Walk-on-spheres method

Last updated

In mathematics, the walk-on-spheres method (WoS) is a numerical probabilistic algorithm, or Monte-Carlo method, used mainly in order to approximate the solutions of some specific boundary value problem for partial differential equations (PDEs). [1] [2] The WoS method was first introduced by Mervin E. Muller in 1956 to solve Laplace's equation, [1] and was since then generalized to other problems.

Contents

It relies on probabilistic interpretations of PDEs, and simulates paths of Brownian motion (or for some more general variants, diffusion processes), by sampling only the exit-points out of successive spheres, rather than simulating in detail the path of the process. This often makes it less costly than "grid-based" algorithms, and it is today one of the most widely used "grid-free" algorithms for generating Brownian paths.

Informal description

Let be a bounded domain in with a sufficiently regular boundary , let h be a function on , and let be a point inside .

Consider the Dirichlet problem:

It can be easily shown [lower-alpha 1] that when the solution exists, for :

where W is a d-dimensional Wiener process, the expected value is taken conditionally on {W0 = x}, and τ is the first-exit time out of Ω.

To compute a solution using this formula, we only have to simulate the first exit point of independent Brownian paths since with the law of large numbers:

The WoS method provides an efficient way of sampling the first exit point of a Brownian motion from the domain, by remarking that for any (d  1)-sphere centred on x, the first point of exit of W out of the sphere has a uniform distribution over its surface. Thus, it starts with x(0) equal to x, and draws the largest sphere centered on x(0) and contained inside the domain. The first point of exit x(1) from is uniformly distributed on its surface. By repeating this step inductively, the WoS provides a sequence (x(n)) of positions of the Brownian motion.

According to intuition, the process will converge to the first exit point of the domain. However, this algorithm takes almost surely an infinite number of steps to end. For computational implementation, the process is usually stopped when it gets sufficiently close to the border, and returns the projection of the process on the border. This procedure is sometimes called introducing an -shell, or -layer. [4]

Formulation of the method

Illustration of a run of the Walk-on-spheres algorithm on an arbitrary domain
O
{\displaystyle \Omega }
with an
e
{\displaystyle \varepsilon }
-shell Walk on Spheres illustration.jpg
Illustration of a run of the Walk-on-spheres algorithm on an arbitrary domain with an -shell

Choose . Using the same notations as above, the Walk-on-spheres algorithm is described as follows:

  1. Initialize :
  2. While :
    1. Set .
    2. Sample a vector uniformly distributed over the unit sphere, independently from the preceding ones.
    3. Set
  3. When :
  4. , the orthogonal projection of on
  5. Return

The result is an estimator of the first exit point from of a Wiener process starting from , in the sense that they have close probability distributions (see below for comments on the error).

Comments and practical considerations

Radius of the spheres

In some cases the distance to the border might be difficult to compute, and it is then preferable to replace the radius of the sphere by a lower bound of this distance. One must then ensure that the radius of the spheres stays large enough so that the process reaches the border. [1]

Bias in the method and GFFP

The Walk-on-spheres method is used until the process gets
d
{\displaystyle \delta }
-close to the border. Then the sphere is replaced by its "intersection" with the boundary of the domain. Improved WoS method illustration.png
The Walk-on-spheres method is used until the process gets -close to the border. Then the sphere is replaced by its "intersection" with the boundary of the domain.

As it is a Monte-Carlo method, the error of the estimator can be decomposed into the sum of a bias, and a statistical error. The statistical error is reduced by increasing the number of paths sampled, or by using variance reduction methods.

The WoS theoretically provides exact (or unbiased) simulations of the paths of the Brownian motion. However, as it is formulated here, the -shell introduced to ensure that the algorithm terminates also adds an error, usually of order . [4] This error has been studied, and can be avoided in some geometries by using Green's Functions First Passage method: [5] one can change the geometry of the "spheres" when close enough to the border, so that the probability of reaching the border in one step becomes positive. This requires the knowledge of Green's functions for the specific domains. (see also Harmonic measure)

When it is possible to use it, the Green's function first-passage (GFFP) method is usually preferred, as it is both faster and more accurate than the classical WoS. [4]

Complexity

It can be shown that the number of steps taken for the WoS process to reach the -shell is of order . [2] Therefore, one can increase the precision to a certain extent without making the number of steps grow notably.

As it is commonly the case for Monte-Carlo methods, this algorithm performs particularly well when the dimension is higher than , and one only needs a small set of values. Indeed, the computational cost increases linearly with the dimension, whereas the cost of grid dependant methods increase exponentially with the dimension. [2]

Variants, extensions

This method has been largely studied, generalized and improved. For example, it is now extensively used for the computation of physical properties of materials (such as capacitance, electrostatic internal energy of molecules, etc.). Some notable extensions include:

Elliptic equations

The WoS method can be modified to solve more general problems. In particular, the method has been generalized to solve Dirichlet problems for equations of the form [6] (which include the Poisson and linearized Poisson−Boltzmann equations) or for any elliptic partial differential equation with constant coefficients. [7]

More efficient ways of solving the linearized Poisson–Boltzmann equation have also been developed, relying on Feynman−Kac representations of the solutions. [8]

Time dependency

Again, within a regular enough border, it possible to use the WoS method to solve the following problem :

of which the solution can be represented as: [9]

,

where the expectation is taken conditionally on

To use the WoS through this formula, one needs to sample the exit-time from each sphere drawn, which is an independent variable with Laplace transform (for a sphere of radius ): [10]

The total time of exit from the domain can be computed as the sum of the exit-times from the spheres. The process also has to be stopped when it has not exited the domain at time .

Other extensions

The WoS method has been generalized to estimate the solution to elliptic partial differential equations everywhere in a domain, rather than at a single point. [11]

The WoS method has also been generalized in order to compute hitting times for processes other than Brownian motions. For example, hitting times of Bessel processes can be computed via an algorithm called "Walk on moving spheres". [12] This problem has applications in mathematical finance.

The WoS can be adapted to solve the Poisson and Poisson–Boltzmann equation with flux conditions on the boundary. [13]

Finally, WoS can be used to solve problems where coefficients vary continuously in space, via connections with the volume rendering equation. [14]

See also

Notes

  1. The link was first established by Kakutani for the 2-dimensional Brownian motion, [3] it can now be seen as a trivial case of the Feynman−Kac formula.

Related Research Articles

<span class="mw-page-title-main">Fokker–Planck equation</span> Partial differential equation

In statistical mechanics, the Fokker–Planck equation is a partial differential equation that describes the time evolution of the probability density function of the velocity of a particle under the influence of drag forces and random forces, as in Brownian motion. The equation can be generalized to other observables as well.

In mathematics, the Poisson summation formula is an equation that relates the Fourier series coefficients of the periodic summation of a function to values of the function's continuous Fourier transform. Consequently, the periodic summation of a function is completely defined by discrete samples of the original function's Fourier transform. And conversely, the periodic summation of a function's Fourier transform is completely defined by discrete samples of the original function. The Poisson summation formula was discovered by Siméon Denis Poisson and is sometimes called Poisson resummation.

The Feynman–Kac formula named after Richard Feynman and Mark Kac, establishes a link between parabolic partial differential equations (PDEs) and stochastic processes. In 1947 when Kac and Feynman were both on Cornell faculty, Kac attended a presentation of Feynman's and remarked that the two of them were working on the same thing from different directions. The Feynman–Kac formula resulted, which proves rigorously the real case of Feynman's path integrals. The complex case, which occurs when a particle's spin is included, is still unproven.

<span class="mw-page-title-main">Propagator</span> Function in quantum field theory showing probability amplitudes of moving particles

In quantum mechanics and quantum field theory, the propagator is a function that specifies the probability amplitude for a particle to travel from one place to another in a given period of time, or to travel with a certain energy and momentum. In Feynman diagrams, which serve to calculate the rate of collisions in quantum field theory, virtual particles contribute their propagator to the rate of the scattering event described by the respective diagram. These may also be viewed as the inverse of the wave operator appropriate to the particle, and are, therefore, often called (causal) Green's functions.

The Hamiltonian constraint arises from any theory that admits a Hamiltonian formulation and is reparametrisation-invariant. The Hamiltonian constraint of general relativity is an important non-trivial example.

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics. The Hamilton–Jacobi equation is particularly useful in identifying conserved quantities for mechanical systems, which may be possible even when the mechanical problem itself cannot be solved completely.

In statistics, econometrics and signal processing, an autoregressive (AR) model is a representation of a type of random process; as such, it is used to describe certain time-varying processes in nature, economics, etc. The autoregressive model specifies that the output variable depends linearly on its own previous values and on a stochastic term ; thus the model is in the form of a stochastic difference equation. Together with the moving-average (MA) model, it is a special case and key component of the more general autoregressive–moving-average (ARMA) and autoregressive integrated moving average (ARIMA) models of time series, which have a more complicated stochastic structure; it is also a special case of the vector autoregressive model (VAR), which consists of a system of more than one interlocking stochastic difference equation in more than one evolving random variable.

The Havriliak–Negami relaxation is an empirical modification of the Debye relaxation model in electromagnetism. Unlike the Debye model, the Havriliak–Negami relaxation accounts for the asymmetry and broadness of the dielectric dispersion curve. The model was first used to describe the dielectric relaxation of some polymers, by adding two exponential parameters to the Debye equation:

<span class="mw-page-title-main">Inelastic mean free path</span>

The inelastic mean free path (IMFP) is an index of how far an electron on average travels through a solid before losing energy.

<span class="mw-page-title-main">Duffing equation</span> Non-linear second order differential equation and its attractor

The Duffing equation, named after Georg Duffing (1861–1944), is a non-linear second-order differential equation used to model certain damped and driven oscillators. The equation is given by

<span class="mw-page-title-main">Newman–Penrose formalism</span> Notation in general relativity

The Newman–Penrose (NP) formalism is a set of notation developed by Ezra T. Newman and Roger Penrose for general relativity (GR). Their notation is an effort to treat general relativity in terms of spinor notation, which introduces complex forms of the usual variables used in GR. The NP formalism is itself a special case of the tetrad formalism, where the tensors of the theory are projected onto a complete vector basis at each point in spacetime. Usually this vector basis is chosen to reflect some symmetry of the spacetime, leading to simplified expressions for physical observables. In the case of the NP formalism, the vector basis chosen is a null tetrad: a set of four null vectors—two real, and a complex-conjugate pair. The two real members asymptotically point radially inward and radially outward, and the formalism is well adapted to treatment of the propagation of radiation in curved spacetime. The Weyl scalars, derived from the Weyl tensor, are often used. In particular, it can be shown that one of these scalars— in the appropriate frame—encodes the outgoing gravitational radiation of an asymptotically flat system.

In mathematics, the theory of optimal stopping or early stopping is concerned with the problem of choosing a time to take a particular action, in order to maximise an expected reward or minimise an expected cost. Optimal stopping problems can be found in areas of statistics, economics, and mathematical finance. A key example of an optimal stopping problem is the secretary problem. Optimal stopping problems can often be written in the form of a Bellman equation, and are therefore often solved using dynamic programming.

In perturbation theory, the Poincaré–Lindstedt method or Lindstedt–Poincaré method is a technique for uniformly approximating periodic solutions to ordinary differential equations, when regular perturbation approaches fail. The method removes secular terms—terms growing without bound—arising in the straightforward application of perturbation theory to weakly nonlinear problems with finite oscillatory solutions.

In game theory, a stochastic game, introduced by Lloyd Shapley in the early 1950s, is a repeated game with probabilistic transitions played by one or more players. The game is played in a sequence of stages. At the beginning of each stage the game is in some state. The players select actions and each player receives a payoff that depends on the current state and the chosen actions. The game then moves to a new random state whose distribution depends on the previous state and the actions chosen by the players. The procedure is repeated at the new state and play continues for a finite or infinite number of stages. The total payoff to a player is often taken to be the discounted sum of the stage payoffs or the limit inferior of the averages of the stage payoffs.

In mathematics, some boundary value problems can be solved using the methods of stochastic analysis. Perhaps the most celebrated example is Shizuo Kakutani's 1944 solution of the Dirichlet problem for the Laplace operator using Brownian motion. However, it turns out that for a large class of semi-elliptic second-order partial differential equations the associated Dirichlet boundary value problem can be solved using an Itō process that solves an associated stochastic differential equation.

<span class="mw-page-title-main">Anatoly Karatsuba</span> Russian mathematician

Anatoly Alexeyevich Karatsuba was a Russian mathematician working in the field of analytic number theory, p-adic numbers and Dirichlet series.

The narrow escape problem is a ubiquitous problem in biology, biophysics and cellular biology.

<span class="mw-page-title-main">Stochastic gradient Langevin dynamics</span>

Stochastic gradient Langevin dynamics (SGLD) is an optimization and sampling technique composed of characteristics from Stochastic gradient descent, a Robbins–Monro optimization algorithm, and Langevin dynamics, a mathematical extension of molecular dynamics models. Like stochastic gradient descent, SGLD is an iterative optimization algorithm which uses minibatching to create a stochastic gradient estimator, as used in SGD to optimize a differentiable objective function. Unlike traditional SGD, SGLD can be used for Bayesian learning as a sampling method. SGLD may be viewed as Langevin dynamics applied to posterior distributions, but the key difference is that the likelihood gradient terms are minibatched, like in SGD. SGLD, like Langevin dynamics, produces samples from a posterior distribution of parameters based on available data. First described by Welling and Teh in 2011, the method has applications in many contexts which require optimization, and is most notably applied in machine learning problems.

The redundancy principle in biology expresses the need of many copies of the same entity to fulfill a biological function. Examples are numerous: disproportionate numbers of spermatozoa during fertilization compared to one egg, large number of neurotransmitters released during neuronal communication compared to the number of receptors, large numbers of released calcium ions during transient in cells and many more in molecular and cellular transduction or gene activation and cell signaling. This redundancy is particularly relevant when the sites of activation is physically separated from the initial position of the molecular messengers. The redundancy is often generated for the purpose of resolving the time constraint of fast-activating pathways. It can be expressed in terms of the theory of extreme statistics to determine its laws and quantify how shortest paths are selected. The main goal is to estimate these large numbers from physical principles and mathematical derivations.

<span class="mw-page-title-main">Lorentz oscillator model</span> Theoretical model describing the optical response of bound charges

The Lorentz oscillator model describes the optical response of bound charges. The model is named after the Dutch physicist Hendrik Antoon Lorentz. It is a classical, phenomenological model for materials with characteristic resonance frequencies for optical absorption, e.g. ionic and molecular vibrations, interband transitions (semiconductors), phonons, and collective excitations.

References

  1. 1 2 3 Muller, Mervin E. (Sep 1956). "Some continuous Monte-Carlo Methods for the Dirichlet Problem". The Annals of Mathematical Statistics. 27 (3): 569–589. doi: 10.1214/aoms/1177728169 . JSTOR   2237369.
  2. 1 2 3 Sabelfeld, Karl K. (1991). Monte Carlo methods in boundary value problems. Berlin [etc.]: Springer-Verlag. ISBN   978-3540530015.
  3. Kakutani, Shizuo (1944). "Two-dimensional Brownian motion and harmonic functions". Proceedings of the Imperial Academy. 20 (10): 706–714. doi: 10.3792/pia/1195572706 .
  4. 1 2 3 Mascagni, Michael; Hwang, Chi-Ok (June 2003). "ϵ-Shell error analysis for "Walk On Spheres" algorithms". Mathematics and Computers in Simulation. 63 (2): 93–104. doi:10.1016/S0378-4754(03)00038-7.
  5. Given, James A.; Hubbard, Joseph B.; Douglas, Jack F. (1997). "A first-passage algorithm for the hydrodynamic friction and diffusion-limited reaction rate of macromolecules". The Journal of Chemical Physics. 106 (9): 3761. Bibcode:1997JChPh.106.3761G. doi:10.1063/1.473428.
  6. Elepov, B.S.; Mikhailov, G.A. (January 1969). "Solution of the dirichlet problem for the equation Δu  cu = q by a model of "walks on spheres"". USSR Computational Mathematics and Mathematical Physics. 9 (3): 194–204. doi:10.1016/0041-5553(69)90070-6.
  7. Booth, Thomas E (February 1981). "Exact Monte Carlo solution of elliptic partial differential equations". Journal of Computational Physics. 39 (2): 396–404. Bibcode:1981JCoPh..39..396B. doi:10.1016/0021-9991(81)90159-5.
  8. Hwang, Chi-Ok; Mascagni, Michael; Given, James A. (March 2003). "A Feynman–Kac path-integral implementation for Poisson's equation using an h-conditioned Green's function". Mathematics and Computers in Simulation. 62 (3–6): 347–355. CiteSeerX   10.1.1.123.3156 . doi:10.1016/S0378-4754(02)00224-0.
  9. Gobet, Emmanuel (2013). Méthodes de Monte-Carlo et processus stochastiques du linéaire au non-linéaire. Palaiseau: Editions de l'Ecole polytechnique. ISBN   978-2-7302-1616-6.
  10. Salminen, Andrei N. Borodin; Paavo (2002). Handbook of Brownian motion : facts and formulae (2. ed.). Basel [u.a.]: Birkhäuser. ISBN   978-3-7643-6705-3.
  11. Booth, Thomas (August 1982). "Regional Monte Carlo solution of elliptic partial differential equations" (PDF). Journal of Computational Physics. 47 (2): 281–290. Bibcode:1982JCoPh..47..281B. doi:10.1016/0021-9991(82)90079-1.
  12. Deaconu, Madalina; Herrmann, Samuel (December 2013). "Hitting time for Bessel processes—walk on moving spheres algorithm (WoMS)". The Annals of Applied Probability. 23 (6): 2259–2289. arXiv: 1111.3736 . doi:10.1214/12-AAP900. S2CID   25036031.
  13. Simonov, Nikolai A. (2007). "Random Walks for Solving Boundary-Value Problems with Flux Conditions". Numerical Methods and Applications. Lecture Notes in Computer Science. Vol. 4310. pp. 181–188. CiteSeerX   10.1.1.63.3780 . doi:10.1007/978-3-540-70942-8_21. ISBN   978-3-540-70940-4.{{cite book}}: Missing or empty |title= (help)
  14. Sawhney, Rohan; Seyb, Dario; Jarosz, Wojciech; Crane, Keenan (July 2022). "Grid-Free Monte Carlo for PDEs with Spatially Varying Coefficients". ACM Transactions on Graphics. 41 (4): 1–17. arXiv: 2201.13240 . doi:10.1145/3528223.3530134. S2CID   246430740.

Further reading