This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the eigenvalues of the spheroidal wave equation

Harald Schmid [email protected] University of Applied Sciences Amberg-Weiden, Amberg, Germany
Abstract.

This paper presents some new results on the eigenvalues of the spheroidal wave equation. We study the angular and Coulomb spheroidal wave equation as a special case of a more general linear Hamiltonian system depending on three parameters. We prove that the eigenvalues of this system satisfy a first-order quasilinear partial differential equation with respect to the parameters. This relation offers a new insight on how the eigenvalues of the spheroidal wave equation depend on the spheroidal parameter. Apart from analytical considerations, the PDE we obtain can also be used for a numerical computation of spheroidal eigenvalues.

Key words and phrases:
spheroidal eigenvalues, Coulomb spheroidal wave equation, linear Hamiltonian system, inviscid forced Burgers’ equation, deformation method
1991 Mathematics Subject Classification:
33E10, 33F05, 34L15, 35F20

1. Introduction

The angular spheroidal wave equation

ddx((1x2)ddxw(x))+(λ+γ2(1x2)μ21x2)w(x)=0,1<x<1\frac{\mathrm{d}}{\mathrm{d}x}\left((1-x^{2})\frac{\mathrm{d}}{\mathrm{d}x}w(x)\right)+\left(\lambda+\gamma^{2}(1-x^{2})-\frac{\mu^{2}}{1-x^{2}}\right)w(x)=0,\quad-1<x<1 (1)

appears in many fields of physics and engineering like quantum mechanics, electromagnetism, signal processing etc. Here μ\mu is supposed to be some given real number, and since only μ2\mu^{2} occurs in (1), we may assume μ0\mu\geq 0 without loss of generality. In many applications the so-called spheroidal parameter (or size parameter) γ\gamma is either real or purely imaginary, so that γ2\gamma^{2} is a real number which may also be negative. If μ\mu is an integer and γ2\gamma^{2} is real, then the separation of the Helmholtz equation in prolate (γ2>0\gamma^{2}>0) or oblate (γ2<0\gamma^{2}<0) spheroidal coordinates results in a second order ODE of the form (1). A slightly more general differential equation is the Coulomb spheroidal wave equation (CSWE)

ddx((1x2)ddxw(x))+(λ+βx+γ2(1x2)μ21x2)w(x)=0\frac{\mathrm{d}}{\mathrm{d}x}\left((1-x^{2})\frac{\mathrm{d}}{\mathrm{d}x}w(x)\right)+\left(\lambda+\beta x+\gamma^{2}(1-x^{2})-\frac{\mu^{2}}{1-x^{2}}\right)w(x)=0 (2)

It differs from (1) only by the presence of a linear term βx\beta x with some additional parameter β\beta, which we also assume to be real. The numbers λ\lambda\in\mathbb{C} for which (2) has a nontrivial bounded solution w(x)w(x) on (1,1)(-1,1) are the eigenvalues (or characteristic values) of the CSWE, and the corresponding eigenfunctions w(x)w(x) are called Coulomb spheroidal functions. They appear in astrophysics and molecular physics and provide, for example, exact wave functions for a one-electron diatomic molecule with fixed nuclei (see [1, Chapter 9] or [2]). In our subsequent considerations, it does not make much difference whether we include the term βx\beta x or not. Therefore, in this paper we will mainly deal with equation (2), while the results we obtain are obviously applicable to the angular spheroidal wave equation (1) as well.

For technical reasons we introduce the parameters u1u_{1}, u2u_{2}, u3u_{3} which are related to β\beta, γ2\gamma^{2}, λ\lambda in (2) by

γ2=u1,β=u32(μ+1)u2,λ=u3+μ(μ+1)\gamma^{2}=u_{1},\quad\beta=-u_{3}-2(\mu+1)u_{2},\quad\lambda=u_{3}+\mu(\mu+1) (3)

Moreover, we set α:=12(μ+1)\alpha:=\frac{1}{2}(\mu+1). With these parameters in mind, we will see that the 2×22\times 2 differential system

(0110)y(z)(22u2αz+α1z2u2αz+α1z2(u1+u22)+u3z)y(z)=Λ(11z001z)y(z),0<z<1\begin{pmatrix}[r]0&-1\\[4.30554pt] 1&0\end{pmatrix}y^{\prime}(z)-\begin{pmatrix}2&2u_{2}-\frac{\alpha}{z}+\frac{\alpha}{1-z}\\[4.30554pt] 2u_{2}-\frac{\alpha}{z}+\frac{\alpha}{1-z}&2(u_{1}+u_{2}^{2})+\frac{u_{3}}{z}\end{pmatrix}y(z)=\Lambda\begin{pmatrix}\frac{1}{1-z}&0\\[4.30554pt] 0&\frac{1}{z}\end{pmatrix}y(z),\quad 0<z<1 (4)

is related to the Coulomb spheroidal wave equation by the following means: In section 2 we show that (4), in combination with appropriate boundary conditions, can be written as an eigenvalue problem for a self-adjoint differential operator T=T(u1,u2,u3)T=T(u_{1},u_{2},u_{3}). The eigenvalues Λ\Lambda of TT depend analytically on the parameters (u1,u2,u3)(u_{1},u_{2},u_{3}), and in addition they have the following properties:

  1. (a)

    λ=u3+μ(μ+1)\lambda=u_{3}+\mu(\mu+1) is an eigenvalue of the CSWE (2) for the parameters γ2=u1\gamma^{2}=u_{1}, β=u32(μ+1)u2\beta=-u_{3}-2(\mu+1)u_{2} if and only if Λ=0\Lambda=0 is an eigenvalue of the linear Hamiltonian system (4).

  2. (b)

    The eigenvalues Λ=Λ(u1,u2,u3)\Lambda=\Lambda(u_{1},u_{2},u_{3}) of (4) satisfy the first-order quasilinear partial differential equation

    2u1Λu1+((Λ+2)(u1+u22)+Λ+u2+u3)Λu2+((2Λ+u3+2)(2Λu22μ1)2(μ+1)(u1+u22)+2μ)Λu3=(1+2μ2Λu2)(Λ+2)2μ\begin{split}2u_{1}\frac{\partial\Lambda}{\partial u_{1}}&+\left((\Lambda+2)(u_{1}+u_{2}^{2})+\Lambda+u_{2}+u_{3}\right)\frac{\partial\Lambda}{\partial u_{2}}\\[-4.30554pt] &+\left((2\Lambda+u_{3}+2)(2\Lambda u_{2}-2\mu-1)-2(\mu+1)(u_{1}+u_{2}^{2})+2\mu\right)\frac{\partial\Lambda}{\partial u_{3}}\\ &=(1+2\mu-2\Lambda u_{2})(\Lambda+2)-2\mu\end{split} (5)

According to (a), the eigenvalue problem (4) may be regarded as a generalization of the Coulomb spheroidal wave equation, and the PDE (5) describes an analytic relation between the eigenvalues and the parameters, where the eigenfunctions need not to be known. A similar observation has already been made in [3] for the Chandrasekhar-Page equation, and [4] provides a method by which one can establish such a relation between the eigenvalues and the parameters in terms of a PDE for other linear Hamiltonian systems as well. In the present paper we will use this \qqdeformation method from [4] to prove assertion (b). Finally, in section 3 we solve (5) by the method of characteristics, where we consider the prolate spheroidal wave equation (β=0\beta=0, γ2>0\gamma^{2}>0) in more detail. In this case, we can associate a linear 2×22\times 2 differential system to (1) which contains only two parameters in addition to the eigenvalue parameter, so that the partial differential equation for the eigenvalues can be further simplified.

2. A linear Hamiltonian system associated to the CSWE

In this section we study the linear 2×22\times 2 Hamiltonian system (4) on the interval (0,1)(0,1). It depends on three parameters u1,u2,u3u_{1},u_{2},u_{3}\in\mathbb{R}, where α=12(μ+1)12\alpha=\frac{1}{2}(\mu+1)\geq\frac{1}{2} with some constant μ0\mu\geq 0, and the prime denotes the derivative with respect to zz. If we define

J:=(0110),H(z,𝐮):=(2αz+α1z+2u2αz+α1z+2u2u3z+2(u1+u22)),W(z):=(11z001z)J:=\begin{pmatrix}[r]0&-1\\[4.30554pt] 1&0\end{pmatrix},\quad H(z,\mathbf{u}):=\begin{pmatrix}2&-\frac{\alpha}{z}+\frac{\alpha}{1-z}+2u_{2}\\[4.30554pt] -\frac{\alpha}{z}+\frac{\alpha}{1-z}+2u_{2}&\frac{u_{3}}{z}+2(u_{1}+u_{2}^{2})\end{pmatrix},\quad W(z):=\begin{pmatrix}\frac{1}{1-z}&0\\[4.30554pt] 0&\frac{1}{z}\end{pmatrix}

and

τy:=W(z)1(Jy(z)H(z,𝐮)y(z))\tau y:=W(z)^{-1}\left(Jy^{\prime}(z)-H(z,\mathbf{u})y(z)\right) (6)

where 𝐮:=(u1,u2,u3)3\mathbf{u}:=(u_{1},u_{2},u_{3})\in\mathbb{R}^{3} combines the three parameters, then (4) can be written in the form τy=Λy\tau y=\Lambda y. Moreover, as W(z)=W(z)>0W(z)^{\ast}=W(z)>0 and H(z,𝐮)=H(z,𝐮)H(z,\mathbf{u})^{\ast}=H(z,\mathbf{u}) for all z(0,1)z\in(0,1), τ\tau defines for fixed parameters 𝐮3\mathbf{u}\in\mathbb{R}^{3} a formally self-adjoint differential expression in the Hilbert space

LW2((0,1),2):={f:(0,1)2|01f(z)W(z)f(z)dz<}\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2}):=\Big{\{}f:(0,1)\longrightarrow\mathbb{C}^{2}\ \big{|}\ \int_{0}^{1}f(z)^{\ast}W(z)f(z)\,\mathrm{d}z<\infty\Big{\}}

of square-integrable vector functions to the weight function W(z)W(z) with scalar product

f,gW:=01f(z)W(z)g(z)dz\langle f,g\rangle_{W}:=\int_{0}^{1}f(z)^{\ast}W(z)\,g(z)\,\mathrm{d}z

The maximal operator Ty:=τyTy:=\tau y generated by τ\tau has the domain (cf. [5, Section 3])

D(T):={yLW2((0,1),2):yACloc((0,1),2) and τyLW2((0,1),2)}D(T):=\left\{y\in\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2})\ :\ y\in\mathrm{AC_{loc}}((0,1),\mathbb{C}^{2})\mbox{ and }\tau y\in\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2})\right\} (7)

whereas the domain of the minimal operator T0y:=τyT_{0}y:=\tau y associated to τ\tau is given by

D(T0):={yD(T):y has compact support in (0,1)}D(T_{0}):=\big{\{}y\in D(T)\,:\,y\mbox{ has compact support in }(0,1)\big{\}}

In order to establish the self-adjointness of TT, we first prove that τ\tau is in the limit point case at z=0z=0 and z=1z=1. To this end, we consider the differential equation τy=Λy\tau y=\Lambda y in the complex plane. It is equivalent to the 2×22\times 2 system

y(z)=(1z(αu3+Λ0α)+11z(α0Λα)+(2u22(u1+u22)22u2))y(z)y^{\prime}(z)=\left(\frac{1}{z}\begin{pmatrix}-\alpha&u_{3}+\Lambda\\[4.30554pt] 0&\alpha\end{pmatrix}+\frac{1}{1-z}\begin{pmatrix}\alpha&0\\[4.30554pt] -\Lambda&-\alpha\end{pmatrix}+\begin{pmatrix}2u_{2}&2(u_{1}+u_{2}^{2})\\[4.30554pt] -2&-2u_{2}\end{pmatrix}\right)y(z) (8)

with regular singular points at z=0z=0 and z=1z=1. If 𝔇0:={z:|z|<1}\mathfrak{D}_{0}:=\{z\in\mathbb{C}:|z|<1\} denotes the unit disk centered at 0, then [3, Lemma 6] implies that for fixed (Λ,𝐮)×3(\Lambda,\mathbf{u})\in\mathbb{C}\times\mathbb{R}^{3} there exists a fundamental matrix of the form Y0(z)=F0(z,Λ,𝐮)zAzQ0(Λ,𝐮)Y_{0}(z)=F_{0}(z,\Lambda,\mathbf{u})z^{A}z^{Q_{0}(\Lambda,\mathbf{u})}, z𝔇0z\in\mathfrak{D}_{0}, where F0:𝔇0××3M2()F_{0}:\mathfrak{D}_{0}\times\mathbb{C}\times\mathbb{R}^{3}\longrightarrow\mathrm{M}_{2}(\mathbb{C}) is an analytic matrix function,

F0(0,Λ,𝐮)=(1u3+Λ2α01),A:=(α00α),Q0(Λ,𝐮)=(00q0(Λ,𝐮)0)F_{0}(0,\Lambda,\mathbf{u})=\begin{pmatrix}1&\frac{u_{3}+\Lambda}{2\alpha}\\[4.30554pt] 0&1\end{pmatrix},\quad A:=\begin{pmatrix}-\alpha&0\\[4.30554pt] 0&\alpha\end{pmatrix},\quad Q_{0}(\Lambda,\mathbf{u})=\begin{pmatrix}0&0\\[4.30554pt] q_{0}(\Lambda,\mathbf{u})&0\end{pmatrix}

and q0:×3q_{0}:\mathbb{C}\times\mathbb{R}^{3}\longrightarrow\mathbb{C} is an analytic function with q00q_{0}\equiv 0 if 2α=μ+12\alpha=\mu+1 is not an integer. Furthermore, the transformation y~(z)=y(1z)\tilde{y}(z)=y(1-z) yields

y~(z)=(1z(α0Λα)+11z(αu3Λ0α)+(2u22(u1+u22)22u2))y~(z)\tilde{y}^{\prime}(z)=\left(\frac{1}{z}\begin{pmatrix}-\alpha&0\\[4.30554pt] \Lambda&\alpha\end{pmatrix}+\frac{1}{1-z}\begin{pmatrix}\alpha&-u_{3}-\Lambda\\[4.30554pt] 0&-\alpha\end{pmatrix}+\begin{pmatrix}-2u_{2}&-2(u_{1}+u_{2}^{2})\\[4.30554pt] 2&2u_{2}\end{pmatrix}\right)\tilde{y}(z)

and hence (8) also possesses a fundamental matrix of the form Y1(z)=F1(z,Λ,𝐮)(1z)A(1z)Q1(Λ,𝐮)Y_{1}(z)=F_{1}(z,\Lambda,\mathbf{u})(1-z)^{A}(1-z)^{Q_{1}(\Lambda,\mathbf{u})}, z𝔇1:={z:|z1|<1}z\in\mathfrak{D}_{1}:=\{z\in\mathbb{C}:|z-1|<1\}, where F1:𝔇1××3M2()F_{1}:\mathfrak{D}_{1}\times\mathbb{C}\times\mathbb{R}^{3}\longrightarrow\mathrm{M}_{2}(\mathbb{C}) is an analytic matrix function,

F1(1,Λ,𝐮)=(10Λ2α1),Q1(Λ,𝐮)=(00q1(Λ,𝐮)0)F_{1}(1,\Lambda,\mathbf{u})=\begin{pmatrix}1&0\\[4.30554pt] -\frac{\Lambda}{2\alpha}&1\end{pmatrix},\quad Q_{1}(\Lambda,\mathbf{u})=\begin{pmatrix}0&0\\[4.30554pt] q_{1}(\Lambda,\mathbf{u})&0\end{pmatrix}

and q1:×3q_{1}:\mathbb{C}\times\mathbb{R}^{3}\longrightarrow\mathbb{C} being an analytic function satisfying q10q_{1}\equiv 0 if 2α=μ+12\alpha=\mu+1 is not an integer. For fixed parameters (Λ,𝐮)×3(\Lambda,\mathbf{u})\in\mathbb{C}\times\mathbb{R}^{3}, the solution

y0(z):=Y0(z)e1=F0(z)(zα00zα)(10q0logz1)(10)y_{0}(z):=Y_{0}(z)e_{1}=F_{0}(z)\begin{pmatrix}z^{-\alpha}&0\\[4.30554pt] 0&z^{\alpha}\end{pmatrix}\begin{pmatrix}1&0\\[4.30554pt] q_{0}\log z&1\end{pmatrix}\begin{pmatrix}1\\[4.30554pt] 0\end{pmatrix}

of (8) asymptotically behaves like y0(z)zαe1y_{0}(z)\sim z^{-\alpha}e_{1} as z0z\to 0 (here and in the following, {e1,e2}\{e_{1},e_{2}\} denotes the standard basis of 2\mathbb{C}^{2}). Since 2α12\alpha\geq 1 and y0(z)W(z)y0(z)z2αy_{0}(z)^{\ast}W(z)y_{0}(z)\sim z^{-2\alpha} as z0z\to 0, the function y0(z)W(z)y0(z)y_{0}(z)^{\ast}W(z)y_{0}(z) is not integrable at z=0z=0, and hence y0y_{0} does not lie left in LW2((0,1),2)\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2}). Moreover, for the solution

y1(z):=Y1(z)e1=F1(z)((1z)α00(1z)α)(10q1log(1z)1)(10)y_{1}(z):=Y_{1}(z)e_{1}=F_{1}(z)\begin{pmatrix}(1-z)^{-\alpha}&0\\[4.30554pt] 0&(1-z)^{\alpha}\end{pmatrix}\begin{pmatrix}1&0\\[4.30554pt] q_{1}\log(1-z)&1\end{pmatrix}\begin{pmatrix}1\\[4.30554pt] 0\end{pmatrix}

we get y1(z)W(z)y1(z)(1z)2α1y_{1}(z)^{\ast}W(z)y_{1}(z)\sim(1-z)^{-2\alpha-1} as z1z\to 1. Therefore, y1(z)W(z)y1(z)y_{1}(z)^{\ast}W(z)y_{1}(z) is not integrable at z=1z=1, and thus y1y_{1} does not lie right in LW2((0,1),2)\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2}). According to Weyl’s alternative (see e. g. [5, Theorem 5.6]), the differential expression τ\tau is in the limit point case at both endpoints. Hence, the closure of its associated minimal operator T0¯\overline{T_{0}} is a self-adjoint extension of T0T_{0} (cf. [6, Theorem 5.4]), and because of T0¯=T0¯=T\overline{T_{0}}=\overline{T_{0}}^{\ast}=T (see [5, Theorem 3.9]), this extension coincides with the maximal operator TT, so that TT is actually a self-adjoint differential operator in the Hilbert space LW2((0,1),2)\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2}).

In the following, we will study the eigenvalues of the self-adjoint operator T=T(𝐮)T=T(\mathbf{u}) associated to τ\tau and their dependence on the parameters 𝐮3\mathbf{u}\in\mathbb{R}^{3}. For this purpose, we introduce the functions η0(z,Λ,𝐮):=Y0(z,Λ,𝐮)e2\eta_{0}(z,\Lambda,\mathbf{u}):=Y_{0}(z,\Lambda,\mathbf{u})e_{2} and η1(z,Λ,𝐮):=Y1(z,Λ,𝐮)e2\eta_{1}(z,\Lambda,\mathbf{u}):=Y_{1}(z,\Lambda,\mathbf{u})e_{2}, which asymptotically behave like

η0(z,Λ,𝐮)zα(u3+Λ2α1)as z0,η1(z,Λ,𝐮)(1z)α(01)as z1\eta_{0}(z,\Lambda,\mathbf{u})\sim z^{\alpha}\begin{pmatrix}\frac{u_{3}+\Lambda}{2\alpha}\\[4.30554pt] 1\end{pmatrix}\quad\mbox{as $z\to 0$},\quad\eta_{1}(z,\Lambda,\mathbf{u})\sim(1-z)^{\alpha}\begin{pmatrix}0\\[4.30554pt] 1\end{pmatrix}\quad\mbox{as $z\to 1$} (9)

Note that y0y_{0}, η0\eta_{0} as well as y1y_{1}, η1\eta_{1} form a fundamental system of (4). Since the solutions y0(z)=Y0(z)e1y_{0}(z)=Y_{0}(z)e_{1} and y1(z)=Y1(z)e1y_{1}(z)=Y_{1}(z)e_{1} are not square-integrable with respect to W(z)W(z), Λ\Lambda\in\mathbb{R} is an eigenvalue of T(𝐮)T(\mathbf{u}) for fixed 𝐮3\mathbf{u}\in\mathbb{R}^{3} if and only if there exists a nontrivial solution yy of (4) which is a constant multiple of both η0\eta_{0} and η1\eta_{1}. Additionally, by virtue of α12\alpha\geq\frac{1}{2} and the asymptotic behavior of the fundamental solutions at the boundary points, a nontrivial solution yy of (4) is an eigenfunction if and only if (z(1z))1/2y(z)(z(1-z))^{-1/2}y(z) is bounded on (0,1)(0,1). Furthermore, we can write η0\eta_{0} as a linear combination

η0(z,Λ,𝐮)=Θ(Λ,𝐮)y1(z,Λ,𝐮)+Ω(Λ,𝐮)η1(z,Λ,𝐮)\eta_{0}(z,\Lambda,\mathbf{u})=\Theta(\Lambda,\mathbf{u})y_{1}(z,\Lambda,\mathbf{u})+\Omega(\Lambda,\mathbf{u})\eta_{1}(z,\Lambda,\mathbf{u}) (10)

with Θ\Theta, Ω\Omega called connection coefficients. Since for any fixed z0(0,1)z_{0}\in(0,1)

(Θ(Λ,𝐮)Ω(Λ,𝐮))=Y1(z0,Λ,𝐮)1Y0(z0,Λ,𝐮)e2\begin{pmatrix}\Theta(\Lambda,\mathbf{u})\\[2.15277pt] \Omega(\Lambda,\mathbf{u})\end{pmatrix}=Y_{1}(z_{0},\Lambda,\mathbf{u})^{-1}Y_{0}(z_{0},\Lambda,\mathbf{u})e_{2}

where the fundamental matrices depend analytically on the parameters, also Θ=Θ(Λ,𝐮)\Theta=\Theta(\Lambda,\mathbf{u}) and Ω=Ω(Λ,𝐮)\Omega=\Omega(\Lambda,\mathbf{u}) depend analytically on (Λ,𝐮)×3(\Lambda,\mathbf{u})\in\mathbb{C}\times\mathbb{R}^{3}. Due to our considerations above, Λ\Lambda\in\mathbb{R} is an eigenvalue of T(𝐮)T(\mathbf{u}) if and only if Θ(Λ,𝐮)=0\Theta(\Lambda,\mathbf{u})=0. As Θ(Λ,𝐮)\Theta(\Lambda,\mathbf{u}) is an entire function for fixed 𝐮3\mathbf{u}\in\mathbb{R}^{3}, the zeros of Θ(Λ,𝐮)\Theta(\Lambda,\mathbf{u}) are isolated, and thus also the eigenvalues of T(𝐮)T(\mathbf{u}) are isolated with multiplicity 11.

In a first step, we will figure out how the eigenvalues of (2) and (4) are related.

Lemma 1.

Λ=0\Lambda=0 is an eigenvalue of the self-adjoint operator T(u1,u2,u3)T(u_{1},u_{2},u_{3}) associated to the linear Hamiltonian system (4) if and only if λ=u3+μ(μ+1)\lambda=u_{3}+\mu(\mu+1) is an eigenvalue of the Coulomb spheroidal wave equation (2) for the parameters γ2=u1\gamma^{2}=u_{1} and β=u32(μ+1)u2\beta=-u_{3}-2(\mu+1)u_{2}.

Proof.

The system (8) with Λ=0\Lambda=0 is equivalent to

y(z)=(1z(αu30α)+11z(α00α)+(2u22(u1+u22)22u2))y(z)y^{\prime}(z)=\left(\frac{1}{z}\begin{pmatrix}-\alpha&u_{3}\\[4.30554pt] 0&\alpha\end{pmatrix}+\frac{1}{1-z}\begin{pmatrix}\alpha&0\\[4.30554pt] 0&-\alpha\end{pmatrix}+\begin{pmatrix}2u_{2}&2(u_{1}+u_{2}^{2})\\[4.30554pt] -2&-2u_{2}\end{pmatrix}\right)y(z)

If we introduce the vector function

y^(x)=(v(x)w(x)):=2(1x2)12y(x+12),x(1,1)\hat{y}(x)=\begin{pmatrix}v(x)\\[2.15277pt] w(x)\end{pmatrix}:=2(1-x^{2})^{-\frac{1}{2}}y\left(\tfrac{x+1}{2}\right),\quad x\in(-1,1) (11)

then this differential system takes the form

y^(x)=(11+x(α12u30α12)+11x(α+1200α+12)+(u2u1+u221u2))y^(x)\hat{y}^{\prime}(x)=\left(\frac{1}{1+x}\begin{pmatrix}-\alpha-\frac{1}{2}&u_{3}\\[4.30554pt] 0&\alpha-\frac{1}{2}\end{pmatrix}+\frac{1}{1-x}\begin{pmatrix}\alpha+\frac{1}{2}&0\\[4.30554pt] 0&-\alpha+\frac{1}{2}\end{pmatrix}+\begin{pmatrix}u_{2}&u_{1}+u_{2}^{2}\\[4.30554pt] -1&-u_{2}\end{pmatrix}\right)\hat{y}(x) (12)

Since y^(x)=(z(1z))1/2y(z)\hat{y}(x)=(z(1-z))^{-1/2}y(z) with z=12(x+1)(0,1)z=\frac{1}{2}(x+1)\in(0,1), we obtain that Λ=0\Lambda=0 is an eigenvalue of T(𝐮)T(\mathbf{u}) if and only if (12) has a nontrivial solution y^(x)\hat{y}(x) which is bounded on (1,1)(-1,1). Using α=12(μ+1)\alpha=\frac{1}{2}(\mu+1), we can write (12) in terms of its components v(x)v(x) and w(x)w(x):

v(x)\displaystyle v^{\prime}(x) =(u2+(μ+2)x1x2)v(x)+(u1+u22+u31+x)w(x)\displaystyle=\left(u_{2}+\frac{(\mu+2)x}{1-x^{2}}\right)v(x)+\left(u_{1}+u_{2}^{2}+\frac{u_{3}}{1+x}\right)w(x)
w(x)\displaystyle w^{\prime}(x) =v(x)(u2+μx1x2)w(x)\displaystyle=-v(x)-\left(u_{2}+\frac{\mu x}{1-x^{2}}\right)\,w(x)

Solving the second equation for v(x)v(x) yields

v(x)=w(x)(u2+μx1x2)w(x)v(x)=-w^{\prime}(x)-\left(u_{2}+\frac{\mu x}{1-x^{2}}\right)\,w(x) (13)

and inserting this expression into the first equation gives the differential equation

ddx((1x2)ddxw(x))+(u3+μ(μ+1)(u3+2(μ+1)u2)x+u1(1x2)μ21x2)w(x)=0\frac{\mathrm{d}}{\mathrm{d}x}\left((1-x^{2})\frac{\mathrm{d}}{\mathrm{d}x}w(x)\right)+\left(u_{3}+\mu(\mu+1)-(u_{3}+2(\mu+1)u_{2})x+u_{1}(1-x^{2})-\frac{\mu^{2}}{1-x^{2}}\right)w(x)=0

for w(x)w(x) which is a CSWE (2) with parameters (3). Conversely, any bounded solution w(x)w(x) of this ODE is a Coulomb spheroidal function having the form w(x)=(1x2)μ/2φ(x)w(x)=(1-x^{2})^{\mu/2}\varphi(x) with some entire function φ(x)\varphi(x), cf. [7, Section 3.2, Satz 1]. If we define v(x)v(x) by (13), then v(x)=(1x2)μ/2φ(x)u2(1x2)μ/2φ(x)v(x)=-(1-x^{2})^{\mu/2}\varphi^{\prime}(x)-u_{2}(1-x^{2})^{\mu/2}\varphi(x) is also bounded on (1,1)(-1,1), and hence the function y^\hat{y} defined by (11) is a bounded solution of (12) which implies that Λ=0\Lambda=0 is an eigenvalue of (4). ∎

In the next step we examine which way the eigenvalues Λ\Lambda of (4) depend on the parameters (u1,u2,u3)(u_{1},u_{2},u_{3}).

Theorem 2.

Let T=T(𝐮)T=T(\mathbf{u}) be the self-adjoint realization of the differential expression (6) with domain D(T)D(T) given by (7), and suppose that Λ0\Lambda_{0} is an eigenvalue of T=T(𝐮0)T=T(\mathbf{u}_{0}) for some given parameter vector 𝐮03\mathbf{u}_{0}\in\mathbb{R}^{3}. Then there exists a domain U3U\subset\mathbb{R}^{3} with 𝐮0U\mathbf{u}_{0}\in U and eigenvalues Λ(𝐮)\Lambda(\mathbf{u}) of T(𝐮)T(\mathbf{u}) with Λ(𝐮0)=Λ0\Lambda(\mathbf{u}_{0})=\Lambda_{0}, such that Λ=Λ(𝐮)\Lambda=\Lambda(\mathbf{u}) depends analytically on 𝐮=(u1,u2,u3)U\mathbf{u}=(u_{1},u_{2},u_{3})\in U. In addition, Λ\Lambda satisfies the first-order partial differential equation (5) on UU.

Proof.

We first prove the analytic dependence of the eigenvalues on the parameters 𝐮=(u1,u2,u3)\mathbf{u}=(u_{1},u_{2},u_{3}) in a neighborhood of 𝐮0=:(a,b,c)\mathbf{u}_{0}=:(a,b,c). For this, let

B(z,𝐮):=H(z,𝐮)H(z,𝐮0)=(02(u2b)2(u2b)2(u1+u22ab2)+u3cz)B(z,\mathbf{u}):=H(z,\mathbf{u})-H(z,\mathbf{u}_{0})=\begin{pmatrix}0&2(u_{2}-b)\\[4.30554pt] 2(u_{2}-b)&2(u_{1}+u_{2}^{2}-a-b^{2})+\frac{u_{3}-c}{z}\end{pmatrix}

If A(𝐮)A(\mathbf{u}) denotes the symmetric multiplication operator A(𝐮)y:=W(z)1B(z,𝐮)yA(\mathbf{u})y:=-W(z)^{-1}B(z,\mathbf{u})y in LW2((0,1),2)\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2}), then T(𝐮)=T(𝐮0)+A(𝐮)T(\mathbf{u})=T(\mathbf{u}_{0})+A(\mathbf{u}). Since

W(z)1/2B(z,𝐮)W(z)1/2=(02(u2b)z(1z)2(u2b)z(1z)2(u1+u22ab2)z+u3c)W(z)^{-1/2}B(z,\mathbf{u})W(z)^{-1/2}=\begin{pmatrix}0&2(u_{2}-b)\sqrt{z(1-z)}\\[4.30554pt] 2(u_{2}-b)\sqrt{z(1-z)}&2(u_{1}+u_{2}^{2}-a-b^{2})z+u_{3}-c\end{pmatrix}

we have |W(z)1/2B(z,𝐮)W(z)1/2|C|W(z)^{-1/2}B(z,\mathbf{u})W(z)^{-1/2}|\leq C for all z(0,1)z\in(0,1) with some constant C=C(𝐮)0C=C(\mathbf{u})\geq 0, and

AyW2\displaystyle\|Ay\|_{W}^{2} =01(W(z)1B(z,𝐮)y(z))W(z)(W(z)1B(z,𝐮)y(z))dz\displaystyle=\int_{0}^{1}\left(W(z)^{-1}B(z,\mathbf{u})\,y(z)\right)^{\ast}W(z)\left(W(z)^{-1}B(z,\mathbf{u})\,y(z)\right)\,\mathrm{d}z
=01|W(z)1/2B(z,𝐮)W(z)1/2W(z)1/2y(z)|2dz01C2|W(z)1/2y(z)|2dz=C2yW2\displaystyle=\int_{0}^{1}|W(z)^{-1/2}B(z,\mathbf{u})W(z)^{-1/2}W(z)^{1/2}y(z)|^{2}\,\mathrm{d}z\leq\int_{0}^{1}C^{2}|W(z)^{1/2}y(z)|^{2}\,\mathrm{d}z=C^{2}\|y\|_{W}^{2}

implies that A=A(𝐮)A=A(\mathbf{u}) is even a bounded operator on LW2((0,1),2)\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2}) for all 𝐮3\mathbf{u}\in\mathbb{C}^{3}. From [8, Chap. V, § 4, Theorem 4.3 and Chap. VII, § 2, Theorem 2.6] it follows that T(𝐮)T(\mathbf{u}) defines a holomorphic family of self-adjoint operators, where its domain D(T)D(T) is independent of 𝐮3\mathbf{u}\in\mathbb{C}^{3}. As shown in [8, Chap. VII, § 2, Sec. 4 and § 3, Sec. 1 – 2] there exists a domain U3U\subset\mathbb{R}^{3} with 𝐮0U\mathbf{u}_{0}\in U as well as an analytical function Λ=Λ(u1,u2,u3)\Lambda=\Lambda(u_{1},u_{2},u_{3}) with Λ(𝐮0)=Λ0\Lambda(\mathbf{u}_{0})=\Lambda_{0}, such that Λ(𝐮)\Lambda(\mathbf{u}) is a simple eigenvalue of T(𝐮)T(\mathbf{u}) for all 𝐮U\mathbf{u}\in U.

Now, if Λ(𝐮)\Lambda(\mathbf{u}) is an eigenvalue of T(𝐮)T(\mathbf{u}), then each corresponding eigenfunction is a constant multiple of the fundamental solutions η0\eta_{0} and η1\eta_{1}, respectively. Hence, η0(z)=Ω(𝐮)η1(z)\eta_{0}(z)=\Omega(\mathbf{u})\cdot\eta_{1}(z) for all z(0,1)z\in(0,1) with some factor Ω(𝐮)0\Omega(\mathbf{u})\neq 0 which depends analytically on 𝐮U\mathbf{u}\in U, and therefore η0(z)=zα(1z)αh(z,𝐮)\eta_{0}(z)=z^{\alpha}(1-z)^{\alpha}h(z,\mathbf{u}), where h:𝔇×U2h:\mathfrak{D}\times U\longrightarrow\mathbb{C}^{2} is an analytic function on some domain [0,1]𝔇[0,1]\subset\mathfrak{D}\subset\mathbb{C} which satisfies

h(0,𝐮)=(u3+Λ(𝐮)2α1)andh(1,𝐮)=(0Ω(𝐮))h(0,\mathbf{u})=\begin{pmatrix}\frac{u_{3}+\Lambda(\mathbf{u})}{2\alpha}\\[4.30554pt] 1\end{pmatrix}\quad\mbox{and}\quad h(1,\mathbf{u})=\begin{pmatrix}0\\[4.30554pt] \Omega(\mathbf{u})\end{pmatrix}

due to (9). Obviously η0LW2((0,1),2)\eta_{0}\in\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2}), and

ν(𝐮):=η0,η0W=01h(z,𝐮)(zμ+1(1z)μ00zμ(1z)μ+1)h(z,𝐮)dz\nu(\mathbf{u}):=\langle\eta_{0},\eta_{0}\rangle_{W}=\int_{0}^{1}h(z,\mathbf{u})^{\ast}\begin{pmatrix}z^{\mu+1}(1-z)^{\mu}&0\\[4.30554pt] 0&z^{\mu}(1-z)^{\mu+1}\end{pmatrix}h(z,\mathbf{u})\,\mathrm{d}z

defines a smooth function ν:U(0,)\nu:U\longrightarrow(0,\infty). Setting ψ(z,𝐮):=ν(𝐮)1/2h(z,𝐮)\psi(z,\mathbf{u}):=\nu(\mathbf{u})^{-1/2}h(z,\mathbf{u}), then

y(z,𝐮):=ν(𝐮)1/2η0(z,Λ(𝐮),𝐮)=zα(1z)αψ(z,𝐮)y(z,\mathbf{u}):=\nu(\mathbf{u})^{-1/2}\eta_{0}(z,\Lambda(\mathbf{u}),\mathbf{u})=z^{\alpha}(1-z)^{\alpha}\psi(z,\mathbf{u})

is an eigenfunction corresponding to Λ(𝐮)\Lambda(\mathbf{u}) for which y,yW=1\langle y,y\rangle_{W}=1 holds for all 𝐮U\mathbf{u}\in U, and y=y(z,𝐮)y=y(z,\mathbf{u}) depends infinitely differentiable on (z,𝐮)(0,1)×U(z,\mathbf{u})\in(0,1)\times U. Since the matrix functions W:(0,1)M2()W:(0,1)\longrightarrow\mathrm{M}_{2}(\mathbb{C}) and H:(0,1)×3M2()H:(0,1)\times\mathbb{R}^{3}\longrightarrow\mathrm{M}_{2}(\mathbb{C}) are also infinitely differentiable with respect to all variables, the assumptions (a), (b), (c) in [4] are fulfilled. In addition, if we introduce the matrix function

G(z,Λ,𝐮):=(2z2zu2+Λu2μ122zu2+Λu2μ122(z1)(u1+u22))G(z,\Lambda,\mathbf{u}):=\begin{pmatrix}2z&2zu_{2}+\Lambda u_{2}-\mu-\frac{1}{2}\\[4.30554pt] 2zu_{2}+\Lambda u_{2}-\mu-\frac{1}{2}&2(z-1)(u_{1}+u_{2}^{2})\end{pmatrix}

then a direct calculation gives

Gz+(ΛW+H)JGGJ(ΛW+H)=k=13fkHuk+gW\frac{\partial G}{\partial z}+(\Lambda W+H)JG-GJ(\Lambda W+H)=\sum_{k=1}^{3}f_{k}\frac{\partial H}{\partial u_{k}}+gW (14)

where the functions fk=fk(Λ,𝐮)f_{k}=f_{k}(\Lambda,\mathbf{u}) and g=g(Λ,𝐮)g=g(\Lambda,\mathbf{u}) on the right hand side are defined by

f1\displaystyle f_{1} :=2u1\displaystyle:=2u_{1}
f2\displaystyle f_{2} :=(Λ+2)(u1+u22)+Λ+u2+u3\displaystyle:=(\Lambda+2)(u_{1}+u_{2}^{2})+\Lambda+u_{2}+u_{3}
f3\displaystyle f_{3} :=(2Λ+u3+2)(2Λu22μ1)2(μ+1)(u1+u22)+2μ\displaystyle:=(2\Lambda+u_{3}+2)(2\Lambda u_{2}-2\mu-1)-2(\mu+1)(u_{1}+u_{2}^{2})+2\mu
g\displaystyle g :=(1+2μ2Λu2)(Λ+2)2μ\displaystyle:=(1+2\mu-2\Lambda u_{2})(\Lambda+2)-2\mu

Here, G:(0,1)××UM2()G:(0,1)\times\mathbb{C}\times U\longrightarrow\mathrm{M}_{2}(\mathbb{C}) and fk,g:×Uf_{k},g:\mathbb{C}\times U\longrightarrow\mathbb{C} are differentiable functions, and therefore also assumption (d) in [4] is satisfied. In addition to the \qqdeformation equation (14), the function

k=13fk(Λ(𝐮),𝐮)yuk+JG(z,Λ(𝐮),𝐮)y=zα(1z)α(k=13fk(Λ(𝐮),𝐮)ψuk+JG(z,Λ(𝐮),𝐮)ψ)\displaystyle\sum_{k=1}^{3}f_{k}(\Lambda(\mathbf{u}),\mathbf{u})\frac{\partial y}{\partial u_{k}}+JG(z,\Lambda(\mathbf{u}),\mathbf{u})y=z^{\alpha}(1-z)^{\alpha}\left(\sum_{k=1}^{3}f_{k}(\Lambda(\mathbf{u}),\mathbf{u})\frac{\partial\psi}{\partial u_{k}}+JG(z,\Lambda(\mathbf{u}),\mathbf{u})\psi\right)

lies in LW2((0,1),2)\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2}). Finally, from [4, Theorem 2.2] it follows that Λ=Λ(u1,u2,u3)\Lambda=\Lambda(u_{1},u_{2},u_{3}) solves

k=13fk(Λ,u1,u2,u3)Λuk=g(Λ,u1,u2,u3)\sum_{k=1}^{3}f_{k}(\Lambda,u_{1},u_{2},u_{3})\frac{\partial\Lambda}{\partial u_{k}}=g(\Lambda,u_{1},u_{2},u_{3})

on UU, and this is exactly the first-order quasilinear PDE (5). ∎

The eigenvalues of (4) coincide with the zeros of the function Θ(Λ,𝐮)\Theta(\Lambda,\mathbf{u}), and this is one of the connection coefficients appearing in (10). The question arises how this function can be calculated numerically for given parameter values 𝐮3\mathbf{u}\in\mathbb{R}^{3} and Λ\Lambda\in\mathbb{C}. In [9], the Coulomb spheroidal wave equation has already been associated to a 2×22\times 2 differential system, which is, however, different from the linear Hamiltonian system studied in the present paper. Nevertheless, the method suggested in [9] for calculating the connection coefficient Θ\Theta can also be applied to the system (4). To this end, we apply the transformation y^(z)=(1z)α1y(z)\hat{y}(z)=(1-z)^{\alpha-1}y(z), which turns (4) resp. (8) into the differential system

y^(z)=(1zA0+1z1A1+C)y^(z)\hat{y}^{\prime}(z)=\left(\frac{1}{z}\,A_{0}+\frac{1}{z-1}\,A_{1}+C\right)\hat{y}(z) (15)

where

A0=(αu3+Λ0α),A1=(10Λ2α1),C=(2u22(u1+u22)22u2)A_{0}=\begin{pmatrix}-\alpha&u_{3}+\Lambda\\[4.30554pt] 0&\alpha\end{pmatrix},\quad A_{1}=\begin{pmatrix}-1&0\\[4.30554pt] \Lambda&2\alpha-1\end{pmatrix},\quad C=\begin{pmatrix}2u_{2}&2(u_{1}+u_{2}^{2})\\[4.30554pt] -2&-2u_{2}\end{pmatrix}

For fixed (λ,𝐮)×3(\lambda,\mathbf{u})\in\mathbb{C}\times\mathbb{R}^{3} it has the Floquet solution

η^0(z):=(1z)α1η0(z)=zαh(z),h(z)=k=0zkdk,d0:=(u3+Λ2α1)\hat{\eta}_{0}(z):=(1-z)^{\alpha-1}\eta_{0}(z)=z^{\alpha}h(z),\quad h(z)=\sum_{k=0}^{\infty}z^{k}d_{k},\quad d_{0}:=\begin{pmatrix}\frac{u_{3}+\Lambda}{2\alpha}\\[4.30554pt] 1\end{pmatrix}

where hh is holomorphic in 𝔇0\mathfrak{D}_{0}, and d0d_{0} is an eigenvector of A0A_{0} corresponding to the eigenvalue α\alpha. Finally, once we have determined the series coefficients dkd_{k} using the recurrence relation

bk:=(2(ku2u3)k(k+μ+1)2k(u1+u22)2u2u3k(k+μ+1)2k2u2k)bk1(0(μ+1)u3k(k+μ+1)0μ+1k)dk1,dk:=dk1+bkb_{k}:=\begin{pmatrix}\frac{2(ku_{2}-u_{3})}{k(k+\mu+1)}&\frac{2k(u_{1}+u_{2}^{2})-2u_{2}u_{3}}{k(k+\mu+1)}\\[4.30554pt] -\frac{2}{k}&-\frac{2u_{2}}{k}\end{pmatrix}b_{k-1}-\begin{pmatrix}0&\frac{(\mu+1)u_{3}}{k(k+\mu+1)}\\[4.30554pt] 0&\frac{\mu+1}{k}\end{pmatrix}d_{k-1},\quad d_{k}:=d_{k-1}+b_{k} (16)

for k=1,2,3,k=1,2,3,\ldots starting with b0:=d0b_{0}:=d_{0}, then we can compute Θ\Theta with arbitrary precision by means of

Lemma 3.

Let μ0\mu\geq 0 and 𝐮=(u1,u2,u3)3\mathbf{u}=(u_{1},u_{2},u_{3})\in\mathbb{R}^{3} be fixed. Then for each Λ\Lambda\in\mathbb{C} we have

Θ(Λ,𝐮):=limke1Tdk(Λ,𝐮)\Theta(\Lambda,\mathbf{u}):=\lim_{k\to\infty}e_{1}^{\operatorname{T}}d_{k}(\Lambda,\mathbf{u})

More precisely, we obtain e1Tdk=Θ+𝒪(kδμ2)e_{1}^{\operatorname{T}}d_{k}=\Theta+\text{O}(k^{\delta-\mu-2}) as kk\to\infty for arbitrary small δ>0\delta>0.

This assertion, as well as formula (16), is obtained by a similar reasoning as given in [9, Section 3], and therefore the proof is omitted here. If we set Λ=0\Lambda=0 and calculate the zeros 𝐮=(u1,u2,u3)\mathbf{u}=(u_{1},u_{2},u_{3}) of the function Θ(0,𝐮)\Theta(0,\mathbf{u}), then we get the eigenvalues λ=u3+μ(μ+1)\lambda=u_{3}+\mu(\mu+1) of the Coulomb spheroidal wave equation (2) for the parameters γ2=u1\gamma^{2}=u_{1} and β=u32(μ+1)u2\beta=-u_{3}-2(\mu+1)u_{2} according to Lemma 1. This approach, like the method presented in [9], provides a simple and efficient way to obtain the eigenvalues of a CSWE. However, with increasing parameter values, numerical effects such as rounding errors or digit cancellation may occur, which require some computational effort to overcome. In this context the partial differential equation (5) may be helpful. It not only provides us with new insights into the analytical dependence of the eigenvalues Λ=Λ(𝐮)\Lambda=\Lambda(\mathbf{u}) on the parameters, but can also be used to determine the Coulomb spheroidal eigenvalues numerically. For this purpose, we need to examine the PDE and its characteristics in more detail.

3. Analysis of the PDE for the eigenvalues

The PDE (5) looks very complicated at first sight, but it has some remarkable properties that simplify the calculation of the solution by the method of characteristics. The characteristic curves related to the quasilinear PDE (5) satisfy the (nonlinear) first-order differential system

du1dt=2u1(t)du2dt=(Λ(t)+2)(u1(t)+u2(t)2)+Λ(t)+u2(t)+u3(t)du3dt=(2Λ(t)+u3(t)+2)(2Λ(t)u2(t)2μ1)2(μ+1)(u1(t)+u2(t)2)+2μdΛdt=(1+2μ2Λ(t)u2(t))(Λ(t)+2)2μ\begin{split}\frac{\mathrm{d}u_{1}}{\mathrm{d}t}&=2u_{1}(t)\\ \frac{\mathrm{d}u_{2}}{\mathrm{d}t}&=(\Lambda(t)+2)(u_{1}(t)+u_{2}(t)^{2})+\Lambda(t)+u_{2}(t)+u_{3}(t)\\ \frac{\mathrm{d}u_{3}}{\mathrm{d}t}&=(2\Lambda(t)+u_{3}(t)+2)(2\Lambda(t)u_{2}(t)-2\mu-1)-2(\mu+1)(u_{1}(t)+u_{2}(t)^{2})+2\mu\\ \frac{\mathrm{d}\Lambda}{\mathrm{d}t}&=(1+2\mu-2\Lambda(t)u_{2}(t))(\Lambda(t)+2)-2\mu\end{split} (17)

We can choose the parameter for the characteristic curve in such a way that the initial conditions are given at t=0t=0. If we define the scalar function

Ψ(t):=Λ(t)(u1(t)+u2(t)21)2(μ+1)u2(t)u3(t)\Psi(t):=\Lambda(t)(u_{1}(t)+u_{2}(t)^{2}-1)-2(\mu+1)u_{2}(t)-u_{3}(t)

then (17) and a straightforward calculation yields ddtΨ=Ψ\frac{\mathrm{d}}{\mathrm{d}t}\Psi=\Psi. In particular, if Ψ(0)=0\Psi(0)=0, then Ψ0\Psi\equiv 0, and therefore u3(t)=Λ(t)(u1(t)+u2(t)21)2(μ+1)u2(t)u_{3}(t)=\Lambda(t)(u_{1}(t)+u_{2}(t)^{2}-1)-2(\mu+1)u_{2}(t) holds for all tt along such a characteristic curve. If, in addition, Λ(t0)=0\Lambda(t_{0})=0 for some t0t_{0}\in\mathbb{R}, then we get u3(t0)=2(μ+1)u2(t0)u_{3}(t_{0})=-2(\mu+1)u_{2}(t_{0}). According to Lemma 1, Λ(t0)=0\Lambda(t_{0})=0 also implies that λ=u3(t0)+μ(μ+1)\lambda=u_{3}(t_{0})+\mu(\mu+1) is an eigenvalue of the Coulomb spheroidal wave equation (2) for the parameters γ2=u1(t0)\gamma^{2}=u_{1}(t_{0}) and β=u3(t0)2(μ+1)u2(t0)=0\beta=-u_{3}(t_{0})-2(\mu+1)u_{2}(t_{0})=0, where β=0\beta=0 in turn reduces the CSWE to the ordinary (angular) spheroidal wave equation. All in all, the above considerations provide an alternative method for the numerical computation of spheroidal eigenvalues:

Theorem 4.

Suppose that u1(t)u_{1}(t), u2(t)u_{2}(t) and Λ(t)\Lambda(t) are solutions of the differential system

du1dt=2u1(t)du2dt=2(Λ(t)+1)(u1(t)+u2(t)2)(2μ+1)u2(t)dΛdt=(1+2μ2Λ(t)u2(t))(Λ(t)+2)2μ\begin{split}\frac{\mathrm{d}u_{1}}{\mathrm{d}t}&=2u_{1}(t)\\ \frac{\mathrm{d}u_{2}}{\mathrm{d}t}&=2(\Lambda(t)+1)(u_{1}(t)+u_{2}(t)^{2})-(2\mu+1)u_{2}(t)\\ \frac{\mathrm{d}\Lambda}{\mathrm{d}t}&=(1+2\mu-2\Lambda(t)u_{2}(t))(\Lambda(t)+2)-2\mu\end{split} (18)

on an interval 0I0\in I\subset\mathbb{R} satisfying u1(0)=au_{1}(0)=a, u2(0)=bu_{2}(0)=b, Λ(0)=Λ0\Lambda(0)=\Lambda_{0} with some given initial values a,ba,b\in\mathbb{R} such that Θ(Λ0,a,b,(a+b21)Λ02(μ+1)b)=0\Theta(\Lambda_{0},a,b,(a+b^{2}-1)\Lambda_{0}-2(\mu+1)b)=0. If Λ(t0)=0\Lambda(t_{0})=0 holds for some t0It_{0}\in I, then λ=(μ+1)(μ2u2(t0))\lambda=(\mu+1)(\mu-2u_{2}(t_{0})) is an eigenvalue of the angular spheroidal wave equation (1) for the parameter γ2=u1(t0)\gamma^{2}=u_{1}(t_{0}).

Proof.

If we set c:=(a+b21)Λ02(μ+1)bc:=(a+b^{2}-1)\Lambda_{0}-2(\mu+1)b, then Θ(Λ0,a,b,c)=0\Theta(\Lambda_{0},a,b,c)=0, and hence Λ0\Lambda_{0} is an isolated eigenvalue of the operator T(𝐮0)T(\mathbf{u}_{0}) for 𝐮0:=(a,b,c)\mathbf{u}_{0}:=(a,b,c). According to Theorem 2, we can find an analytic continuation Λ0:U\Lambda_{0}:U\longrightarrow\mathbb{R} on some domain U3U\subset\mathbb{R}^{3} with 𝐮0U\mathbf{u}_{0}\in U, such that Λ0(𝐮)\Lambda_{0}(\mathbf{u}) is an eigenvalue of T(𝐮)T(\mathbf{u}) satisfying, in addition, the partial differential equation (5). Defining u3(t):=Λ(t)(u1(t)+u2(t)21)2(μ+1)u2(t)u_{3}(t):=\Lambda(t)(u_{1}(t)+u_{2}(t)^{2}-1)-2(\mu+1)u_{2}(t), then we can write the second equation in (18) as

du2dt=(Λ(t)+2)(u1(t)+u2(t)2)+u2(t)+u3(t)+Λ(t)\frac{\mathrm{d}u_{2}}{\mathrm{d}t}=(\Lambda(t)+2)(u_{1}(t)+u_{2}(t)^{2})+u_{2}(t)+u_{3}(t)+\Lambda(t)

Moreover, the differential equations (18) imply

du3dt\displaystyle\frac{\mathrm{d}u_{3}}{\mathrm{d}t} =dΛdt(u1(t)+u2(t)21)+Λ(t)(du1dt+2u2(t)du2dt)2(μ+1)du2dt\displaystyle=\frac{\mathrm{d}\Lambda}{\mathrm{d}t}(u_{1}(t)+u_{2}(t)^{2}-1)+\Lambda(t)\left(\frac{\mathrm{d}u_{1}}{\mathrm{d}t}+2u_{2}(t)\frac{\mathrm{d}u_{2}}{\mathrm{d}t}\right)-2(\mu+1)\frac{\mathrm{d}u_{2}}{\mathrm{d}t}
=(2Λ(t)+u3(t)+2)(2Λ(t)u2(t)2μ1)2(μ+1)(u1(t)+u2(t)2)+2μ\displaystyle=(2\Lambda(t)+u_{3}(t)+2)(2\Lambda(t)u_{2}(t)-2\mu-1)-2(\mu+1)(u_{1}(t)+u_{2}(t)^{2})+2\mu

Hence, the functions u1(t)u_{1}(t), u2(t)u_{2}(t), u3(t)u_{3}(t), Λ(t)\Lambda(t) satisfy the characteristic equations (17), and from the initial values u1(0)=au_{1}(0)=a, u2(0)=bu_{2}(0)=b, u3(0)=cu_{3}(0)=c, Λ(0)=Λ0=Λ0(a,b,c)\Lambda(0)=\Lambda_{0}=\Lambda_{0}(a,b,c) it follows that this characteristic curve lies completely on the integral surface Λ0(𝐮)\Lambda_{0}(\mathbf{u}). In particular, Λ(t)=Λ0(u1(t),u2(t),u3(t))\Lambda(t)=\Lambda_{0}(u_{1}(t),u_{2}(t),u_{3}(t)), and therefore Λ(t)\Lambda(t) is an eigenvalue of (4) for the parameter values u1(t)u_{1}(t), u2(t)u_{2}(t), u3(t)u_{3}(t). Finally, if Λ(t0)=0\Lambda(t_{0})=0, then u3(t0)=2(μ+1)u2(t0)u_{3}(t_{0})=-2(\mu+1)u_{2}(t_{0}) implies β=u3(t0)2(μ+1)u2(t0)=0\beta=-u_{3}(t_{0})-2(\mu+1)u_{2}(t_{0})=0, and λ=(μ+1)(μ2u2(t0))\lambda=(\mu+1)(\mu-2u_{2}(t_{0})) is an eigenvalue of (1) for the spheroidal parameter γ2=u1(t0)\gamma^{2}=u_{1}(t_{0}). ∎

We may, for example, choose b=0b=0 in Theorem 4, and then we need to find a zero Λ0\Lambda_{0} of the function Θ~(Λ):=Θ(Λ,a,0,(a1)Λ)\tilde{\Theta}(\Lambda):=\Theta(\Lambda,a,0,(a-1)\Lambda) for some given initial value aa\in\mathbb{R}. Then we have to solve the differential system (18) with u1(0)=au_{1}(0)=a, u2(0)=0u_{2}(0)=0, Λ(0)=Λ0\Lambda(0)=\Lambda_{0} on some interval II with 0I0\in I by an appropriate numerical method, and we follow the path of this characteristic curve until we get Λ(t0)=0\Lambda(t_{0})=0 for some t0It_{0}\in I. Finally, we arrive at an eigenvalue λ=(μ+1)(μ2u2(t0))\lambda=(\mu+1)(\mu-2u_{2}(t_{0})) of the angular spheroidal wave equation (1) for γ2=u1(t0)\gamma^{2}=u_{1}(t_{0}). In the appendix, this method is illustrated by a concrete numerical example, namely for the values μ=1\mu=1 and a=5a=5.

The differential system (17) in Theorem 4 may also be regarded as the characteristic equation of a PDE for a function Λ(u1,u2)\Lambda(u_{1},u_{2}) with only two variables u1u_{1} and u2u_{2}. The question arises whether one can associate a more simple linear Hamiltonian system with the spheroidal wave equation (β=0\beta=0), which then yields a PDE for the eigenvalues that is less complicated than (5). In the remainder of this paper we address this problem, where we restrict our consideration to the case of the prolate spheroidal wave equation (2) with β=0\beta=0 and γ2>0\gamma^{2}>0. For this purpose, we fix u3=Λ(u1+u221)2(μ+1)u2u_{3}=\Lambda(u_{1}+u_{2}^{2}-1)-2(\mu+1)u_{2} in (4) resp. (8), and we assume u1=γ2>0u_{1}=\gamma^{2}>0, in which case also u1+u22>0u_{1}+u_{2}^{2}>0 holds.

First, let us introduce new parameters that make the coefficient matrix of the differential system (8) appear more symmetric, and these are

v1:=u2u1+u22,v2:=u1+u22,ζ:=Λ(u1+u22)2αu2u1+u22v_{1}:=\frac{u_{2}}{\sqrt{u_{1}+u_{2}^{2}}},\quad v_{2}:=\sqrt{u_{1}+u_{2}^{2}},\quad\zeta:=\frac{\Lambda(u_{1}+u_{2}^{2})-2\alpha u_{2}}{\sqrt{u_{1}+u_{2}^{2}}} (19)

where v1(1,1)v_{1}\in(-1,1) and v2(0,)v_{2}\in(0,\infty). Note that conversely

u1+u22=v22,u2=v1v2,Λ=ζ+2αv1v2,u3+Λ=(ζ2αv1)v2u_{1}+u_{2}^{2}=v_{2}^{2},\quad u_{2}=v_{1}v_{2},\quad\Lambda=\frac{\zeta+2\alpha v_{1}}{v_{2}},\quad u_{3}+\Lambda=(\zeta-2\alpha v_{1})v_{2}

By means of the transformation

y^(z)=(100v2)y(z)\hat{y}(z)=\begin{pmatrix}1&0\\[4.30554pt] 0&v_{2}\end{pmatrix}y(z) (20)

the differential system (4) resp. (8) is equivalent to the 2×22\times 2 system

y^(z)=(1z(αζ2αv10α)+11z(α0ζ2αv1α)+(2v1v22v22v22v1v2))y^(z)\hat{y}^{\prime}(z)=\left(\frac{1}{z}\begin{pmatrix}-\alpha&\zeta-2\alpha v_{1}\\[4.30554pt] 0&\alpha\end{pmatrix}+\frac{1}{1-z}\begin{pmatrix}\alpha&0\\[4.30554pt] -\zeta-2\alpha v_{1}&-\alpha\end{pmatrix}+\begin{pmatrix}2v_{1}v_{2}&2v_{2}\\[4.30554pt] -2v_{2}&-2v_{1}v_{2}\end{pmatrix}\right)\hat{y}(z)

which can be written in the form τ^y^=ζy^\hat{\tau}\hat{y}=\zeta\hat{y}, where

τ^y^:=W(z)1(Jy^(z)H^(z)y^(z)),H^(z):=(2v2+2αv11zαz+α1z+2v1v2αz+α1z+2v1v22v22αv1z)\hat{\tau}\hat{y}:=W(z)^{-1}\left(J\hat{y}^{\prime}(z)-\hat{H}(z)\hat{y}(z)\right),\quad\hat{H}(z):=\begin{pmatrix}2v_{2}+\frac{2\alpha v_{1}}{1-z}&-\frac{\alpha}{z}+\frac{\alpha}{1-z}+2v_{1}v_{2}\\[4.30554pt] -\frac{\alpha}{z}+\frac{\alpha}{1-z}+2v_{1}v_{2}&2v_{2}-\frac{2\alpha v_{1}}{z}\end{pmatrix}

By a similar reasoning as for (6), it follows that the differential expression τ^\hat{\tau} is in the limit point case at z=0z=0 and z=1z=1. Hence, the closure of its associated minimal operator is the only self-adjoint realization of τ^\hat{\tau} in the Hilbert space LW2((0,1),2)\mathrm{L}_{W}^{2}((0,1),\mathbb{C}^{2}). We denote this self-adjoint operator by T^=T^(v1,v2)\hat{T}=\hat{T}(v_{1},v_{2}). By virtue of (20), we have y,yW<\langle y,y\rangle_{W}<\infty if and only if y^,y^W<\langle\hat{y},\hat{y}\rangle_{W}<\infty. Hence, Λ\Lambda is an eigenvalue of T(𝐮)T(\mathbf{u}) if and only if ζ\zeta is an eigenvalue of T^(v1,v2)\hat{T}(v_{1},v_{2}).

Now, like in Theorem 2, suppose that ζ=ζ(v1,v2)\zeta=\zeta(v_{1},v_{2}) is an eigenvalue of T^(v1,v2)\hat{T}(v_{1},v_{2}), which depends analytically on the parameters (v1,v2)V(v_{1},v_{2})\in V in some domain V(1,1)×(0,)V\subset(-1,1)\times(0,\infty). If we introduce the matrix function

G^(z,ζ,v1,v2):=(v2zv1v2z+α(v121)12v1v2v1v2z+α(v121)12v1v2v2(z1))\hat{G}(z,\zeta,v_{1},v_{2}):=\begin{pmatrix}v_{2}z&v_{1}v_{2}z+\alpha(v_{1}^{2}-1)-\frac{1}{2}v_{1}v_{2}\\[4.30554pt] v_{1}v_{2}z+\alpha(v_{1}^{2}-1)-\frac{1}{2}v_{1}v_{2}&v_{2}(z-1)\end{pmatrix}

then a direct calculation shows that the deformation equation

G^z+(ζW+H^)JG^G^J(ζW+H^)=f^1Hv1+f^2Hv2+g^W\frac{\partial\hat{G}}{\partial z}+(\zeta W+\hat{H})J\hat{G}-\hat{G}J(\zeta W+\hat{H})=\hat{f}_{1}\,\frac{\partial H}{\partial v_{1}}+\hat{f}_{2}\,\frac{\partial H}{\partial v_{2}}+\hat{g}W

is fulfilled with the functions

f^1:=(ζ+v2)(1v12),f^2:=(ζv1+v1v2+12)v2,g^:=4α2v1(1v12)ζv1v2\displaystyle\hat{f}_{1}:=(\zeta+v_{2})(1-v_{1}^{2}),\quad\hat{f}_{2}:=(\zeta v_{1}+v_{1}v_{2}+\tfrac{1}{2})v_{2},\quad\hat{g}:=4\alpha^{2}v_{1}(1-v_{1}^{2})-\zeta v_{1}v_{2}

From [4, Theorem 2.2] it follows that the eigenvalues ζ(v1,v2)\zeta(v_{1},v_{2}) satisfy the first-order quasilinear PDE

(ζ+v2)(1v12)ζv1+(ζv1+v1v2+12)v2ζv2=4α2v1(1v12)ζv1v2(\zeta+v_{2})(1-v_{1}^{2})\frac{\partial\zeta}{\partial v_{1}}+(\zeta v_{1}+v_{1}v_{2}+\tfrac{1}{2})v_{2}\frac{\partial\zeta}{\partial v_{2}}=4\alpha^{2}v_{1}(1-v_{1}^{2})-\zeta v_{1}v_{2} (21)

Using Lemma 1 with u3=Λ(u1+u221)2(μ+1)u2u_{3}=\Lambda(u_{1}+u_{2}^{2}-1)-2(\mu+1)u_{2}, we obtain that λ=Λ(u1+u221)+(μ+1)(μ2u2)\lambda=\Lambda(u_{1}+u_{2}^{2}-1)+(\mu+1)(\mu-2u_{2}) is an eigenvalue of (2) for the parameters γ2=u1\gamma^{2}=u_{1}, β=Λ(u1+u221)\beta=-\Lambda(u_{1}+u_{2}^{2}-1) if and only if Λ=0\Lambda=0 is an eigenvalue of T(𝐮)T(\mathbf{u}). In this case λ=(μ+1)(μ2u2)\lambda=(\mu+1)(\mu-2u_{2}) and β=0\beta=0, i.e., the CSWE reduces to the angular spheroidal wave equation (1). Note that Λ=0\Lambda=0 is equivalent to ζ=2αv1=(μ+1)v1\zeta=-2\alpha v_{1}=-(\mu+1)v_{1}. Since u1=(1v12)v22u_{1}=(1-v_{1}^{2})v_{2}^{2} and u2=v1v2u_{2}=v_{1}v_{2}, it follows that λ=(μ+1)(μ2v1v2)\lambda=(\mu+1)(\mu-2v_{1}v_{2}) is an eigenvalue of (1) for γ2=(1v12)v22\gamma^{2}=(1-v_{1}^{2})v_{2}^{2} if and only if ζ=(μ+1)v1\zeta=-(\mu+1)v_{1} is an eigenvalue of T^(v1,v2)\hat{T}(v_{1},v_{2}).

Finally, let us simplify the PDE (21) even further. To this end, we need to modify the parameters introduced in (19) once more. We set

v1=tanhu,v2=etcoshu,ζ=ωetcoshuv_{1}=\tanh u,\quad v_{2}=\mathrm{e}^{t}\cosh u,\quad\zeta=\omega-\mathrm{e}^{t}\cosh u

With these parameters (t,u)2(t,u)\in\mathbb{R}^{2} we convert the eigenvalue problem τ^y^=ζy^\hat{\tau}\hat{y}=\zeta\hat{y} to

W(z)1(Jy^(z)Φ(z,t,u)y^(z))=ωy^(z),0<z<1W(z)^{-1}\left(J\hat{y}^{\prime}(z)-\Phi(z,t,u)\hat{y}(z)\right)=\omega\hat{y}(z),\quad 0<z<1 (22)

where now ω\omega is the eigenvalue parameter, and

Φ(z,t,u):=(2etcoshu+2αtanhuetcoshu1zαz+α1z+2etsinhuαz+α1z+2etsinhu2etcoshu2αtanhu+etcoshuz)\Phi(z,t,u):=\begin{pmatrix}2\mathrm{e}^{t}\cosh u+\frac{2\alpha\tanh u-\mathrm{e}^{t}\cosh u}{1-z}&-\frac{\alpha}{z}+\frac{\alpha}{1-z}+2\mathrm{e}^{t}\sinh u\\[4.30554pt] -\frac{\alpha}{z}+\frac{\alpha}{1-z}+2\mathrm{e}^{t}\sinh u&2\mathrm{e}^{t}\cosh u-\frac{2\alpha\tanh u+\mathrm{e}^{t}\cosh u}{z}\end{pmatrix}
Theorem 5.

Let ω=ω(t,u)\omega=\omega(t,u) be an eigenvalue of the linear Hamiltonian system (22), and suppose that ω(t,u)\omega(t,u) depends analytically on the parameters (t,u)S(t,u)\in S in some domain S2S\subset\mathbb{R}^{2}. Then ω(t,u)\omega(t,u) solves the quasilinear partial differential equation

ωt+2ωωu=2(μ+1)2tanhucosh2u+e2tsinh(2u)+etcoshu\frac{\partial\omega}{\partial t}+2\omega\cdot\frac{\partial\omega}{\partial u}=2(\mu+1)^{2}\frac{\tanh u}{\cosh^{2}u}+\mathrm{e}^{2t}\sinh(2u)+\mathrm{e}^{t}\cosh u (23)

on SS. Moreover, λ=(μ+1)(μ2etsinhu)\lambda=(\mu+1)(\mu-2\mathrm{e}^{t}\sinh u) is an eigenvalue of the spheroidal wave equation (1) for γ=et\gamma=\mathrm{e}^{t} if and only if the equation ω(t,u)=etcoshu(μ+1)tanhu\omega(t,u)=\mathrm{e}^{t}\cosh u-(\mu+1)\tanh u is satisfied for some (t,u)S(t,u)\in S.

Proof.

Note that ω=ζ+v2\omega=\zeta+v_{2}, and that v1v_{1} does not depend on tt. Hence,

ωt\displaystyle\frac{\partial\omega}{\partial t} =ζv2v2t+v2t=etcoshuζv2+etcoshu\displaystyle=\frac{\partial\zeta}{\partial v_{2}}\frac{\partial v_{2}}{\partial t}+\frac{\partial v_{2}}{\partial t}=\mathrm{e}^{t}\cosh u\,\frac{\partial\zeta}{\partial v_{2}}+\mathrm{e}^{t}\cosh u
ωu\displaystyle\frac{\partial\omega}{\partial u} =ζv1v1u+ζv2v2u+v2u=1cosh2uζv1+etsinhuζv2+etsinhu\displaystyle=\frac{\partial\zeta}{\partial v_{1}}\frac{\partial v_{1}}{\partial u}+\frac{\partial\zeta}{\partial v_{2}}\frac{\partial v_{2}}{\partial u}+\frac{\partial v_{2}}{\partial u}=\frac{1}{\cosh^{2}u}\,\frac{\partial\zeta}{\partial v_{1}}+\mathrm{e}^{t}\sinh u\,\frac{\partial\zeta}{\partial v_{2}}+\mathrm{e}^{t}\sinh u

and therefore

ωt+2ωωu\displaystyle\frac{\partial\omega}{\partial t}+2\omega\cdot\frac{\partial\omega}{\partial u} =2ωcosh2uζv1+(etcoshu+2ωetsinhu)ζv2+etcoshu+2ωetsinhu\displaystyle=\frac{2\omega}{\cosh^{2}u}\,\frac{\partial\zeta}{\partial v_{1}}+(\mathrm{e}^{t}\cosh u+2\omega\,\mathrm{e}^{t}\sinh u)\frac{\partial\zeta}{\partial v_{2}}+\mathrm{e}^{t}\cosh u+2\omega\,\mathrm{e}^{t}\sinh u

From

2ωcosh2u=2(ζ+v2)(1v12),etcoshu+2ωetsinhu=2(ζv1+v1v2+12)v2\frac{2\omega}{\cosh^{2}u}=2(\zeta+v_{2})(1-v_{1}^{2}),\quad\mathrm{e}^{t}\cosh u+2\omega\,\mathrm{e}^{t}\sinh u=2(\zeta v_{1}+v_{1}v_{2}+\tfrac{1}{2})v_{2}

and the partial differential equation (21) it follows that

ωt+2ωωu\displaystyle\frac{\partial\omega}{\partial t}+2\omega\cdot\frac{\partial\omega}{\partial u} =2(ζ+v2)(1v12)ζv1+2(ζv1+v1v2+12)v2ζv2+2(ζv1+v1v2+12)v2\displaystyle=2(\zeta+v_{2})(1-v_{1}^{2})\frac{\partial\zeta}{\partial v_{1}}+2(\zeta v_{1}+v_{1}v_{2}+\tfrac{1}{2})v_{2}\frac{\partial\zeta}{\partial v_{2}}+2(\zeta v_{1}+v_{1}v_{2}+\tfrac{1}{2})v_{2}
=8α2v1(1v12)2ζv1v2+2(ζv1+v1v2+12)v2\displaystyle=8\alpha^{2}v_{1}(1-v_{1}^{2})-2\zeta v_{1}v_{2}+2(\zeta v_{1}+v_{1}v_{2}+\tfrac{1}{2})v_{2}
=8α2v1(1v12)+(2v1v2+1)v2=2(μ+1)2tanhucosh2u+2e2tsinhucoshu+etcoshu\displaystyle=8\alpha^{2}v_{1}(1-v_{1}^{2})+(2v_{1}v_{2}+1)v_{2}=2(\mu+1)^{2}\frac{\tanh u}{\cosh^{2}u}+2\mathrm{e}^{2t}\sinh u\cosh u+\mathrm{e}^{t}\cosh u

and this is exactly the PDE (23). As we have already noted, λ=(μ+1)(μ2v1v2)=(μ+1)(μ2etsinhu)\lambda=(\mu+1)(\mu-2v_{1}v_{2})=(\mu+1)(\mu-2\mathrm{e}^{t}\sinh u) is an eigenvalue of (1) for γ2=(1v12)v22=e2t\gamma^{2}=(1-v_{1}^{2})v_{2}^{2}=\mathrm{e}^{2t} if and only if ω=ζ+v2=(μ+1)tanhu+etcoshu\omega=\zeta+v_{2}=-(\mu+1)\tanh u+\mathrm{e}^{t}\cosh u is an eigenvalue of T^(v1,v2)\hat{T}(v_{1},v_{2}) and hence of (22). ∎

It is worth mentioning that the partial differential equation (23) is basically a forced inviscid Burgers’ equation. In fact, if we set t=s2t=\frac{s}{2}, then (23) becomes

ωs+ωωu=(μ+1)2tanhucosh2u+12essinh(2u)+12es2coshu\frac{\partial\omega}{\partial s}+\omega\cdot\frac{\partial\omega}{\partial u}=(\mu+1)^{2}\frac{\tanh u}{\cosh^{2}u}+\tfrac{1}{2}\mathrm{e}^{s}\sinh(2u)+\tfrac{1}{2}\mathrm{e}^{\frac{s}{2}}\cosh u

and this quasilinear PDE can be written in the form ωs+ωωu=f(s,u)\frac{\partial\omega}{\partial s}+\omega\cdot\frac{\partial\omega}{\partial u}=f(s,u). The forced inviscid Burgers’ equation is commonly encountered in fluid mechanics or gas dynamics, where it is used, for example, to study nonlinear interactions of dispersive waves. It is quite surprising that this PDE is related to the linear Hamiltonian system (22) and thus as well to the eigenvalues of the spheroidal wave equation.

4. Conclusion

Calculating the eigenvalues of a spheroidal wave equation is known to be a tricky task. There is an ongoing effort in finding new techniques to compute these eigenvalues in a convenient way and with high precision (see, for example, [1], [9], [10], [11], [12], [13] and the references therein). In the present work, we used an approach via a 2×22\times 2 differential system. Apart from some numerical issues, we have mainly studied the analytical dependence of the eigenvalues on the parameters γ\gamma and β\beta (or u1u_{1}, u2u_{2}, u3u_{3}, respectively). Results such as those in Theorem 5 and relations like the PDE (5) may help to obtain new insights into the analytic properties of the angular or Coulomb spheroidal eigenvalues. The CSWE, in turn, is a special case of the confluent Heun differential equation (CHE), which is likewise a subject of current research. The methods presented here may also apply to the CHE and to even more general eigenvalue problems.

Appendix: A numerical example

In section 3, as an application of Theorem 4, it has been suggested to compute an eigenvalue of the spheroidal wave equation by starting from a zero of the function Θ~\tilde{\Theta} and then following a characteristic curve of the PDE (5) until a zero of the function Λ(t)\Lambda(t) is located. As a concrete numerical example, let us consider the function Θ~(Λ):=Θ(Λ,a,0,(a1)Λ)\tilde{\Theta}(\Lambda):=\Theta(\Lambda,a,0,(a-1)\Lambda) in the case μ=1\mu=1 for a=5a=5. The limit formula in Lemma 3 provides the function shown in Fig. 1. It has a zero at Λ0=0.8417200168449013\Lambda_{0}=-0.8417200168449013, which can be determined, for instance, with the secant method. Here, all results are displayed rounded to 1616 decimal places, while the calculations were internally done with higher precision. Solving the differential system (18) for the initial values u1(0)=5u_{1}(0)=5, u2(0)=0u_{2}(0)=0, Λ(0)=Λ0\Lambda(0)=\Lambda_{0} by means of the Runge-Kutta method, we obtain u1(t)u_{1}(t), u2(t)u_{2}(t) and Λ(t)\Lambda(t). The function Λ(t)\Lambda(t) is illustrated in Fig. 2; it has a zero at t0=0.2793371978706399t_{0}=0.2793371978706399. Evaluating u1u_{1}, u2u_{2} at t0t_{0} finally gives the eigenvalue λ=2(12u2(t0))=5.2736106330552739\lambda=2(1-2u_{2}(t_{0}))=-5.2736106330552739 for the spheroidal wave equation (1) with parameter γ2=u1(t0)=8.7417666942941543\gamma^{2}=u_{1}(t_{0})=8.7417666942941543.

Refer to caption
Figure 1. The function Θ~(Λ)\tilde{\Theta}(\Lambda) for μ=1\mu=1 and a=5a=5 with a zero at Λ0=0.841720\Lambda_{0}=-0.841720\ldots
Refer to caption
Figure 2. The solution Λ(t)\Lambda(t) of (18) for μ=1\mu=1 starting from u1(0)=5u_{1}(0)=5, u2(0)=0u_{2}(0)=0 and Λ(0)=0.841720\Lambda(0)=-0.841720\ldots

References

  • [1] P. E. Falloon, Theory and computation of spheroidal harmonics with general arguments, Master’s thesis, The University of Western Australia, Department of Physics (2001).
  • [2] T. Kereselidze, J. F. Ogilvie, The hydrogen-atom problem and Coulomb Sturmian functions in spheroidal coordinates, Vol. 77 of Advances in Quantum Chemistry, Academic Press, 2018, pp. 391–421. doi:10.1016/bs.aiq.2018.02.002.
  • [3] D. Batic, H. Schmid, M. Winklmeier, On the eigenvalues of the Chandrasekhar-Page angular equation, Journal of Mathematical Physics 46 (1) (2005) 012504. doi:10.1063/1.1818720.
  • [4] H. Schmid, On the deformation of linear Hamiltonian systems, Journal of Mathematical Analysis and Applications 499 (2) (2021) 125051. doi:10.1016/j.jmaa.2021.125051.
  • [5] J. Weidmann, Spectral Theory of Ordinary Differential Operators, vol. 1258 of Lecture Notes in Mathematics, Springer, Berlin – Heidelberg, 1987. doi:10.1007/BFb0077960.
  • [6] H. Sun, Y. Shi, Self-adjoint extensions for linear Hamiltonian systems with two singular endpoints, Journal of Functional Analysis 259 (2010) 2003–2027. doi:10.1016/j.jfa.2010.06.008.
  • [7] J. Meixner, F. W. Schäfke, Mathieusche Funktionen und Sphäroidfunktionen, Springer, New York, 1954, in German. doi:10.1007/978-3-662-00941-3.
  • [8] T. Kato, Perturbation theory for linear operators, 2nd Edition, vol. 1258 of Lecture Notes in Mathematics, Springer, Berlin – Heidelberg – New York, 1995. doi:10.1007/978-3-642-66282-9.
  • [9] H. Schmid, Computation of the eigenvalues for the angular and Coulomb spheroidal wave equation, Applied Numerical Mathematics 185 (2023) 101–119. doi:10.1016/j.apnum.2022.11.018.
  • [10] C. Flammer, Spheroidal Wave Functions, Stanford University Press, Stanford, CA, 1957.
  • [11] D. B. Hodge, Eigenvalues and eigenfunctions of the spheroidal wave equation, Journal of Mathematical Physics 11 (8) (1970) 2308. doi:10.1063/1.1665398.
  • [12] P. Kirby, Calculation of spheroidal wave functions, Computer Physics Communications 175 (7) (2006) 465–472. doi:10.1016/j.cpc.2006.06.006.
  • [13] S. L. Skorokhodov, Calculation of eigenvalues and eigenfunctions of the Coulomb spheroidal wave equation, Matematicheskoe modelirovanie 27 (7) (2015) 111–116, in Russian.