy of |$X_{\sigma_{B_n}}$| under |$\mathbb{P}_{\rho_{n-1}}$|, where |$B_n = \{x\in\mathbb{R}^d\colon |x - \rho_{n-1}|< r_n\}$| and |$\sigma_{B_n} = \inf\{t>0: X_t\not\in B_n\}$|. The algorithm comes to an end at the random index |$N = \min\{n\geq 0\colon \rho_n\not\in D\}$|, again using the standard understanding that |$\min\emptyset {:=} \infty$|. See, for example, the depiction in Fig. 3.

Steps of the walks-on-sphere algorithm until exiting the convex domain |$D$| in the stable setting. In this realization, |$N = 3$|.
Even though the domain |$D$| may be unbounded, our main result predicts that, irrespective of the point of issue of the algorithm, there will always be at most a geometrically distributed number of steps (whose parameter also does not depend on the point of issue) before the algorithm ends.
There are a number of remarks that we can make from the conclusion above.
Although |${\it{\Gamma}}$| has the same distribution for each |$x\in D$|, it is not the same random variable for each |$x\in D$|. As we shall see in the proof of the above theorem, the inequality |$N\leq {\it{\Gamma}}$| is derived by comparing each step of the walk-on-spheres algorithm with a sequence of Bernoulli random variables. This sequence of Bernoulli random variables are defined up to null sets, which may be different under each |$\mathbb{P}_x$|. Therefore, although the distribution of |${\it{\Gamma}}$| does not depend on |$x$|, its null sets do.
- The stochastic domination in Theorem 4.1 is much stronger than the usual comparison of the mean number of steps. Indeed, although it immediately implies that |$\mathbb{E}_x[N] = 1/p$|, we can also deduce that there is an exponentially decaying tail in the distribution of the number of steps. Specifically, for any |$x\in D$|,\[ \mathbb{P}(N >n ) \leq \mathbb{P}({\it{\Gamma}} >n ) = (1-p)^n, \qquad n\in\mathbb{N}. \]
- The randomness in the geometric random variables |${\it{\Gamma}}$| is heavily correlated to |$N$|. The fact that each of the |${\it{\Gamma}}$| are geometrically distributed has the advantage that\[ \sup_{x\in D}\mathbb{E}_x[N] \leq \sup_{x\in D}\mathbb{E}_x[{\it{\Gamma}}] = \frac{1}{p}. \]
However, it is less clear as to what kind of distributional properties can be said of the random variable |$ \sup_{x\in D}{\it{\Gamma}}, $| which a.s. upper bounds |$\sup_{x\in D}N$|.
Finally, it is worth stating formally that the walk-on-spheres algorithm is unbiased and therefore, providing |$\mathbb{E}_x[{g}(X_{\tau_D})]<\infty$|, the strong law of large numbers applies and a straightforward Monte Carlo simulation of the solution to (1.7) is possible. Moreover, providing |$\mathbb{E}_x[{g}(X_{\tau_D})^2]<\infty$|, the central limit theorem offers the rate of convergence.
The first part is a straightforward consequence of the earlier mentioned strong law of large numbers and the fact that Theorem 1.3 ensures that |$\mathbb{E}_x[{g}(\rho_N)]=\mathbb{E}_x[{g}(X_{\tau_D})]<\infty$|. For the second part, we need to show that (4.2) implies |$\mathbb{E}_x[{g}(\rho_N)^2]=\mathbb{E}_x[{g}(X_{\tau_D})^2]<\infty$|. However, if we consider the computation in (A.2) of the appendix, which shows that |$\mathbb{E}_x[{g}(X_{\tau_D})]<\infty$| when |${g}$| is continuous and in |$L^1_{\alpha}(D^\mathrm{c})$|, then it is easy to see that the same statement holds replacing |${g}$| by |${g}^2$|. Under finiteness of the second moment, the central limit theorem completes the proof. □
We now return to the proof of Theorem 4.1. Our approach is to break it into several parts. For convenience, we shall henceforth write |$X^{(x)}=(X^{(x)}(t)\colon t\geq 0)$| to indicate the dependency of |$X$| on its initial position |$X_0 = x$| (equivalent to writing |$(X, \mathbb{P}_x)$|). For any |$x = (x_1, \dots, x_d)\in\mathbb{R}^d$|, such that |$x_1>0$|, we have |$V(x) = \{(z_1, \dots, z_d)\in\mathbb{R}^d \colon z_1>0\}$| for the open half-space containing |$x$| and denote its boundary |$\partial V(x) = \{(z_1, \dots, z_d)\in\mathbb{R}^d \colon z_1=0\}$|. For any Borel set |$A\subset\mathbb{R}^d$|, we write |$ \sigma_A = \inf\{t>0 \colon X_t\not\in A\}. $| We will typically use in place of |$A$| the set |$V(x)$| as well as |$B(x,1) = \{z\in\mathbb{R}^d\colon |z- x|<1\}$|, the unit ball centred at |$x\in\mathbb{R}^d$|. Finally write |${\rm\bf i} = (1,0, \dots, 0)\in\mathbb{R}^d$|.
Without loss of generality (by appealing to the spatial homogeneity of |$X$|, which allows us to appropriately choose our coordinate system) suppose that |$x= |x|\,{ \rm\bf i}\in D$| is such that |$\partial V(x)$| is a tangent hyperplane to both |$D$| and |$B_1$|. Then, |$X^{(x)}_{\sigma_{B_1}}$| is equal in distribution to |$|x|\, X^{(\rm\bf i)}_{\sigma_{B({\rm\bf i},1)}}$| and |$X^{(x)}_{\sigma_{V(x)}}$| is equal in distribution to |$|x|\,X^{(\rm\bf i)}_{\sigma_{V(\mathbf{i}) }}$|.
An important consequence of the previous result is the comparison between the first exit from the largest sphere in |$D$| centred at |$x$| and the first exit from the tangent hyperplane to the latter sphere. Recall that |$B_n = \{z\in\mathbb{R}^d\colon |z - \rho_{n-1}|< r_n\}$| denotes the |$n$|th sphere.
The inequality follows from the inclusion |$D\subset V(x)$|. The formula for |$p(\alpha,d)$| uses the coordinate system and scaling property of stable processes in Lemma 4.3 as well as the identity for the first exit from a sphere given by Theorem 3.1. □
We are now ready to prove our main result.
5. Nonconvex domains
The key element in the proof of Theorem 4.1 is the comparison of the event that the next step of the walk-on-spheres exits the domain |$D$| with the event that the next step of the walk-on-spheres exits a larger, more regular domain. More precisely, the aforesaid regular domain is taken to be the half-space that contains |$D$| with boundary hyperplane that is tangent to both the current maximal sphere and |$D$|. It is the use of a half-space that allows us to work with unbounded domains, but that forces the assumption that |$D$| is convex. With a little more care, we can remove the need for convexity without disturbing the main idea of the proof. However, this will come at the cost of insisting that |$D$| is bounded. It does, however, open the possibility that |$D$| is not a connected domain. We give two results in this respect.
For the first one, we introduce the following definition, which has previously been used in the potential analysis of stable processes; (see, for example, Chen & Song, 1998).
It is well known that, for example, bounded |$C^{1,1}$| domains satisfy (UECC). We need a slightly more restrictive class of domains than those respecting UECC. See Fig. 4.
We say that |$D$| satisfies the regularized uniform exterior-cone condition (RUECC), if it is UECC, and the following additional condition holds: for each |$x\in D$|, suppose that |$\partial(x)$| is the closest point on the boundary of |$D$| to |$x$|. Then, the isometric cone that qualifies |$D$| as UECC can be placed with its vertex at |$\partial(x)$| and symmetrically oriented around the line that passes through |$x$| and |$\partial(x)$|.
Intuitively, |$N(\varepsilon)$| is the step that exits the inner |$\varepsilon$|-thickened boundary of |$D$|.
Reviewing the line of reasoning in the proof of Theorem 4.1, we see that this comparison of events on the first step can be repeated at each surviving step of the algorithm to deduce the claimed result. □
The |$\mathcal{O}(\varepsilon^{-\alpha})$| bound in (5.1) can be compared with the bounds achieved by Binder & Braverman (2012) for the classical walk-on-spheres with Brownian motion for domains with more general geometries than convex. The worst case in Binder & Braverman (2012) is |$\mathcal{O}(\varepsilon^{2-4/a})$| for a parameter |$a>0$| (describing the domain’s thickness or fractal boundary). Notably in the limit |$\alpha\to 2$| (|$X$| converges to Brownian motion) and |$a \to \infty$| (the domain loses regularity), the two agree with an |$\mathcal{O}(\varepsilon^{-2})$| bound.
6. Fractional Poisson problem
The combinations of Theorem 2.10 and 3.2 in Bucur (2016) treat the case that |$D$| is a ball. In the more general setting, among others, Bogdan & Byczkowski (1999), Ros-Oton & Serra (2014) and Ros-Oton (2016) (see also citations therein) offer results in this direction, albeit from a more analytical perspective. We give a new probabilistic proof of Theorem 6.1 in the appendix using a method that combines the idea of walk-on-spheres with the version of Theorem 6.1 when |$D$| is a ball. It is for this reason that the (otherwise unclear) need for the assumption that |${f}\in C^{\alpha +\varepsilon}(\overline{D})$| enters. Note in particular that Theorem 1.3 follows as a corollary.
Although the above identity is presented in a probabilistic context, it has a much older history in the analysis literature. Known as Boggio’s formula, the original derivation in the setting of potential theory dates back to Boggio (1905). See the discussion in Delaurentis & Romero (1990) and Bucur (2016).
In the next result, we will write as a slight abuse of notation |$V_r(x,{f}(\cdot)) = \int_{|y-x|<r}{f}(y)\,V_r(x,{\rm d}y)$| for bounded measurable |${f}$|.
The proof is now complete. □
Theorem 6.1 and Lemma 6.3 ensure that the strong law of large numbers may be invoked. For the central limit theorem, we need |$\mathbb{E}_x[\chi^2]<\infty$|. Taking account of the fact that |$\chi$| is the sum of two terms, the Cauchy–Schwarz inequality ensures that |$\mathbb{E}_x[\chi^2] $| is finite if |$\mathbb{E}_x\left[{g}(\rho_{N})^2\right] $| and |$ \mathbb{E}_x\left[\left( \sum_{n=0}^{N-1} r_n^{\alpha}\, V_{1}(0, {f}(\rho_n + r_n\cdot))\right)^2\right]$| are finite. Recall that |$\mathbb{E}_x\left[{g}(\rho_{N})^2\right] = \mathbb{E}_x[{g}(X_{\sigma_D})^2]$| and, from Corollary 4.2, that (6.7) is sufficient to ensure that this expectation is bounded.
However, the latter expectation can be bounded by |$\mathbb{E}_x[\sigma_{B^*}^2]$|, where |$B^* = B(x, R)$| for some suitably large |$R$|, such that |$D$| is compactly embedded in |$B^*$|. Moreover, appealing to Getoor (1961), we know that |$\mathbb{E}_x[\sigma_{B^*}^2]$| is bounded. □
7. Numerical experiments
In the following section, all of the routines associated with the simulations are publicly available at the following repository: https://bitbucket.org/wos_paper/wos_repo.
This gives two terms: one can be evaluated directly (by storing |$\mathbb{E} [1-I(R^2; 1-\alpha/2, \alpha/2)]$|) and the second can be evaluated using a Monte Carlo method, but with smaller variance (as the quantity in square brackets in (7.5) is |${\mathcal{O}{\left({{r_n}}\right)}}$|). It is worth noting that a similar mixed approach using the trapezoidal rule over |$\theta$| and randomizing |$r$| as earlier for evaluating the left-hand side of (7.4) was also tested. However, the results showed that the pure Monte Carlo approach is, in comparison with the mixed one, superior with regard to accuracy and computational cost. With this view, we decided to focus on the first one.
Accuracy of this algorithm and its feasibility of implementation were checked with model solutions to problems of the type (6.1) and they are presented below in order.
7.1 Free-space Green’s function

Example simulation for (1.7) with exterior data |${g}(x)=G(x,y)$| with |$y=(2,0)$| on the domain given by the unit ball centred at the origin, based on |$10^6$| samples. The left-hand plot shows the relative error and the right-hand plot shows the sample variance. The sample variance is larger for small |$\alpha$| as the process stops further away from the boundary and can see the singularity at |$(2,0)$| in the exterior data. Accordingly, the relative error is higher as we are using a fixed number of samples.
7.2 Gaussian data
This integral can be computed numerically via a quadrature approximation. Here, instead of a fixed number of samples, the number of samples is taken adaptively based on a tolerance |$\varepsilon$| for the computed sample standard deviation. Figure 6 shows the results with |$y=(2,0)$| and tolerance |$\varepsilon=10^{-4}$| for the evaluation of |$u(0.6,0.6)$| as previously. The estimator standard deviation and absolute error exhibit no obvious trend, whereas the sample variance peaks at about |$\alpha=0.6$|. Also, at this value, the largest number of samples is needed to satisfy the tolerance. Despite the sample variance decreasing after |$\alpha=0.6$|, there is an increasing trend in the amount of work required. This implies that the increase in the number of steps with |$\alpha$| (see Fig. 10) dominates, and therefore a solution point of accuracy |$10^{-4}$| is computationally more costly for larger values of |$\alpha$|.

Example simulation with the walk-on-spheres algorithm for (1.7) based on desired tolerance of |$10^{-4}$|. From top left to bottom right, we see the standard deviation of the estimator, the sample variance, the absolute error (using a quadrature approximation for (7.6) for the reference value) and the amount work (number of samples |$\times$| mean number of steps).
7.3 Nonconstant source term

Example simulation with the walk-on-spheres algorithm for (6.1) based on desired tolerance of |$10^{-3}$| and |$n_{R,{\it{\Theta}}}=1000$|. From top left to bottom right, we see the standard deviation of the estimator, the sample variance, the absolute error and the amount work (number of samples |$\times$| mean number of steps).
7.4 Distribution of the number of steps in convex and nonconvex domains
In previous sections, a large focus was put on deriving upper bounds and limiting distributions for |$N$|. Here, we provide numerical support for these theoretical results. The walk-on-spheres algorithm was simulated inside a unit-ball domain centred at the origin as well as inside a domain of a hundred touching unit balls centred at points |$(i,j)$|, |$i,j=-10,\dots,10$|, the so-called ‘Swiss cheese’ domain as shown in Figure 8. The first represents a convex domain, whereas the latter a nonconvex one. The algorithm was started at a point |$x=(\sqrt{0.29},-\sqrt{0.7})$| which lies very close to the boundary in both domains. This point was chosen, as numerical simulations in a unit-ball domain revealed that the mean number of steps decreases with increasing distance from the boundary of the starting point. Theorem 4.1 states that |$N$| is stochastically dominated by a geometric distribution with parameter |$p(\alpha,d)$|. In two dimensions, we are able to numerically compute |$p(\alpha,2)$|, because it is the solution to (1.7) with |$D = B(0, 1)$|, |${g}(x)=\mathbf{1}_{\{x_1<-1\}}(x)$| and zero source term |${f}=0$| as deduced from Corollary 4.4. We computed values of |$p(\alpha,2)$| for different |$\alpha$| to accuracy |$10^{-4}$|.
The left-hand histogram in Fig. 9 confirms the stochastic dominance of |${\it{\Gamma}}$| and an exponentially decaying tail as stated in Remark 2 of Theorem 4.1. However, the right-hand histogram shows that this statement fails in the particular example of the Swiss cheese domain. Moreover, the plot of the mean number of steps against |$\alpha$| in Fig. 10 shows the observed value of |$\mathbb{E}_x[N]$| is bounded above by |$1/p(\alpha,2)$| for the unit-ball domain. On the other hand, this is not the case for the Swiss cheese domain, where the observed value of |$\mathbb{E}_x[N]$| exceeds |$1/p(\alpha,2)$| for |$\alpha$| in the range |$(0.3,1.6)$|.

Histogram of the proportion of runs of the walk-on-spheres algorithm with |$\alpha=1$|, for which |$N>n$| for the unit-ball domain (left) and the Swiss cheese domain (right). The dotted curve shows the tail of Geom(|$p(1,2)$|), this is |$(1-p(1,2))^n$|, as in Remark 2 of Theorem 4.1.

Mean number of steps for the walk-on-spheres algorithm started at |$x=(\sqrt{0.29},-\sqrt{0.7})$| inside the circle domain (left) and inside the Swiss cheese domain (right). The dashed curve on both plots is |$1/p(\alpha,2)$| as in Corollary 4.4.
An explanation for why this is happening might be as follows. At larger values of |$\alpha$|, the path of |$X$| starts resembling that of a Brownian motion (albeit with a countable infinity of arbitrarily small discontinuities). The process |$X$| is started inside a ball in the Swiss cheese. When it exits this ball, its exit position is relatively close to the boundary with high probability. Therefore, the exit point of the ball containing the point of issue is more likely to be in the ‘cheese’ (which would cause an end to the algorithm) and less likely to be inside another vacuous ball. Accordingly, |$\mathbb{E}_x[N]$| does not deviate largely from the example of a single ball. However, for small values of |$\alpha$|, exit points from the sphere containing the point of issue have a higher probability to be far from the boundary, landing inside another vacuous ball, thereby requiring the algorithm to continue. In that case, the comparison with the case of exiting a single sphere breaks down.
Acknowledgements
We would like to thank Mateusz Kwaśniki for pointing out a number of references to us and Alexander Freudenberg for a close reading of an earlier version of this article.
Funding
Engineering and Physical Sciences Research Council (EPSRC) grant (EP/L002442/1 to A.E.K.).
Appendix. Proof of Theorem 6.1
Our proof of Theorem 6.1 uses heavily the joint conclusion of Theorems 2.10 and 3.2 in Bucur (2016), namely that the Theorem 6.1 is true in the case that |$D$| is a ball. Our proof is otherwise constructive proving existence and uniqueness separately.
Existence: On account of the fact that |$D$| is bounded, we can define a ball of sufficiently large radius |$R>0$|, say |$B^* = B(X_0, R)$|, centred at |$X_0$|, such that |$D$| is a subset of |$B^*$| and hence |$\sigma_D \leq \sigma_{B^*}$| a.s., irrespective of the initial position of |$X$|, where |$\sigma_{B^*} = \inf\{t>0\colon X_t\not\in B^*\}$|. In particular, thanks to stationary and independent increments, this upper bound for |$\sigma_D$| does not depend on |$X_0$| in law and |$\sup_{x\in D}\mathbb{E}_x[\sigma_D]\leq \mathbb{E}_0[\sigma_{B^*}]<\infty$|.
The boundedness of |${f}$| on |$D$| and the uniform finite mean of |$\sigma_D$| ensures that the second expectation in the definition of |$\upsilon$| is bounded on |$\overline{D}$|. We claim that |$\upsilon$| is continuous in |$\mathbb{R}^d$| and belongs to |$L^1_\alpha(\mathbb{R}^d)$|. Continuity of |$\upsilon$| follows, thanks to the path regularity of |$X$|, the continuity of |${g}$|, the openness of |$D$| and the fact that |$\omega\mapsto X_{\sigma_D}(\omega)$| and |$\omega\mapsto\int_0^{\sigma_D(\omega)} {f}(X_s(\omega))\,{\rm d}s$| are continuous in the Skorohod topology (for which it is important that |$\omega\mapsto\sigma_D(\omega)$| is finite). Continuity is also a consequence of the classical potential analytic point of view, seeing the identity for |$\upsilon$| in (6.3) in terms of Riesz potentials (see, for example, the classical texts of Landkof, 1972; Bliedtner & Hansen, 1986).
To check that |$\upsilon\in L^1_\alpha(\mathbb{R}^d)$|, we need some estimates. For |$x\in D^{\rm c}$|, |$\upsilon(x) = {g}(x)$| and hence, as |${g}\in L^1_\alpha(D^\mathrm{c})$|, it suffices to check that |$ \int_{D}|\upsilon(x)|/(1+|x|^{\alpha + d})\,{\rm d}x<\infty.$| However, this is trivial on account of the boundedness and continuity of |$\upsilon$| on |$\overline D$|.
On account of the fact that |$\mathbb{P}_x(\sigma_D =0) = 1$| for all |$x\in D^{\rm c}$|, it follows that |$\upsilon = {g}$| on |$D^{\rm c}$| and hence (A.5) is identical to (6.1).