Skip to main content
Log in

Highly localized RBF Lagrange functions for finite difference methods on spheres

  • Published:
BIT Numerical Mathematics Aims and scope Submit manuscript

Abstract

The aim of this paper is to show how rapidly decaying RBF Lagrange functions on the sphere can be used to create a numerically feasible, stable finite difference method based on radial basis functions (an RBF-FD-like method). For certain classes of PDEs this approach leads to rigorous convergence estimates for stencils which grow moderately with increasing discretization fineness.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Briefly, by scaling \(k\rightarrow k(\cdot /\sigma )\), we rescale the Fourier transform, \({\widehat{k}} \rightarrow \sigma ^d {\widehat{k}}(\sigma \cdot )\). The positive definiteness of \(k(\cdot /\sigma )\) is a consequence of the positivity of its Fourier transform, which is unchanged by rescaling.

References

  1. Chu, T., Schmidt, O.T.: RBF-FD discretization of the Navier-Stokes equations using staggered nodes. arXiv:2206.06495 (2022)

  2. Davydov, O.: Error bounds for a least squares meshless finite difference method on closed manifolds. Adv. Comput. Math. 49(4), 1–42 (2023)

    Article  MathSciNet  Google Scholar 

  3. de Boor, C., Ron, A.: Fourier analysis of the approximation power of principal shift-invariant spaces. Constr. Approx. 8(4), 427–462 (1992)

    Article  MathSciNet  Google Scholar 

  4. Floater, M.S., Iske, A.: Multistep scattered data interpolation using compactly supported radial basis functions. J. Comput. Appl. Math. 73(1–2), 65–78 (1996)

    Article  MathSciNet  Google Scholar 

  5. Flyer, N., Wright, G.B.: A radial basis function method for the shallow water equations on a sphere. In: Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, pp. rspa–2009. The Royal Society (2009)

  6. Fornberg, B., Flyer, N.: Solving PDEs with radial basis functions. Acta Numer. 24, 215–258 (2015)

    Article  MathSciNet  Google Scholar 

  7. Fornberg, B., Flyer, N.: A Primer on Radial Basis Functions with Applications to the Geosciences. Society for Industrial and Applied Mathematics, Philadelphia, PA (2015)

    Book  Google Scholar 

  8. Fuselier, E., Hangelbroek, T., Narcowich, F.J., Ward, J.D., Wright, G.B.: Localized bases for kernel spaces on the unit sphere. SIAM J. Numer. Anal. 51(5), 2538–2562 (2013)

    Article  MathSciNet  Google Scholar 

  9. Hangelbroek, T., Narcowich, F.J., Rieger, C., Ward, J.D.: An inverse theorem for compact Lipschitz regions in \({\mathbb{R} }^d\) using localized kernel bases. Math. Comp. 87(312), 1949–1989 (2018)

    Article  MathSciNet  Google Scholar 

  10. Hangelbroek, T., Narcowich, F.J., Ward, J.D.: Polyharmonic and related kernels on manifolds: interpolation and approximation. Found. Comput. Math. 12(5), 625–670 (2012)

    Article  MathSciNet  Google Scholar 

  11. Hangelbroek, T.: Polyharmonic approximation on the sphere. Constr. Approx. 33(1), 77–92 (2011)

    Article  MathSciNet  Google Scholar 

  12. Hangelbroek, T., Narcowich, F.J., Rieger, C., Ward, J.D.: Direct and inverse results on bounded domains for meshless methods via localized bases on manifolds. In: Contemporary Computational Mathematics—A Celebration of the 80th Birthday of Ian Sloan. Vol. 1, 2, pp. 517–543. Springer, Cham (2018)

  13. Hangelbroek, T., Ron, A.: Nonlinear approximation using Gaussian kernels. J. Funct. Anal. 259(1), 203–219 (2010)

    Article  MathSciNet  Google Scholar 

  14. Hubbert, S., Morton, T.M.: Lp-error estimates for radial basis function interpolation on the sphere. J. Approx. Theory 129(1), 58–77 (2004)

    Article  MathSciNet  Google Scholar 

  15. Keiner, J., Kunis, S., Potts, D.: Efficient reconstruction of functions on the sphere from scattered data. J. Fourier Anal. Appl. 13(4), 435–458 (2007)

    Article  MathSciNet  Google Scholar 

  16. Le Gia, Q.T., Sloan, I.H., Wendland, H.: Multiscale RBF collocation for solving PDEs on spheres. Numer. Math. 121(1), 99–125 (2012)

    Article  MathSciNet  Google Scholar 

  17. Le Gia, Q.T., Sloan, I.H., Wendland, H.: Zooming from global to local: a multiscale RBF approach. Adv. Comput. Math. 43(3), 581–606 (2017)

    Article  MathSciNet  Google Scholar 

  18. Mhaskar, H., Narcowich, F., Prestin, J., Ward, J.: \({L}^p\) Bernstein estimates and approximation by spherical basis functions. Math. Comput. 79(271), 1647–1679 (2010)

    Article  Google Scholar 

  19. Morton, T.M., Neamtu, M.: Error bounds for solving pseudodifferential equations on spheres by collocation with zonal kernels. J. Approx. Theory 114(2), 242–268 (2002)

    Article  MathSciNet  Google Scholar 

  20. Narcowich, F., Petrushev, P., Ward, J.: Decomposition of Besov and Triebel-Lizorkin spaces on the sphere. J. Funct. Anal. 238(2), 530–564 (2006)

    Article  MathSciNet  Google Scholar 

  21. Narcowich, F.J., Petrushev, P., Ward, J.D.: Localized tight frames on spheres. SIAM J. Math. Anal. 38(2), 574–594 (2006)

    Article  MathSciNet  Google Scholar 

  22. Narcowich, F.J., Rowe, S.T., Ward, J.D.: A novel Galerkin method for solving PDEs on the sphere using highly localized kernel bases. Math. Comp. 86(303), 197–231 (2017)

    Article  MathSciNet  Google Scholar 

  23. Narcowich, F.J., Schaback, R., Ward, J.D.: Multilevel interpolation and approximation. Appl. Comput. Harmon. Anal. 7(3), 243–261 (1999)

    Article  MathSciNet  Google Scholar 

  24. Narcowich, F., Ward, J., Wendland, H.: Sobolev bounds on functions with scattered zeros, with applications to radial basis function surface fitting. Math. Comput. 74(250), 743–763 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  25. Narcowich, F.J., Sun, X., Ward, J.D., Wendland, H.: Direct and inverse Sobolev error estimates for scattered data interpolation via spherical basis functions. Found. Comput. Math. 7(3), 369–390 (2007)

    Article  MathSciNet  Google Scholar 

  26. Narcowich, F.J., Ward, J.D.: Scattered data interpolation on spheres: error estimates and locally supported basis functions. SIAM J. Math. Anal. 33(6), 1393–1410 (2002)

    Article  MathSciNet  Google Scholar 

  27. Powell, M.J.D.: The theory of radial basis function approximation in 1990. In: Advances in Numerical Analysis, Vol. II (Lancaster, 1990), Oxford Sci. Publ., pages 105–210. Oxford Univ. Press, New York (1992)

  28. Schaback, R.: Improved error bounds for scattered data interpolation by radial basis functions. Math. Comp. 68(225), 201–216 (1999)

    Article  ADS  MathSciNet  Google Scholar 

  29. Shankar, V., Fogelson, A.L.: Hyperviscosity-based stabilization for radial basis function-finite difference (RBF-FD) discretizations of advection-diffusion equations. J. Comput. Phys. 372, 616–639 (2018)

    Article  ADS  MathSciNet  CAS  PubMed  PubMed Central  Google Scholar 

  30. Tominec, I., Larsson, E., Heryudono, A.: A least squares radial basis function finite difference method with improved stability properties. SIAM J. Sci. Comput. 43(2), A1441–A1471 (2021)

    Article  MathSciNet  Google Scholar 

  31. Tominec, I., Nazarov, M., Larsson, E.: Stability estimates for radial basis function methods applied to time-dependent hyperbolic PDEs. arXiv:2110.14548 (2021)

  32. Wendland, H.: Scattered Data Approximation, vol. 17. Cambridge University Press (2004)

Download references

Acknowledgements

Research of Dr. Hangelbroek was supported by grants DMS-1716927 and DMS-2010051 from the National Science Foundation. Drs. Narcowich and Ward were supported by grant DMS-1813091 from the National Science Foundation.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to T. Hangelbroek.

Ethics declarations

Conflict of interest

The authors have no conflicts of interest to declare.

Additional information

Communicated by Elisabeth Larsson.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

A Kernels on the sphere

In this section, we consider SBFs of the form \(\varPhi (x,y) = \phi (x\cdot y)\). Many SBFs are the restriction to \({\mathbb {S}}^2\times {\mathbb {S}}^2\) from radial, translation invariant kernels (RBFs) on \({\mathbb {R}}^3\): i.e., restrictions of kernels of the form \(\varPsi (x,y) = \psi (|x-y|)\). By writing \( |x-y| = \sqrt{2-2t} \), we may write the restricted RBF in terms of a single real variable \(t = x\cdot y\), so

$$\begin{aligned} \phi (t) = \psi (\sqrt{2-2t}). \end{aligned}$$
(A.1)

The restricted polyharmonic splines are a well-known class of conditionally positive SBFs given by

$$\begin{aligned} \phi (t)= \left\{ \begin{array}{ll }(1-t)^{s} \log (1- t), &{} s \in {\mathbb {N}}, \\ (1-t)^{s}, &{} s \in {\mathbb {N}} - \frac{1}{2}, \end{array} \right. \end{aligned}$$

for all \(t\in [-1,1]\). In terms of \(R(t):= \sqrt{1-t}\) (which is invertible on \([-1,1]\)), we can express \(\phi (t)\) as

$$\begin{aligned} \phi (t)= \left\{ \begin{array}{ll }R(t)^{2s} \log R(t), &{} s \in {\mathbb {N}}, \\ R(t)^{2s}, &{} s \in {\mathbb {N}} - \frac{1}{2}. \end{array} \right. \end{aligned}$$

In this appendix, we focus on strictly positive definite SBFs. In contrast to the restricted polyharmonic splines, which are discussed throughout the body of the article, such kernels require tuning of a shape parameter.

1.1 A.1 Strictly positive definite kernels and shape parameters

By rescaling the RBF by \(\sigma >0\) as \(\varPsi _{\sigma }(x,y) = \psi (\frac{|x-y|}{\sigma })\), we obtain a new positive definite radial basis function.Footnote 1 This means that the RBF has a parameter (roughly the reciprocal of the “shape parameter” \(\epsilon \) defined below) which must me be set. By changing this parameter, one alters the scaling of the RBF; this is desirable (indeed, necessary) for many practical problems: e.g., for a finite point set \(X_N\subset {\mathbb {R}}^d\), the collocation matrix

$$\begin{aligned} {{\varvec{\Psi }}}_{X_N} =\bigl (\varPsi _{\sigma }(x_j,x_k)\bigr )_{j,k} \end{aligned}$$

will have a large condition number if \(\sigma \) is large relative to the nearest neighbor distance \(q= \min _{j\ne k} |x_j-x_k|\). Thus, in order to treat (e.g., by interpolation, quadrature, etc.) densely sampled data, one may wish to choose a small value of \(\sigma \) to stabilize the problem. This is also the case when using the SBF obtained by restricting the RBF \(\varPsi _{\sigma }\).

Choosing small \(\sigma \) may come at a cost, however: the shape parameter affects the native space, both by modifying the inner product, and in some cases by altering the underlying set. The known convergence results for RBF approximation and interpolation are most often assume a fixed shape parameter. Indeed, for most RBFs (which have a positive, continuous Fourier transform and therefore lack a Strang-Fix condition [3]), choosing \(\sigma \propto h\), will lead to non-convergence [27]. To date there are few direct approximation results which deal with multi-scale RBF approximation [4, 13, 17, 23].

In what follows, we’ll use the notation \(R(t):= \sqrt{1-t}\) (which is invertible on \([-1,1]\)) and the modified “shape parameter” \(\epsilon := \frac{\sqrt{2}}{\sigma }\), which yields the relation, via (A.1), \(\phi (t) = \psi (\epsilon R(t))\).

Let us now give a few examples of zonal kernels, along with their native spaces.

B Differential equations on the sphere: assembling the FD matrix

In this section, we consider how to assemble the kernel FD matrix. Generally, this requires some understanding of the expression of \({\mathscr {L}}\) in spherical coordinates (a different method to construct \({\textbf{M}}_{X_N}\) in Cartesian coordinates has been given in [5]) and how to use it to obtain an expression for \({\mathscr {L}}^{(1)} \varPhi \).

The basic challenge is to assemble the Kansa type matrix \({\textbf{K}}_{X_N} = \bigl ({\mathscr {L}}^{(1)} \varPhi (x_j,x_k)\bigr )_{j,k}\) for a few first and second order linear differential operators \({\mathscr {L}}\).

1.1 B.1 Working in coordinates

In what follows, we use spherical coordinates \((\theta ,\varphi )\in [0,2\pi )\times [0,\pi ]\) to describe points on the sphere. This involves the convention \(x =\cos \theta \sin \varphi \), \(y= \sin \theta \sin \varphi \) and \(z=\cos \varphi \).

The surface gradient is \(\nabla f(\theta , \varphi ) = \frac{\partial f}{\partial \varphi } \varvec{\varphi } + \frac{1}{\sin ^2 \varphi }\frac{\partial f}{\partial \theta } \varvec{\theta }\). Here \(\varvec{\varphi }\) and \(\varvec{\theta }\) are the basic tangent vectors for the spherical coordinate system:

$$\begin{aligned} \varvec{\varphi } = \begin{pmatrix}\cos \theta \cos \varphi \\ \sin \theta \cos \varphi \\ - \sin \varphi \end{pmatrix} \quad \text { and } \quad \varvec{\theta } = \begin{pmatrix}-\sin \theta \sin \varphi \\ \cos \theta \sin \varphi \\ 0\end{pmatrix}. \end{aligned}$$

The spherical divergence operator applied to a vector field \(F = F_{\varphi } \varvec{\varphi }+F_{\theta } \varvec{\theta }\) gives

$$\begin{aligned} \textrm{div} F = \frac{1}{\sin \varphi }\bigl (\frac{d}{d\theta } (\sin \varphi F_{\theta }) +\frac{d}{d\varphi } (\sin \varphi F_{\varphi })\bigr ) = \frac{dF_{\theta }}{d\theta } + \cot \varphi F_{\varphi } + \frac{d F_{\varphi }}{d \varphi }. \end{aligned}$$
(B.1)

An example of divergence-free vector field (tangent to the sphere) is, for an angle \(\alpha \), is

$$\begin{aligned} \varvec{u} =u_{\varphi }\varvec{\varphi }+u_{\theta }\varvec{\theta } = - \sin \alpha \cos \theta \varvec{\varphi } + (\cos \alpha +\sin \alpha \cot \varphi \sin \theta )\varvec{\theta }. \end{aligned}$$

The fact that this is divergence-free is evident from (B.1). Writing \(x= \cos \theta \sin \varphi \), \(y= \sin \theta \sin \varphi \) and \(z=\cos \varphi \), we have

$$\begin{aligned} \varvec{u}(x,y,z) = \begin{pmatrix} -\sin \varphi \sin \theta \cos \alpha - \cos \varphi \sin \alpha \\ \sin \varphi \cos \theta \cos \alpha \\ \sin \varphi \cos \theta \sin \alpha \end{pmatrix} = \begin{pmatrix} -y \cos \alpha - z\sin \alpha \\ x\cos \alpha \\ x \sin \alpha \end{pmatrix}. \end{aligned}$$

An example of a transport term We consider a first order operator of the form \({\mathscr {L}}= \varvec{u}\cdot \nabla \), where \(\nabla \) is the “surface” gradient and \(\varvec{u} = u_\phi \varvec{\phi } + u_{\theta } \varvec{\theta }\) is a (tangent) vector field. Carrying out the (Cartesian) inner product simply produces

$$\begin{aligned} {\mathscr {L}}f = u_{\varphi }\frac{\partial f}{\partial \varphi }\langle \varvec{\varphi },\varvec{\varphi }\rangle + u_{\theta }\frac{1}{\sin ^2\varphi } \frac{\partial f}{\partial \theta } \langle \varvec{\theta },\varvec{\theta }\rangle = u_{\varphi }\frac{\partial f}{\partial \varphi } +u_{\theta } \frac{\partial f}{\partial \theta }. \end{aligned}$$

1.2 B.2 First order operators

We apply this to a zonal kernel \(\varPhi (x,y) = \phi (x\cdot y)\) in the first argument. So we can write the dot product as

$$\begin{aligned} x\cdot y= & {} \sin \varphi _1 \sin \varphi _2 \bigl (\cos \theta _1 \cos \theta _2 + \sin \theta _1 \sin \theta _2\bigr ) + \cos \varphi _1 \cos \varphi _2\\= & {} \sin \varphi _1 \sin \varphi _2 \bigl (\cos (\theta _1 - \theta _2)\bigr ) + \cos \varphi _1 \cos \varphi _2 \end{aligned}$$

For first order differential problems, we need \({\mathscr {L}}^{(1)}\phi (x\cdot y)\), which requires an expression for \(\phi '(t)\). Indeed, note that

$$\begin{aligned} \frac{\partial }{\partial \varphi }^{(1)} \phi (x\cdot y)= & {} \left[ \cos \varphi _1 \sin \varphi _2 \bigl ( \cos (\theta _1 - \theta _2)\bigr ) - \sin \varphi _1 \cos \varphi _2\right] \phi '(x\cdot y) = (\varvec{\varphi }\cdot y) \phi '(x\cdot y) \nonumber \\ \end{aligned}$$
(B.2)
$$\begin{aligned} \frac{\partial }{\partial \theta }^{(1)} \phi (x\cdot y)= & {} -\sin \varphi _1 \sin \varphi _2 \bigl (\sin (\theta _1 - \theta _2)\bigr ) \phi '(x\cdot y) = (\varvec{\theta }\cdot y) \phi '(x\cdot y) \end{aligned}$$
(B.3)

From this, it follows that the surface gradient is

$$\begin{aligned} \nabla ^{(1)} \phi (x\cdot y)= & {} (\varvec{\varphi } \cdot y) \phi '(x\cdot y) \varvec{\varphi } + \frac{1}{\sin ^2 \varphi }(\varvec{\theta } \cdot y) \phi '(x\cdot y) \varvec{\theta }. \end{aligned}$$
(B.4)

For the transport term we get in this way the formula

$$\begin{aligned} {\mathscr {L}}^{(1)} \phi (x \cdot y) = u_{\varphi } (\varvec{\varphi } \cdot y) \phi '(x\cdot y) + u_{\theta } (\varvec{\theta } \cdot y) \phi '(x\cdot y). \end{aligned}$$

1.3 B.3 s order operators

In principle, the kernel derivative formulas (B.2), (B.3) and (B.4) along with (B.1) are sufficient to calculate second order operators in divergence form \({\mathscr {L}}= \textrm{div} (\varvec{a} \nabla f)\) for a sufficiently smooth tensor field \(\varvec{a}\), although this may be too cumbersome to carry out by hand.

Laplace-Beltrami For \({\mathscr {L}}= \Delta \) (so \(\varvec{a} = \textrm{Id}\)), it is much easier to use the rotation invariance of \(\Delta \). In that case, we can write \(x\cdot y = \cos (\vartheta )\), with \(\vartheta \) the solid angle between x and y (equivalently, we can perform a rotation mapping y to the north pole). In this case, rotation invariance gives \(\Delta ^{(1)} \phi (x\cdot y) = \Delta \phi (\cos \vartheta ) \) and

$$\begin{aligned} \Delta \phi (\cos \vartheta ) =\frac{1}{\sin \vartheta } \frac{\partial }{\partial \vartheta } \bigl (\sin \vartheta \frac{\partial }{\partial \vartheta } \phi (\cos \vartheta )\bigr ) = -2\cos \vartheta \phi '(\cos \vartheta ) +(1-\cos ^2\vartheta )\phi ''(\cos \vartheta ). \end{aligned}$$

which simplifies to

$$\begin{aligned} \Delta ^{(1)} \phi (x\cdot y) = (1-(x\cdot y)^2)\phi ''(x\cdot y) -2(x\cdot y)\phi '(x\cdot y). \end{aligned}$$

A second example: In the basis \(\{ \varvec{\varphi }, \varvec{\theta } \}\), we consider the tensor \(\varvec{a}\) given by

$$\begin{aligned} \varvec{a}= a(\theta ,\varphi ) \begin{pmatrix} 1 &{} 0 \\ 0 &{} \sin ^2\varphi \end{pmatrix}, \end{aligned}$$

i.e. \(\varvec{a}\) lies parallel to the metric tensor of the unit sphere. In this case, we can simplify the second order differential operator \({\mathscr {L}}\). Because \(\nabla f = \frac{\partial f}{\partial \theta } \varvec{\varphi } + \frac{1}{\sin ^2\varphi } \frac{\partial f}{\partial \varphi } \varvec{\theta }\), we can write

$$\begin{aligned} {\mathscr {L}}f= & {} \textrm{div}(\varvec{a} \nabla f) \\= & {} \frac{1}{\sin \varphi } \frac{\partial }{\partial \varphi } \sin \varphi \, a(\theta ,\varphi ) \frac{\partial f}{\partial \varphi } + \frac{\partial }{\partial \theta } a(\theta ,\varphi ) \frac{\partial f}{\partial \theta }\\= & {} a(\theta ,\varphi ) \left( \Delta f - \cot ^2 \varphi \frac{\partial ^2 f}{\partial \theta ^2}\right) + \frac{\partial a}{\partial \varphi } \frac{\partial f}{\partial \varphi }+ \frac{\partial a}{\partial \theta }\frac{\partial f}{\partial \theta } \end{aligned}$$

For the kernel function \(\phi (x\cdot y)\) centered at a fixed y we therefore get, similarly as for the Laplace-Beltrami operator and the surface gradient the formula

$$\begin{aligned} {\mathscr {L}}^{(1)} \phi (x \cdot y)= & {} a(\theta ,\varphi ) \left( \Delta ^{(1)} \phi (x \cdot y) - \cot ^2 \varphi \Big ( \phi ''(x \cdot y) (\varvec{\theta } \cdot y)^2 + \phi '(x \cdot y) ( \frac{\partial \varvec{\theta }}{\partial \theta } \cdot y) \Big ) \right) \\{} & {} + \left( \frac{\partial a}{\partial \varphi } \varvec{\varphi } \cdot y + \frac{\partial a}{\partial \theta }\varvec{\theta } \cdot y \right) \phi '(x \cdot y). \end{aligned}$$
Table 1 Some SBFs and their native spaces

Considering this with \(a(\varphi ,\theta )= (1-\frac{1}{2} \cos \varphi )\) gives an elliptic PDE considered in [22]. This simplifies to

$$\begin{aligned} {\mathscr {L}}^{(1)} \phi (x \cdot y)= & {} (1-\frac{1}{2} \cos \varphi ) \left( \Delta ^{(1)} \phi (x \cdot y) - \cot ^2 \varphi \Big ( \phi ''(x \cdot y) (\varvec{\theta } \cdot y)^2 + \phi '(x \cdot y) ( \frac{\partial \varvec{\theta }}{\partial \theta } \cdot y) \Big ) \right) \\{} & {} + \frac{1}{2} \sin {\varphi } \, (\varvec{\varphi } \cdot y) \phi '(x \cdot y), \end{aligned}$$

where \((\varphi ,\theta )\) are the spherical coordinates corresponding to x (Tables 1, 2).

Table 2 Derivatives of SBFs appearing in Table 1

1.4 B.4 Derivatives of well known SBFs

Given an SBF of the form \(\phi (t) = \psi (\epsilon R(t))\) (as described in A), the first and second derivatives can be written as:

$$\begin{aligned} \phi ' = - \frac{\epsilon }{2R} \psi '(\epsilon R) =: \epsilon ^2 w(\epsilon R)\qquad \text {and} \qquad \phi '' = -\frac{\epsilon ^3}{2R}w'(\epsilon R). \end{aligned}$$
(B.5)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Erb, W., Hangelbroek, T., Narcowich, F.J. et al. Highly localized RBF Lagrange functions for finite difference methods on spheres. Bit Numer Math 64, 16 (2024). https://doi.org/10.1007/s10543-024-01016-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10543-024-01016-x

Keywords

Mathematics Subject Classification

Navigation