Skip to main content
Log in

A spectral algorithm to simulate nonstationary random fields on spheres and multifractal star-shaped random sets

  • Original Paper
  • Published:
Stochastic Environmental Research and Risk Assessment Aims and scope Submit manuscript

Abstract

An extension of the turning arcs algorithm is proposed for simulating a random field on the two-dimensional sphere with a second-order dependency structure associated with a locally varying Schoenberg sequence. In particular, the correlation range as well as the fractal index of the simulated random field, obtained as a weighted sum of Legendre waves with random degrees, may vary from place to place on the spherical surface. The proposed algorithm is illustrated with numerical examples, a by-product of which is a closed-form expression for two new correlation functions (exponential-Bessel and hypergeometric models) on the sphere, together with their respective Schoenberg sequences. The applicability of our findings is also described via the emulation of three-dimensional multifractal star-shaped random sets.

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.

Fig. 1

Similar content being viewed by others

References

  • Adler RJ (1981) The geometry of random fields. Wiley & Sons, New York

    Google Scholar 

  • Alegria A, Cuevas F, Diggle P, Porcu E (2018) A family of covariance functions for random fields on spheres. CSGB Research Reports, Department of Mathematics, Aarhus University, Aarhus

    Google Scholar 

  • Alegría A, Emery X, Lantuéjoul C (2020) The turning arcs: a computationally efficient algorithm to simulate isotropic vector-valued Gaussian random fields on the \(d\)-sphere. Stat Comput. https://doi.org/10.1007/s11222-020-09952-8 In press

    Article  Google Scholar 

  • Alegría A, Porcu E, Furrer R, Mateu J (2019) Covariance functions for multivariate Gaussian fields evolving temporally over planet earth. Stoch Environ Res Risk Assess 33(8–9):1593–1608

    Google Scholar 

  • Anh VV, Broadbridge P, Olenko A, Wang YG (2018) On approximation for fractional stochastic partial differential equations on the sphere. Stoch Environ Res Risk Assess 32(9):2585–2603

    Google Scholar 

  • Arroyo D, Emery X (2017) Spectral simulation of vector random fields with stationary gaussian increments in \(d\)-dimensional euclidean spaces. Stoch Environ Res Risk Assess 31(7):1583–1592

    Google Scholar 

  • Berg C, Porcu E (2017) From Schoenberg coefficients to Schoenberg functions. Constr Approx 45(2):217–241

    Google Scholar 

  • Bingham N (1978) Tauberian theorems for Jacobi series. Proc Lond Math Soc 3(2):285–309

    Google Scholar 

  • Brafman F (1951) Generating functions of Jacobi and related polynomials. Proc Am Math Soc 2(6):942–949

    Google Scholar 

  • Castruccio S, Stein ML (2013) Global space-time models for climate ensembles. Ann Appl Stat 7(3):1593–1611

    Google Scholar 

  • Cheng D, Xiao Y (2016) Excursion probability of Gaussian random fields on sphere. Bernoulli 22(2):1113–1130

    Google Scholar 

  • Chilès J-P, Delfiner P (2012) Geostatistics: modeling spatial uncertainty, 2nd edn. John Wiley & Sons, New Jersey

    Google Scholar 

  • de Fouquet C (1994) Reminders on the conditioning kriging. In: Armstrong M, Dowd P A (eds) Geostatistical simulations. Kluwer Academic, Dordrecht, pp 131–145

    Google Scholar 

  • Devroye L (1986) Non-uniform random variate generation. Springer, New York

    Google Scholar 

  • Emery X (2008) Statistical tests for validating geostatistical simulation algorithms. Compute Geosci 34(1):1610–1620

    Google Scholar 

  • Emery X, Arroyo D (2018) On a continuous spectral algorithm for simulating non-stationary Gaussian random fields. Stoch Environ Res Risk Assess 32(4):905–919

    Google Scholar 

  • Emery X, Arroyo D, Porcu E (2016) An improved spectral turning-bands algorithm for simulating stationary vector Gaussian random fields. Stoch Environ Res Risk Assess 30(7):1863–1873

    Google Scholar 

  • Emery X, Lantuéjoul C (2006) TBSIM: A computer program for conditional simulation of three-dimensional Gaussian random fields via the turning bands method. Comput Geosci 32(10):1615–1628

    Google Scholar 

  • Emery X, Lantuéjoul C (2008) A spectral approach to simulating intrinsic random fields with power and spline generalized covariances. Comput Geosci 12(1):121–132

    Google Scholar 

  • Emery X, Porcu E (2019) Simulating isotropic vector-valued Gaussian random fields on the sphere through finite harmonics approximations. Stoch Environ Res Risk Assess 33(8–9):1659–1667

    Google Scholar 

  • Fouedjio F, Desassis N, Rivoirard J (2016) A generalized convolution model and estimation for non-stationary random functions. Spat Stat 16:35–52

    Google Scholar 

  • Gneiting T (2013) Strictly and non-strictly positive definite functions on spheres. Bernoulli 19(4):1327–1349

    Google Scholar 

  • Gneiting T, Ševčíková H, Percival DB (2012) Estimators of fractal dimension: assessing the roughness of time series and spatial data. Stat Sci 27(2):247–277

    Google Scholar 

  • Guella J, Menegatto V (2018) Unitarily invariant strictly positive definite kernels on spheres. Positivity 22(1):91–103

    Google Scholar 

  • Guinness J, Fuentes M (2016) Isotropic covariance functions on spheres: some properties and modeling considerations. J Multivar Anal 143:143–152

    Google Scholar 

  • Hansen LV, Thorarinsdottir TL, Ovcharov E, Gneiting T, Richards D (2015) Gaussian random particles with flexible Hausdorff dimension. Adv Appl Probab 47(2):307–327

    Google Scholar 

  • Heaton M, Katzfuss M, Berrett C, Nychka D (2014) Constructing valid spatial processes on the sphere using kernel convolutions. Environmetrics 25(1):2–15

    Google Scholar 

  • Hobolth A (2003) The spherical deformation model. Biostatistics 4(4):583–595

    Google Scholar 

  • Jones RH (1963) Stochastic processes on a sphere. Ann Inst Math Stat 34:213–218

    Google Scholar 

  • Jun M, Stein M (2008) Nonstationary covariance models for global data. Ann Appl Stat 2(4):1271–1289

    Google Scholar 

  • Kent JT, Dryden IL, Anderson CR (2000) Using circulant symmetry to model featureless objects. Biometrika 87(3):527–544

    Google Scholar 

  • Kucinskas AB, Turcotte DL, Huang J, Ford PG (1992) Fractal analysis of Venus topography in Tinatin Planitia and Ovda Regio. J Geophys Res Planets 97(E8):13635–13641

    Google Scholar 

  • Lang A, Schwab C (2015) Isotropic Gaussian random fields on the sphere: regularity, fast simulation and stochastic partial differential equations. Ann Appl Probab 25(6):3047–3094

    Google Scholar 

  • Lantuéjoul C (1994) Non conditional simulation of stationary isotropic multigaussian random functions. In: Armstrong M, Dowd P ,A (eds) Geostatistical simulations. Kluwer Academic, Dordrecht, pp 147–177

    Google Scholar 

  • Lantuéjoul C (2002) Geostatistical simulation: models and algorithms. Springer, Berlin

    Google Scholar 

  • Lantuéjoul C, Freulon X, Renard D (2019) Spectral simulation of isotropic Gaussian random fields on a sphere. Math Geosci 51(8):999–1020

    Google Scholar 

  • Malyarenko A (2004) Abelian and Tauberian theorems for random fields on two-point homogeneous spaces. Theory Probab Math Stat 69:115–127

    Google Scholar 

  • Mantoglou A, Wilson JL (1982) The turning bands method for simulation of random fields using line generation by a spectral method. Water Resour Res 18(5):1379–1394

    Google Scholar 

  • Marinucci D, Peccati G (2011) Random fields on the sphere: representation, limit theorems and cosmological applications. Cambridge University Press, Cambridge

    Google Scholar 

  • Matheron G (1973) The intrinsic random functions and their applications. Adv Appl Probab 5(3):439–468

    Google Scholar 

  • Moller J, Nielsen M, Porcu E, Rubak E (2018) Determinantal point process models on the sphere. Bernoulli 24(2):1171–1201

    Google Scholar 

  • Nott DJ, Dunsmuir WT (2002) Estimation of nonstationary spatial covariance structure. Biometrika 89(4):819–829

    Google Scholar 

  • Olver FW, Lozier DM, Boisvert RF, Clark CW (2010) NIST handbook of mathematical functions. Cambridge University Press, Cambridge

    Google Scholar 

  • Peron A, Porcu E, Emery X (2018) Admissible nested covariance models over spheres cross time. Stoch Environ Res Risk Assess 32(11):3053–3066

    Google Scholar 

  • Porcu E, Alegria A, Furrer R (2018) Modeling temporally evolving and spatially globally dependent data. Int Stat Rev 86(2):344–377

    Google Scholar 

  • Porcu E, Castruccio S, Alegria A, Crippa P (2019) Axially symmetric models for global data: a journey between geostatistics and stochastic generators. Environmetrics 30(1):e2555

    Google Scholar 

  • Sánchez L, Emery X, Séguret S (2019) 5D geostatistics for directional variables: application in geotechnics to the simulation of the linear discontinuity frequency. Comput Geosci 133:104325

    Google Scholar 

  • Schoenberg IJ (1942) Positive definite functions on spheres. Duke Math J 9(1):96–108

    Google Scholar 

  • Schreiner M (1997) Locally supported kernels for spherical spline interpolation. J Approx Theory 89(2):172–194

    Google Scholar 

  • Sedivy R, Mader RM (1997) Fractals, chaos, and cancer: do they coincide? Cancer Investig 15(6):601–607

    CAS  Google Scholar 

  • Stoyan D, Stoyan H (1994) Fractals, random shapes, and point fields. John Wiley & Sons, Chichester

    Google Scholar 

  • Zhou B, Wang J, Wang H (2017) Three-dimensional sphericity, roundness and fractal dimension of sand particles. Géotechnique 68(1):18–30

    Google Scholar 

  • Ziegel J (2013) Stereological modelling of random particles. Commun Stat Theory Methods 42(7):1428–1442

    Google Scholar 

  • Ziegel J (2014) Convolution roots and differentiability of isotropic positive definite functions on spheres. Proc Am Math Soc 142(6):2063–2077

    Google Scholar 

Download references

Acknowledgements

The authors acknowledge the funding of the National Agency for Research and Development of Chile, through grants CONICYT/FONDECYT/REGULAR/No. 1170290 (X. Emery), CONICYT PIA AFB180004 (X. Emery) and CONICYT/FONDECYT/INICIACIÓN/No. 11190686 (A. Alegría).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xavier Emery.

Additional information

Publisher's Note

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

Appendices

Appendices

A. Proof of Proposition 1

The positive semidefiniteness of (7) is a direct consequence of the addition theorem for spherical harmonic functions, which is described below. The set of spherical harmonic functions, \(\{{Y}_{nm}: n\in {\mathbb {N}}, m = -n, \ldots , n\}\), is an orthogonal basis of the Hilbert space of complex-valued square integrable functions on \({\mathbb {S}}^2\). Explicit expressions for these functions can be found in Olver et al. (2010, formula 14.30.1) and Marinucci and Peccati (2011). The addition theorem for spherical harmonic functions (Olver et al. 2010, formula 14.30.9) establishes that

$$\begin{aligned} {P}_n(\varvec{x}_1^\top \varvec{x}_2)& = \frac{4\pi }{2n+1} \sum _{m=0}^n\bigg [ \hbox {Re} \left\{ {Y}_{nm}(\varvec{x}_1) \right\} \hbox {Re} \left\{ {Y}_{nm}(\varvec{x}_2)\right\} \\&+\hbox {Im} \left\{ {Y}_{nm}(\varvec{x}_1)\right\} \hbox {Im} \left\{ {Y}_{nm}(\varvec{x}_2)\right\} \bigg ],\\&\quad \varvec{x}_1, \varvec{x}_2\in {\mathbb {S}}^2, \end{aligned}$$

where \(\hbox {Re}\) and \(\hbox {Im}\) represent real and imaginary parts, respectively. A straightforward calculation shows that, for any \(k\in {\mathbb {N}}^{*}\), and for any system of points \(\varvec{x}_1,\ldots , \varvec{x}_k \in {\mathbb {S}}^2\) and constants \(a_1,\ldots ,a_k \in {\mathbb {R}}\),

$$\begin{aligned} &\sum _{i,j=1}^k a_i a_j C(\varvec{x}_i,\varvec{x}_j)\\& = \sum _{i,j=1}^k a_i a_j \sum _{n=0}^\infty \beta _n(\varvec{x}_i,\varvec{x}_j) {P}_n(\varvec{x}_i^\top \varvec{x}_j)\\& = \sum _{n=0}^\infty \frac{4\pi }{2n+1} \sum _{m=0}^n \left[ \sum _{i,j=1}^k \left\{ {c}_{nm,i} c_{nm,j} + {d}_{nm,i} d_{nm,j} \right\} \beta _n(\varvec{x}_i,\varvec{x}_j)\right] , \end{aligned}$$

where \(c_{nm,i} = a_i \hbox {Re}\left\{ {Y}_{nm}(\varvec{x}_i)\right\}\) and \(d_{nm,i} = a_i \hbox {Im}\left\{ {Y}_{nm}(\varvec{x}_i)\right\}\). The last expression is clearly nonnegative due to the positive semidefiniteness of the functions \(\beta _n\), and the exchange order of summations is well justified by dominated convergence.

B. Proof of Proposition 2

The basic random field defined in (24) clearly has a zero expectation. On the other hand, in order to obtain its covariance function, we use the same arguments as in Alegría et al. (2020). Indeed, note that

$$\begin{aligned} {\mathbb {E}}\{{Z}(\varvec{x}_1) {Z}(\varvec{x}_2)\}& = {\mathbb {E}}(\varepsilon ^2) \sum _{n=0}^\infty \{b_n(\varvec{x}_1) b_n(\varvec{x}_2)\}^{1/2} (2n+1)\\&\times \int _{{\mathbb {S}}^2} {P}_n(\varvec{\omega }^\top \varvec{x}_1) {P}_n(\varvec{\omega }^\top \varvec{x}_2) U(\hbox {d}{\varvec{\omega }}), \\&\quad \varvec{x}_1,\varvec{x}_2\in {\mathbb {S}}^2, \end{aligned}$$

where U is the uniform probability measure on \({\mathbb {S}}^2\). The result follows from the duplication equation for Legendre polynomials (see, e.g., (Ziegel 2014, equation 2.4)): for any \(n,k\in {\mathbb {N}}\),

$$\begin{aligned} \int _{{\mathbb {S}}^2} {P}_n(\varvec{\omega }^\top \varvec{x}_1) {P}_k(\varvec{\omega }^\top \varvec{x}_2) U(\mathrm{d}\varvec{\omega })& = \frac{\delta _{n,k}}{2n+1} {P}_n(\varvec{x}_1^\top \varvec{x}_2), \\&\quad \varvec{x}_1,\varvec{x}_2\in {\mathbb {S}}^2, \end{aligned}$$

where \(\delta _{n,k}\) denotes the Kronecker delta.

C. Assessment of the central limit approximation

Starting with the well-known Berry–Esséen inequality, Alegría et al. (2020) showed that the Kolmogorov–Smirnov distance between the marginal distribution of \(\widetilde{Z}(\varvec{x})\) as defined in (25) and a Gaussian distribution is upper bounded as follows:

$$\begin{aligned}&\underset{z\in {\mathbb {R}}}{\sup } \Bigg |{\mathbb {P}} \left( \frac{\widetilde{Z}(\varvec{x})}{C(\varvec{x},\varvec{x})^{1/2}} < z \right) - G(z) \Bigg |\le \frac{\xi \, {\mathbb {E}}( |\varepsilon |^3)}{C(\varvec{x},\varvec{x})^{3/2} \, L^{1/2}} \\&\times \sum _{n}\frac{b_{n}(\varvec{x})^{3/2} \, (2n + 1)^{3/2} \, {\mathbb {E}} \left( |P_n(\varvec{\omega }^T \varvec{x}) |^3 \right) }{\zeta _n^{1/2}}, \end{aligned}$$

where the sum is extended over all the integers n such that \(\zeta _n\) is positive, G is the standard Gaussian cumulative distribution function, \(\xi\) is a constant between 0.4097 and 0.4748, and \({\mathbb {E}} \left( |P_n(\varvec{\omega }^T \varvec{x}) |^3 \right)\) does not depend on \(\varvec{x}\) and behaves as \({\mathcal {O}}(n^{-3/2})\) at large n. Now, in the four examples presented in Sect. 4.2, the simulated random field \(\widetilde{Z}(\varvec{x})\) has a unit variance, \(\varepsilon\) has a Rademacher distribution and \(\{\zeta _{n}: n \in {\mathbb {N}}\}\) is the probability mass sequence of a shifted zeta distribution with parameter 2, so that \(C(\varvec{x},\varvec{x})=1\), \({\mathbb {E}}( |\varepsilon |^3)=1\) and \(\zeta _n^{-1/2} = \pi \, (n+1) / \sqrt{6}\). Accordingly:

$$\begin{aligned} \underset{z\in {\mathbb {R}}}{\sup } \Bigg |{\mathbb {P}} \left( \widetilde{Z}(\varvec{x}) < z \right) - G(z) \Bigg |\le \frac{\xi }{L^{1/2}} \sum _{n} b_{n}(\varvec{x})^{3/2} \, \tau _n, \end{aligned}$$

with \(\tau _n = (2n + 1)^{3/2} \, {\mathbb {E}} \left( |P_n(\varvec{\omega }^T \varvec{x}) |^3 \right) \, \pi \, (n+1) / \sqrt{6} = {\mathcal {O}}(n)\) as \(n \rightarrow +\infty\). Furthermore, one has:

  • \(b_{n}(\varvec{x}) = {\mathcal {O}}(n^{-2\nu (\varvec{x})-1})\) with \(\nu (\varvec{x}) \in [0.2,1.8]\) (Legendre-Matérn model);

  • \(b_{n}(\varvec{x}) = {\mathcal {O}}(a(\varvec{x})^n)\) with \(a(\varvec{x}) \in [0.1,0.9]\) (multiquadric model);

  • \(b_{n}(\varvec{x}) = {\mathcal {O}}\left(\frac{a(\varvec{x})^n}{n!}\right)\) with \(a(\varvec{x}) \in [0.1,8.0]\) (exponential-Bessel model);

  • \(b_{n}(\varvec{x}) = {\mathcal {O}}(a(\varvec{x})^n \, n^{\nu (\varvec{x})-1})\), with \(a(\varvec{x}) \in [0.1,0.9]\) and \(\nu (\varvec{x}) \in [1,19]\) (hypergeometric model).

As a result, in all the four cases, the sequence \(\{b_{n}(\varvec{x})^{3/2} \, \tau _n: n \in {\mathbb {N}}\}\) is summable, hence, the Berry–Esséen upper bound is finite and proportional to \(L^{-1/2}\). By increasing L, it is possible to ensure that the distance between the marginal distribution of the simulated random field and a standard Gaussian distribution is less that any given positive threshold.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Emery, X., Alegría, A. A spectral algorithm to simulate nonstationary random fields on spheres and multifractal star-shaped random sets. Stoch Environ Res Risk Assess 34, 2301–2311 (2020). https://doi.org/10.1007/s00477-020-01855-4

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00477-020-01855-4

Keywords

Navigation