Skip to main content
Log in

Strict positive definiteness under axial symmetry on the sphere

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

Abstract

Axial symmetry for covariance functions defined over spheres has been a very popular assumption for climate, atmospheric, and environmental modeling. For Gaussian random fields defined over spheres embedded in a three-dimensional Euclidean space, maximum likelihood estimation techiques as well kriging interpolation rely on the inverse of the covariance matrix. For any collection of points where data are observed, the covariance matrix is determined through the realizations of the covariance function associated with the underlying Gaussian random field. If the covariance function is not strictly positive definite, then the associated covariance matrix might be singular. We provide conditions for strict positive definiteness of any axially symmetric covariance function. Furthermore, we find conditions for reducibility of an axially symmetric covariance function into a geodesically isotropic covariance. Finally, we provide conditions that legitimate Fourier inversion in the series expansion associated with an axially symmetric covariance function.

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.

Institutional subscriptions

Similar content being viewed by others

References

  • Abramowitz M, Stegun IA (1964) Handbook of mathematical functions: with formulas, graphs, and mathematical tables, vol 55. Courier Corporation, North Chelmsford

    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

  • Barbosa VS, Menegatto VA (2017) Strict positive definiteness on products of compact two-point homogeneous spaces. Integral Transforms Spec Funct 28(1):56–73

    Article  Google Scholar 

  • Beatson RK, zu Castell W (2017) Dimension hopping and families of strictly positive definite zonal basis functions on spheres. J Approx Theory 221:22–37

    Article  Google Scholar 

  • Berg C, Christensen JPR, Ressel P (1984) Harmonic analysis on semigroups: Theory of positive definite and related functions, vol 100. Graduate texts in mathematics. Springer, New York

    Book  Google Scholar 

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

    Article  Google Scholar 

  • Chen D, Menegatto VA, Sun X (2003) A necessary and sufficient condition for strictly positive definite functions on spheres. Proc Am Math Soc 131:2733–2740

    Article  Google Scholar 

  • Clarke J, Alegria A, Porcu E (2018) Regularity properties and simulations of gaussian random fields on the sphere cross time. Electron J Stat 1:399–426

    Article  Google Scholar 

  • Daley DJ, Porcu E (2013) Dimension walks and schoenberg spectral measures. Proc Am Math Soc 141:1813–1824

    Google Scholar 

  • De Iaco S, Posa D (2018) Strict positive definiteness in geostatistics. Stoch Environ Res Risk Assess 32:577–590

    Article  Google Scholar 

  • Emery X, Porcu E, Bissiri PG (2019) A semiparametric class of axially symmetric random fields on the sphere. Stoch Environ Res Risk Assess 33:1863–1874

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Guella JC, Menegatto VA, Peron AP (2016a) An extension of a theorem of schoenberg to a product of spheres. Banach J Math Anal 10(4):671–685

    Article  Google Scholar 

  • Guella JC, Menegatto VA, Peron AP (2016b) Strictly positive definite kernels on a product of spheres II. In: SIGMA, vol 12(103)

  • Guella JC, Menegatto VA, Peron AP (2017) Strictly positive definite kernels on a product of circles. Positivity 21(1):329–342

    Article  Google Scholar 

  • Hitczenko M, Stein ML (2012) Some theory for anisotropic processes on the sphere. Stat Methodol 9:211–227

    Article  Google Scholar 

  • Huang C, Zhang H, Robeson S (2012) A simplified representation of the covariance structure of axially symmetric processes on the sphere. Stat Probab Lett 82:1346–1351

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Jun M, Stein ML (2007) An approach to producing space-time covariance functions on spheres. Technometrics 49:468–479

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Lang A, Schwab C (2013) Isotropic random fields on the sphere: regularity, fast simulation and stochastic partial differential equations. Ann Appl Prob 25:3047–3094

    Article  Google Scholar 

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

    Book  Google Scholar 

  • Menegatto VA (1994) Strictly positive definite kernels on the hilbert sphere. Appl Anal 55:91–101

    Article  Google Scholar 

  • Menegatto VA (1995) Strictly positive definite kernels on the circle. Rocky Mt J Math 25:1149–1163

    Article  Google Scholar 

  • Menegatto VA, Oliveira CP, Peron AP (2006) Strictly positive definite kernels on subsets of the complex plane. Comput Math Appl 51:1233–1250

    Article  Google Scholar 

  • Myers D (1992) Kriging, cokriging, radial basis functions and the role of positive definiteness. Comput Math Appl 24:139–148

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Porcu E, Castruccio S, Alegría A, Crippa P (2019) Axially symmetric models for global data: a journey between geostatistics and stochastic generators. Environmetrics 30:1327–1349

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Stein ML (1999) Statistical interpolation of spatial data: some theory for kriging. Springer, New York

    Book  Google Scholar 

  • Stein ML (2007) Spatial variation of total column ozone on a global scale. Ann Appl Stat 1:191–210

    Article  Google Scholar 

  • Xu Y, Cheney EW (1992) Strictly positive definite functions on spheres. Proc Am Math Soc 116:977–981

    Article  Google Scholar 

Download references

Acknowledgements

Emilio Porcu has been partially supported by CONICYT/FONDECYT/REGULAR/No. 1170290. Ana Paula Peron has been partially supported by FAPESP # 2016/09906-0 and CNPq # 203033/2019-1.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emilio Porcu.

Additional information

Publisher's Note

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

Appendix: Mathematical Proofs and Lemmas

Appendix: Mathematical Proofs and Lemmas

1.1 Proof of Proposition 1

Proof

We start by making use of the assumption \(c_m(n,n')=c(n)\delta _{nn'}\) to write

$$\begin{aligned}&K(L,L^{\prime },\Delta \ell ) \\&\quad = \sum _{m=-\infty }^\infty \sum _{n,n'=\left| {m}\right| }^\infty e^{{\mathsf {i}} m \Delta \ell }\bar{P}^m_{n}(\sin L)\bar{P}^m_{n'}(\sin L^{\prime })c(n)\delta _{nn'} \\&\quad =\sum _{m=-\infty }^\infty \sum _{n=\left| {m}\right| }^\infty e^{{\mathsf {i}} m\Delta \ell }\bar{P}^m_{n}(\sin L)\bar{P}^m_n(\sin L^{\prime }) c(n). \end{aligned}$$
(23)

The two sums in (23) can be interchanged provided

$$\begin{aligned} \sum _{n=0}^\infty \sum _{m=-n}^n \left| {\bar{P}^m_{n}(\sin L)\bar{P}^m_n(\sin L^{\prime })}\right| c(n)<\infty . \end{aligned}$$
(24)

To show it, we note that Cauchy-Schwartz inequality in concert with (12) show that, for every \(n\ge 0\),

$$\begin{aligned}&\sum _{m=-n}^n \left| {\bar{P}^m_{n}(\sin L)\bar{P}^m_n(\sin L^{\prime })}\right| \\&\quad \le \sqrt{\sum _{m=-n}^n (\bar{P}^m_{n}(\sin L))^2 \sum _{m=-n}^n (\bar{P}^m_n(\sin L^{\prime }))^2} =n+\frac{1}{2}. \end{aligned}$$
(25)

The fact that \(\textstyle \sum _{n=0}^\infty (n+1/2)c(n)<\infty\) proves that (24) holds true, so that the interchange is legitimate. Therefore, we can swap the two sums in (23) and write:

$$\begin{aligned} K(L,L^{\prime },\Delta \ell )=\sum _{n=0}^\infty c(n) \sum _{m=-n}^{n}e^{{\mathsf {i}} m \Delta \ell } \bar{P}^m_n(\sin L)\bar{P}^m_n(\sin L^{\prime }). \end{aligned}$$
(26)

By (10),

$$\begin{aligned} \bar{P}^{-m}_n(\sin L )\bar{P}^{-m}_n(\sin L^{\prime })= \bar{P}^m_n(\sin L )\bar{P}^m_n(\sin L^{\prime } ) \end{aligned}$$
(27)

so that:

$$\begin{aligned}&c(n)\{e^{-{\mathsf {i}} m\Delta \ell }\bar{P}^{-m}_n(\sin L )\bar{P}^{-m}_n(\sin L^{\prime } )\\&\qquad + e^{{\mathsf {i}} m\Delta \ell }\bar{P}^m_n(\sin L )\bar{P}^m_n(\sin L^{\prime } )\}\\&\quad = 2c(n)\bar{P}^m_n(\sin L )\bar{P}^m_n(\sin L^{\prime } )\cos (m\Delta \ell ), \end{aligned}$$

and therefore (26) yields:

$$\begin{aligned}&K(L,xL^{\prime }, \Delta \ell )\\&\quad =\sum _{n=0}^\infty c(n) \left\{ 2\sum _{m=1}^{n} \bar{P}^m_n(\sin L)\bar{P}^m_n(\sin L^{\prime } )\cos (m \Delta \ell )\right. \\&\qquad \qquad \qquad \left. + \bar{P}^0_n(\sin L )\bar{P}^0_n(\sin L^{\prime } ) \right\} , \end{aligned}$$

which by (27) and recalling that the cosine is an even function, yields:

$$\begin{aligned}&K(L,L^{\prime },\Delta \ell )\\&\quad =\sum _{n=0}^\infty c(n) \sum _{m=-n}^{n} \bar{P}^m_n(\sin L )\bar{P}^m_n(\sin L^{\prime } )\cos (m\Delta \ell )\\&\quad =\sum _{n=0}^\infty c(n)(2n+1)\frac{1}{2} \sum _{m=-n}^{n} \frac{(n-m)!}{(n+m)!}\\&\qquad {P}^m_n(\sin L ){P}^m_n(\sin L^{\prime })\cos (m \Delta \ell ), \end{aligned}$$

which by the spherical harmonics addition theorem (3) yields

$$\begin{aligned} K(L,L^{\prime }, \Delta \ell )= \sum _{n=0}^\infty c(n)\left( n+\frac{1}{2}\right) P_n(\langle \varvec{s}, \varvec{s}^{\prime } \rangle ), \end{aligned}$$
(28)

where \(\varvec{s},\varvec{s}^{\prime }\) belong to \({\mathbb {S}}^2\), so that \(\varvec{s}=(L,\ell )\) and \(\varvec{s}^{\prime }=(L^{\prime }, \ell ^{\prime })\). \(\square\)

1.2 Auxiliary Lemmas

To provide a constructive proof of Theorem 2, two auxiliary results are required.

Lemma 4

Fix\(m\in {\mathbb {Z}}\)and let\(w_1,\ldots ,w_h\)behdistinct points in\((-1,1)\)such that

$$\begin{aligned} \sum _{r=1}^h z_{r}\bar{P}^m_n(w_r)=0 \end{aligned}$$
(29)

for every\(n\in \{\left| {m}\right| ,\left| {m}\right| +1, \ldots \}\)and for some complexh-ple\((z_{1}, \ldots , z_{h})\). Then, \(z_{r}=0\)for every\(r=1,\ldots ,h\).

Proof

The identity (29) implies that

$$\begin{aligned} \sum _{r=1}^h {\text {Re}}(z_{r})\bar{P}^m_n(w_r)&=0 \end{aligned}$$
(30)
$$\begin{aligned} \sum _{r=1}^h {\text {Im}}(z_{r})\bar{P}^m_n(w_r)&=0 \end{aligned}$$
(31)

for every \(n\in \{\left| {m}\right| ,\left| {m}\right| +1, \ldots \}\).

At this stage, recall that

$$\begin{aligned} \bar{P}^m_n(x)=k_{m,n}(1-x^2)^{\left| {m}\right| /2}\frac{d^{\left| {m}\right| }}{d x^{\left| {m}\right| }}P_n(x) \end{aligned}$$

where \(k_{m,n}\) is a non zero constant, for each \(n\in \{\left| {m}\right| ,\left| {m}\right| +1,\ldots \}\) and \(x\in (-1,1)\).

Consider now a polynomial p of degree at most \(h-1\) such that

$$\begin{aligned} p(w_r)={\text {Re}}(z_r)/\{(1-w_r^2)^{\left| {m}\right| /2}\}, \end{aligned}$$
(32)

for \(r=1,\ldots ,h\), and let q be a polynomial of degree at most \(h+\left| {m}\right| -1\) such that the \(\left| {m}\right|\)-th partial derivative of q is p. Since the ordinary Legendre polynomials \(P_0,\ldots , P_{\left| {m}\right| +h-2}\) are known to generate the \((\left| {m}\right| +h-1)\)-dimensional polynomial space, there exist \(d_0,\ldots ,d_{\left| {m}\right| +h-2}\in {\mathbb {R}}\) such that

$$\begin{aligned} d_0P_0+\cdots + d_{\left| {m}\right| +h-2}P_{\left| {m}\right| +h-2}=q, \end{aligned}$$

which taking the \(\left| {m}\right|\)–th derivative of both sides and multiplying by \((1-x^2)^{\left| {m}\right| /2}\), yields that:

$$\begin{aligned} \frac{d_{\left| {m}\right| }}{k_{m,\left| {m}\right| }} \bar{P}^m_{\left| {m}\right| }(x)+\cdots + \frac{d_{\left| {m}\right| +h-2}}{k_{m,\left| {m}\right| +h-2}} \bar{P}^m_{\left| {m}\right| +h-2}(x)=(1-x^2)^{\left| {m}\right| /2}p(x) \end{aligned}$$
(33)

for every \(x\in (-1,1)\). Hence, we can consider the following function

$$\begin{aligned} f(x)=\frac{d_{\left| {m}\right| }}{k_{m,\left| {m}\right| }}P^m_{\left| {m}\right| }(x)+\cdots + \frac{d_{\left| {m}\right| +h-2}}{k_{m,\left| {m}\right| +h-2}}P^m_{\left| {m}\right| +h-2}(x) \end{aligned}$$

knowing that by (32) and (33) we have that

$$\begin{aligned} f(w_r)={\text {Re}}(z_r) \end{aligned}$$
(34)

as \(r=1,\ldots ,h\). By (30), we have that

$$\begin{aligned} \sum _{r=1}^h {\text {Re}}(z_{r})f(w_r)=0, \end{aligned}$$

which by (34) yields that

$$\begin{aligned} \sum _{r=1}^h {\text {Re}}(z_{r})^2=0, \end{aligned}$$
(35)

We have just proved that (30) implies (35). Similarly, we can prove that (31) implies

$$\begin{aligned} \sum _{r=1}^h {\text {Im}}(z_{r})^2=0 \end{aligned}$$

and the proof is complete. \(\square\)

Lemma 5

LetKbe an axially symmetric covariance function as in (17) such that for each\(m\ge 0\), \(c_m:\{m,m+1,\ldots \}\times \{m,m+1,\ldots \}\rightarrow {\mathbb {C}}\)is a strictly positive definite function. Then, for\(a=(a_1,\ldots ,a_k)\in {\mathbb {R}}^k\)and\(\varvec{s}_1,\ldots ,\varvec{s}_k\)distinct points in\({\mathbb {S}}^2\)such that the longitude and the latitude of\(\varvec{s}_j\)are\(\ell _j\in (-\pi ,\pi ]\)and\(L_j\in [-\pi /2,\pi /2]\), respectively (\(j=1,\ldots ,k\)), the following assertions are equivalent:

  1. (i)

    \(a^TKa=0\), that is,

    $$\begin{aligned} \sum _{j,h=1}^k a_jK(L_j,L_h, \Delta \ell _{jh})a_h=0, \qquad (\Delta \ell _{jh} := \ell _j -\ell _h). \end{aligned}$$
    (36)
  2. (ii)

    The equality

    $$\begin{aligned} \sum _{j=1}^k a_j e^{ {\mathsf {i}} m \ell _j}\bar{P}^m_n(\sin L_j )=0. \end{aligned}$$
    (37)

    holds for each\(m\in {\mathbb {Z}}\)and each\(n\in \{\left| {m}\right| ,\left| {m}\right| +1,\ldots \}\),

Proof

Note that

$$\begin{aligned} a^TKa&:= \sum _{j,h=1}^k a_jK(L_j,L_h, \Delta \ell _{jh})a_h \\&= \sum _{j,h=1}^k a_ja_h \sum _{m=-\infty }^\infty \sum _{n,n'=\left| {m}\right| }^\infty \\&\quad e^{{\mathsf {i}} m \Delta \ell _{jh}} \bar{P}^m_n(\sin L_j )\bar{P}^m_{n'}(\sin L_h ) c_m(n,n') \\&=\sum _{m=-\infty }^\infty \sum _{n,n'=\left| {m}\right| }^\infty c_m(n,n')\left\{ \sum _{j=1}^k a_j e^{ {\mathsf {i}} m \ell _j}\bar{P}^m_n(\sin L_j) \right\} \\&\quad \left\{ \sum _{h=1}^k a_h e^{- {\mathsf {i}} m \ell _h}\bar{P}^m_n(\sin L_h ) \right\} . \end{aligned}$$
(38)

Hence, it is clear that (ii) implies (i). Now suppose that (i) holds. Combining (36) and (38), we have that:

$$\begin{aligned} \sum _{m=-\infty }^\infty \sum _{n,n'=\left| {m}\right| }^\infty c_m(n,n')b_{n,m}b^*_{n',m}=0 \end{aligned}$$
(39)

where

$$\begin{aligned} b_{n,m}=\sum _{j=1}^k a_j e^{ {\mathsf {i}} m \ell _j}\bar{P}^m_n(\sin L_j ). \end{aligned}$$
(40)

Since \(c_m\) is positive definite (for every m), then

$$\begin{aligned} \sum _{n,n'=\left| {m}\right| }^N c_m(n,n')b_{n,m}b^*_{n',m}\ge 0, \end{aligned}$$

for every N and every m. Thus, letting N diverge to infinity we have that

$$\begin{aligned} \sum _{n,n'=\left| {m}\right| }^\infty c_m(n,n')b_{n,m}b^*_{n',m}\ge 0, \end{aligned}$$

for every m. Therefore, by (39), we have that for every \(m\in {\mathbb {Z}}\),

$$\begin{aligned} \sum _{n,n'=\left| {m}\right| }^\infty c_m(n,n')b_{n,m}b^*_{n',m}=0. \end{aligned}$$
(41)

At this stage, for each \(m,N\in {\mathbb {Z}}\) with \(m,N> 0\), let \(C_m(N)\) be the \(N\times N\) matrix whose \((n-m+1,n'-m+1)\) entry is \(c_m(n,n')\) (\(n,n'= m,\ldots , N+m-1\)).

For each \(m,N\in {\mathbb {Z}}\) with \(N, m>0\), the matrix \(C_m(N)\) must be positive definite since by assumption, for each \(m\in {\mathbb {Z}}\), \(c_m(n,n')\) is strictly positive definite. Hence, by Cholesky decomposition, there is a complex \(N\times N\) lower triangular matrix \(A_m(N)\) such that \(C_m(N)=A_m(N) A_m(N)^*\) the diagonal entries of \(A_m(N)\) are real and positive and the matrix \(A_0(N)\) is real. Denoting by \(a^{(m,N)}_{nr}\) the \((n-m+1,r-m+1)\) entry of the matrix \(A_m(N)\) if \(N,m> 0\), the Cholesky decomposition of the matrix \(C_m(N)\) can be written in the following form:

$$\begin{aligned} c_m(n,n')=\sum _{r=m}^{n \wedge n'} a^{(m,N)}_{nr} a^{(m,N)*}_{n'r}, \end{aligned}$$

for \(n,n'=m, \ldots , N+m-1\), where \(x\wedge y\) denotes the minimum among x and y. If \(M \ge N\), then by the Cholesky decomposition of the matrix \(C_m(M)\) we have that:

$$\begin{aligned} \sum _{r=m}^{n \wedge n'} a^{(m,M)}_{nr} a^{(m,M)*}_{n'r}=\sum _{r=m}^{n \wedge n'} a^{(m,N)}_{nr} a^{(m,N)*}_{n'r}, \end{aligned}$$

for \(n,n'=m, \ldots , N+m-1\). Note the Cholesky decomposition of \(C_m(N)\) is unique being \(C_m(N)\) positive definite and therefore \(a^{(m,N)}_{nr}=a^{(m,M)}_{nr}\) if \(m\le r \le n \le N+m-1 \le M+m-1\). In particular, we have that \(a^{(m,n-m+1)}_{nr}=a^{(m,M)}_{nr}\) for any quartet (mrnM) where \(m\le r \le n \le M-m+1\). Therefore, we can write \(a^{(m)}_{nr}\) for \(a^{(m,n-m+1)}_{nr}\) and

$$\begin{aligned} c_m(n,n')=\sum _{r=\left| {m}\right| }^{n \wedge n'} a^{(m)}_{nr} a^{(m)*}_{n'r}, \end{aligned}$$
(42)

where for each r, \(a^{(m)}_{rr}>0\). Note that (42) holds also for \(m<0\) if we let \(a^{(-m)}_{nr}=a^{(m)*}_{nr}\). Substituting (42) in (41) we obtain that, for every \(m\in {\mathbb {Z}}\),

$$\begin{aligned} \sum _{n,n'=\left| {m}\right| }^\infty \sum _{r=\left| {m}\right| }^{n \wedge n'} a^{(m)}_{nr} a^{(m)*}_{n'r}b_{n,m}b^*_{n',m}=0, \end{aligned}$$

that is equivalent to:

$$\begin{aligned} \sum _{r=\left| {m}\right| }^{\infty } \sum _{n,n'= r}^\infty a^{(m)}_{nr} a^{(m)*}_{n'r}b_{n,m}b^*_{n',m}=0, \end{aligned}$$

which in turn is equivalent to:

$$\begin{aligned} \sum _{r=\left| {m}\right| }^{\infty } \left| {\sum _{n=r}^\infty a^{(m)}_{nr} b_{n,m}}\right| ^2=0, \end{aligned}$$

which implies that for every \(r\ge \left| {m}\right|\)

$$\begin{aligned} \sum _{n=r}^\infty a^{(m)}_{nr} b_{n,m}=0. \end{aligned}$$
(43)

Thus, for every fixed \(m\in {\mathbb {Z}}\) and every fixed \(r\ge \left| {m}\right|\) we can write:

$$\begin{aligned} \sum _{n=r+s}^\infty a^{(m)}_{n,r+s} b_{n,m}=0, \end{aligned}$$
(44)

for \(s=0,1,2,\ldots\) Recalling that \(a^{(m)}_{r+s,r+s}>0\) for each s, we can define a sequence \((d_s)_{s=0}^\infty\) according to the recursion:

$$\begin{aligned} d_0=1,\, d_1=-\frac{a^{(m)}_{r+1,r}}{a^{(m)}_{r+1,r+1}}, \ldots , d_s=-\frac{1}{a^{(m)}_{r+s,r+s}}\sum _{l=0}^{s-1} d_l a^{(m)}_{r+s,r+l}, \ldots \end{aligned}$$

This ensures that:

$$\begin{aligned} \sum _{s=0}^{t} d_s a_{r+t,r+s}=0 \end{aligned}$$

for \(t=1,2,\ldots\) and therefore by (44),

$$\begin{aligned} 0&= \sum _{s=0}^\infty d_s \sum _{n=r+s}^\infty a^{(m)}_{n,r+s} b_{n,m} =\sum _{n=r}^\infty b_{n,m} \sum _{s=0}^{n-r} d_s a^{(m)}_{n,r+s}\\&= b_{r,m} d_0 a^{(m)}_{r,r} + \sum _{n=r+1}^\infty b_{n,m} \sum _{s=0}^{n-r} d_s a^{(m)}_{n,r+s}\\&= b_{r,m} a^{(m)}_{r,r} + \sum _{t=1}^\infty b_{r+t,m} \sum _{s=0}^{t} d_s a^{(m)}_{r+t,r+s} = b_{r,m} a^{(m)}_{r,r}. \end{aligned}$$

We have just proved that \(b_{r,m} a^{(m)}_{r,r}=0\) for any \(m\in {\mathbb {Z}}\) and any \(r\ge \left| {m}\right|\). At this stage, recall once again that \(a^{(m)}_{r,r}>0\) for any (mr). Therefore, for any \(m\in {\mathbb {Z}}\) and any \(r\ge \left| {m}\right|\), \(b_{r,m}=0\), namely by (40)

$$\begin{aligned} \sum _{j=1}^k a_j e^{ {\mathsf {i}} m \ell _j}\bar{P}^m_n(\sin L_j )= 0 \end{aligned}$$
(45)

for each \(m\in {\mathbb {Z}}\) and each \(n\in \{\left| {m}\right| ,\left| {m}\right| +1,\ldots \}\), proving (ii) . \(\square\)

Now we are ready to prove our main result.

1.3 Proof of Theorem 2

Proof

We aim at proving that if \((a_1,\ldots ,a_k)\in {\mathbb {R}}^k\), and \(\varvec{s}_1,\ldots ,\varvec{s}_k\) are k distinct points in \({\mathbb {S}}^2\) such that the longitude and the latitude of \(\varvec{s}_j\) are \(\ell _j\in (-\pi ,\pi ]\) and \(L_j\in [-\pi /2,\pi /2]\), respectively (\(j=1,\ldots ,k\)), and

$$\begin{aligned} \sum _{j,h=1}^k a_jK(L_j,L_h, \Delta \ell _{jh})a_h=0, \end{aligned}$$
(46)

then \(a_j=0\) for \(j=1,\ldots ,k\).

By Lemma 5, letting \(m=0\), we obtain:

$$\begin{aligned} \sum _{j=1}^k a_j \bar{P}_n(\sin (L_j))= 0 \end{aligned}$$
(47)

for every \(n\in \{0,1,2,\ldots \}\), where \(\bar{P}_n\) denotes the ordinary normalized Legendre polynomial of degree n.

By denoting \(L_{(1)},\ldots , L_{(h)}\) the distinct elements among \(L_1,\ldots ,L_k\) and letting \(a_{(r)}=\sum _{j\in B_r} a_j\) where \(B_r=\{j=1,\ldots , k: L_j=L_{(r)}\}\), as \(r=1,\ldots , h\), the equation (47) becomes:

$$\begin{aligned} \sum _{j=1}^h a_{(j)} \bar{P}_n(\sin L_{(j)})= 0 \end{aligned}$$
(48)

for every \(n\in \{0,1,2,\ldots \}\).

At this stage, consider the following function:

$$\begin{aligned} p(L)=\sum _{n\in E} d_{n} \bar{P}_n(\sin L ), \end{aligned}$$

where E is a finite or infinite subset of \(\{0,1,2,\ldots \}\), so that by (48),

$$\begin{aligned} \sum _{j=1}^h a_{(j)} p(L_{(j)})=0. \end{aligned}$$
(49)

There exist \(\{d_{n}\in {\mathbb {R}}: n\in E \}\) for some set E such that \(p(L_{(j)})=a_{(j)}\) for \(j=1,\ldots ,h\). Indeed, we can just take \(E=\{0,1,\ldots ,h-1\}\) recalling that \(\bar{P}_0\),..., \(\bar{P}_{h-1}\) generate the h–dimensional polynomial space and that interpolation of arbitrary data at h nodes is possible.

Hence, by (49), \(\textstyle \sum _{j=1}^h a_{(j)}^2=0\) and therefore \(a_{(j)}=0\), namely

$$\begin{aligned} \sum _{j\in B_r} a_j=0, \quad j=1,\ldots ,h. \end{aligned}$$
(50)

Now note that if \(L_j\in \{-\pi /2,\pi /2\}\) then \(\varvec{s}_j\) is a pole and therefore \(L_i\ne L_j\) for any \(i\ne j\) since \(\varvec{s}_1,\ldots , \varvec{s}_n\) are distinct points (i.e., \(B_r=\{j\}\) if \(L_j\in \{-\pi /2,\pi /2\}\)). Hence, by (50), we have just proved that if \(L_j\in \{-\pi /2,\pi /2\}\) then \(a_j=0\).

Hence , the equation (45) becomes:

$$\begin{aligned} \sum _{r=1}^h \sum _{j\in B_r} a_j e^{ {\mathsf {i}} m \ell _j} \bar{P}^m_n(\sin L_{(r)} )= 0, \end{aligned}$$
(51)

for each \(m\in {\mathbb {Z}}\) and each \(n\in \{\left| {m}\right| ,\left| {m}\right| +1,\ldots \}\). Since \(\overline{P}_n^m(\pm 1) =0\), (51) becomes:

$$\begin{aligned} \sum _{r\in A} \sum _{j\in B_r} a_j e^{{\mathsf {i}} m \ell _j}\bar{P}^m_n(\sin L_{(r)} )= 0 \end{aligned}$$
(52)

where \(A=\{r=1,\ldots ,h: L_{r}\notin \{-\pi /2,\pi /2\}\}\), for each \(m\in {\mathbb {Z}}\) and each \(n\in \{\left| {m}\right| ,\left| {m}\right| +1,\ldots \}\). Moreover, since we have just proved that \(a_j=0\) if \(L_j=\pm \pi /2\), then we need to show that \(a_j=0\) for all \(j\in B_r\) and for all \(r\in A\).

By Lemma 4, (52) implies that:

$$\begin{aligned} \sum _{j\in B_r} a_j e^{{\mathsf {i}} m \ell _j}=0 \end{aligned}$$
(53)

for every \(m\in {\mathbb {Z}}\) and \(r\in A\). At this stage, observe that \(\ell _i \ne \ell _j\) when \(i,j\in B_r\) because in this case \(L_i=L_j\) and \(\varvec{s}_i\ne \varvec{s}_j\).

Moreover, the matrix of the system (53) is Vandermonde–like associated to the distinct points \(e^{{\mathsf {i}} \ell _j}\), \(j\in B_r\) for each \(r\in A\). Therefore, (53) implies that \(a_j=0\) for each \(j=1,\ldots ,k\) such that \(L_j\notin \{-\pi /2,\pi /2\}\) and the proof is complete. \(\square\)

1.4 Proof of Theorem 3

Proof

In order to evaluate the integral in the right hand side of (19), we can plug in (17) and swap the double series and the triple integral if the following is verified:

$$\begin{aligned}&\int _{-\pi /2}^{\pi /2}\int _{-\pi /2}^{\pi /2} \int _{-\pi }^\pi \sum _{n,n'=0}^\infty \sum _{m=-(n\wedge n')}^{n\wedge n'} \\&\quad g^{(m,n,n')}_{m_0,n_0,n'_0}(\Delta \ell ,L,L^{\prime }) {\mathrm {d}}\Delta \ell {\mathrm {d}}L {\mathrm {d}}L^{\prime } <\infty , \end{aligned}$$
(54)

where:

$$\begin{aligned}&g^{(m,n,n')}_{m_0,n_0,n'_0}(\Delta \ell ,L,L^{\prime })\\&\quad =\left| {e^{im \Delta \ell } \bar{P}^m_{n}(\sin L)\bar{P}^m_{n'}(\sin L^{\prime })c_m(n,n') e^{- {\mathsf {i}} m_0 \Delta \ell }\bar{P}^{m_0}_{n_0}(\sin L) \bar{P}^{m_0}_{n'_0}(\sin L^{\prime }) \cos L\cos L^{\prime }}\right| \end{aligned}$$

for every \(m_0\in {\mathbb {Z}}\) and \(n_0,n_0'\ge \left| {m_0}\right|\). Trivially,

$$\begin{aligned}&g^{(m,n,n')}_{m_0,n_0,n'_0}(\Delta \ell ,L,L^{\prime }) \\&\quad \le \left| {\bar{P}^m_{n}(\sin L)\bar{P}^m_{n'}(\sin L^{\prime })c_m(n,n') \bar{P}^{m_0}_{n_0}(\sin L) \bar{P}^{m_0}_{n'_0}(\sin L^{\prime })}\right| . \end{aligned}$$
(55)

At this stage, note that by (12),

$$\begin{aligned} \left| {\bar{P}^{m}_{n}(\sin L)}\right| \le \sqrt{n+\frac{1}{2}}, \end{aligned}$$

for every \(m,n\in {\mathbb {Z}}\) with \(n\ge \left| {m}\right|\), and therefore (55) yields:

$$\begin{aligned}&g^{(m,n,n')}_{m_0,n_0,n'_0}(\Delta \ell ,L,L^{\prime })\\&\quad \le \left| {\bar{P}^m_{n}(\sin L)\bar{P}^m_{n'}(\sin L^{\prime })c_m(n,n') \sqrt{(n_0+1/2)(n'_0+1/2)}}\right| , \end{aligned}$$

which, by the Cauchy–Schwartz inequality and (12), in turn yields:

$$\begin{aligned}&\sum _{m=-( n\wedge n')}^{n\wedge n'} g^{(m,n,n')}_{m_0,n_0,n'_0}(\Delta \ell ,L,L^{\prime })\\&\quad \le \sqrt{\left( n_0+\frac{1}{2}\right) \left( n'_0+\frac{1}{2}\right) \left( n+\frac{1}{2}\right) \left( n'+\frac{1}{2}\right) } \\&\qquad \times \sup _{m\in \{-(n\wedge n'), \ldots , n\wedge n'\}} \left| {c_m(n,n')}\right| , \end{aligned}$$

which by (18) implies that (54) is verified. It is therefore possible to plug (17) into the integral in (19) and swap the double sum with the triple integral. In this way, due to the orthogonality of the associated Legendre polynomials \(\{\bar{P}^m_{n}(\sin L): n\in {\mathbb {N}}\}\) given by (6) and of \(\{e^{im \ell }: m\in {\mathbb {Z}}\}\), and due to (7), (19) is obtained. \(\square\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bissiri, P.G., Peron, A.P. & Porcu, E. Strict positive definiteness under axial symmetry on the sphere. Stoch Environ Res Risk Assess 34, 723–732 (2020). https://doi.org/10.1007/s00477-020-01796-y

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00477-020-01796-y

Keywords

Navigation