1 Introduction

Consider a Borel probability measure \(\mu :\mathscr {B}(\mathbb {R}^d)\rightarrow [0,1]\) on \(\mathbb {R}^d\), where \(\mathscr {B}(\mathbb {R}^d)\) denotes the Borel sigma algebra on \(\mathbb {R}^d\). For fixed \(n\in \mathbb {N}\), we aim to allocate a suitable n-point set \(\{x_1,\ldots ,x_n\}\subset \mathbb {R}^d\) such that the normalized atomic measure

$$\begin{aligned} \nu _n:=\frac{1}{n}\sum _{j=1}^n\delta _{x_j} \end{aligned}$$
(1.1)

approximates \(\mu \). Here, \(\delta _{x_j}:\mathscr {B}(\mathbb {R}^d)\rightarrow \{0,1\}\) denotes the point measure localized at \(x_j\). To quantify the \(L_2\)-discrepancy between \(\mu \) and \(\nu _n\), select a measure \(\beta \) on \(\mathscr {B}(\mathbb {R}^d)\) with \(\mu ,\delta _x\in L_2(\mathscr {B}(\mathbb {R}^d),\beta )\), for all \(x\in \mathbb {R}^d\), and consider

$$\begin{aligned} \mathscr {D}_{\beta }(\mu ,\nu _n):=\Vert \mu -\nu _n\Vert ^2_{L_2(\mathscr {B}(\mathbb {R}^d),\beta )} =\int _{\mathscr {B}(\mathbb {R}^d)} \left| \mu (B)- \nu _n(B)\right| ^2\textrm{d} \beta (B), \end{aligned}$$
(1.2)

cf. [40, 42, 43], see Sect. 2 for explicit examplesFootnote 1. For fixed \(n\in \mathbb {N}\), we aim to minimize \(\mathscr {D}_{\beta }(\mu ,\nu _n)\) among all n-point sets \(\{x_1,\ldots ,x_n\}\subset \mathbb {R}^d\). The present manuscript is concerned with discretizations of (1.2) that facilitate numerical minimization.

The associated discrepancy kernel \(K_\beta :\mathbb {R}^d\times \mathbb {R}^d\rightarrow \mathbb {R}\) is defined by

$$\begin{aligned} K_\beta (x,y):= \langle \delta _x,\delta _y\rangle _{L_2(\mathscr {B}(\mathbb {R}^d),\beta )} =\int _{\mathscr {B}(\mathbb {R}^d)} \delta _x(B)\delta _y(B) \textrm{d} \beta (B), \end{aligned}$$
(1.3)

and we assume it is continuous. Fubini’s Theorem and \(\mu (B)=\int _{\mathbb {R}^d} \delta _x(B)\textrm{d}\mu (x)\) applied to

$$\begin{aligned} \Vert \mu -\nu _n\Vert ^2_{L_2(\mathscr {B}(\mathbb {R}^d),\beta )}=\Vert \mu \Vert ^2_{L_2(\mathscr {B}(\mathbb {R}^d),\beta )}-2\langle \mu ,\nu _n\rangle _{L_2(\mathscr {B}(\mathbb {R}^d),\beta )}+\Vert \nu _n\Vert ^2_{L_2(\mathscr {B}(\mathbb {R}^d),\beta )} \end{aligned}$$

yield that (1.2) is identical to

$$\begin{aligned} \mathscr {D}_{\beta }(\mu ,\nu _n)= & {} \iint \limits _{\mathbb {R}^d\times \mathbb {R}^d} K_{\beta }(x,y)\textrm{d}\mu (x) \textrm{d}\mu (y) \nonumber \\{} & {} -2\sum _{j=1}^n\int _{\mathbb {R}^d}\frac{K_{\beta }(x,x_j)}{n}\textrm{d}\mu (x) +\sum _{i,j=1}^n \frac{K_{\beta }(x_i,x_j)}{n^2} . \end{aligned}$$
(1.4)

If a compact set \(\mathbb {X}\subset \mathbb {R}^d\) is known in advance such that \(\text {supp}(\mu )\subset \mathbb {X}\), then we shall restrict the minimization to \(\{x_1,\ldots ,x_n\}\subset \mathbb {X}\), so that only the restricted kernel \(K_\beta |_{\mathbb {X}\times \mathbb {X}}\) matters. By endowing \(\mathbb {X}\) with a finite Borel measure \(\sigma _\mathbb {X}\) having full support, Mercer’s Theorem yields an orthonormal basis \(\{\phi _l\}_{l=0}^\infty \) for \(L_2(\mathbb {X},\sigma _\mathbb {X})\) and coefficients \((a_l)_{l=0}^\infty \) such that the spectral decomposition

$$\begin{aligned} K_\beta |_{\mathbb {X}\times \mathbb {X}}(x,y) = \sum _{l=0}^\infty a_l \phi _l(x)\overline{\phi _l(y)},\quad x,y\in \mathbb {X}, \end{aligned}$$
(1.5)

holds with absolute and uniform convergence. We call \((a_l)_{l=0}^\infty \) the Fourier coefficients of the kernel \(K_\beta |_{\mathbb {X}\times \mathbb {X}}\). If \(\text {supp}(\mu ),\text {supp}(\nu _n)\subset \mathbb {X}\), then the Fourier expansion of the \(L_2\)-discrepancy (1.4) is

$$\begin{aligned}{} & {} \mathscr {D}_{\beta }(\mu ,\nu _n) = \sum _{l=0}^\infty a_l \left| \hat{\mu }_{l}-\hat{\nu }_{n,l} \right| ^2, \nonumber \\{} & {} \hat{\mu }_l{:}{=}\int _{\mathbb {X}}\overline{ \phi _l(x)}\textrm{d}\mu (x),\quad \hat{\nu }_{n,l}:=\frac{1}{n}\sum _{j=1}^n\overline{\phi _{l}(x_j)}, \end{aligned}$$
(1.6)

where the Fourier coefficients \(\hat{\mu }_l\) and \(\hat{\nu }_{n,l}\) of the measures \(\mu \) and \(\nu _n\), respectively, are well-defined if \(a_l\ne 0\). Truncation of the discretization (1.6) enables the use of the nonequispaced fast Fourier transform, thereby offering more efficient minimization of \(\mathscr {D}_\beta (\mu ,\nu _n)\), cf. [31, 33]. Thus, we aim to

A) compute \((a_l)_{l=0}^\infty \) and \((\phi _l)_{l=0}^\infty \) in the Fourier expansion (1.5) of \(K_\beta |_{\mathbb {X}\times \mathbb {X}}\).

The \(L_2\)-discrepancy \(\mathscr {D}_\beta (\mu ,\nu _n)\) also coincides with the worst case integration error

$$\begin{aligned} \mathscr {D}_\beta (\mu ,\nu _n) = \sup _{\Vert f\Vert _{\mathscr {H}_\beta (\mathbb {X})}\le 1 }\left| \int _\mathbb {X}f(x)\textrm{d}\mu (x) - \frac{1}{n}\sum _{j=1}^n f(x_j) \right| ^2 \end{aligned}$$
(1.7)

with respect to the reproducing kernel Hilbert space \(\mathscr {H}_\beta (\mathbb {X})\) generated by \(K_\beta |_{\mathbb {X}\times \mathbb {X}}\), cf. [12, 13, 30, 31]. To specify \(\mathscr {H}_\beta (\mathbb {X})\), we aim to

B) identify \(\mathscr {H}_\beta (\mathbb {X})\) with a classical function space.

Fourier decay properties generally quantify Sobolev smoothness. To accomplish (B), we aim to determine the asymptotics of \(K_\beta |_{\mathbb {X}\times \mathbb {X}}\)’s Fourier coefficients \((a_l)_{l=0}^\infty \) in (1.5).

For \(\mathbb {X}=\mathbb {S}^{d-1}\) and a particular choice of \(\beta \), the kernel \(K_{\beta }|_{\mathbb {S}^{d-1}\times \mathbb {S}^{d-1}}\) essentially coincides with the Euclidean distance, see [12, 13]. The Fourier expansion is determined in [10], and the decay of the Fourier coefficients yields that \(K_{\beta }|_{\mathbb {S}^{d-1}\times \mathbb {S}^{d-1}}\) reproduces the Sobolev space \(\mathscr {H}_\beta (\mathbb {S}^{d-1})=\mathbb {H}^{\frac{d}{2}}(\mathbb {S}^{d-1})\). For the sphere and the torus, the nonequispaced fast Fourier transform is available, and both (A) and (B) are discussed in [33, 34].

This manuscript is dedicated to derive analogous results for other compact sets \(\mathbb {X}\). We focus on the unit ball, the special orthogonal group, and the Grassmannian manifold,

$$\begin{aligned} \mathbb {B}^d&:=\{x\in \mathbb {R}^d:\Vert x\Vert \le 1\},\\ \textrm{SO}(d)&:=\{x\in \mathbb {R}^{d\times d} : \det (x)=1,\, x^{-1}=x^\top \},\\ \mathcal {G}_{k,d}&:=\{x\in \mathbb {R}^{d\times d} : x^\top =x,\, x^2=x,\, \textrm{trace}(x)=k\}. \end{aligned}$$

We achieve goal (A) for \(\mathbb {X}=\mathbb {B}^d\) with odd d. Both goals, (A) and (B), are achieved for \(\textrm{SO}(3)\) and \(\mathcal {G}_{2,4}\). We also provide numerical experiments. For \(\textrm{SO}(3)\), the computations are based on the nonequispaced fast Fourier transform designed in [32, 44]. For \(\mathcal {G}_{2,4}\), we derive the nonequispaced fast Fourier transform by parametrization via the double covering \(\mathbb {S}^2\times \mathbb {S}^2\) and developing the respective transform there.

For \(\textrm{SO}(d)\) and \(\mathcal {G}_{k,d}\) with fixed k and d, in principle, one could still be able to compute the Fourier expansion in goal (A). However, one may be faced with rather complicated expressions. In our present computations for \(\textrm{SO}(3)\) and \(\mathcal {G}_{2,4}\), the structural relations to the unit sphere enabled the use of Chebychev and Legendre polynomials, which reduced the complexity. Nonetheless, we do accomplish (B) for the general cases \(\textrm{SO}(d)\) and \(\mathcal {G}_{k,d}\).

2 Two Introductory Examples

We first present a well-known elementary example on the interval [0, s], for which both aims A) and B) are achieved. Second, to support our perspective on discrepancy, we prove that the so-called Askey function is a discrepancy kernel of the form (1.3).

2.1 The Brownian Motion Kernel on [0, s]

Let \(\textrm{d}r\) be the Lebesgue measure on \([0,\infty )\). The mapping \( h:[0,\infty )\rightarrow \mathscr {B}(\mathbb {R})\) defined by \(r\mapsto [r,\infty ) \) induces the pushforward measure \(\beta :=h_*(\textrm{d}r)\) that induces the discrepancy

$$\begin{aligned} \mathscr {D}_\beta (\mu ,\nu _n)= \int _{0}^\infty |\mu ([r,\infty ))-\nu _n([r,\infty ))|^2 \textrm{d}r. \end{aligned}$$

The associated discrepancy kernel \(K_\beta :\mathbb {R}\times \mathbb {R}\rightarrow \mathbb {R}\) isFootnote 2

$$\begin{aligned} K_\beta (x,y) = \int _0^\infty \delta _x([r,\infty ))\delta _y([r,\infty )) \textrm{d}r =\min (x,y)_+, \end{aligned}$$

so that \(\mathscr {D}_\beta (\delta _x,\delta _y)=|x-y|\) for \(x,y\in [0,\infty )\). The restriction of the kernel \(K_\beta \) to \([0,s]\times [0,s]\) has the Fourier expansion

$$\begin{aligned} K_\beta (x,y)&=\sum _{\begin{array}{c} m\in \mathbb {N}\\ m \text { odd} \end{array}} \frac{4s^2}{m^2\pi ^2}\cdot \frac{\sin (\frac{\pi }{2s} mx)}{\sqrt{\frac{s}{2}}}\cdot \frac{\sin (\frac{\pi }{2s} my)}{\sqrt{\frac{s}{2}}},\quad x,y\in [0,s], \end{aligned}$$

with respect to the Lebesgue measure \(\sigma _{[0,s]}\) on [0, s]. The reproducing kernel Hilbert space is

$$\begin{aligned} \mathscr {H}_{\beta }([0,s])=\{f:[0,s]\rightarrow \mathbb {C}\;:\; f \text { is absolutely continuous, } f(0)=0,\; f'\in L_2([0,s])\}, \end{aligned}$$

where the inner product between f and g is given by \(\langle f',g'\rangle _{L_2([0,s])}\), cf. [3, 22] and [43, Section 9.5.5]. Note that \(K_\beta |_{[0,1]\times [0,1]}\) is often called the Brownian motion kernel and \(\mathscr {H}_{\beta }([0,s])\) is continuously embedded into the Sobolev space \(\mathbb {H}^1([0,s])\).

2.2 Askey’s Function and Its Restrictions

Many positive definite kernels in the literature are of the form (1.3) and, hence, are discrepancy kernels. For odd d, Askey’s kernel function \((x,y)\mapsto (1-\Vert x-y\Vert )^{\frac{d+1}{2}}_+\) is positive definite, cf. [29]. In the following, we shall check that it is of the form (1.3).

Denote the Euclidean ball of radius s centered at \(z\in \mathbb {R}^d\) by

$$\begin{aligned} \mathbb {B}^d_{s}(z):=\{x\in \mathbb {R}^d:\Vert x-z\Vert \le {s}\}, \end{aligned}$$

with the conventions \(\mathbb {B}^d_{s}:=\mathbb {B}^d_{s}(0)\) and \(\mathbb {B}^d:=\mathbb {B}^d_1\). Fix \(r>0\) and consider the discrepancy

$$\begin{aligned} \mathscr {D}_{d,r}(\mu ,\nu _n):=\frac{1}{\textrm{vol}(\mathbb {B}^d_{\frac{r}{2}})} \int _{\mathbb {R}^d} \left| \mu (\mathbb {B}^d_{\frac{r}{2}}(z))- \nu _n(\mathbb {B}^d_{\frac{r}{2}}(z))\right| ^2\textrm{d} z, \end{aligned}$$
(2.1)

where \(\textrm{vol}(\mathbb {B}^d_{\frac{r}{2}})=\frac{\pi ^{d/2}}{\Gamma (\frac{d}{2}+1)}(\frac{r}{2})^d\). The associated discrepancy kernel is

$$\begin{aligned} K_{d,r}(x,y)=\frac{1}{\textrm{vol}(\mathbb {B}^d_{\frac{r}{2}})}\int _{\mathbb {R}^d} \delta _x(\mathbb {B}^d_{\frac{r}{2}}(z))\delta _y(\mathbb {B}^d_{\frac{r}{2}}(z)) \textrm{d} z.\nonumber \\ \end{aligned}$$
(2.2)

In order to additionally integrate over r, recall the (generalized) hypergeometric functions

where \(f_1,\ldots ,f_k,g_1,\ldots ,g_l,z\in \mathbb {R}\) and \((f)_n:=f\cdot (f+1)\cdots (f+n-1)\) denotes the Pochhammer symbol with \((f)_0:=1\). We consider \(G_d:[0,\infty )\rightarrow \mathbb {R}\) given by

Since d is odd, either \(\frac{d+1}{4}\) or \(\frac{d-1}{4}\) is a natural number, so that the series terminates and \(G_d\) is a polynomial in \(r^{2}\) on [0, 1]. By integration with respect to \(G_d\), we obtain the \(L_2\)-discrepancy and the associated discrepancy kernel

$$\begin{aligned} \mathscr {D}_{d}(\mu ,\nu _n):= \int _0^\infty \!\!\mathscr {D}_{d,r}(\mu ,\nu _n)\textrm{d} G_d(r),\quad \text { and }\quad K_{d}(x,y)= \int _0^\infty \!K_{d,r}(x,y) \textrm{d} G_d(r), \end{aligned}$$

respectively. It turns out that \(K_d\) coincides with Askey’s function.

Theorem 2.1

Let d be odd. The discrepancy kernel \(K_d\) satisfies

$$\begin{aligned} K_d(x,y)=(1-\Vert x-y\Vert )^{\frac{d+1}{2}}_+,\quad x,y\in \mathbb {R}^d. \end{aligned}$$
(2.3)

The proof is presented in Appendix A. Provided that \(d\ge 3\), Askey’s kernel function reproduces the Sobolev space \(\mathbb {H}^{\frac{d+1}{2}}(\mathbb {R}^d)\) with an equivalent norm, see [49].

3 The Distance Kernel on \(\mathbb {S}^{d-1}\)

This section is dedicated to recall results on discrepancy kernels on the sphere \(\mathbb {S}^{d-1}\subset \mathbb {R}^d\), for \(d\ge 2\), from [12, 13, 31, 46] that shall guide our subsequent investigations.

Denote the geodesic ball of radius r centered at \(z\in \mathbb {S}^{d-1}\) by

$$\begin{aligned} B^{\mathbb {S}^{d-1}}_r(z):= \{x\in \mathbb {S}^{d-1} : \textrm{dist}_{\mathbb {S}^{d-1}}(x,z)\le r\}, \end{aligned}$$

where \(\textrm{dist}_{\mathbb {S}^{d-1}}(x,z)=\arccos (\langle x,z\rangle )\) is the geodesic distance on \(\mathbb {S}^{d-1}\). We define

$$\begin{aligned} h:[0,\pi ]\times \mathbb {S}^{d-1}\rightarrow \mathscr {B}(\mathbb {R}^d),\qquad (r,z)\mapsto B^{\mathbb {S}^{d-1}}_r(z) \end{aligned}$$

and endow \([0,\pi ]\) with the weighted Lebesgue measure \(\sin (r){\mathrm d}r\), whereas \(\mathbb {S}^{d-1}\) carries the normalized, orthogonal invariant surface measure \(\sigma _{\mathbb {S}^{d-1}}\). The push-forward \(\beta _d:=h_*(\sin (r){\mathrm d}r \otimes \sigma _{\mathbb {S}^{d-1}})\) is a measure on \(\mathscr {B}(\mathbb {R}^d)\), so that the associated \(L_2\)-discrepancy is

$$\begin{aligned} \mathscr {D}_{\beta _d}(\mu ,\nu _n)= \int _{0}^\pi \int _{\mathbb {S}^{d-1}} |\mu (B^{\mathbb {S}^{d-1}}_r(z)) -\nu _n(B^{\mathbb {S}^{d-1}}_r(z))|^2 \textrm{d}\sigma _{\mathbb {S}^{d-1}}(z)\sin (r)\textrm{d}r. \end{aligned}$$

The associated discrepancy kernel is

$$\begin{aligned} K_{\beta _d}(x,y)=\int _{0}^\pi \int _{\mathbb {S}^{d-1}}\delta _x(B^{\mathbb {S}^{d-1}}_r(z))\delta _y(B^{\mathbb {S}^{d-1}}_r(z)) \textrm{d}\sigma _{\mathbb {S}^{d-1}}(z)\sin (r)\textrm{d}r,\quad x,y\in \mathbb {R}^{d}.\nonumber \\ \end{aligned}$$
(3.1)

According to [12, 13, 31], see also [2], \(K_{\beta _d}\) satisfies

$$\begin{aligned} K_{\beta _d}(x,y)=1-\frac{\Gamma (\frac{d}{2})}{2\sqrt{\pi }\Gamma (\frac{d+1}{2})}\Vert x-y\Vert ,\qquad x,y\in \mathbb {S}^{d-1}. \end{aligned}$$
(3.2)

If either x or y is not contained in \(\mathbb {S}^{d-1}\), then \(K_{\beta _d}(x,y)=0\).

Choose \(\sigma _\mathbb {X}:=\sigma _{\mathbb {S}^{d-1}}\) for the decomposition (1.5) and let \(\{Y^m_{{l}} : l=1,\ldots ,Z(d,m)\}\subset L_2(\mathbb {S}^{d-1},\sigma _{\mathbb {S}^{d-1}})\) denote the set of orthonormal spherical harmonics of degree m on \(\mathbb {S}^{d-1}\), where \( Z(d,m)=\frac{2m+d-2}{d-2}\left( {\begin{array}{c}m+d-3\\ m\end{array}}\right) . \) For \(\tau >(d-1)/2\), the Sobolev space \(\mathbb {H}^\tau (\mathbb {S}^{d-1})\) is the reproducing kernel Hilbert space associated with the reproducing kernel

$$\begin{aligned} (x,y)\mapsto \sum _{m=0}^\infty (1+m(m+d-2))^{-\tau }\sum _{l=1}^{Z(d,m)}Y^m_{l}(x)\overline{Y^m_{l}(y)},\quad x,y\in \mathbb {S}^{d-1}. \end{aligned}$$
(3.3)

The coefficients in the Fourier expansion

$$\begin{aligned} 1-\frac{\Gamma (\frac{d}{2})}{2\sqrt{\pi }\Gamma (\frac{d+1}{2})}\Vert x-y\Vert = \sum _{m=0}^\infty c_m \sum _{l=1}^{Z(d,m)}Y^m_{l}(x)\overline{Y^m_{l}(y)},\quad x,y\in \mathbb {S}^{d-1}, \end{aligned}$$

satisfy \(|c_m|\sim m^{-d}\), cf. [12]. This is the same asymptotics as the coefficients in (3.3) for \(\tau =d/2\). Therefore, \(K_{\beta _d}|_{\mathbb {S}^{d-1}\times \mathbb {S}^{d-1}}\) reproduces the Sobolev space \(\mathscr {H}_{\beta _d}(\mathbb {S}^{d-1})=\mathbb {H}^{\frac{d}{2}}(\mathbb {S}^{d-1})\) with an equivalent normFootnote 3.

In order to determine the Fourier coefficients of kernels on the sphere that are polynomial in \(\Vert x-y\Vert \), such as \(K_{\beta _d}|_{\mathbb {S}^{d-1}\times \mathbb {S}^{d-1}}\), we require the Fourier coefficients of the monomial terms \( \Vert x-y\Vert ^p\). For any \(p\in \mathbb {N}\), the Fourier expansion

$$\begin{aligned} 2^{-\frac{p}{2}}\Vert x-y\Vert ^p = \sum _{m=0}^\infty a_m(p,\mathbb {S}^{d-1}) \sum _{l=1}^{Z(d,m)}Y^m_{l}(x) \overline{Y^m_{l}(y)},\quad x,y\in \mathbb {S}^{d-1}, \end{aligned}$$
(3.4)

holds with coefficients determined by

$$\begin{aligned} a_m(p,\mathbb {S}^{d-1}):=\frac{1}{Z(d,m)}\iint \limits _{\mathbb {S}^{d-1}\times \mathbb {S}^{d-1}} 2^{-\frac{p}{2}} \Vert x-y\Vert ^p \sum _{l=1}^{Z(d,m)}\overline{Y^m_{l}(x)}Y^m_{l}(y) \textrm{d}\sigma _{\mathbb {S}^{d-1}}(x)\textrm{d}\sigma _{\mathbb {S}^{d-1}}(y).\nonumber \\ \end{aligned}$$
(3.5)

Note that (3.5) is well-defined for the entire range \(p>-(d-1)\) and p is not required to be an integer. For \(p>0\), the following proposition is essentially due to [10], see also [12, 14]. Simple continuation arguments cover the full range of p, and the asymptotics \(\frac{\Gamma (-\frac{p}{2}+m)}{\Gamma (\frac{p}{2}+d-1+m)}=m^{-(p+d-1)}(1+o(1))\) are standard.

Proposition 3.1

( [10]) Suppose \(d\ge 2\). For any \(p >-(d-1)\), we have

$$\begin{aligned} a_m(p,\mathbb {S}^{d-1})&=\frac{2^{d} \Gamma (\frac{d}{2})}{4\sqrt{\pi }}\cdot \frac{2^{p/2} \Gamma (\frac{d}{2}+\frac{p}{2}-\frac{1}{2})}{\Gamma (-\frac{p}{2})}\cdot \frac{\Gamma (-\frac{p}{2}+m)}{\Gamma (\frac{p}{2}+d-1+m)}. \end{aligned}$$
(3.6)

In particular, if \(p\not \in 2\mathbb {N}\), then

$$\begin{aligned} |a_m(p,\mathbb {S}^{d-1})|&=\left| \frac{2^{d} \Gamma (\frac{d}{2})}{4\sqrt{\pi }} \cdot \frac{2^{p/2} \Gamma (\frac{d}{2}+\frac{p}{2}-\frac{1}{2})}{\Gamma (-\frac{p}{2})}\right| m^{-(p+d-1)}(1+o(1)), \end{aligned}$$
(3.7)

and the series (3.4) terminates if \(p\in 2\mathbb {N}\).

For \(p\in 2\mathbb {N}\), the term \(\Gamma (-\frac{p}{2})\) is not well-defined and (3.6) is to be understood with the convention \(\frac{\Gamma (-\frac{p}{2}+m)}{\Gamma (-\frac{p}{2})}=(-\frac{p}{2})_m\). Hence, we observe \(a_m(p,\mathbb {S}^{d-1}) = 0\) for all \(m > p/2\) if \(p\in 2\mathbb {N}\).

It is noteworthy that the kernel \(K_{d,r}\) in (2.2) for \(d=3\) is a discrepancy kernel that does not generate a Sobolev space on \(\mathbb {R}^d\) but its restriction does. The proof of the following proposition is presented in Appendix B.

Proposition 3.2

Let \(r\ge 1\). The reproducing kernel Hilbert space of \(K_{3,r}\), given by (2.2) with \(d=3\), is continuously embedded into \(\mathbb {H}^{2}(\mathbb {R}^3)\), but the reverse embedding does not hold. In contrast, \(K_{3,r}|_{\mathbb {S}^{2}\times \mathbb {S}^{2}}\) reproduces \(\mathbb {H}^{\frac{3}{2}}(\mathbb {S}^{2})\) with an equivalent norm.

To provide numerical examples for \(d=3\), Proposition 3.1 provides the coefficients \((a_m)_{m=0}^\infty \) in the kernel expansion of \(K_{\beta _{3}}\),

$$\begin{aligned} 1-\frac{1}{4}\Vert x-y\Vert = \sum _{m=0}^\infty a_m \sum _{l=1}^{2m+1}Y^m_{l}(x)\overline{Y^m_{l}(y)},\quad x,y\in \mathbb {S}^2. \end{aligned}$$

For \(\text {supp}(\mu ),\text {supp}(\nu _n)\subset \mathbb {S}^2\), the \(L_2\)-discrepancy (1.6) for \(K_{\beta _{3}}\) with \(\mathbb {X}=\mathbb {S}^2\) becomes

$$\begin{aligned} \mathscr {D}_{\beta _{3}}(\mu ,\nu _n) = \sum _{m=0}^\infty a_m \sum _{l=1}^{2m+1} \left| \hat{\mu }^m_{l}- \frac{1}{n}\sum _{j=1}^n\overline{Y^m_{l}(x_j)}\right| ^2, \end{aligned}$$
(3.8)

where \(\hat{\mu }^m_{l}\) denotes the Fourier coefficient of \(\mu \) with respect to \(Y^m_{l}\), cf. (1.6). By truncating this series, the nonequispaced fast Fourier transform on \(\mathbb {S}^2\), cf. [33, 39, 41], enables efficient minimization of

$$\begin{aligned} \sum _{m=0}^M a_m \sum _{l=1}^{2m+1} \left| \hat{\mu }^m_{l}- \frac{1}{n}\sum _{j=1}^n\overline{Y^m_{l}(x_j)}\right| ^2 \end{aligned}$$
(3.9)

among all n-point sets \(\{x_1,\ldots ,x_n\}\subset \mathbb {S}^2\) for fixed n. We are most interested in \(n\gg M\). See Figure 1 for a numerical experiment with \(M=8\) and \(n=50\).

Fig. 1
figure 1

The target measure \(\mu \) is supported on two circles on the sphere \(\mathbb {S}^2\) with weight ratio 9/1. Numerical minimization of (3.9) splits 50 points into 45 points equally distributed on one and 5 points on the other circle

4 Discrepancy Kernels on Compact Sets

Here we discuss discrepancy kernels that extend the kernels of the previous section in a natural way. For \(d\ge 1\), let us define the half-space

$$\begin{aligned} \Omega ^d_r(z):=\{x\in \mathbb {R}^d : \langle z,x\rangle \ge r \}\in \mathscr {B}(\mathbb {R}^d),\quad z\in \mathbb {S}^{d-1},\quad r\in \mathbb {R}. \end{aligned}$$

For fixed \(s>0\), we consider the mapping \( h:[-s,s]\times \mathbb {S}^{d-1}\rightarrow \mathscr {B}(\mathbb {R}^d)\) defined by \((r,z)\mapsto \Omega ^d_r(z) \) and endow \([-s,s]\) with the Lebesgue measure \({\mathrm d}r\). The push-forward measure \(\beta _{d,s}:=h_*({\mathrm d}r\otimes \sigma _{\mathbb {S}^{d-1}})\) leads to the associated \(L_2\)-discrepancy

$$\begin{aligned} \mathscr {D}_{\beta _{d,s}}(\mu ,\nu _n)=\int _{-s}^{s} \int _{\mathbb {S}^{d-1}} \left| \mu (\Omega ^d_r(z)) - \nu _n(\Omega ^d_r(z))\right| ^2\textrm{d} \sigma _{\mathbb {S}^{d-1}}(z) \textrm{d} r. \end{aligned}$$

The associated discrepancy kernel is

$$\begin{aligned} K_{\beta _{d,s}}(x,y)=\int _{-s}^{s} \int _{\mathbb {S}^{d-1}} \delta _x(\Omega ^d_r(z))\delta _y(\Omega ^d_r(z)) \textrm{d} \sigma _{\mathbb {S}^{d-1}}(z) \textrm{d} r,\qquad x,y\in \mathbb {R}^d. \end{aligned}$$
(4.1)

Since \(B^{\mathbb {S}^{d-1}}_r(z) = \Omega ^d_{\cos (r)}(z)\cap \mathbb {S}^{d-1}\), for \(r\in [0,\pi ]\) and \(z\in \mathbb {S}^{d-1}\), we deduce

$$\begin{aligned} K_{\beta _d}|_{\mathbb {S}^{d-1}\times \mathbb {S}^{d-1}}=K_{\beta _{d,1}}|_{\mathbb {S}^{d-1}\times \mathbb {S}^{d-1}},\qquad d\ge 2, \end{aligned}$$

with \(K_{\beta _d}\) as in (3.1). In contrast to \(K_{\beta _d}\), the kernel \(K_{\beta _{d,s}}\) is not identically zero outside of \(\mathbb {S}^{d-1}\times \mathbb {S}^{d-1}\) and makes also sense for \(d=1\).

Example 4.1

For \(d=1\), we have \(\mathbb {S}^0=\{\pm 1\}\), so that the half-spaces are \(\Omega ^1_r(1)=[r,\infty )\) and \(\Omega ^1_r(-1)=(-\infty ,-r]\). Direct calculation of (4.1) yields

$$\begin{aligned} K_{\beta _{1,s}}(x,y)= {\left\{ \begin{array}{ll} s-\frac{1}{2}|x-y|,&{} |x|,|y|\le s,\\ \frac{s}{2}+\frac{xy}{2|y|},&{} |x|\le s\le |y|,\\ s\textrm{H}(xy), &{} s\le |x|,|y|, \end{array}\right. } \end{aligned}$$

where \(\textrm{H}\) is the Heaviside step function.

Proposition 4.2

The Fourier expansion of the kernel \(K_{\beta _{1,s}}|_{[-s,s]\times [-s,s]}\) with respect to the Lebesgue measure \(\sigma _{[-s,s]}\) on \([-s,s]\) is

$$\begin{aligned} K_{\beta _{1,s}}(x,y)= & {} \sum _{\begin{array}{c} m\in \mathbb {N}\\ m \text { odd} \end{array}} \frac{4s^2}{m^2\pi ^2}\cdot \frac{1}{s} \cdot \sin (\frac{\pi }{2s}mx)\sin (\frac{\pi }{2s}my) \\{} & {} + \sum _{\{u>0 \,:\,\tan (u)=\frac{1}{u}\} } \frac{s^2}{u^2}\cdot \frac{1}{s(\sin (u)^2+1)} \cdot \cos (\frac{u}{s}x)\cos (\frac{u}{s}y),\quad x,y\in [-s,s]. \end{aligned}$$

Its reproducing kernel Hilbert space is

$$\begin{aligned} \mathscr {H}_{K_{\beta _{1,s}}}([-s,s])=\{f:[-s,s]\rightarrow \mathbb {C}\;:\; f \text { is absolutely continuous, } f'\in L_2([-s,s])\}, \end{aligned}$$

where the inner product between f and g is given by

$$\begin{aligned} \frac{1}{2s}\big (f(-s)+f(s)\big )\big (\overline{g(-s)+g(s)} \big ) +\langle f',g'\rangle _{L_2([-s,s])}. \end{aligned}$$

Note that \(\mathscr {H}_{K_{\beta _{1,s}}}([-s,s])\) is continuously embedded into \(\mathbb {H}^1([-s,s])\). The proof of Proposition 4.2 is presented in Appendix C. It uses that, up to a constant, \(K_{\beta _{1,s}}|_{[-s,s]\times [-s,s]}\) is the Green’s function of the 1-dimensional harmonic equation \(\Delta u = f\) on \([-s,s]\) with the boundary conditions \(u'(s) = - u'(-s)\) and \(u(s) + u(-s) = -2u'(s)\).

It turns out that \(K_{\beta _{d,s}}\) has a simple form on \(\mathbb {B}^d_{s}\times \mathbb {B}^d_s\).

Theorem 4.3

For \(d\ge 1\), the discrepancy kernel \(K_{\beta _{d,s}}\) satisfies

$$\begin{aligned} K_{\beta _{d,s}}(x,y) = s-\frac{\Gamma (\frac{d}{2})}{2\sqrt{\pi }\Gamma (\frac{d+1}{2})}\Vert x-y\Vert ,\qquad x,y\in \mathbb {B}^d_{s}. \end{aligned}$$
(4.2)

The identity (4.2) for \(x,y\in \mathbb {S}^{d-1}\) with \(s=1\) has been established in [13], see also (3.2). Essentially, the same proof still works for the more general situation. Theorem 4.3 provides a simple form of \(K_{\beta _{d,s}}|_{\mathbb {X}\times \mathbb {X}}\) with \(\mathbb {X}= \mathbb {B}^d_{s}\), which may facilitate further computations. An immediate consequence is \(\mathscr {D}_{\beta _{d,s}}(\delta _x,\delta _y)=\frac{\Gamma (\frac{d}{2})}{\sqrt{\pi }\Gamma (\frac{d+1}{2})}\Vert x-y\Vert \), for \(x,y\in \mathbb {B}^d_{s}\).

5 The Euclidean Ball \(\mathbb {B}^d\)

This section is dedicated to derive the Fourier expansion of the discrepancy kernel \(K_{\beta _{d,1}}\) in (4.1) on \(\mathbb {B}^d\). Proposition 4.2 has covered \(d=1\), and we now derive the spectral decomposition of

$$\begin{aligned} K^{d,p}:\mathbb {B}^d\times \mathbb {B}^d \rightarrow \mathbb {R},\qquad (x,y)\mapsto \Vert x-y\Vert ^p, \end{aligned}$$

for all odd d with odd \(p>1-d\) with respect to the Lebesgue measure \(\sigma _{\mathbb {B}^d}\) on \(\mathbb {B}^d\). The case \(d=3\) with \(p=-1\) is discussed in [37].

Let \(\{\mathcal {C}^{\alpha }_m: m\in \mathbb {N},\; \alpha >-1/2\}\) denote the family of Gegenbauer polynomials with the standard normalization

$$\begin{aligned} \mathcal {C}^{\alpha }_m(1) =\left( {\begin{array}{c}m+2\alpha -1\\ m\end{array}}\right) = \frac{\Gamma (m+2\alpha )}{\Gamma (2\alpha )\Gamma (m+1)},\quad \alpha \ne 0. \end{aligned}$$

By \(\alpha =\frac{d}{2}-1\), the addition theorem for spherical harmonics yields

$$\begin{aligned} \sum _{l=1}^{Z(d,m)} Y^m_{l}(x)\overline{Y^m_{l}(y)} = \tfrac{2m+d-2}{d-2} \mathcal {C}^{\frac{d}{2}-1}_m(\langle x,y\rangle ),\quad x,y\in \mathbb {S}^{d-1}. \end{aligned}$$
(5.1)

For \(m\in \mathbb {N}\), let us define the kernels \(\mathcal {K}^{d,p}_m:[0,1]\times [0,1]\rightarrow \mathbb {R}\),

(5.2)

For \(d\ge 3\) and arbitrary real \(p>1-d\), we deduce from [18] that

$$\begin{aligned} \Vert x-y\Vert ^p = \sum _{m=0}^\infty \mathcal {K}^{d,p}_m(\Vert x\Vert ,\Vert y\Vert ) \,\mathcal {C}^{\frac{d}{2}-1}_m \big (\big \langle \tfrac{x}{\Vert x\Vert },\tfrac{y}{\Vert y\Vert }\big \rangle \big ), \quad x,y\in \mathbb {B}^d,\; (x\ne y \text { if } p < 0).\nonumber \\ \end{aligned}$$
(5.3)

For \(x=0\) or \(y=0\), the right-hand side of (5.3) is well-defined by analytic continuation.

Using the addition formula (5.1) we obtain

$$\begin{aligned} K^{d,p}(x,y) = \sum _{m=0}^\infty \tfrac{d-2}{2m+d-2} \mathcal {K}^{d,p}_m(\Vert x\Vert ,\Vert y\Vert ) \sum _{l=1}^{Z(d,m)}Y^{m}_{l}(\tfrac{x}{\Vert x\Vert }) Y^{m}_{l}(\tfrac{y}{\Vert y\Vert }),\quad x,y\in \mathbb {B}^d. \end{aligned}$$

The Fourier expansion of \({\mathcal {K}}_{m}^{d,p}\) with respect to the measure \(r^{d-1} \mathrm dr\) satisfies

$$\begin{aligned}{} & {} {\mathcal {K}}_{m}^{d,p}(r,s) = \sum _{j=1}^{\infty } \lambda _{m,j}^{d,p} \varphi _{m,j}^{d,p}(r) \varphi _{m,j}^{d,p}(s),\nonumber \\{} & {} \quad \int _{0}^{1} {\mathcal {K}}_{m}^{d,p}(r,s) \varphi _{m,j}^{d,p}(r)r^{d-1} \mathrm d r = \lambda _{m,j}^{d,p} \varphi ^{d,p}_{m,j}(s), \end{aligned}$$
(5.4)

where \(\int _{0}^{1} |\varphi _{m,j}^{d,p}(r)|^{2} r^{d-1}\mathrm dr = 1\). We consider

$$\begin{aligned} \varphi _{m',j',l'}^{d,p}(x):= \varphi _{m',j'}^{d,p}(\Vert x\Vert ) Y^{m'}_{l'}(\tfrac{x}{\Vert x\Vert }), \quad x \in {\mathbb {B}}^{d}, \end{aligned}$$

as well as the integral operator in Mercer’s theorem induced by \(K^{d,p}\),

$$\begin{aligned} T^{d,p} : L_2(\mathbb {B}^d) \rightarrow \mathcal {C}(\mathbb {B}^d), \quad f\mapsto \int _{\mathbb {B}^d} K^{d,p}(\cdot ,y)f(y)\textrm{d} y, \end{aligned}$$
(5.5)

so that direct computations yield

$$\begin{aligned} T^{d,p} \varphi _{m',j',l'}^{d,p}(x) = \lambda _{m',j'}^{d,p} \tfrac{(d-2) \textrm{vol}({\mathbb {S}}^{d-1})}{2m'+d-2} \varphi _{m',j',l'}^{d,p}(x) \end{aligned}$$
(5.6)

with the scaling \(\int _{{\mathbb {B}}^{d}}\left| \varphi _{m',j',l'}^{d,p}(x)\right| ^{2} \mathrm d x = \textrm{vol}({\mathbb {S}}^{d-1})\). This leads to the Fourier expansion

$$\begin{aligned} K^{d,p} (x,y)&= \sum _{m=0}^\infty \sum _{j=1}^\infty \lambda _{m,j}^{d,p} \tfrac{(d-2)\textrm{vol} ({\mathbb {S}}^{d-1})}{2m+d-2} \sum _{l=1}^{Z(d,m)}\frac{\varphi ^{d,p}_{m,j,l}(x) }{\sqrt{\textrm{vol}({\mathbb {S}}^{d-1})} } \frac{\varphi ^{d,p}_{m,j,l}(y)}{\sqrt{\textrm{vol}({\mathbb {S}}^{d-1})}} ,\quad x,y\in \mathbb {B}^d. \end{aligned}$$
(5.7)

Thus, the original problem is reduced to the spectral decomposition of the sequence of kernels \(\mathcal {K}^{d,p}_m\), for \(m\in \mathbb {N}\). The kernel \(\mathcal {K}^{d,p}_m\) induces the integral operator

$$\begin{aligned} T_m^{d,p} : L_2([0,1],r^{d-1}\textrm{d}r) \rightarrow \mathcal {C}([0,1]), \quad f\mapsto \int _0^1 \mathcal {K}^{d,p}_m(\cdot ,r)f(r)r^{d-1}\textrm{d}r, \end{aligned}$$
(5.8)

with eigenvalues \(\lambda ^{d,p}_{m,j}\) and eigenfunctions \(\varphi ^{d,p}_{m,j}\). We now specify these eigenvalues and eigenfunctions. In the following theorem, \(J_{\nu }\) denotes the Bessel function of the first kind of order \(\nu \) and \(\zeta _{k}:=\textrm{e}^{2\pi \textrm{i}/k}\) is the k-th root of unity.

Theorem 5.1

Suppose that both \(d\ge 3\) and \(p>1-d\) are odd and let \(m\in \mathbb {N}\). Then the following holds:

  1. a)

    Any eigenvalue \(\lambda \ne 0\) of \(T_m^{d,p}\) is in a one-to-one correspondence with

    $$\begin{aligned} \omega = \big |\lambda ^{-1}2^{d+p-2}(d+2m-2)(-\tfrac{p}{2})_{\frac{d+p}{2}-1}(\tfrac{d+p}{2}-1)!\big |^{\frac{1}{d+p}} \end{aligned}$$
    (5.9)

    with \(\omega \) satisfying \(\det (A(\omega ))=0\), where

    $$\begin{aligned} A(\omega ) = {\left\{ \begin{array}{ll} \left( \zeta _{d+p}^{-i\ell } J_{m+\frac{d}{2}-i-1}(\zeta _{d+p}^{\ell }\omega ) \right) _{i=1,\,\ell =0}^{\frac{d+p}{2},\,\frac{d+p}{2}-1}, &{} (-\tfrac{p}{2} )_{\frac{d+p}{2}-1} \lambda > 0,\\ \left( \zeta _{2(d+p)}^{-i(2\ell +1)}J_{m+\frac{d}{2}-i-1}(\zeta _{2(d+p)}^{2\ell +1}\omega ) \right) _{i=1,\,\ell =0}^{\frac{d+p}{2},\,\frac{d+p}{2}-1}, &{} (-\tfrac{p}{2} )_{\frac{d+p}{2}-1} \lambda < 0. \end{array}\right. }\nonumber \\ \end{aligned}$$
    (5.10)
  2. b)

    The eigenfunctions are exactly

    $$\begin{aligned} r\mapsto {\left\{ \begin{array}{ll} \sum _{\ell =1}^{\frac{d+p}{2}} c_\ell \,r^{1-\frac{d}{2}}J_{m+\frac{d}{2}-1}(\zeta ^{\ell -1}_{d+p}\omega r ), &{} (-\tfrac{p}{2} )_{\frac{d+p}{2}-1} \lambda > 0,\\ \sum _{\ell =1}^{\frac{d+p}{2}} c_\ell \,r^{1-\frac{d}{2}}J_{m+\frac{d}{2}-1}(\zeta ^{2\ell -1}_{2(d+p)}\omega r ), &{} (-\tfrac{p}{2} )_{\frac{d+p}{2}-1} \lambda < 0, \end{array}\right. } \end{aligned}$$

    where \(c\in \mathbb {R}^{\frac{d+p}{2}}\) is in the nullspace of \(A(\omega )\).

Remark 5.2

Computer experiments seem to indicate that the nullspace of \(A(\omega )\) is one-dimensional if \(\det (A(\omega ))=0\). In that case, the function

$$\begin{aligned} r\mapsto {\left\{ \begin{array}{ll} \sum _{\ell =1}^{\frac{d+p}{2}} (-1)^{\ell } A_{[1,\ell ]}(\omega ) \,r^{1-\frac{d}{2}}J_{m+\frac{d}{2}-1} (\zeta ^{\ell -1}_{d+p}\omega r ),&{} (-\tfrac{p}{2} )_{\frac{d+p}{2}-1} \lambda > 0,\\ \sum _{\ell =1}^{\frac{d+p}{2}} (-1)^{\ell } A_{[1,\ell ]}(\omega ) \,r^{1-\frac{d}{2}}J_{m+\frac{d}{2}-1} (\zeta ^{2\ell -1}_{2(d+p)}\omega r ),&{} (-\tfrac{p}{2} )_{\frac{d+p}{2}-1} \lambda < 0, \end{array}\right. }\nonumber \\ \end{aligned}$$
(5.11)

where \(A_{[1,\ell ]}(\omega )\) denotes the \((1,\ell )\) minor of \(A(\omega )\), spans the eigenspace associated with \(\lambda \).

Appendix D is dedicated to the proof of Theorem 5.1. The proof reveals strong ties with polyharmonic operators on the unit ball and higher order differential operators on the interval [0, 1]. We refer to [1] for structurally related spectral decompositions of polyharmonic operators on [0, 1] with homogeneous Neumann boundary conditions.

Corollary 5.3

(\(d=3\), \(p=1\)) The nonzero eigenvalues of \(T^{3,1}_m\), for \(m\in \mathbb {N}\setminus \{0\}\), are exactly \(\lambda =-\omega ^{-4}(4m+2)\), where \(\omega \) is a positive solution of the equation

$$\begin{aligned} J_{m-\frac{1}{2}}(\omega ) J_{m-\frac{3}{2}}( \mathrm i \omega ) - \mathrm i J_{m-\frac{1}{2}}( \mathrm i \omega ) J_{m-\frac{3}{2}}(\omega )= 0. \end{aligned}$$

The corresponding eigenspaces are 1-dimensional with the representative

$$\begin{aligned} f_{\lambda }(r)= r^{-\frac{1}{2}}\left( J_{m+\frac{1}{2}}(\omega r)\mathrm i J_{m-\frac{1}{2}}(\mathrm i \omega ) + J_{m+\frac{1}{2}}(\mathrm i \omega r)J_{m-\frac{1}{2}}(\omega )\right) . \end{aligned}$$
(5.12)

The formulas in Corollary 5.3 are derived from Theorem 5.1. Since \(J_{m-\frac{1}{2}}(\textrm{i}\omega )\ne 0\), for all \(\omega \in \mathbb {R}\setminus \{0\}\), the eigenspaces are 1-dimensional and (5.12) is not the zero-function.

In view of (4.2) in Theorem 4.3 we are particularly interested in kernels of the form \(c-\Vert x-y\Vert \). In this case, the expansion holds with \(-\mathcal {K}^{d,1}_m\) for \(m\ge 1\) and \(c-\mathcal {K}^{d,1}_0\) for \(m=0\).

6 The Rotation Group \(\textrm{SO}(3)\)

In this section we derive the Fourier expansion of the discrepancy kernel on the special orthogonal group \(\textrm{SO}(3)\). The eigenfunctions turn out to be classical functions but the coefficients and their decay rates need to be determined. We also provide numerical experiments by using the nonequispaced fast Fourier transform on \(\textrm{SO}(3)\).

6.1 Fourier Expansion on \(\textrm{SO}(3)\)

By identifying \(\mathbb {R}^{d\times d}\) with \(\mathbb {R}^{d^2}\), Theorem 4.3 applies to subsets of \(\mathbb {R}^{d\times d}\) endowed with the trace inner product

$$\begin{aligned} \langle x,y\rangle _{\textrm{F}}:=\textrm{trace}(x^\top y),\quad x,y\in \mathbb {R}^{d\times d}, \end{aligned}$$

and the induced Frobenius norm \(\Vert \cdot \Vert _{\textrm{F}}\) on \(\mathbb {R}^{d\times d}\). In this way, \(\textrm{SO}(3)\) is contained in \(\mathbb {B}^{9}_{\sqrt{3}}\), and it is natural to consider \({s}=\sqrt{3}\). We endow \(\textrm{SO}(3)\) with the normalized Haar measure \(\sigma _{\textrm{SO}(3)}\). Let \(\{\mathcal {D}^m_{k,l}:k,l=-m,\ldots ,m\}\) denote the orthonormal Wigner \(\mathcal {D}\)-functions on \(\textrm{SO}(3)\), which are closely related to the irreducible representations of \(\textrm{SO}(3)\) and provide an orthonormal basis for \(L_2(\textrm{SO}(3))\), cf. [48]. For \(p>0\), the Fourier expansion

$$\begin{aligned} 2^{-\frac{p}{2}}\Vert x-y\Vert ^p_{\textrm{F}} =\sum _{m=0}^\infty a_m(p,\textrm{SO}(3)) \sum _{k,l=-m}^m \mathcal {D}^m_{k,l}(x) \overline{\mathcal {D}^m_{k,l}(y)},\quad x,y\in \textrm{SO}(3),\nonumber \\ \end{aligned}$$
(6.1)

holds and, analogous to (3.5), the coefficients are

$$\begin{aligned} a_m(p,\textrm{SO}(3))\!:=\frac{1}{(2m+1)^2}\!\!\!\!\!\iint \limits _{\textrm{SO}(3)\times \textrm{SO}(3)} \!\!\!\!\!\!\!2^{-\frac{p}{2}}\Vert x-y\Vert _{\textrm{F}}^p\!\!\! \sum _{k,l=-m}^m\!\!\! \overline{\mathcal {D}^m_{k,l}(x)}\mathcal {D}^m_{k,l}(y)\textrm{d}\sigma _{\textrm{SO}(3)}(x)\textrm{d}\sigma _{\textrm{SO}(3)}(y).\nonumber \\ \end{aligned}$$
(6.2)

We now compute these coefficients for the entire range \(p>-3\).

Proposition 6.1

For \(p > -3\), the coefficients (6.2) are

$$\begin{aligned} a_m(p,\textrm{SO}(3)) = \frac{2^p\Gamma (\frac{p}{2}+\frac{3}{2})}{\sqrt{\pi }\Gamma (-\frac{p}{2})} \cdot \frac{\Gamma (-\frac{p}{2}+m)}{\Gamma (\frac{p}{2}+2+m)}\cdot \frac{1}{(m+\frac{1}{2})}. \end{aligned}$$
(6.3)

In particular, if \(p\not \in 2\mathbb {N}\), then

$$\begin{aligned} |a_m(p,\textrm{SO}(3))| = \left| \frac{2^p\Gamma (\frac{p}{2}+\frac{3}{2})}{\sqrt{\pi } \Gamma (-\frac{p}{2})}\right| m^{-(p+3)}(1+o(1)),\qquad m\in \mathbb {N}, \end{aligned}$$

and the series (6.1) terminates if \(p\in 2\mathbb {N}\).

The proof is given in Appendix E. For \(p\in 2\mathbb {N}\), we again apply the convention \(\frac{\Gamma (-\frac{p}{2}+m)}{\Gamma (-\frac{p}{2})}=(-\frac{p}{2})_m\) in (6.3), so that \(a_m(p,\textrm{SO}(3))=0\) for all \(m>\frac{p}{2}\) if \(p\in 2\mathbb {N}\). Provided that \(\tau >3/2\), the Sobolev space \(\mathbb {H}^\tau (\textrm{SO}(3))\) is the reproducing kernel Hilbert space associated with the reproducing kernel

$$\begin{aligned} (x,y)\mapsto \sum _{m=0}^\infty (1+m(m+1))^{-\tau }\sum _{k,l=-m}^m \mathcal {D}^m_{k,l}(x) \overline{\mathcal {D}^m_{k,l}(y)},\quad x,y\in \textrm{SO}(3). \end{aligned}$$

The choice \(p=1\) in Proposition 6.1 implies that the kernel \(K_{\beta _{9,s}}|_{\textrm{SO}(3)\times \textrm{SO}(3)}\) reproduces the Sobolev space \(\mathscr {H}_{K_{\beta _{9,s}}}(\textrm{SO}(3))=\mathbb {H}^{2}(\textrm{SO}(3))\) with an equivalent norm provided that \(s\ge \sqrt{3}\).

For \(d\ge 2\), we shall observe that \(K_{\beta _{d^2,\sqrt{d}}}|_{\textrm{SO}(d)\times \textrm{SO}(d)}\) reproduces the Sobolev space \(\mathbb {H}^{\tau }(\textrm{SO}(d))\), for \(\tau =\frac{d(d-1)+2}{4}\), with an equivalent norm. Indeed, Theorem 4.3 and Sect. 3 yield that \(K_{\beta _{d^2,\sqrt{d}}}|_{\mathbb {S}^{d^2-1}\times \mathbb {S}^{d^2-1}}\) reproduces \(\mathbb {H}^{\frac{d^2}{2}}(\mathbb {S}^{d^2-1})\) with equivalent norms. Rescaling implies that \(K_{\beta _{d^2,\sqrt{d}}}|_{\sqrt{d}\,\mathbb {S}^{d^2-1}\times \sqrt{d}\,\mathbb {S}^{d^2-1}}\) reproduces \(\mathbb {H}^{\frac{d^2}{2}}({\sqrt{d}\,}\mathbb {S}^{d^2-1})\). Since \(\textrm{SO}(d)\subset \sqrt{d}\,\mathbb {S}^{d^2-1}\), results on restricting kernels in [27] lead to \(\tau =\frac{d^2}{2}-\frac{1}{2}\left( (d^2-1)- \dim (\textrm{SO}(d))\right) \), where \(\dim (\textrm{SO}(d))=\frac{d(d-1)}{2}\).

6.2 Numerical Examples on \(\textrm{SO}(3)\)

Proposition 6.1 yields the coefficients of the kernel expansion

$$\begin{aligned} K_{\beta _{9,\sqrt{3}}}(x,y) = \sum _{m=0}^\infty a_m \sum _{k,l=-m}^m \mathcal {D}^m_{k,l}(x)\overline{\mathcal {D}^m_{k,l}(y)},\quad x,y\in \textrm{SO}(3). \end{aligned}$$

For \(\text {supp}(\mu ),\text {supp}(\nu _n)\subset \textrm{SO}(3)\), the \(L_2\)-discrepancy (1.6) for \(K_{\beta _{9,\sqrt{3}}}\) becomes

$$\begin{aligned} \mathscr {D}_{\beta _{9,\sqrt{3}}}(\mu ,\nu _n) = \sum _{m=0}^\infty a_m \sum _{k,l=-m}^m \left| \hat{\mu }^m_{k,l}- \frac{1}{n}\sum _{j=1}^n \overline{\mathcal {D}^m_{k,l}(x_j)}\right| ^2, \end{aligned}$$
(6.4)

where \(\hat{\mu }^m_{k,l}\) denotes the Fourier coefficient of \(\mu \) with respect to \(\mathcal {D}^m_{k,l}\), cf. (1.6). We truncate the series (6.4) at \(M=8\) and minimize

$$\begin{aligned} \sum _{m=0}^M a_m \sum _{k,l=-m}^m \left| \hat{\mu }^m_{k,l} - \frac{1}{n}\sum _{j=1}^n \overline{\mathcal {D}^m_{k,l}(x_j)}\right| ^2 \end{aligned}$$
(6.5)

among all n-point sets \(\{x_1,\ldots ,x_n\}\subset \textrm{SO}(3)\) for fixed \(n=30\). We efficiently solve the least squares minimization by using the nonequispaced fast Fourier transform on \(\textrm{SO}(3)\), cf. [32, 44]. Figure 2 shows the minimizing points mapped onto \(\mathbb {B}^3\).

Fig. 2
figure 2

We use the standard parametrization of \(\textrm{SO}(3)\) by \(\mathbb {S}^3\) via unit quaternions, which is then mapped into \(\mathbb {R}^3\) by stereographic projection and \(\mathbb {R}^3\) is parametrized by \(\mathbb {B}^3\) via \(x\mapsto \frac{x}{\Vert x\Vert }\tan (\frac{\pi }{2}\Vert x\Vert )\). The target measure \(\mu \) is supported on two disjoint parts with weight ratio 9/1 colored in darker blue by the cylindrical surface and the great circle. Numerical minimization of (6.5) splits 30 points in \(\textrm{SO}(3)\) into 27 points on the inner surface and 3 points on the great circle. We plotted 6 points on the great circle but antipodal points correspond to the same point in \(\textrm{SO}(3)\)

7 The Grassmannian \(\mathcal {G}_{2,4}\)

First, the Fourier expansion of the discrepancy kernel on \(\mathcal {G}_{2,4}\) is computed. To prepare for developing the nonequispaced fast Fourier transform on \(\mathcal {G}_{2,4}\), we then explicitly parametrize the Grassmannian \(\mathcal {G}_{2,4}\) by its double covering \(\mathbb {S}^2\times \mathbb {S}^2\). Next, we derive the nonequispaced fast Fourier transform on \(\mathbb {S}^2\times \mathbb {S}^2\) and provide numerical minimization experiments on \(\mathcal {G}_{2,4}\).

7.1 Fourier Expansion on \(\mathcal {G}_{2,4}\)

Theorem 4.3 also applies to the Grassmannian

$$\begin{aligned} \mathcal {G}_{2,4}:=\{x\in \mathbb {R}^{4\times 4} : x^\top = x,\; x^2=x,\; {\text {rank}}(x)=2\} \end{aligned}$$

with \(s=\sqrt{2}\) when \(\mathbb {R}^{4\times 4}\) is identified with \(\mathbb {R}^{16}\). To derive the Fourier expansion on \(\mathcal {G}_{2,4}\), we require some preparations. We shall use integer partitions \(\lambda =(\lambda _1,\lambda _2)\in \mathbb {N}^2\) with \(\lambda _1\ge \lambda _2\ge 0\). We also denote \(|\lambda |:=\lambda _1+\lambda _2\). The orthogonal group \(\textrm{O}(4)\) acts transitively on \(\mathcal {G}_{2,4}\) by conjugation and induces the irreducible decomposition

$$\begin{aligned} L_{2}(\mathcal {G}_{2,4},\sigma _{\mathcal {G}_{2,4}}) = \bigoplus _{\lambda _1\ge \lambda _2\ge 0} H_{\lambda }(\mathcal {G}_{2,4}),\qquad H_{\lambda }(\mathcal {G}_{2,4}) \perp H_{\lambda '}(\mathcal {G}_{2,4}), \quad \lambda \ne \lambda ', \end{aligned}$$
(7.1)

where \(\sigma _{\mathcal {G}_{2,4}}\) is the normalized orthogonally invariant measure on \(\mathcal {G}_{2,4}\) and \(H_{\lambda }(\mathcal {G}_{2,4})\) is equivalent to the irreducible representation \({\mathcal {H}}_{2\lambda }^{4}\) of \(\textrm{O}(4)\) with type \(2\lambda \), cf. [7, 35]. The normalized eigenfunctions of the Laplace–Beltrami operator on \(\mathcal {G}_{2,4}\) form an orthonormal basis for \(L_{2}(\mathcal {G}_{2,4})\), and each \(H_\lambda (\mathcal {G}_{2,4})\) is contained in the eigenspace \(E_{\alpha _\lambda }\) associated with the eigenvalue \(\alpha _\lambda = 4(\lambda _1^2+\lambda _2^2+\lambda _1)\), cf. [6,7,8, 23, 35, 45].

Let \(Q_\lambda \) be the reproducing kernel of \(H_\lambda (\mathcal {G}_{2,4})\). Any orthonormal basis \(\{\varphi _{\lambda ,l}\}_{l=1}^{\dim (\mathcal {H}^{4}_{2\lambda })}\) for \(H_{\lambda }(\mathcal {G}_{2,4})\) yields the spectral decomposition

$$\begin{aligned} Q_\lambda (x,y) = \sum _{l=1}^{\dim (\mathcal {H}^{4}_{2\lambda })}\varphi _{\lambda ,l}(x) \overline{\varphi _{\lambda ,l}(y)},\qquad x,y\in \mathcal {G}_{2,4},\quad \lambda _1\ge \lambda _2\ge 0. \end{aligned}$$
(7.2)

The orthogonal decomposition (7.1) leads to the Fourier expansion

$$\begin{aligned} 2^{-\frac{p}{2}}\Vert x -y\Vert _{\textrm{F}}^{p}=\sum _{\lambda _1\ge \lambda _2\ge 0}a_\lambda (p,\mathcal {G}_{2,4}) Q_\lambda (x,y),\qquad x,y \in \mathcal {G}_{2,4},\quad p>0. \end{aligned}$$
(7.3)

The coefficients \(a_\lambda (p,\mathcal {G}_{2,4})\) in (7.3) are

$$\begin{aligned} a_\lambda (p,\mathcal {G}_{2,4}) := \frac{1}{\dim (\mathcal {H}^{4}_{2\lambda })}\iint \limits _{\mathcal {G}_{2,4} \times \mathcal {G}_{2,4}}2^{-\frac{p}{2}}\Vert x -y\Vert _{\textrm{F}}^{p} \overline{Q_\lambda (x,y)} \textrm{d}\sigma _{\mathcal {G}_{2,4}}(x)\textrm{d}\sigma _{\mathcal {G}_{2,4}}(y).\nonumber \\ \end{aligned}$$
(7.4)

In order to determine \(a_\lambda (p,\mathcal {G}_{2,4}) \), we shall make use of the hypergeometric coefficients \( (f)_{(\lambda _1,\lambda _2)} := (f)_{\lambda _1} (f-\tfrac{1}{2})_{\lambda _2}\). Also recall the notation \(|\lambda |=\lambda _1+\lambda _2\) and \(\Vert \lambda \Vert =\sqrt{\lambda _1^2+\lambda _2^2}\).

Theorem 7.1

For \(p>-4\), we have

(7.5)

In particular, if \(p\not \in 2\mathbb {N}\), then

$$\begin{aligned} |a_\lambda (p,\mathcal {G}_{2,4})|= \left| \frac{\Gamma (\frac{p}{2}+2)}{2\Gamma (-\frac{p}{2})}\right| \Vert \lambda \Vert ^{-(p+4)} (1+o(1)), \qquad \lambda _1\ge \lambda _2\ge 0, \end{aligned}$$
(7.6)

and the series (7.3) terminates if \(p\in 2\mathbb {N}\).

The proof of this theorem is contained in Section F.1 of Appendix F. If \(p\in 2\mathbb {N}\), then \( a_\lambda (p,\mathcal {G}_{2,4})=0\) for all \(|\lambda |>\frac{p}{2}\). For \(\tau >2\), the Sobolev space \(\mathbb {H}^\tau (\mathcal {G}_{2,4})\) is the reproducing kernel Hilbert space with associated reproducing kernel

$$\begin{aligned} (x,y)\mapsto \sum _{\lambda _1\ge \lambda _2\ge 0}(1+4\lambda _1^2+4\lambda _2^2+4\lambda _1)^{-\tau }Q_\lambda (x,y),\quad x,y\in \mathcal {G}_{2,4}, \end{aligned}$$
(7.7)

cf. [11, 16]. Since the coefficients in (7.7) behave asymptotically as \(\Vert \lambda \Vert ^{-2\tau }\), the choice \(p=1\) in Theorem 7.1 implies that the kernel \(K_{\beta _{16,s}}|_{\mathcal {G}_{2,4}\times \mathcal {G}_{2,4}}\) reproduces the Sobolev space \(\mathscr {H}_{K_{\beta _{16,s}}}(\mathcal {G}_{2,4})=\mathbb {H}^{\frac{5}{2}}(\mathcal {G}_{2,4})\) with an equivalent norm provided that \(s\ge \sqrt{2}\).

Analogous to \(\textrm{SO}(d)\) at the end of Sect. 6.1, we deduce with [27] that, for \(d\ge 2\), \(K_{\beta _{d^2,\sqrt{k}}}|_{\mathcal {G}_{k,d}\times \mathcal {G}_{k,d}}\) reproduces the Sobolev space \(\mathbb {H}^{\frac{k(d-k)+1}{2}}(\mathcal {G}_{k,d})\) with an equivalent norm.

7.2 Parametrization of \(\mathcal {G}_{2,4}\) by

To derive the nonequispaced fast Fourier transform on \(\mathcal {G}_{2,4}\), we shall first explicitly construct the parametrization of \(\mathcal {G}_{2,4}\) by its double covering \(\mathbb {S}^2\times \mathbb {S}^2\). We denote the \(d\times d\)-identity matrix by \(I_d\), and the cross-product between two vectors \(x,y\in \mathbb {S}^2\) is denoted by \(x\times y\in \mathbb {R}^3\). The mapping \(\mathcal {P}:\mathbb {S}^2\times \mathbb {S}^2\rightarrow \mathcal {G}_{2,4}\) given by

$$\begin{aligned} \begin{aligned} (x,y) \mapsto&\frac{1}{2} \begin{pmatrix} 1 + x^\top y &{}&{} - (x \times y)^\top \\ - x \times y &{} &{}x y^\top + y x^\top + (1-x^\top y) I_3 \\ \end{pmatrix} \end{aligned} \end{aligned}$$
(7.8)

is surjective and, for all \(x,y,u,v\in \mathbb {S}^2\),

$$\begin{aligned} \mathcal {P}(u,v)=\mathcal {P}(x,y)\quad \text { if and only if } \quad (u,v) \in \{\pm (x,y) \}, \end{aligned}$$
(7.9)

see Section F.2 and Theorem F.4 of Appendix F. In order to specify the inverse map, note that can be identified with \( \mathcal {M}(3):=\{xy^\top \in \mathbb {R}^{3\times 3} : x,y\in \mathbb {S}^2\}. \) We now define \(\mathcal {L}:\mathcal {G}_{2,4}\rightarrow \mathcal {M}(3)\),

$$\begin{aligned} \mathcal {L}(P) := \left( \begin{array}{ccc} \frac{1}{2}(P_{11}+P_{22}-P_{33}-P_{44}) &{} P_{23} - P_{14} &{} P_{24} + P_{13} \\ P_{23} + P_{14} &{} \frac{1}{2}(P_{11}-P_{22}+P_{33}-P_{44}) &{} P_{34}- P_{12} \\ P_{24} - P_{13} &{} P_{34} + P_{12} &{} \frac{1}{2}(P_{11}-P_{22}-P_{33}+P_{44}) \\ \end{array}\right) ,\nonumber \\ \end{aligned}$$
(7.10)

and direct computations lead to

$$\begin{aligned} \mathcal {L}(\mathcal {P}(x,y)) = xy^\top ,\quad x,y\in \mathbb {S}^2. \end{aligned}$$
(7.11)

The right-hand side determines x and y up to the ambiguity (7.9). Under the Frobenius norm, \(\mathcal {P}\) is distance preserving in the sense

$$\begin{aligned} \Vert \mathcal {P}(x,y)-\mathcal {P}(u,v)\Vert _{\textrm{F}} = \Vert xy^\top - uv^\top \Vert _{\textrm{F}},\quad x,y,u,v\in \mathbb {S}^2. \end{aligned}$$
(7.12)

The latter follows from (F.20) in Lemma F.6 in Section F.2 of Appendix F.

We shall now check how the spherical harmonics \(Y^m_l\) on \(\mathbb {S}^2\) relate to the eigenfunctions \(\varphi _{\lambda ,l}\in H_\lambda (\mathcal {G}_{2,4})\) of the Laplace–Beltrami operator on \(\mathcal {G}_{2,4}\), cf. (7.2). The functions \(Y_{k,l}^{m,n}:\mathcal {G}_{2,4}\rightarrow {\mathbb {C}}\) given by

$$\begin{aligned} Y_{k,l}^{m,n}(\mathcal {P}(x,y)) := Y_k^m(x) \cdot Y_l^n(y) \end{aligned}$$
(7.13)

are well-defined for \(m+n \in 2{\mathbb {N}}\), the latter taking into account the ambiguity (7.9).

Theorem 7.2

For \(m_\lambda :=(\lambda _1+\lambda _2)\) and \(n_\lambda :=(\lambda _1-\lambda _2)\), we have

$$\begin{aligned} H_{\lambda }(\mathcal {G}_{2,4}) = \textrm{span}\{ Y_{k,l}^{m_\lambda ,n_\lambda }, Y_{l,k}^{n_\lambda ,m_\lambda } : k=-m_\lambda ,\dots ,m_\lambda ,\; l=-n_\lambda ,\dots ,n_\lambda \}.\nonumber \\ \end{aligned}$$
(7.14)

The proof is presented at the end of Section F.2 of Appendix F. Note that the geodesic distance on \(\mathcal {G}_{2,4}\) is \(\textrm{dist}_{\mathcal {G}_{2,4}}(P,Q)=\sqrt{2}\sqrt{\theta _1^2+\theta _2^2}\), where \(\theta _1,\theta _2\in [0,\pi /2]\) are the principal angles determined by the two largest eigenvalues \(\cos ^2(\theta _1)\) and \(\cos ^2(\theta _2)\) of the matrix PQ. Aside from (7.12), \(\mathcal {P}\) is also distance-preserving with respect to the respective geodesic distances, i.e.,Footnote 4

This equality follows from (F.23) in Lemma F.6 in the appendix via further direct calculations.

The identity (7.14) provides explicit expressions for the orthonormal basis \(\{\varphi _{\lambda ,l}\}_{l=1}^{\dim (\mathcal {H}^{4}_{2\lambda })}\) of \(H_{\lambda }(\mathcal {G}_{2,4})\) that is used to construct the reproducing kernel \(Q_\lambda \) in (7.2). It also provides a fast Fourier transform on \(\mathcal {G}_{2,4}\) from the respective transform on \(\mathbb {S}^2\times \mathbb {S}^2\) that is developed in the subsequent section.

7.3 Nonequispaced Fast Fourier Transform on \(\mathcal {G}_{2,4}\)

The nonequispaced fast (spherical) Fourier transform on \(\mathbb {S}^2\) has been developed in [39, 41] under the acronym nfsft. Here, we shall derive the analogous transform on \(\mathbb {S}^2\times \mathbb {S}^2\), which induces the nonequispaced fast Fourier transform on \(\mathcal {G}_{2,4}\) via the mapping \(\mathcal {P}\) and (7.14) with (7.13).

For a given finite set of coefficients \(f^{m_1,m_2}_{k,l}\in \mathbb {C}\), \(m_1,m_2=0,\ldots ,M\), \(k=-m_1,\ldots ,m_1\), \(l=-m_2,\ldots ,m_2\), we aim to evaluate

$$\begin{aligned} F(x,y):=\sum _{m_1,m_2=0}^M \sum _{k=-m_1}^{m_1} \sum _{l=-m_2}^{m_2} f^{m_1,m_2}_{k,l} Y^{m_1}_k(x)Y^{m_2}_l(y) \end{aligned}$$
(7.15)

at n scattered locations \((x_j,y_j)_{j=1}^n\subset \mathbb {S}^2\times \mathbb {S}^2\). Direct evaluation of (7.15) leads to \(O(n M^4)\) operations. We shall now derive an approximative algorithm that is more efficient for \(n\gg M\).

By following the ideas in [39, 41], switching to spherical coordinates reveals that (7.15) is a 4-dimensional trigonometric polynomial. This enables the use of the 4-dimensional nonequispaced fast Fourier transform nfft to significantly reduce the complexity. In spherical coordinates the spherical harmonics are trigonometric polynomials such that

$$\begin{aligned} Y^m_k(z(\theta ,\varphi )) =e^{\textrm{i}k\varphi } \sum _{k'=-m}^m c^m_{k,k'}e^{\textrm{i}k'\theta }, \qquad z(\theta ,\varphi ) = \left( \begin{array}{c} \sin (\theta )\cos (\varphi )\\ \sin (\theta )\sin (\varphi )\\ \cos (\theta ) \end{array} \right) \in \mathbb {S}^2,\nonumber \\ \end{aligned}$$
(7.16)

where \(0\le \theta \le \pi \), \(0\le \varphi \le 2\pi \), and \(c^m_{k,k'}\in \mathbb {C}\) are suitable coefficients that we assume to be given or precomputed. Thus, for \( x = z(\theta _1,\varphi _1)\) and \(y=z(\theta _2,\varphi _2)\), there are coefficients \(b^{k,l}_{k',l'}\in \mathbb {C}\) such that

$$\begin{aligned} F(x,y)=\sum _{k,l,k',l'=-M}^M b^{k,l}_{k',l'} e^{\textrm{i}k\varphi _1}e^{\textrm{i}l\varphi _2} e^{\textrm{i} k'\theta _1}e^{\textrm{i} l'\theta _2}. \end{aligned}$$
(7.17)

We check in Sect. 1 of Appendix F that the set of coefficients \(b^{k,l}_{k',l'}\) can be computed by \(O(M^5)\) operations provided that the numbers \(c^m_{k,k'}\) in (7.16) are given. The expression (7.17) can be evaluated at n scattered locations by the nonequispaced discrete Fourier transform ndft with \(O(n M^4)\) operations, cf. [39, 41]. An efficient approximative algorithm is the nonequispaced fast Fourier transform nfft that requires only \(O(M^4\log (M)+n|\log (\epsilon )|^4)\) operations with accuracy \(\epsilon \), see [39, 41] for details on accuracy. Thus, our algorithm for evaluating (7.15) at n scattered locations requires \(O(M^5+n|\log (\epsilon )|^4)\) operations. For \(n\gg M\), this is a significant reduction in complexity compared to the original \(O(nM^4)\) operations. We shall choose \(n\sim M^4\) in the subsequent section, so that the complexity is reduced from \(O(M^8)\) to \(O(M^5+M^4|\log (\epsilon )|^4)\) operations. For potential further reduction, we refer to Remark F.7 in the appendix.

7.4 Numerical example on \(\mathcal {G}_{2,4}\)

By Theorem 7.1, we can calculate the coefficients of the kernel expansion

$$\begin{aligned} K_{\beta _{16,\sqrt{2}}}(x,y) = \sum _{\lambda _1\ge \lambda _2\ge 0} a_\lambda \sum _{l=1}^{\dim (\mathcal {H}^{d}_{2\lambda })} \varphi _{\lambda ,l}(x)\overline{\varphi _{\lambda ,l}(y)},\quad x,y\in \mathcal {G}_{2,4}. \end{aligned}$$

The eigenfunctions \(\varphi _{\lambda ,l}\) are given by the tensor products of spherical harmonics in (7.13), cf. Theorem 7.2. For \(\text {supp}(\mu ),\text {supp}(\nu _n)\subset \mathcal {G}_{2,4}\), the \(L_2\)-discrepancy (1.6) of the kernel \(K_{\beta _{16,\sqrt{2}}}|_{\mathcal {G}_{2,4}\times \mathcal {G}_{2,4}}\) is

$$\begin{aligned} \mathscr {D}_{\beta _{16,\sqrt{2}}}(\mu ,\nu _n) = \sum _{\lambda _1\ge \lambda _2\ge 0} a_\lambda \sum _{l=1}^{\dim (\mathcal {H}^{d}_{2\lambda })} \left| \hat{\mu }_{\lambda ,l} - \frac{1}{n}\sum _{j=1}^n \overline{\varphi _{\lambda ,l}(x_j)}\right| ^2, \end{aligned}$$
(7.18)

where \(\hat{\mu }_{\lambda ,l}\) is the Fourier coefficient of \(\mu \) with respect to \(\varphi _{\lambda ,l}\), cf. (1.6).

Let us consider \(\mu =\sigma _{\mathcal {G}_{2,4}}\). According to [11] (see also [15, 16]), the lower bound

$$\begin{aligned} n^{-5/4}\lesssim \mathscr {D}_{\beta _{16,\sqrt{2}}}(\mu ,\nu _n) \end{aligned}$$

holds for all n-point sets \(\{x_1,\ldots ,x_n\}\subset \mathcal {G}_{2,4}\). We truncate the series (7.18) and let \(\nu ^{M}_n=\frac{1}{n}\sum _{j=1}^n\delta _{x^M_j}\) denote a minimizer of

$$\begin{aligned} \sum _{\lambda +\lambda _2\le M} a_\lambda \sum _{l=1}^{\dim (\mathcal {H}^{d}_{2\lambda })} \left| \hat{\mu }_{\lambda ,l} - \frac{1}{n}\sum _{j=1}^n \overline{\varphi _{\lambda ,l}(x_j)}\right| ^2 \end{aligned}$$
(7.19)

among all n-point sets \(\{x_1,\ldots ,x_n\}\subset \mathcal {G}_{2,4}\). A suitable choice \(n\sim M^4 \) leads to the optimal rate

$$\begin{aligned} \mathscr {D}_{\beta _{16,\sqrt{2}}}(\mu ,\nu ^{M}_n)\sim n^{-5/4}, \end{aligned}$$
(7.20)

cf. [11, 25]. Note that we can efficiently solve the least squares minimization (7.19) by using the nonequispaced fast Fourier transform on \(\mathcal {G}_{2,4}\) derived from the nonequispaced fast Fourier transform on \(\mathbb {S}^2\times \mathbb {S}^2\) of Sect. 7.3 and applying Theorem  7.2. Figure 3 shows logarithmic plots of the number of points versus the \(L_2\)-discrepancy. We observe a line with slope \(-5/4\) as predicted by (7.20).

Fig. 3
figure 3

Logarithmic plot of the number of points n versus the \(L_2\)-discrepancy \(\mathscr {D}_{\beta _{16},\sqrt{2}}(\mu ,\nu ^{M}_n)\) on \(\mathcal {G}_{2,4}\), where \(\nu ^{M}_n\) is derived from numerical minimization