Skip to main content
Log in

Covariance functions for multivariate Gaussian fields evolving temporally over planet earth

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

Abstract

The construction of valid and flexible cross-covariance functions is a fundamental task for modeling multivariate space–time data arising from, e.g., climatological and oceanographical phenomena. Indeed, a suitable specification of the covariance structure allows to capture both the space–time dependencies between the observations and the development of accurate predictions. For data observed over large portions of planet earth it is necessary to take into account the curvature of the planet. Hence the need for random field models defined over spheres across time. In particular, the associated covariance function should depend on the geodesic distance, which is the most natural metric over the spherical surface. In this work, we propose a flexible parametric family of matrix-valued covariance functions, with both marginal and cross structure being of the Gneiting type. We also introduce a different multivariate Gneiting model based on the adaptation of the latent dimension approach to the spherical context. Finally, we assess the performance of our models through the study of a bivariate space–time data set of surface air temperatures and precipitable water content.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  • Abramowitz M, Stegun IA (eds) (1970) Handbook of mathematical functions. Dover, New York

    Google Scholar 

  • Alonso-Malaver C, Porcu E, Giraldo R (2015) Multivariate and multiradial Schoenberg measures with their dimension walks. J Multivar Anal 133:251–265

    Article  Google Scholar 

  • Anderes EB, Stein ML (2011) Local likelihood estimation for nonstationary random fields. J Multivar Anal 102(3):506–520

    Article  Google Scholar 

  • Apanasovich TV, Genton MG (2010) Cross-covariance functions for multivariate random fields based on latent dimensions. Biometrika 97:15–30

    Article  Google Scholar 

  • Apanasovich TV, Genton MG, Sun Y (2012) A valid Matérn class of cross-covariance functions for multivariate random fields with any number of components. J Am Stat Assoc 107(497):180–193

    Article  Google Scholar 

  • Banerjee S (2005) On geodetic distance computations in spatial modeling. Biometrics 61(2):617–625

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Bochner S (1955) Harmonic analysis and the theory of probability. California monographs in mathematical sciences. University of California Press, Berkeley

    Google Scholar 

  • Cramer H (1940) On the theory of stationary random processes. Ann Math 41(1):215–230

    Article  Google Scholar 

  • Cressie N (1993) Statistics for spatial data, revised edition. Wiley, New York

    Google Scholar 

  • Cressie N, Huang H (1999) Classes of nonseparable, spatiotemporal stationary covariance functions. J Am Stat Assoc 94:1330–1340

    Article  Google Scholar 

  • Daley D, Porcu E (2014) Dimension walks and Schoenberg spectral measures. Proc Am Math Soc 142(5):1813–1824

    Article  Google Scholar 

  • Daley D, Porcu E, Bevilacqua M (2015) Classes of compactly supported covariance functions for multivariate random fields. Stoch Environ Res Risk Assess 29(4):1249–1263

    Article  Google Scholar 

  • De Iaco S, Posa D (2013) Positive and negative non-separability for space–time covariance models. J Stat Plan Inference 143(2):378–391

    Article  Google Scholar 

  • De Iaco S, Posa D, Myers D (2013) Characteristics of some classes of space–time covariance functions. J Stat Plan Inference 143(11):2002–2015

    Article  Google Scholar 

  • Feller W (1966) An introduction to probability theory and its applications. Number v. 2 in wiley mathematical statistics series. Wiley, New York

    Google Scholar 

  • Genton MG, Kleiber W (2015) Cross-covariance functions for multivariate geostatistics. Stat Sci 30(2):147–163

    Article  Google Scholar 

  • Gneiting T (2002) Stationary covariance functions for space–time data. J Am Stat Assoc 97(458):590–600

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Gneiting T, Genton MG, Guttorp P (2007) Geostatistical space–time models, stationarity, separability and full symmetry. In: Finkenstadt B, Held L, Isham V (eds) Statistical methods for spatio-temporal systems. Chapman & Hall/CRC, Boca Raton, pp 151–175

    Google Scholar 

  • Gneiting T, Kleiber W, Schlather M (2010) Matérn cross-covariance functions for multivariate random fields. J Am Stat Assoc 105:1167–1177

    Article  CAS  Google Scholar 

  • Hannan E (2009) Multiple time series. Wiley series in probability and statistics. Wiley, New York

    Google Scholar 

  • Hansen LV, Thorarinsdottir TL, Gneiting T, Richards D (2011) Lévy particles: modelling and simulating star-shaped random sets. CSGB research reports, Department of Mathematics, Aarhus University

  • Heaton MJ, Datta A, Finley AO, Furrer R, Guinness J, Guhaniyogi R, Gerber F, Gramacy RB, Hammerling D, Katzfuss M, Lindgren F, Nychka DW, Sun F, Zammit-Mangion A (2018) A case study competition among methods for analyzing large spatial data. J Agric Biol Environ Stat. https://doi.org/10.1007/s13253-018-00348-w

    Article  Google Scholar 

  • Huang C, Zhang H, Robeson S (2011) On the validity of commonly used covariance and variogram functions on the sphere. Math Geosci 43:721–733

    Article  Google Scholar 

  • Kalnay E, Kanamitsu M, Kistler R, Collins W, Deaven D, Gandin L, Iredell M, Saha S, White G, Woollen J et al (1996) The NCEP/NCAR 40-year reanalysis project. Bull Am Meteorol Soc 77(3):437–472

    Article  Google Scholar 

  • Krupskii P, Genton MG (2019) A copula model for non-gaussian multivariate spatial data. J Multivar Anal 169:264–277

    Article  Google Scholar 

  • Ma C (2017) Time varying isotropic vector random fields on spheres. J Theor Probab 30(4):1763–1785

    Article  Google Scholar 

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

    Book  Google Scholar 

  • Porcu E, Schilling RL (2011) From Schoenberg to Pick–Nevanlinna: towards a complete picture of the variogram class. Bernoulli 17(1):441–455

    Article  Google Scholar 

  • Porcu E, Zastavnyi V (2011) Characterization theorems for some classes of covariance functions associated to vector valued random fields. J Multivar Anal 102(9):1293–1301

    Article  Google Scholar 

  • Porcu E, Gregori P, Mateu J (2006) Nonseparable stationary anisotropic space–time covariance functions. Stoch Environ Res Risk Assess 21(2):113–122

    Article  Google Scholar 

  • Porcu E, Bevilacqua M, Genton MG (2016) Spatio-temporal covariance and cross-covariance functions of the great circle distance on a sphere. J Am Stat Assoc 111(514):888–898

    Article  CAS  Google Scholar 

  • Rodrigues A, Diggle PJ (2010) A class of convolution-based models for spatio-temporal processes with non-separable covariance structure. Scand J Stat 37(4):553–567

    Article  Google Scholar 

  • Sampson PD, Guttorp P (1992) Nonparametric estimation of nonstationary spatial covariance structure. J Am Stat Assoc 87(417):108–119

    Article  Google Scholar 

  • Schoenberg IJ (1938) Metric spaces and completely monotone functions. Ann Math 39(4):811–841

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Wackernagel H (2003) Multivariate geostatistics: an introduction with applications, 3rd edn. Springer, New York

    Book  Google Scholar 

  • Xu G, Genton MG (2017) Tukey g-and-h random fields. J Am Stat Assoc 112(519):1236–1249

    Article  CAS  Google Scholar 

  • Yaglom AM (1987) Correlation theory of stationary and related random functions. Volume I: basic results. Springer, New York

    Book  Google Scholar 

  • Zhang H, Cai W (2015) When doesn’t cokriging outperform kriging? Stat Sci 30(2):176–180

    Article  Google Scholar 

  • Zhang H, Wang Y (2010) Kriging and cross-validation for massive spatial data. Environmetrics 21(3–4):290–304

    CAS  Google Scholar 

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

    Article  Google Scholar 

Download references

Acknowledgements

Alfredo Alegría’s work was initiated when he was a Ph.D. student at Universidad Técnica Federico Santa María. Alfredo Alegría was supported by Beca CONICYT-PCHA/Doctorado Nacional/2016-21160371. Emilio Porcu is partially supported by Proyecto Fondecyt Regular Number 1130647. Reinhard Furrer acknowledges support of the Swiss National Science Foundation SNSF-144973 and SNSF-175529. Jorge Mateu is partially supported by Grant MTM2016-78917-R.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alfredo Alegría.

Additional information

Publisher's Note

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

Appendices

Appendix 1: Background on positive definite functions

We start by defining (semi) positive definite functions, that arise in statistics as the covariances of Gaussian RFs as well as the characteristic functions of probability distributions. Let \({\mathcal {E}}\) be a non-empty set and \(m\in {\mathbb {N}}\). We say that the matrix-valued function \({\mathbf{F}}:{\mathcal {E}}\times {\mathcal {E}} \rightarrow {\mathbb {R}}^{m\times m}\) is (semi) positive definite if for all integer \(n\ge 1\), \(\{e_1,\ldots ,e_n\} \subset {\mathcal {E}}\) and \(\{\varvec{a}_1,\ldots ,\varvec{a}_n\}\subset {\mathbb {R}}^m\), the following inequality holds:

$$\begin{aligned} \sum _{\ell =1}^{n} \sum _{r=1}^n \varvec{a}_{\ell }^\top {\mathbf{F}}(e_\ell ,e_r) \varvec{a}_{r} \ge 0. \end{aligned}$$
(6.1)

We denote as \({\mathcal {P}}^m({\mathcal {E}})\) the class of such mappings \({\mathbf{F}}\) satisfying Eq. (6.1).

Next, we focus on the cases where \({\mathcal {E}}\) is either \({\mathbb {R}}^d\), \({\mathbb {S}}^d\) or \({\mathbb {S}}^d \times {\mathbb {R}}^k\), for \(d,k\in {\mathbb {N}}\). For a clear presentation of the results, Table 7 summarizes the notation introduced along this “Appendices 1, 2 and 3”.

1.1 Matrix-valued positive definite functions on Euclidean spaces: the classes \(\Phi ^{m}_{d,{\mathcal {S}}}\) and \(\Phi ^m_{d,{\mathcal {I}}}\)

This section summarizes matrix-valued positive definite functions on the Euclidean space \({\mathcal {E}} = {\mathbb {R}}^d\). Specifically, we expose some characterizations for the stationary and Euclidean isotropic members of the class \({\mathcal {P}}^m({\mathbb {R}}^d)\).

Table 7 Summary of the notation used along the “Appendices 1, 2 and 3

We say that \({\mathbf{F}} \in {\mathcal {P}}^m({\mathbb {R}}^d)\) is stationary if there exists a mapping \(\tilde{\varvec{\varphi }}:{\mathbb {R}}^d \rightarrow {\mathbb {R}}^{m\times m}\) such that

$$\begin{aligned} {\mathbf{F}}(\varvec{x},\varvec{y}) = \tilde{\varvec{\varphi }}(\varvec{x}-\varvec{y}) = [\tilde{\varphi }_{ij}(\varvec{x}-\varvec{y})]_{i,j=1}^m, \quad \varvec{x},\varvec{y}\in {\mathbb {R}}^d. \end{aligned}$$
(6.2)

We call \(\Phi _{d,{\mathcal {S}}}^m\) the class of continuous mappings \(\tilde{\varvec{\varphi }}\) such that \({\mathbf{F}}\) in (6.2) is positive definite. Cramér’s Theorem (Cramer 1940) establishes that \(\tilde{\varvec{\varphi }}\in \Phi _{d,{\mathcal {S}}}^m\) if and only if it can be represented through

$$\begin{aligned} \tilde{\varvec{\varphi }}(\varvec{h}) = \int _{{\mathbb {R}}^d} \exp \{ \imath \varvec{h}^\top \varvec{\omega } \} \text {d}\tilde{\varvec{\Lambda }}_d(\varvec{\omega }), \quad \varvec{h}\in {\mathbb {R}}^d, \end{aligned}$$
(6.3)

where \(\imath = \sqrt{-1} \in {\mathbb {C}}\) and \(\tilde{\varvec{\Lambda }}_d: {\mathbb {R}}^d\rightarrow {\mathbb {C}}^{m\times m}\) is a matrix-valued mapping, with increments being Hermitian and positive definite matrices, and whose elements, \(\tilde{\Lambda }_{d,ij}(\cdot )\), for \(i,j=1,\ldots ,m\), are functions of bounded variation (see Wackernagel 2003). In particular, the diagonal terms, \(\tilde{\Lambda }_{d,ii}(\varvec{\omega })\), are real, non-decreasing and bounded, whereas the off-diagonal elements are generally complex-valued. Cramer’s Theorem is the multivariate version of the celebrated Bochner’s Theorem (Bochner 1955). If the elements of \(\tilde{\varvec{\Lambda }}_d(\cdot )\) are absolutely continuous, then Eq. (6.3) simplifies to

$$\begin{aligned} \tilde{\varvec{\varphi }}(\varvec{h}) = \int _{{\mathbb {R}}^d} \exp \{ \imath \varvec{h}^\top \varvec{\omega } \} \tilde{\varvec{\lambda }}_d(\varvec{\omega })\text {d}\varvec{\omega }, \quad \varvec{h}\in {\mathbb {R}}^d, \end{aligned}$$

with \(\tilde{\varvec{\lambda }}_d(\varvec{\omega })=[\tilde{\lambda }_{d,ij}(\varvec{\omega })]_{i,j=1}^m\) being Hermitian and positive definite, for any \(\varvec{\omega }\in {\mathbb {R}}^d\). The mapping \(\tilde{\varvec{\lambda }}_d(\varvec{\omega })\) is known as the matrix-valued spectral density and classical Fourier inversion yields

$$\begin{aligned} \tilde{\varvec{\lambda }}_d(\varvec{\omega }) = \frac{1}{(2\pi )^d} \int _{{\mathbb {R}}^d} \exp \{ - \imath \varvec{h}^\top \varvec{\omega } \} \tilde{\varvec{\varphi }}(\varvec{h}) \text {d}\varvec{h}, \quad \varvec{\omega }\in {\mathbb {R}}^d. \end{aligned}$$

Finally, the following inequality between the elements of \(\tilde{\varvec{\varphi }}\) is true

$$\begin{aligned} |\tilde{\varphi }_{ij}(\varvec{h})|^2 \le \tilde{\varphi }_{ii}(\varvec{0}) \tilde{\varphi }_{jj}(\varvec{0}),\quad {\text { for all }} \varvec{h}\in {\mathbb {R}}^d. \end{aligned}$$

However, the maximum value of the mapping \(\tilde{\varphi }_{ij}(\varvec{h})\), with \(i\ne j\), is not necessarily reached at \(\varvec{h}=\varvec{0}\). In general, \(\tilde{\varphi }_{ij}\) is not itself a scalar-valued positive definite function when \(i\ne j\).

Consider an element \({\mathbf{F}}\) in \({\mathcal {P}}^m({\mathbb {R}}^d)\) and suppose that there exists a continuous and bounded mapping \(\varvec{\varphi }:{\mathbb {R}}_+ \rightarrow {\mathbb {R}}^{m\times m}\) such that

$$\begin{aligned} {\mathbf{F}}(\varvec{x},\varvec{y}) = \varvec{\varphi }(\Vert \varvec{x}-\varvec{y}\Vert ), \quad \varvec{x},\varvec{y}\in {\mathbb {R}}^d. \end{aligned}$$

Then, \({\mathbf{F}}\) is called stationary and Euclidean isotropic (or radial). We denote as \(\Phi ^m_{d,{\mathcal {I}}}\) the class of bounded, continuous, stationary and Euclidean isotropic mappings \(\varvec{\varphi }(\cdot )=[\varphi _{ij}(\cdot )]_{i,j=1}^m\).

When \(m=1\), characterization of the class \(\Phi _{d,{\mathcal {I}}}\) was provided through the celebrated paper by Schoenberg (1938). Alonso-Malaver et al. (2015) characterize the class \(\Phi ^m_{d,{\mathcal {I}}}\) through the continuous members \(\varvec{\varphi }\) having representation

$$\begin{aligned} \varvec{\varphi }(r) = \int _{[0,\infty )} \Omega _d(r\omega ) \text {d}\varvec{\Lambda }_d(\omega ), \quad r\ge 0, \end{aligned}$$

where \(\varvec{\Lambda }_d: [0,\infty )\rightarrow {\mathbb {R}}^{m\times m}\) is a matrix-valued mapping, with increments being positive definite matrices, and elements \(\Lambda _{d,ij}(\cdot )\) of bounded variation, for each \(i,j=1,\ldots ,m\). Here, the function \(\Omega _d(\cdot )\) is defined as

$$\begin{aligned} \Omega _d(z) = \Gamma (d/2)(z/2)^{-(d-2)/2} J_{(d-2)/2}(z), \quad z \ge 0, \end{aligned}$$

with \(\Gamma\) being the Gamma function and \(J_\nu\) the Bessel function of the first kind of degree \(\nu\) (see Abramowitz and Stegun 1970). If the elements of \(\varvec{\Lambda }_d(\cdot )\) are absolutely continuous, then we have an associated spectral density \(\varvec{\lambda }_d:[0,\infty )\rightarrow {\mathbb {R}}^{m\times m}\) as in the stationary case, which is called, following Daley and Porcu (2014), a d-Schoenberg matrix.

The classes \(\Phi ^m_{d,{\mathcal {I}}}\) are non-increasing in d, and the following inclusion relations are strict

$$\begin{aligned} \Phi ^m_{\infty ,{\mathcal {I}}} := \bigcap _{d=1}^\infty \Phi ^m_{d,{\mathcal {I}}} \subset \cdots \subset \Phi ^m_{2,{\mathcal {I}}} \subset \Phi ^m_{1,{\mathcal {I}}}. \end{aligned}$$

The elements \(\varvec{\varphi }\) in the class \(\Phi ^m_{\infty ,{\mathcal {I}}}\) can be represented as

$$\begin{aligned} \varvec{\varphi }(r) = \int _{[0,\infty )} \exp (-r^2\omega ^2) \text {d}\varvec{\Lambda }(\omega ), \quad r\ge 0, \end{aligned}$$

where \(\varvec{\Lambda }\) is a matrix-valued mapping with similar properties as \(\varvec{\Lambda }_d\).

1.2 Matrix-valued positive definite functions on \({\mathbb {S}}^d\): the class \(\Psi _{d,{\mathcal {I}}}^m\)

In this section, we pay attention to matrix-valued positive definite functions on the unit sphere. Consider \({\mathbf{F}} \in {\mathcal {P}}^m({\mathbb {S}}^d)\). We say that \({\mathbf{F}}\) is geodesically isotropic if there exists a bounded and continuous mapping \(\varvec{\psi }:[0,\pi ]\rightarrow {\mathbb {R}}^{m\times m}\) such that

$$\begin{aligned} {\mathbf{F}}(\varvec{x},\varvec{y}) = \varvec{\psi }(\theta (\varvec{x},\varvec{y})), \quad \varvec{x},\varvec{y}\in {\mathbb {S}}^d. \end{aligned}$$

The continuous mappings \(\varvec{\psi }\) are the elements of the class \(\Psi _{d,{\mathcal {I}}}^m\) and the following inclusion relations are true:

$$\begin{aligned} \Psi _{\infty ,{\mathcal {I}}}^m = \bigcap _{d= 1}^\infty \Psi _{d,{\mathcal {I}}}^m \subset \cdots \subset \Psi _{2,{\mathcal {I}}}^m \subset \Psi _{1,{\mathcal {I}}}^m, \end{aligned}$$

where \(\Psi _{\infty ,{\mathcal {I}}}^m\) is the class of geodesically isotropic positive definite functions being valid on the Hilbert sphere \({\mathbb {S}}^\infty = \{ (x_n)_{n\in {\mathbb {N}}} \in {\mathbb {R}}^{{\mathbb {N}}} : \sum _{n\in {\mathbb {N}}} x_n^2 =1 \}\).

The elements of the class \(\Psi _{d,{\mathcal {I}}}^m\) have an explicit connection with Gegenbauer (or ultraspherical) polynomials (Abramowitz and Stegun 1970). Here, \({\mathcal {G}}_n^{\lambda }\) denotes the \(\lambda\)-Gegenbauer polynomial of degree n, which is defined implicitly through the expression

$$\begin{aligned} \frac{1}{(1+r^2-2r\cos \theta )^\lambda } = \sum _{n=0}^\infty r^n {\mathcal {G}}_n^{\lambda }(\cos \theta ), \quad \theta \in [0,\pi ], \quad r\in (-1,1). \end{aligned}$$

In particular, \({\mathcal {T}}_n := {\mathcal {G}}_n^{0}\) and \({\mathcal {P}}_n := {\mathcal {G}}_n^{1/2}\) are respectively the Chebyshev and Legendre polynomials of degree n.

The following result (Hannan 2009; Yaglom 1987) offers a complete characterization of the classes \(\Psi _{d,{\mathcal {I}}}^m\) and \(\Psi _{\infty ,{\mathcal {I}}}^m\), and corresponds to the multivariate version of Schoenberg’s Theorem (Schoenberg 1942). Equalities and summability conditions for matrices must be understood in a componentwise sense.

Theorem 6.1

Letdandmbe positive integers.

  1. (1)

    The mapping\(\varvec{\psi }\)is a member of the class\(\Psi _{d,{\mathcal {I}}}^m\)if and only if it admits the representation

    $$\begin{aligned} \varvec{\psi }(\theta ) = \sum _{n=0}^\infty {\mathbf{B}}_{n,d} \frac{{\mathcal {G}}_n^{(d-1)/2}(\cos \theta )}{{\mathcal {G}}_n^{(d-1)/2}(1)}, \quad \theta \in [0,\pi ], \end{aligned}$$

    where\(\{\mathbf{B }_{n,d}\}_{n=0}^\infty\)is a sequence of symmetric, positive definite and summable matrices.

  2. (2)

    The mapping \(\varvec{\psi }\)is a member of the class\(\Psi _{\infty ,{\mathcal {I}}}^m\)if and only if it can be represented as

    $$\begin{aligned} \varvec{\psi }(\theta ) = \sum _{n=0}^\infty {\mathbf{B}}_{n} (\cos \theta )^n, \quad \theta \in [0,\pi ], \end{aligned}$$

    where\(\{\mathbf{B }_{n}\}_{n=0}^\infty\)is a sequence of symmetric, positive definite and summable matrices.

Using orthogonality properties of Gegenbauer polynomials (Abramowitz and Stegun 1970) and through classical Fourier inversion we can prove that

$$\begin{aligned} {\mathbf{B}}_{0,1} & = \frac{1}{\pi }\int _0^\pi \varvec{\psi }(\theta )\text {d}\theta , \\ {\mathbf{B}}_{n,1} & = \frac{2}{\pi }\int _0^\pi \cos (n\theta )\varvec{\psi }(\theta )\text {d}\theta , \quad {\text { for }} n\ge 1, \end{aligned}$$

whereas for \(d\ge 2\), we have

$$\begin{aligned} {\mathbf{B}}_{n,d} = \frac{(2n+d-1)}{2^{3-d}\pi } \frac{[\Gamma ((d-1)/2)]^2}{\Gamma (d-1)} \int _{0}^\pi {\mathcal {G}}_{n}^{(d-1)/2}(\cos \theta ) (\sin \theta )^{d-1} \varvec{\psi }(\theta ) \text {d}\theta , \quad n\ge 0, \end{aligned}$$

where integration is taken componentwise. The matrices \(\{\mathbf{B }_{n,d}\}_{n=0}^\infty\) are called Schoenberg’s matrices. For the case \(m=1\), such result is reported by Gneiting (2013).

1.3 Matrix-valued positive definite functions on \({\mathbb {S}}^d\times {\mathbb {R}}^k\): the class \(\Upsilon _{d,k}^m\)

Let d, k and m be positive integers. We now focus on the class of matrix-valued positive definite functions on \({\mathbb {S}}^d \times {\mathbb {R}}^k\), being bounded, continuous, geodesically isotropic in the spherical component and stationary in the Euclidean one. The case \(k=1\) is particularly important, since \({\mathcal {P}}^m({\mathbb {S}}^d\times {\mathbb {R}})\) can be interpreted as the class of admissible space–time covariances for multivariate Gaussian RFs, with spatial locations on the unit sphere.

Consider \({\mathbf{F}} \in {\mathcal {P}}^m({\mathbb {S}}^d\times {\mathbb {R}}^k)\) and suppose that there exists a bounded and continuous mapping \({\mathbf{C}}:[0,\pi ]\times {\mathbb {R}}^k \rightarrow {\mathbb {R}}^{m\times m}\) such that

$$\begin{aligned} {\mathbf{F}}((\varvec{x},\varvec{t}),(\varvec{y},\varvec{s})) = {\mathbf{C}}(\theta (\varvec{x},\varvec{y}),\varvec{t}-\varvec{s}), \quad \varvec{x},\varvec{y}\in {\mathbb {S}}^d, \varvec{t},\varvec{s}\in {\mathbb {R}}^k. \end{aligned}$$

Such mappings \({\mathbf{C}}\) are the elements of the class \(\Upsilon _{d,k}^m\). These classes are non-increasing in d and we have the inclusions

$$\begin{aligned} \Upsilon _{\infty ,k}^m := \bigcap _{d=1}^\infty \Upsilon _{d,k}^m \subset \cdots \subset \Upsilon _{2,k}^m \subset \Upsilon _{1,k}^m. \end{aligned}$$

Ma (2017) proposes the generalization of Theorem 6.1 to the space–time case. Theorem 6.2 below offers a complete characterization of the class \(\Upsilon _{d,k}^m\) and \(\Upsilon _{\infty ,k}^m\), for any \(m\ge 1\). Again, equalities and summability conditions must be understood in a componentwise sense.

Theorem 6.2

Let d, k and m be positive integers and \({\mathbf{C}}: [0,\pi ] \times \mathbb {R}^k \rightarrow {\mathbb {R}}^{m\times m}\)a continuous matrix-valued mapping, with\({\text {C}}_{ii}(0,\varvec{0})<\infty\), for all\(i=1,\ldots ,m\).

  1. (1)

    The mapping\(\mathbf{C }\)belongs to the class\(\Upsilon _{d,k}^m\)if and only if

    $$\begin{aligned} \mathbf{C }(\theta ,\varvec{u}) = \sum _{n=0}^{\infty } \tilde{\varvec{\varphi }}_{n,d}(\varvec{u}) \frac{{\mathcal {G}}_n^{(d-1)/2}(\cos \theta )}{{\mathcal {G}}_n^{(d-1)/2}(1)}, \quad (\theta ,\varvec{u}) \in [0,\pi ] \times \mathbb {R}^k, \end{aligned}$$

    with\(\left\{ \tilde{\varvec{\varphi }}_{n,d}(\cdot ) \right\} _{n=0}^{\infty }\)being a sequence of members of the class\(\Phi _{k,{\mathcal {S}}}^m\), with the additional requirement that the sequence of matrices\(\{ \tilde{\varvec{\varphi }}_{n,d}(\varvec{0})\}_{n=0}^\infty\)is summable.

  2. (2)

    The mapping\(\mathbf{C }\)belongs to the class\(\Upsilon _{\infty ,k}^m\)if and only if

    $$\begin{aligned} \mathbf{C }(\theta ,\varvec{u}) = \sum _{n=0}^{\infty } \tilde{\varvec{\varphi }}_n(\varvec{u}) (\cos \theta )^n, \quad (\theta ,\varvec{u}) \in [0,\pi ] \times \mathbb {R}^k, \end{aligned}$$

    with\(\left\{ \tilde{\varvec{\varphi }}_n(\cdot ) \right\} _{n=0}^{\infty }\)being a sequence of members of the class\(\Phi _{k,{\mathcal {S}}}^m\), with the additional requirement that the sequence of matrices\(\{\tilde{\varvec{\varphi }}_n(\varvec{0})\}_{n=0}^\infty\)is summable.

Again, using orthogonality arguments, we have

$$\begin{aligned} \tilde{\varvec{\varphi }}_{0,1}(\varvec{u}) & = \frac{1}{\pi }\int _0^\pi \mathbf{C }(\theta ,\varvec{u}) \text {d}\theta , \\ \tilde{\varvec{\varphi }}_{n,1}(\varvec{u}) & = \frac{2}{\pi }\int _0^\pi \cos (n\theta )\mathbf{C }(\theta ,\varvec{u}) \text {d}\theta , \quad {\text { for }} n\ge 1, \end{aligned}$$

whereas for \(d\ge 2\),

$$\begin{aligned} \tilde{\varvec{\varphi }}_{n,d}(\varvec{u}) & = \frac{(2n+d-1)}{2^{3-d}\pi } \frac{[\Gamma ((d-1)/2)]^2}{\Gamma (d-1)}\nonumber \\&\int _{0}^{\pi } {\mathcal {G}}_{n}^{(d-1)/2}(\cos \theta ) (\sin \theta )^{d-1} \mathbf{C }(\theta ,\varvec{u}) \mathrm{d} \theta , \quad n\ge 0. \end{aligned}$$

Appendix 2: Proof of Theorem 3.1

In order to illustrate the results following subsequently, a technical Lemma will be useful. We do not provide a proof because it is obtained following the same arguments as in Porcu and Zastavnyi (2011).

Lemma 7.1

Let m, d and k be strictly positive integers. Let\((X,{\mathcal{B}}, \mu )\)be a measure space, for\(X \subset \mathbb {R}\)and\({\mathcal B}\)being the Borel sigma algebra. Let\(\varvec{\psi }: [0,\pi ] \times X \rightarrow \mathbb {R}\) and\(\varvec{\varphi }: [0,\infty ) \times X \rightarrow \mathbb {R}\) be continuous mappings satisfying

  1. 1.

    \(\varvec{\psi }(\cdot ,\xi ) \in \Psi _{d,{\mathcal{I}}}^m\) a.e.\(\xi \in X\);

  2. 2.

    \(\varvec{\psi }(\theta ,\cdot ) \in L_1(X,{\mathcal{B}},\mu )\) for any\(\theta \in [0,\pi ]\);

  3. 3.

    \(\varvec{\varphi }(\cdot ,\xi ) \in \Phi _{k,{\mathcal{I}}}^m\) a.e.\(\xi \in X\);

  4. 4.

    \(\varvec{\varphi }(u, \cdot ) \in L_1(X,{\mathcal{B}},\mu )\) for any\(u \in [0,\infty )\).

Let\(\mathbf{C } :[0,\pi ] \times [0,\infty )\rightarrow {\mathbb {R}}^{m\times m}\)be the mapping defined through

$$\begin{aligned} \mathbf{C } (\theta ,u) = \int _X \varvec{\psi }(\theta ,\xi ) \varvec{\varphi }(u, \xi ) \mu (\mathrm{d} \xi ), \quad (\theta ,u) \in [0,\pi ] \times [0,\infty ). \end{aligned}$$

Then, \(\mathbf{C }\)is continuous and bounded. Further,\(\mathbf{C }\)belongs to the class\(\Upsilon _{d,k}^m\).

Of course, Lemma 7.1 is a particular case of the scale mixtures introduced in Sect. 2.

Proof of Theorem 3.1

Let \((X, {\mathcal{B}}, \mu )\) as in Lemma 7.1 and consider \(X = \mathbb {R}_+\) with \(\mu\) the Lebesgue measure. We offer a proof of the constructive type. Let us define the function \(\varvec{\psi }(\theta ,\xi )\) with members \(\psi _{ij}(\cdot ,\cdot )\) defined through

$$\begin{aligned} \psi _{ij}(\theta ,\xi ) = { \sigma _{ii}\sigma _{jj} \rho _{ij}} \left( 1 - \frac{\theta }{\xi c_{ij} } \right) _{+}^{n+1}, \quad (\theta , \xi ) \in [0,\pi ] \times X, \quad i,j=1,\ldots ,m, \end{aligned}$$

where, as asserted, the constants \(\sigma _{ii}\), \(\rho _{ij}\) and \(c_{ij}\) are determined according to condition (3.1). Let us now define the mapping \((u,\xi ) \mapsto \varphi (u,\xi ):= \xi ^{n+1} (1 - \xi f(u))_{+}^{\ell }\), with \((u,\xi ) \in [0,\infty ) \times X\). It can be verified that both \(\varvec{\psi }\) and \(\varphi\) satisfy requirements 1–4 in Lemma 7.1. In particular, Condition 1 yields thanks to Lemmas 3 and 4 in Gneiting (2013), as well as Theorem 1 in Daley et al. (2015). Also, arguments in Porcu et al. (2016) show that Condition 3 holds for any \(\ell \ge 1\). We can now apply Lemma 7.1, so that we have that

$$\begin{aligned} {\text {C}}_{i,j, n,\ell }(\theta ,u):= \int _X \varvec{\psi }(\theta ,\xi ) \varphi (u,\xi ) \mathrm{d} \xi , \quad [0, \pi ] \times [0,\infty ) \end{aligned}$$

is a member of the class \(\Upsilon _{2n+1,1}^m\) for any \(\ell \ge 1\). Pointwise application of an elegant scale mixture argument as in Proposition 1 of Porcu et al. (2016) shows that

$$\begin{aligned} {\text {C}}_{i,j,n,\ell } (\theta ,u) = {\mathcal {B}}(n+2,\ell +1) \frac{ \rho _{ij} \sigma _{ii} \sigma _{jj}}{f(u)^{n+2}} \left( 1- \frac{\theta f(u)}{c_{ij}} \right) _+^{n+\ell +1}, \quad (\theta ,u) \in [0,\pi ] \times [0,\infty ), \end{aligned}$$

where \({\mathcal {B}}\) denotes the Beta function (Abramowitz and Stegun 1970). We now omit the factor \({\mathcal {B}}(n+2,\ell +1)\) since it does not affect positive definiteness. Now, standard convergence arguments show that

$$\begin{aligned} \lim _{\ell \rightarrow \infty } {\text {C}}_{i,j,n,\ell }(\theta / \ell ,u) = \frac{ \rho _{ij} \sigma _{ii} \sigma _{jj}}{f(u)^{n+2}} \exp \left( - \frac{\theta f(u)}{c_{ij}} \right) , (\theta ,u) \in [0,\pi ] \times [0,\infty ), \end{aligned}$$

with the convergence being uniform in any compact set. The proof is then completed in view of Bernstein’s theorem (Feller 1966). \(\square\)

Appendix 3: Proof of Theorem 3.2

Before we state the proof of Theorem 3.2, we need to introduce three auxiliary lemmas.

Lemma 8.1

Let \({\text {C}}: [0,\pi ]\times {\mathbb {R}}^k \rightarrow {\mathbb {R}}\) be a continuous, bounded and integrable function, for some positive integer k. Then \({\text {C}} \in \Upsilon _{d,k}\), for \(d\ge 1\), if and only if the mapping \(\psi _{\varvec{\omega }}:[0,\pi ] \rightarrow {\mathbb {R}}\) defined as

$$\begin{aligned} \psi _{\varvec{\omega }}(\theta ) = \frac{1}{(2\pi )^k}\int _{{\mathbb {R}}^k} \exp \{-\imath \varvec{\omega }^\top \varvec{v} \} {\text {C}}(\theta ,\varvec{v}) \text {d}\varvec{v}, \quad \theta \in [0,\pi ], \end{aligned}$$
(8.1)

belongs to the class \(\Psi _{d,{\mathcal {I}}}\), for all \(\varvec{\omega }\in {\mathbb {R}}^k\).

We do not report the proof of Lemma 8.1 since the arguments are the same as in the proof of Lemma 8.2 below. Note that this lemma is a spherical version of the result given by Cressie and Huang (1999).

Lemma 8.2

Let\({\text {C}}: [0,\pi ]\times {\mathbb {R}}^l \times {\mathbb {R}}^k \rightarrow {\mathbb {R}}\)be a continuous, bounded and integrable function, for some positive integers l and k. Then , \({\text {C}} \in \Upsilon _{d,k+l}\), with\(d\ge 1\), if and only if the mapping\({\text {C}}_{\varvec{\omega }}:[0,\pi ] \times {\mathbb {R}}^l \rightarrow {\mathbb {R}}\)defined as

$$\begin{aligned}&{\text {C}}_{\varvec{\omega }}(\theta ,\varvec{u}) = \frac{1}{(2\pi )^k}\int _{{\mathbb {R}}^k} \exp \{-\imath \varvec{\omega }^\top \varvec{v} \} {\text {C}}(\theta ,\varvec{u},\varvec{v}) \text {d}\varvec{v}, \\&\quad (\theta ,\varvec{u})\in [0,\pi ] \times {\mathbb {R}}^l, \end{aligned}$$
(8.2)

belongs to the class\(\Upsilon _{d,l}\), for all\(\varvec{\omega }\in {\mathbb {R}}^k\).

Proof of Lemma 8.2

Suppose that \({\text {C}} \in \Upsilon _{d,k+l}\), then the characterization of Berg and Porcu (2017) implies that

$$\begin{aligned} {\text {C}}(\theta ,\varvec{u},\varvec{v}) = \sum _{n=0}^\infty \tilde{\varphi }_{n,d}(\varvec{u},\varvec{v}) \frac{{\mathcal {G}}_n^{(d-1)/2}(\cos \theta )}{{\mathcal {G}}_n^{(d-1)/2}(1)}, \end{aligned}$$

where \(\{\tilde{\varphi }_{n,d}(\cdot ,\cdot )\}_{n=0}^\infty\) is a sequence of functions in \(\Phi _{k+l,{\mathcal {S}}}\), with \(\sum _{n=0}^\infty \tilde{\varphi }_{n,d}(\varvec{0},\varvec{0}) < \infty\). Therefore,

$$\begin{aligned}&{\text {C}}_{\varvec{\omega }}(\theta ,\varvec{u})\nonumber \\&\quad = \frac{1}{(2\pi )^k} \int _{{\mathbb {R}}^k} \exp \{-\imath \varvec{\omega }^\top \varvec{v} \} \left( \sum _{n=0}^\infty \tilde{\varphi }_{n,d}(\varvec{u},\varvec{v}) \frac{{\mathcal {G}}_n^{(d-1)/2}(\cos \theta )}{{\mathcal {G}}_n^{(d-1)/2}(1)} \right) \text {d}\varvec{v}\\&\quad = \sum _{n=0}^\infty \left( \frac{1}{(2\pi )^k}\int _{{\mathbb {R}}^k} \exp \{-\imath \varvec{\omega }^\top \varvec{v} \} \tilde{\varphi }_{n,d}(\varvec{u},\varvec{v}) \text {d}\varvec{v} \right) \frac{{\mathcal {G}}_n^{(d-1)/2}(\cos \theta )}{{\mathcal {G}}_n^{(d-1)/2}(1)}, \end{aligned}$$

where the last step is justified by dominated convergence. We need to prove that for each fixed \(\varvec{\omega }\in {\mathbb {R}}^k\), the sequence of functions

$$\begin{aligned} \varvec{u} \mapsto \tilde{\lambda }_{n,d}(\varvec{u};\varvec{\omega }) : = \frac{1}{(2\pi )^k} \int _{{\mathbb {R}}^k} \exp \{-\imath \varvec{\omega }^\top \varvec{v} \} \tilde{\varphi }_{n,d}(\varvec{u},\varvec{v}) \text {d}\varvec{v}, \quad n\ge 0, \end{aligned}$$

belongs to the class \(\Phi _{l,{\mathcal {S}}}\), a.e. \(\varvec{\omega }\in {\mathbb {R}}^k\). In fact, we have that

$$\begin{aligned}&\frac{1}{(2\pi )^{l}} \int _{{\mathbb {R}}^l} \exp \{-\imath \varvec{\tau }^\top \varvec{u} \} \tilde{\lambda }_{n,d}(\varvec{u};\varvec{\omega }) \text {d}\varvec{u} \\&\quad = \frac{1}{(2\pi )^{k+l}} \int _{{\mathbb {R}}^l} \int _{{\mathbb {R}}^k} \exp \{-\imath \varvec{\tau }^\top \varvec{u} - \imath \varvec{\omega }^\top \varvec{v} \} \tilde{\varphi }_{n,d}(\varvec{u},\varvec{v}) \text {d}\varvec{v} \text {d}\varvec{u}. \end{aligned}$$
(8.3)

Since \(\tilde{\varphi }_{n,d}(\cdot ,\cdot )\) belongs to \(\Phi _{k+l,{\mathcal {S}}}\), Bochner’s Theorem implies that the right side in Eq. (8.3) is non-negative everywhere. This implies that \(\tilde{\lambda }_{n,d}(\cdot ;\varvec{\omega })\) belongs to the class \(\Phi _{l,{\mathcal {S}}}\). Also, direct inspection shows that \(\sum _{n=0}^\infty \tilde{\lambda }_{n,d}(\varvec{0};\varvec{\omega }) < \infty\), for all \(\varvec{\omega }\in {\mathbb {R}}^k\). The necessary part is completed.

On the other hand, suppose that for each \(\varvec{\omega }\in {\mathbb {R}}^k\) the function \({\text {C}}_{\varvec{\omega }}(\theta ,\varvec{u})\) belongs to the class \(\Upsilon _{d,l}\), then there exists a sequence of mappings \(\{ \tilde{\lambda }_{n,d}(\cdot ;\varvec{\omega }) \}_{n=0}^\infty\) in \(\Phi _{l,{\mathcal {S}}}\) for each \(\varvec{\omega }\in {\mathbb {R}}^k\), such that

$$\begin{aligned} {\text {C}}_{\varvec{\omega }}(\theta ,\varvec{u}) = \sum _{n=0}^\infty \tilde{\lambda }_{n,d}(\varvec{u};\varvec{\omega }) \frac{{\mathcal {G}}_n^{(d-1)/2}(\cos \theta )}{{\mathcal {G}}_n^{(d-1)/2}(1)}. \end{aligned}$$

Thus,

$$\begin{aligned} {\text {C}}(\theta ,\varvec{u},\varvec{v}) & = \int _{{\mathbb {R}}^k} \exp \{ \imath \varvec{\omega }^\top \varvec{v} \} \left( \sum _{n=0}^\infty \tilde{\lambda }_{n,d}(\varvec{u};\varvec{\omega }) \frac{{\mathcal {G}}_n^{(d-1)/2}(\cos \theta )}{{\mathcal {G}}_n^{(d-1)/2}(1)} \right) \text {d}\varvec{\omega }\\ & = \sum _{n=0}^\infty \left( \int _{{\mathbb {R}}^k} \exp \{ \imath \varvec{\omega }^\top \varvec{v} \} \tilde{\lambda }_{n,d}(\varvec{u};\varvec{\omega }) \text {d}\varvec{\omega } \right) \frac{{\mathcal {G}}_n^{(d-1)/2}(\cos \theta )}{{\mathcal {G}}_n^{(d-1)/2}(1)}. \end{aligned}$$

We conclude the proof by invoking again Bochner’s Theorem and the result of Berg and Porcu (2017). \(\square\)

Lemma 8.3

Let d and k be two positive integers. Consider g and f be completely monotone and Bernstein functions, respectively. Then,

$$\begin{aligned} {\text {K}}(\theta ,\varvec{v}) = \frac{1}{ \left\{ f(\theta ) |_{[0,\pi ]} \right\} ^{k/2} } g\left( \frac{\Vert \varvec{v}\Vert ^2}{f(\theta ) |_{[0,\pi ]} } \right) , \quad (\theta ,\varvec{v}) \in [0,\pi ]\times {\mathbb {R}}^k, \end{aligned}$$

belongs to the class\(\Upsilon _{d,k}\), for any positive integer d.

Proof of Lemma 8.3

By Lemma 8.1, we must show that \(\psi _{\varvec{\omega }}\), defined through Eq. (8.1), belongs to the class \(\Psi _{d,{\mathcal {I}}}\), for all \(\varvec{\omega }\in {\mathbb {R}}^k\). In fact, we can assume that \({\text {C}}\) is integrable, since the general case is obtained with the same arguments given by Gneiting (2002). Bernstein’s Theorem establishes that g can be represented as the Laplace transform of a bounded measure G, then

$$\begin{aligned} \psi _{\varvec{\omega }} (\theta ) & = \int _{{\mathbb {R}}^k} \exp \{-\imath \varvec{\omega }^\top \varvec{v} \} \frac{1}{ \left\{ f(\theta ) |_{[0,\pi ]} \right\} ^{k/2} } \int _{[0,\infty )}\\&\exp \left\{ - \frac{r\Vert \varvec{v}\Vert ^2}{ f(\theta ) |_{[0,\pi ]} } \right\} \text {d}G(r) \text {d}\varvec{v}\\ & = \pi ^{k/2} \int _{[0,\infty )} \exp \left\{ - \frac{\Vert \varvec{\omega }\Vert ^2}{4r} f(\theta ) |_{[0,\pi ]} \right\} \text {d}\tilde{G}(r), \end{aligned}$$

where the last equality follows from Fubini’s Theorem and \(\text {d}G(r) = r^{k/2} \text {d}\tilde{G}(r)\). In addition, the composition between a negative exponential and a Bernstein function is completely monotone on the real line (Feller 1966). Then, for any \(\varvec{\omega }\) and r, the mapping \(\theta \mapsto \exp \{ - \Vert \varvec{\omega }\Vert ^2 f(\theta )|_{[0,\pi ]} /(4r)\}\) is the restriction of a completely monotone function to the interval \([0,\pi ]\). Theorem 7 in Gneiting (2013) implies that such mapping, and thus \(\psi _{\varvec{\omega }}\), belongs to the class \(\Psi _{d,{\mathcal {I}}}\), for any \(d\in {\mathbb {N}}\) and \(\varvec{\omega }\in {\mathbb {R}}^k\). \(\square\)

Proof of Theorem 3.2

By Lemma 8.2, we must show that (8.2) belongs to the class \(\Upsilon _{d,l}\), for all \(d\in {\mathbb {N}}\). In fact, assuming again that \({\text {C}}\) is integrable and invoking Bernstein’s Theorem we have

$$\begin{aligned} {\text {C}}_{\varvec{\omega }} (\theta ,\varvec{u}) & = \int _{{\mathbb {R}}^k} \exp \{-\imath \varvec{\omega }^\top \varvec{v} \} \frac{1}{ \{f_2( \theta )|_{[0,\pi ]}\}^{l/2} \left\{ f_1\left[ \frac{\Vert \varvec{u}\Vert ^2}{ f_2( \theta )|_{[0,\pi ]} } \right] \right\} ^{k/2}} \int _{[0,\infty )} \\&\exp \left\{ - \frac{r \Vert \varvec{v}\Vert ^2}{ f_1\left[ \frac{\Vert \varvec{u}\Vert ^2}{ f_2( \theta )|_{[0,\pi ]} } \right] } \right\} \text {d}G(r) \text {d}\varvec{v}\\ & = \pi ^{k/2} \frac{1}{\{f_2( \theta )|_{[0,\pi ]}\}^{l/2} } \int _{[0,\infty )} \\&\exp \left\{ - \frac{\Vert \varvec{\omega }\Vert ^2}{4r} f_1\left[ \frac{\Vert \varvec{u}\Vert ^2}{ f_2( \theta )|_{[0,\pi ]} } \right] \right\} \text {d}\tilde{G}(r), \end{aligned}$$

where the last equality follows from Fubini’s Theorem and \(\text {d}G(r) = r^{k/2} \text {d}\tilde{G}(r)\). In addition, for any \(\varvec{\omega }\), the mapping

$$\begin{aligned} g_{\varvec{\omega }}(\cdot ) := \int _{[0,\infty )} \exp \left\{ - \frac{ \Vert \varvec{\omega }\Vert ^2}{4r} f_1(\cdot ) \right\} \text {d}\tilde{G}(r) \end{aligned}$$

is completely monotone (Feller 1966). Therefore,

$$\begin{aligned} {\text {C}}_{\varvec{\omega }} (\theta ,\varvec{u}) = \pi ^{k/2} \frac{1}{\{f_2( \theta )|_{[0,\pi ]}\}^{l/2} } g_{\varvec{\omega }}\left( \frac{\Vert \varvec{u}\Vert ^2}{ f_2( \theta )|_{[0,\pi ]} } \right) , \end{aligned}$$

and by Lemma 8.3, we have that \({\text {C}}_{\varvec{\omega }}\in \Upsilon _{d,l}\), for all \(d\in {\mathbb {N}}\). \(\square\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Alegría, A., Porcu, E., Furrer, R. et al. Covariance functions for multivariate Gaussian fields evolving temporally over planet earth. Stoch Environ Res Risk Assess 33, 1593–1608 (2019). https://doi.org/10.1007/s00477-019-01707-w

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00477-019-01707-w

Keywords

Navigation