INTRODUCTION
Electrostatic problems are subject to the solution of the Poisson equation in an open domain with specific boundary conditions. This kind of problem is often solved using a finite-dimensional approximation. In general, one needs to first build a mesh of the domain that can be used to define basis functions and then compute the solution that satisfies the Poisson equation in a weak sense. Even recent meshless finite element methods (FEMs) still require a careful arrangement of nodes within the domain to discretise the interior of the domain [1, 2]. Consequently, errors appear when one approximates the geometries and underlying physics using basis functions. Moreover, the discretisation of domains or boundaries can be the most expensive procedure or even fail when the geometry is complicated. In addition, comprehensive pre-processing is required for non-standard geometric data, such as point clouds from laser scanning for digital twin applications. For open-domain problems, the domain of interest often needs to be truncated when FEM are employed. This would introduce an error if we simply ground the outermost boundaries far from the region of interest because the potential, in fact, approaches zero at infinity; alternatively, an artificial layer can be introduced as an infinite domain, which would be time-consuming. Even within boundary element methods for which only the boundaries are discretised, one also needs to solve linear systems, which will be a challenge when the number of unknowns is large.
In addition to traditional grid-based numerical methods for partial differential equations (PDEs), novel approaches that incorporate machine learning [3] and neural networks [4], such as methods based on physics-informed neural networks [5, 6], have been proposed. In particular, grid-free Monte Carlo methods have been introduced in the field of geometry processing and rendering [7, 8], which handle the Laplace equation and the Poisson equation. According to Kakutani's principle and the mean value theorem, the walk-on-sphere (WoS) algorithm can determine the expectation of the boundary value of which a random Brownian walk starts from a given point and hits the boundary for the first time to solve a Laplace equation at the given point [9, 10]. Within Monte Carlo methods, no spatial discretisation is required, which shows great advantages when dealing with complex geometries compared to traditional mesh-based methods. Recent efforts have also been devoted to handling other PDEs of complicated physics and inverse problems [11], such as Helmholtz and Yukawa equations [12], Poisson and diffusion equations with varying coefficients for inhomogeneous materials [13], real-time fluid simulation [14], etc. Along with pioneering work [8, 13], many acceleration techniques have been investigated, such as the implementation of the graphics processing unit [15], fast methods to query the closest point within the spatial hierarchy [16], and boundary value caching for the WoS algorithm [17]. In addition, other Monte Carlo grid-free solvers and formulations have been proposed for certain classes of second-order PDEs, including the walk-on-boundary (WoB) solver [18] and the bidirectional formulation for the WoS algorithm [19].
The handling of boundary conditions is critical for Monte Carlo solvers. Similar to the ray tracing of the light transport simulation in rendering [7], random walks are terminated and absorbed at boundaries with Dirichlet boundary conditions. However, it becomes more complicated and a challenge when dealing with other types of boundary conditions. For the Neumann or Robin boundary conditions, it is often required to simulate reflecting random walks (or partially reflecting or absorbed walks) that bounce off the boundary [20–22]. Recently, an efficient and more general walk-on-star (WoSt) has been proposed to handle the Neumann condition [23], which can be viewed as a Monte Carlo estimator for the boundary integral equation of a Laplace problem. For common problems in an open domain, the classical WoS and WoSt algorithms become less efficient because of the enormously useless random walks tending to infinity and the non-recurrent behaviour of Brownian motion. By formulating the appropriate integral equation, the WoB algorithm is reported to handle Dirichlet, Neumann, Robin, and mixed boundary conditions for both the interior and exterior domains [18]. Another efficient approach to exterior problems with pure Dirichlet boundary conditions can be achieved by performing the Kelvin transformation [24, 25]. In this paper, we have demonstrated the spherical inversion of the domain in a more elegant form compared to the previous work [25], which can be generalised to other coordinate transformations. Moreover, we have dealt with a floating potential boundary condition that is neither a pure Dirichlet nor a pure Neumann condition for both the interior and exterior domains. The proposed method has advantages in terms of geometric flexibility and robustness, output sensitivity, and parallelism.
THEORY AND METHODS
Electrostatic BVPs of exterior domains
We start by discussing the common electrostatic boundary value problems (BVPs) with Dirichlet and floating potential boundaries in an open domain Ωinf, as shown in Figure 1, which reads:
[IMAGE OMITTED. SEE PDF]
Coordinate transformation
Using the inversion transformation , that is, the Kelvin transformation, which reads:
Equation (10) is satisfied when the total charge Qm in the m-th metallic region Ωm (or ) are given, where and ∇x denotes the gradient with respect to the transformed coordinates x. Here, dS represents the surface area differential element of in the un-primed coordinates denoted by x; and denotes the unit normal vector of the boundary . Since ψ(x) is formally determined according to Equation (9), Equation (10) can be rewritten as follows:
Direct integral equations for the potential and the gradient
The solution ψ(x) to the linear Poisson equation can be expressed via a boundary integral involving the associated Green's function GΩ(x, y) of a domain Ω with a smooth boundary ∂Ω. According to the Green's second identity, the solution to Equation (7) at an evaluation point x in the interior of can be expressed as follows:
When the domain Ω is deliberately chosen as a ball region centred on point x with radius a, denoted by Ba(x), the analytical expressions of GΩ(x, y) and ∇yGΩ(x, y) can be obtained, as given in Part B in Supporting Information S1. Therefore, Equation (12) reads:
Monte Carlo integration
In general, Equations (13) and (14) can be calculated using a Monte Carlo estimator, which approximates an integral by using random samples of the integrand. In particular, for any integrable function , I = ϕ(x)dx can be approximated by the following:
[IMAGE OMITTED. SEE PDF]
Walk-on-sphere algorithm
Within Monte Carlo estimator Equation (15), the direct integral equation Equation (13) for the potential can be written in an estimated form as follows:
[IMAGE OMITTED. SEE PDF]
Similarly, as illustrated in Figure 3b, the gradient Equation (14) can be estimated as follows:
Note that variance reduction techniques can be applied to estimate the gradient since differentiation would amplify the high-frequency variance [27, 28]. Towards this end, we can replace Equation (18) with the following equation:
Sampling strategies
For Monte Carlo estimators, for example, Equations (17) and (18), the sampling strategies are critical, which determine the probability density functions (PDFs) for the sampling from xi to xi+1 and xi to yi and, in turn, affect the expressions of the estimators. Although Monte Carlo integration and estimators per se are unbiased, proper sampling methods would be critical to reducing variance and improving efficiency.
For the WoS algorithm Equations (17) and (18), the next point xi+1 can be sampled uniformly on the surface of the ball region Ba(xi) centred on xi, which produces the PDF reading as p(xi+1|xi) = 1/|∂Ba(xi)|. For the source term involving the Green's function GB(xi, yi) which will approach infinity when yi approaches the centre xi, ‘importance sampling’ should be considered since the source closer to the centre xi contributes more to Monte Carlo estimation, as suggested in Ref. [8]. can be transformed into the corresponding integration in the local polar coordinates (r, θ), or in the local spherical coordinates (r, θ, ϕ), with and being the Green's functions in a circle and sphere, respectively (see Part B in Supporting Information S1); r = |yi − xi| denotes the distance from the centre xi of the circle (ball), and or would not be singular when r approaches zero. Thus, the ‘normalised Green's function’ can be used as a PDF, which reads p(yi|xi) = GB(xi, yi)/|GB(xi)|, where reads a2/4 and a2/6 for 2D and 3D, respectively. Since GB(xi, yi) is symmetric, we have , where denotes a unit vector that can be sampled uniformly on a unit-sphere or unit-circle. Furthermore, the distance r follows the radial PDF pr(r), which can be sampled by rejection sampling or Ulirich's polar method [8]. In particular, the radial PDF reads and . When handling the source term in Equation (18) to estimate the gradient, we can slightly change the integral as so that or would not be singular when r approaches zero; consequently, similar to the importance sampling for the potential, p(y0|x0) can be the normalised gradient of the Green's function of r∇GB(r), which reads p2D(y0|x0) = 3(1 − r2/a2)/(4πa) and p3D(y0|x0) = (1 − r3/a3)/(3πa). Now, , where r follows the radial PDF , which reads and . In addition to importance sampling of Green's functions, importance sampling of source terms can be used, in particular, for point sources and line sources. Moreover, multiple importance sampling can be used [26].
Monte Carlo methods for floating potentials
We first consider the boundary condition of the floating potential in the untransformed domain, which is governed by Equation (4). Taking the i-th free-standing potential for example, the integral can also be approximated using Monte Carlo estimation according to Equation (15), which reads:
[IMAGE OMITTED. SEE PDF]
Here, the subscript k denotes the k-th estimate of the gradient at the point . Furthermore, the potential on the surface of the ball centred at the point can be recursively estimated using the WoS algorithm Equation (17), which reads:
Note that the random walk path of the k-th estimate starting from the point would terminate at the ɛ-shell of the m-th Dirichlet boundary with floating potential denoted by φm (φ0 denotes φD) after the path.
Now, combining Equation (22) with Equation (26), we can obtain:
Once the floating potentials φm are determined, the potential and the gradient at given points can be readily calculated using the WoS algorithm.
It should be noted that the variance of the Monte Carlo approximation of the integration Equation (22) for the floating potential boundary condition depends on the sample strategies of on the boundary ∂Ωi. In general, we can cast a ray in a random direction from an arbitrary within the domain Ωi [29], resulting in the PDF as , where G0(x, y) denotes the Green's function in free space. Here, is usually chosen as the centre of the domain. When Ωi is non-convex, projecting a ray from xi might have multiple intersections with its boundary ∂Ωi, and only one is chosen as the next point; accordingly, the PDF should be multiplied by 1/Nint, where Nint denotes the number of intersections [18].
Regarding the boundaries of the floating potential in the transformed domain given by Equation (11), we can calculate the floating potential in a similar way. Now, the elements of the linear system read as follows:
Finally, can be solved. Therefore, the potential and/or the gradient of interest can be readily calculated within WoS.
Complexity analysis
We now discuss the space and time complexity of the Monte Carlo method and the WoS algorithm. The procedure includes (i) closest point query of Nelement geometric elements; (ii) single random walk that converges to the ɛ-layer with a thickness of ɛ to the Dirichlet boundary; and (iii) repetition of Nwalk random walks. The time complexity , respectively, reads: (i) if the bounding volume hierarchy is adopted [30]; (ii) [13]; (iii) . Within the Monte Carlo method, it is not necessary to store the intermediate results for each iteration (step) and walk. Therefore, the space complexities are for the geometry and for the field (where Nfield is the number of interested field points) due to the progressive and pointwise (parallelism) nature of the Monte Carlo method. In addition, when handling floating potentials within the proposed Monte Carlo method, the time complexity becomes , where M is the number of electrodes and Ni is the number of samples on the surface of each electrode; and the space complexity remains unchanged.
RESULTS AND DISCUSSION
Impact of the thicknesses of the ɛ-layer and σ-layer
As indicated in Equation (23), the gradient (or the normal derivative) at the point on the boundary ∂Ωi is approximated by the gradient at the offset point , which is located at the distance of σ in the normal direction . This approximation takes into account the evaluation of potential , which can be influenced by the thickness ɛ of the absorbing ɛ-layer, according to the WoS algorithm, as indicated in Equation (16). Subsequently, we examine the impact of parameters ɛ and σ. Note that when the number of random walks N is sufficiently large, for example, N = 107, the sampling error resulting from the Monte Carlo nature of the method could become insignificant in comparison to the errors originated from the ɛ-layer and the σ-layer. The first example is a simplified model of an overhead transmission line above a flat ground, as shown in Figure 5a, for which the infinite domain of interest (shaded region) can be transformed into a bounded region within the Kevin transform, producing an eccentric region to be considered. Figure 5b plots the average number of steps with respect to the ratio ɛ/r0 between the thickness ɛ of the ɛ-layer and the curvature (the radius r0) of the inner wire when random walks end within the ɛ-layer. It is evident that for various curvatures or radii r0, the thinner the ɛ-layer, the more steps will be required to grab the known value on the Dirichlet boundary. Furthermore, the WoS algorithm reaches the Dirichlet boundary in , on average. Figure 5c plots the relative error of the electric potential and of the electric field E with respect to the number of random walks N. It is evident that the relative error is quite small, which decreases dramatically as the number of random walks N increases. Moreover, for a limited maximum number of random walks, variant reduction methods can also be adopted to improve accuracy. Figure 5d depicts the relative error of the potential with respect to the normalised thickness ɛ/r0 of the ɛ-layer. The relative error is shown to be very small even when ɛ/r0 is considerably large, for example, ɛ/r0 = 10−2, indicating that the Monte Carlo method is accurate and feasible for electrostatic problems. As the thickness of the ɛ-layer decreases, the relative errors or derivations (variances) of the solution will approach zero. Figure 5e demonstrates the impact of the thicknesses of the σ-layer on the gradient. Due to the inaccuracy in evaluating the electric potential inside the ɛ-layer, the gradient on the Dirichlet boundary may not be accurate when the thickness of σ-layer is comparable to the thickness of the ɛ-layer [18]. When σ is small, the random walk paths starting from the offset point are more likely to adhere very closely to the boundary; therefore, more random walks may be required to obtain accurate results of the electric field (gradient). Moreover, instead of using the gradient approximation at an offset point as indicated by Equation (23), an exact expression of the boundary derivative may be used, for which the WoB method is incorporated [18].
[IMAGE OMITTED. SEE PDF]
Calculation of capacitance matrix
Having demonstrated the feasibility and precision of the Monte Carlo WoS method, we discuss the floating potential condition by addressing a typical problem of a multi-conductor system, as shown in the inset of Figure 6a. In the model, there are two arrays of electrodes for which the non-excited electrodes satisfy the floating potential condition. When one calculates the mutual capacitance matrix within conventional approaches, such as FEMs or boundary element methods, geometry discretisation is necessary for the electrodes and/or the surrounding space; in addition, each electrode is defined as a port or terminal on which excitation is applied; consequently, the capacitance matrix is subsequently obtained through a steady source sweep [31]. Therefore, multiple simulations of solving linear systems are performed, in fact, which may be time-consuming when the number of electrodes is large and the geometry is complicated. In contrast to conventional methods, as indicated in Equation (30), no linear system is solved to render the capacitance matrix. Furthermore, the precision of the capacitance matrix is determined by the number of random walks Nim that start from ∂Ωi and end in ∂Ωm and the number of sampling points Ni in ∂Ωi, indicating that the precision of Cim can be improved by increasing the number of random walks, as illustrated in Figure 6b. As a reference, the mutual capacitance matrix calculated within the conventional FEM simulation is visualised in Figure 6c. We stress that the convergence of random walks is independent of the size and complexity of the geometry but depends only on the thickness of the ɛ-layer, implying that the proposed method is suitable for calculating the equivalent circuit parameters when the geometry is complicated.
[IMAGE OMITTED. SEE PDF]
Lightning striking near a real tree
Finally, we consider the scenario of a lightning strike near a real tree, as shown in Figure 7a. Real lightning processes can be difficult to model accurately; in practice, a fractal-based random model is often adopted to characterise lightning physics [32]. As for real trees, when fine geometries such as leaves and branches are considered within conventional grid-based numerical methods, tremendous meshes and the number of unknowns will be introduced, which in turn makes the linear system unsolvable. Within the proposed method, no mesh is necessary and no linear system is solved, implying that very complicated geometry, such as real trees, can be handled. Furthermore, as indicated in Equations (19) and (20), the evaluation at each point of interest is independent, implying that one does not need to calculate everywhere as usual and only regions or points of interest are necessary; in addition, progressive results can be rendered as the number of samples increases. Figure 7d (Figure 7g) and Figure 7e (Figure 7h) show the distribution of the electric potential (field) when the number of samples is N = 103 and N = 106, respectively. We note that the potential exhibits a notable variance when N is small; and as the number of samples increases, the results approach the reference (see Figure 7c). We point out that the branches and leaves of the tree may not be modelled as pure metal but rather as dielectrics; it is possible to handle the boundary conditions at dielectric interfaces within Monte Carlo methods, which is currently under development and will be present elsewhere in future work.
[IMAGE OMITTED. SEE PDF]
We note that the computational time for the Monte Carlo method can be longer than that for conventional methods. However, this is not about comparing whether oranges or apples are better. As noted in Ref. [8], discretisation or meshing of complex geometry can be the most expensive procedure of the whole simulation. Therefore, the mesh-free Monte Carlo method may be feasible. Furthermore, variance reduction techniques can be used to obtain accurate results with a small number of samples, as discussed in Part C in Supporting Information S1. Moreover, it is possible to combine the Monte Carlo method with conventional methods, which may become a Swiss-knife tool for electrostatic problems with complex geometries.
CONCLUSION AND PERSPECTIVES
In summary, we have introduced a grid-free Monte Carlo method to handle electrostatic problems with Dirichlet and floating boundary conditions within the WoS algorithm. By applying Green's theorem, the potential at a point of interest can be expressed as an integral or weighted average of the potential on the sphere centred at the given point, which may be evaluated in a recursive manner when the integral is calculated within the Monte Carlo method. According to the Kakutani principle or the Feynman–Kac formula, the solution to a Laplace equation is the expectation of the boundary value for which a random Brownian walk starts from the given point and hits the boundary for the first time, which can be obtained using the WoS algorithm. For the Poisson equation, the source contribution can also be included in each random walk. When floating potential boundary conditions are involved, meshing is still not necessary and only a small linear system is solved. Moreover, when calculating the mutual capacitance matrix of a multi-conductor system, one does not need to solve for swept terminals multiple times, and progressive results towards real values can be obtained as the sampling increases. In addition, only solutions of points or regions of interest need to be evaluated without having to perform a global solution within conventional methods, showing the output sensitivity. Although only Dirichlet boundary conditions along with floating boundary conditions are investigated in this paper, we expect Neumann and Robin boundary conditions to be handled by incorporating other state-of-the-art approaches such as the WoSt [23] and WoB [18] algorithms. Theoretically, provided that an integral equation is derived for the potential or an equivalent source, the Monte Carlo method can be used to handle complex media, such as inhomogeneous media, when the material variations are not large [13], or even with an abrupt change [33, 34]. However, handling a general complex material such as anisotropic media can be a challenge, since the Green's function is complex, which would be an interesting research topic in future work.
ACKNOWLEDGEMENTS
This work was financially supported by the State Grid Corporation of China under grant no. 5500-202155584A-0-5-SF.
CONFLICT OF INTEREST STATEMENT
The authors declare no potential conflict of interests.
DATA AVAILABILITY STATEMENT
The data that support the findings of this study are available on request from the corresponding author.
Seidl, A., Schmidt, T.: Implementation of meshless FEM for engineering applications. Proc. World Acad. Sci. Eng. Technol. 23, 233–237 (2007)
Schaback, R.: Direct discretizations with applications to meshless methods for PDEs. Dolomites Res. Notes Approx. 6(12), 37–50 (2013)
Samaniego, E., et al.: An energy approach to the solution of partial differential equations in computational mechanics via machine learning: concepts, implementation and applications. Comput. Methods Appl. Mech. Eng. 362, 112790 (2020). https://doi.org/10.1016/j.cma.2019.112790
Zang, Y., et al.: Weak adversarial networks for high‐dimensional partial differential equations. J. Comput. Phys. 411, 109409 (2020). https://doi.org/10.1016/j.jcp.2020.109409
Zeng, S., Cai, Y., Zou, Q.: Deep neural networks based temporal‐difference methods for high‐dimensional parabolic partial differential equations. J. Comput. Phys. 468, 111503 (2022). https://doi.org/10.1016/j.jcp.2022.111503
Wang, Y., Zhong, L.: NAS‐PINN: neural architecture search‐guided physics‐informed neural network for solving PDEs. J. Comput. Phys. 496, 112603 (2024) https://doi.org/10.1016/j.jcp.2023.112603
Shirley, P., Wang, C., Zimmerman, K.: Monte Carlo techniques for direct lighting calculations. ACM Trans. Graph. 15(1), 1–36 (1996). https://doi.org/10.1145/226150.226151
Sawhney, R., Crane, K.: Monte Carlo geometry processing: a grid‐free approach to PDE‐based methods on volumetric domains. ACM Trans. Graph. 39(4), 123:1–123:18 (2020). https://doi.org/10.1145/3386569.3392374
Muller, M.E.: Some continuous Monte Carlo methods for the Dirichlet problem. Ann. Math. Stat. 27(3), 569–589 (1956). https://doi.org/10.1214/aoms/1177728169
Sabelfeld, K.K., Simonov, N.A.: Random walks on boundary for solving PDEs, 1st ed. De Gruyter, Berlin, Boston (1994)
Yılmazer, E.F., Vicini, D., Jakob, W.: Solving inverse PDE problems using grid‐free Monte Carlo estimators. arXiv:2208.02114 (2022)
Yang, X., Rasila, A., Sottinen, T.: Walk on spheres algorithm for Helmholtz and Yukawa equations via Duffin correspondence. Methodol. Comput. Appl. Probab. 19(2), 589–602 (2017). https://doi.org/10.1007/s11009‐016‐9504‐9
Sawhney, R., et al.: Grid‐free Monte Carlo for PDEs with spatially varying coefficients. ACM Trans. Graph. 41(4), 53:1–53:17 (2022). https://doi.org/10.1145/3528223.3530134
Rioux‐Lavoie, D., et al.: A Monte Carlo method for fluid simulation. ACM Trans. Graph. 41(6), 240:1–240:16 (2022). https://doi.org/10.1145/3550454.3555450
Mossberg, L.: GPU‐accelerated Monte Carlo geometry processing for gradient‐domain methods. Dissertation, Linköping University (2021)
Pharr, M., Jakob, W., Humphreys, G.: Physically based rendering: from theory to implementation. MIT Press, Cambridge (2023)
Miller, B., et al.: Boundary value caching for walk on spheres. ACM Trans. Graph. 42(4), 82:1–82:11 (2023). https://doi.org/10.1145/3592400
Sugimoto, R., et al.: A practical walk‐on‐boundary method for boundary value problems. ACM Trans. Graph. 42(4), 81:1–81:16 (2023). https://doi.org/10.1145/3592109
Qi, Y., et al.: A bidirectional formulation for walk on spheres. Comput. Graph. Forum 41(4), 51–62 (2022). https://doi.org/10.1111/cgf.14586
Makarov, N.R.: Monte Carlo methods for solving boundary value problems of the second and third kinds. Russ. J. Numer. Anal. Math. Model. 13(2), 117–132 (1998). https://doi.org/10.1515/rnam.1998.13.2.117
Sipin, A.S.: Random walk on balls for a boundary value problems with mixed boundary conditions. AIP Conf. Proc. 2849(1), 100001 (2023). https://doi.org/10.1063/5.0162520
Simonov, N.A.: Walk‐on‐spheres algorithm for solving third boundary value problem. Appl. Math. Lett. 64, 156–161 (2017). https://doi.org/10.1016/j.aml.2016.09.008
Sawhney, R., et al.: Walk on stars: a grid‐free Monte Carlo method for PDEs with Neumann boundary conditions. ACM Trans. Graph. 42(4), 80:1–80:20 (2023). https://doi.org/10.1145/3592398
Bogdan, K., Żak, T.: On Kelvin transformation. J. Theor. Probab. 19(1), 89–120 (2006). https://doi.org/10.1007/s10959‐006‐0003‐8
Nabizadeh, M.S., Ramamoorthi, R., Chern, A.: Kelvin transformations for simulations on infinite domains. ACM Trans. Graph. 40(4), 97:1–97:15 (2021). https://doi.org/10.1145/3476576.3476658
Veach, E., Guibas, L.J.: Optimally combining sampling techniques for Monte Carlo rendering. In: Proceedings of the 22nd Annual Conference on Computer Graphics and Interactive Techniques, pp. 419–428. Association for Computing Machinery (1995)
Alanko, S., Avellaneda, M.: Reducing variance in the numerical solution of BSDEs. Compt. Rendus Math. 351(3), 135–138 (2013). https://doi.org/10.1016/j.crma.2013.02.010
Titsias, M., Shi, J.: Double control variates for gradient estimation in discrete latent variable models. In: Proceedings of the 25th International Conference on Artificial Intelligence and Statistics, vol. 151, pp. 6134–6151. PMLR (2022)
Barill, G., et al.: Fast winding numbers for soups and clouds. ACM Trans. Graph. 37(4), 43:1–43:12 (2018). https://doi.org/10.1145/3197517.3201337
Lauterbach, C., et al.: Fast BVH construction on GPUs. In: Computer Graphics Forum, vol. 28(2), pp. 375–384. Wiley Online Library (2009). https://doi.org/10.1111/j.1467‐8659.2009.01377.x
Zhu, Q., et al.: Four‐plate capacitive power transfer system with different grounding connections. Int. J. Electr. Power Energy Syst. 115, 105494 (2020). https://doi.org/10.1016/j.ijepes.2019.105494
Dellera, L., Garbagnati, E.: Lightning stroke simulation by means of the leader progression model. I. Description of the model and evaluation of exposure of free‐standing structures. IEEE Trans. Power Deliv. 5(4), 2009–2022 (1990). https://doi.org/10.1109/61.103696
Ermakov, S.M., Sipin, A.S.: The “walk in hemispheres” process and its applications to solving boundary value problems. Vestn. St. Petersbg. Univ. Math. 42(3), 155–163 (2009). https://doi.org/10.3103/s1063454109030029
Simonov, N.A.: Random walks for solving boundary‐value problems with flux conditions. In: International Conference on Numerical Methods and Applications, pp. 181–188. Springer (2006)
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2025. This work is published under http://creativecommons.org/licenses/by-nc/4.0/ (the "License"). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Numerical simulation plays a crucial role in the analysis and design of power equipment, such as lightning protection devices, which may become inefficient using traditional grid‐based methods when handling complex geometries of large problems. The authors propose a grid‐free Monte Carlo method to handle electrostatic problems of complex geometry for both the interior and exterior domains, which is governed by the Poisson equation with a floating potential boundary condition that is neither a pure Dirichlet nor a Neumann condition. The potential and gradient at any given point can be expressed in terms of integral equations, which can be estimated recursively within the walk‐on‐sphere algorithm. Numerical examples have been demonstrated, including the evaluation of the mutual capacitance matrix of multi‐conductor structures and lighting striking near real fractal trees. The proposed method shows advantages in terms of geometric flexibility and robustness, output sensitivity, and parallelism, which may become a candidate for game‐changing numerical methods and exhibit great potential applications in high‐voltage engineering.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details









1 School of Electrical Engineering, Xi'an Jiaotong University, Xi'an, Shaanxi, China
2 Tibet Yangbajing High Altitude Electrical Safety and Electromagnetic Environment National Observation and Research Station, China Electric Power Research Institute, Beijing, China