1: \documentclass[a4paper]{article}\usepackage{anysize,times}\usepackage{rsflash}
2:
3: \usepackage{cite}\usepackage{epsfig}
4:
5: \pagestyle{plain}%\pagestyle{empty}
6: \marginsize{1in}{1in}{1in}{1in}
7:
8: \sloppy\flushbottom
9:
10: \begin{document}\begin{titlepage}
11: \large\renewcommand{\thefootnote}{\normalsize\fnsymbol{footnote}}\hfill
12: \begin{tabular}{l}HEPHY-PUB 789/04\\%UWThPh-2004-??\\
13: July 2004\end{tabular}
14: \\[2cm]\Huge\begin{center}{\Huge\bf FACETS OF THE\\SPINLESS SALPETER
15: EQUATION}\\\vspace{2cm}\Large{\bf Wolfgang LUCHA\footnote[1]{\large\ {\em
16: E-mail address\/}: wolfgang.lucha@oeaw.ac.at}}\\[.3cm]\large Institut f\"ur
17: Hochenergiephysik,\\\"Osterreichische Akademie der Wissenschaften,\\
18: Nikolsdorfergasse 18, A-1050 Wien, Austria\\[1cm]\Large{\bf Franz
19: F.~SCH\"OBERL\footnote[2]{\large\ {\em E-mail address\/}:
20: franz.schoeberl@univie.ac.at}}\\[.3cm]\large Institut f\"ur Theoretische
21: Physik, Universit\"at Wien,\\Boltzmanngasse 5, A-1090 Wien, Austria
22:
23: \vspace{1cm}{\large\bf Abstract}\end{center}\large The spinless Salpeter
24: equation represents the simplest, and most straightforward, generalization of
25: the Schr\"odinger equation of standard nonrelativistic quantum theory towards
26: the inclusion of relativistic kinematics. Moreover, it can be also regarded
27: as a well-defined approximation to the Bethe--Salpeter formalism for
28: descriptions of bound states in relativistic quantum field theories. The
29: corresponding Hamiltonian is, in contrast to all Schr\"odinger operators, a
30: nonlocal operator. Because of the nonlocality, constructing analytical
31: solutions for such kind of equation of motion proves difficult. In view of
32: this, different sophisticated techniques have been developed in order to
33: extract rigorous analytical information about these solutions. This review
34: introduces some of these methods and compares their significance by
35: application to interactions relevant in physics.
36:
37: \vfill
38:
39: \noindent{\em PACS numbers\/}: 03.65.Ge, 03.65.Pm, 11.10.St
40: \renewcommand{\thefootnote}{\arabic{footnote}}\normalsize\end{titlepage}
41:
42: \twocolumn
43:
44: \marginsize{1in}{1in}{1in}{1in}
45:
46: \section{Bethe--Salpeter Formalism in the ``Instantaneous Approximation''}
47: Within quantum field theory, the appropriate framework for the description of
48: bound states is the Bethe--Salpeter formalism \cite{BSE}. Therein, all bound
49: states of two particles (in fact, of any two fermionic constituents) are
50: governed by the {\em homogeneous Bethe--Salpeter equation\/}. Here we are
51: interested in a particular well-defined approximation to this formalism,
52: obtained by several simplifying steps:\begin{enumerate}\item The {\em
53: instantaneous approximation\/}, neglecting any retardation effect, considers
54: all interactions of the (two) bound-state constituents in their static
55: limit.\item The additional assumption that all the bound-state constituents
56: propagate as free particles with some effective mass $m$ yields the {\em
57: Salpeter equation\/} \cite{SE}.\item A disregard of all of their spin degrees
58: of freedom focuses on the treatment of scalar bound particles.\item In
59: technical respect, the {\em canonical transformation\/}
60: \begin{equation}\mbox{\boldmath{$x$}}\to\lambda\,\mbox{\boldmath{$x$}}\
61: ,\quad\mbox{\boldmath{$p$}}\to\frac{\mbox{\boldmath{$p$}}}{\lambda}
62: \label{Eq:scale}\end{equation}of position ($\mbox{\boldmath{$x$}}$) and
63: momentum ($\mbox{\boldmath{$p$}}$) variables casts in the case of particles
64: of equal mass $m$ for a scale factor $\lambda=2$ this approach into
65: one-particle form.\end{enumerate}(For more details of the derivation,
66: consult, for instance, Refs.~\cite{Lucha91:BSQ,Resag94,Kopaleishvili01} and
67: references therein.) Refraining from the nonrelativistic limit, we get the
68: (nonlocal!) Hamiltonian\begin{equation}H=T+V\ .\label{Eq:SRH}\end{equation}
69: This operator is composed of the ``square-root operator'' $T$ of the
70: relativistically correct expression for the kinetic or free energy of a
71: particle of mass $m$ and momentum~$\mbox{\boldmath{$p$}},$\begin{equation}
72: T=T(\mbox{\boldmath{$p$}})\equiv\sqrt{\mbox{\boldmath{$p$}}^2+m^2}\
73: ,\label{Eq:RKE}\end{equation}and a (coordinate-dependent) static interaction
74: potential$$V=V(\mbox{\boldmath{$x$}})\ ;$$frequently, the potential
75: $V(\mbox{\boldmath{$x$}})$ is assumed to be a central potential that depends
76: merely on the radial coordinate $r$:$$V=V(r)\ ,\quad
77: r\equiv|\mbox{\boldmath{$x$}}|\ .$$The eigenvalue equation for this
78: particular Hamiltonian,\begin{equation}H\,|\chi_k\rangle=E_k\,|\chi_k\rangle\
79: ,\quad k=0,1,2,\dots\ ,\label{Eq:EVE}\end{equation}defining a complete system
80: of Hilbert-space eigenstates $|\chi_k\rangle$ of $H$ corresponding to its
81: (energy) eigenvalues
82: $E_k,$$$E_k\equiv\frac{\langle\chi_k|\,H\,|\chi_k\rangle}
83: {\langle\chi_k|\chi_k\rangle}\ ,\quad k=0,1,2,\dots\ ,$$is commonly known as
84: the ``spinless Salpeter equation.''
85:
86: \section{The Relativistic Virial Theorem}\label{Sec:RVT}Useful general
87: statements about the solutions of explicit or implicit eigenvalue equations
88: may be proved with the help of virial theorems obtained by generalization
89: \cite{Lucha:RVT} of the well-known result of nonrelativistic quantum theory.
90: (Ref.~\cite{Lucha:RVTs} is a brief review of relativistic virial theorems.)
91: For eigenvalue equations of the form (\ref{Eq:EVE}), the derivations of such
92: virial theorems can be traced back to the (trivial) observation that
93: expectation values taken with respect to given eigenstates $|\chi_k\rangle$
94: of $H$ --- or matrix elements taken with respect to arbitrary pairs of {\em
95: degenerate\/} eigenstates, $|\chi_i\rangle$ and $|\chi_j\rangle,$ of $H,$
96: i.~e., eigenstates satisfying $E_i=E_j$ --- of the commutators $[G,H]$ of the
97: operator $H$ and any other operator $G$ (the domain of which must be assumed
98: to contain the domain of $H$) clearly vanish. Suppressing the subscript that
99: enumerates the eigenstates, this
100: means\begin{equation}\langle\chi|\,[G,H]\,|\chi\rangle=0\ .\label{Eq:EV(C)}
101: \end{equation}For the symmetrized, self-adjoint generator of dilations,
102: \begin{equation}G\equiv\mbox{$\frac{1}{2}$}\,(\mbox{\boldmath{$x$}}\cdot
103: \mbox{\boldmath{$p$}}+\mbox{\boldmath{$p$}}\cdot\mbox{\boldmath{$x$}})\
104: ,\label{Eq:DilGen}\end{equation}and $H$ of the form (\ref{Eq:SRH}) their
105: commutator $[G,H]$ becomes$$[G,H]={\rm
106: i}\left[\mbox{\boldmath{$p$}}\cdot\frac{\partial\,T}
107: {\partial\mbox{\boldmath{$p$}}}(\mbox{\boldmath{$p$}})-\mbox{\boldmath{$x$}}
108: \cdot\frac{\partial\,V}{\partial\mbox{\boldmath{$x$}}}(\mbox{\boldmath{$x$}})
109: \right].$$In this case Eq.~(\ref{Eq:EV(C)}) yields the {\em master virial
110: theorem\/} \cite{Lucha:RVT,Lucha:RVTs}
111: \begin{equation}\left\langle\chi\left|\,\mbox{\boldmath{$p$}}\cdot
112: \frac{\partial\,T}{\partial\mbox{\boldmath{$p$}}}(\mbox{\boldmath{$p$}})\,
113: \right|\chi\right\rangle=\left\langle\chi\left|\,\mbox{\boldmath{$x$}}\cdot
114: \frac{\partial\,V}{\partial\mbox{\boldmath{$x$}}}(\mbox{\boldmath{$x$}})\,
115: \right|\chi\right\rangle;\label{Eq:RVT}\end{equation}this relation expresses
116: the equality of all the expectation values of the momentum-space radial
117: derivative of $T(\mbox{\boldmath{$p$}})$ and the (configuration-space) radial
118: derivative of $V(\mbox{\boldmath{$x$}}),$ it produces the specific virial
119: theorem for a particular~$H.$ For any nonrelativistic (Schr\"odinger)
120: Hamiltonian, i.~e.,$$H=H_{\rm
121: S}=m+\frac{\mbox{\boldmath{$p$}}^2}{2\,m}+V(\mbox{\boldmath{$x$}})\ ,$$
122: Theorem (\ref{Eq:RVT}) entails, retaining the conventional
123: factor~$\mbox{$\frac{1}{2}$},$
124: \begin{equation}\left\langle\chi\left|\,\frac{\mbox{\boldmath{$p$}}^2}{2\,m}\,
125: \right|\chi\right\rangle=\frac{1}{2}\left\langle\chi\left|\,
126: \mbox{\boldmath{$x$}}\cdot\frac{\partial\,V}{\partial\mbox{\boldmath{$x$}}}
127: (\mbox{\boldmath{$x$}})\,\right|\chi\right\rangle.\label{Eq:NRVT}\end{equation}
128: For the semirelativistic ``spinless-Salpeter'' Hamiltonian (\ref{Eq:SRH}),
129: involving the square-root operator of the relativistic kinetic energy
130: (\ref{Eq:RKE}), our master virial theorem (\ref{Eq:RVT}) leads~to
131: \begin{equation}\left\langle\chi\left|\,\frac{\mbox{\boldmath{$p$}}^2}
132: {\sqrt{\mbox{\boldmath{$p$}}^2+m^2}}\,\right|\chi\right\rangle=
133: \left\langle\chi\left|\,\mbox{\boldmath{$x$}}\cdot\frac{\partial\,V}{\partial
134: \mbox{\boldmath{$x$}}}(\mbox{\boldmath{$x$}})\,\right|\chi\right\rangle.
135: \label{Eq:SVT}\end{equation}In the nonrelativistic limit $m\to\infty$ (i.~e.,
136: for $\mbox{\boldmath{$p$}}^2\ll m^2$), this {\em spinless-Salpeter
137: relativistic virial theorem\/}, Eq.~(\ref{Eq:SVT}), necessarily reduces to
138: its nonrelativistic counterpart (\ref{Eq:NRVT}). Similarly, the virial
139: theorem for the Dirac equation \cite{Fock30,Brack83} is easily deduced
140: \cite{Lucha:RVTs} from our master virial theorem (\ref{Eq:RVT}).
141:
142: \section{Bounds to Energy Eigenvalues of a Spinless-Salpeter Hamiltonian}The
143: precise determination of eigenvalues of operators is of particular importance
144: for any formulation of quantum theory. Unfortunately, for most cases it is
145: not possible to determine the point spectrum (the set of all eigenvalues) of
146: a given operator {\em analytically\/}. Several powerful tools, however, allow
147: to derive analytic {\em bounds\/} to eigenvalues; applications of these
148: techniques to the spinless-Salpeter operator (\ref{Eq:SRH}) are reviewed, for
149: instance, in
150: Refs.~\cite{Lucha94:Como,Lucha98:Dubna,Lucha:Oberwoelz,Lucha:Dubrovnik}.
151:
152: \subsection{Minimum--maximum principle}\label{Sec:MMP}The theoretical
153: foundation of any derivation of a system of rigorous upper bounds to the
154: (isolated) eigenvalues of some operator $H$ in Hilbert space and hence the
155: primary tool for any localization of the discrete spectrum of $H$ is the
156: well-known {\em minimum--maximum principle\/}
157: \cite{Reed78,Weinstein72,Thirring90}. Its precise formulation is based on
158: several prerequisites:\begin{itemize}\item Let this operator $H$ be some {\em
159: self-adjoint\/} operator.\item Assume that this operator is {\em bounded from
160: below\/}.\item Define the {\em eigenvalues\/} of $H,$ $E_k,$ $k=0,1,2,\dots,$
161: by the eigenvalue equation, with eigenstates $|\chi_k\rangle,$
162: $$H\,|\chi_k\rangle=E_k\,|\chi_k\rangle\ ,\quad k=0,1,2,\dots\ .$$\item Let
163: these eigenvalues $E_k$ be {\em ordered}, according~to$$E_0\le E_1\le
164: E_2\le\cdots\ .$$\item Consider only the eigenvalues $E_k$ {\em below\/} the
165: onset of the {\em essential spectrum\/} of the above operator $H.$\item
166: Restrict all considerations to some $d$-dimensional {\em subspace\/}
167: $D_d\subset{\cal D}(H)$ of the domain ${\cal D}(H)$ of~$H.$\end{itemize}Then
168: this theorem asserts that every eigenvalue $E_k$ of $H$ --- counting
169: multiplicity of degenerate levels ---
170: satisfies$$E_k\le\displaystyle\sup_{|\psi\rangle\in D_{k+1}}
171: \frac{\langle\psi|\,H\,|\psi\rangle}{\langle\psi|\psi\rangle}\quad\mbox{for
172: all}\ k=0,1,2,\dots\ .$$In the case of one-dimensional subspaces, that is,
173: $d=1,$ the minimum--maximum theorem reduces to {\em Rayleigh's principle\/}:
174: the ground-state eigenvalue $E_0$ of an operator $H$ is less than, or equal
175: to, every expectation value of
176: $H$:$$E_0\le\frac{\langle\psi|\,H\,|\psi\rangle}{\langle\psi|\psi\rangle}\
177: ,\quad|\psi\rangle\in{\cal D}(H)\ .$$Given some operator inequality satisfied
178: by the operator $H,$ the minimum--maximum principle may be employed to
179: derive, by comparison, upper bounds on the (discrete) eigenvalues of $H,$
180: provided that a few assumptions hold:\begin{itemize}\item The operator $H,$
181: exhibiting all properties required by the minimum--maximum principle, is
182: bounded from above by some other operator called ${\cal O},$ i.~e., it is
183: subject to an {\em (operator) inequality\/} of the form$$H\le{\cal O}\
184: .$$Applying both the minimum--maximum principle and this operator inequality,
185: any eigenvalue $E_k$ of $H$ must be bounded from above by the supremum of the
186: expectation values of the operator ${\cal O}$ within the $(k+1)$-dimensional
187: subspace $D_{k+1}$ of ${\cal
188: D}(H)$:\begin{eqnarray}E_k&\equiv&\frac{\langle\chi_k|\,H\,|\chi_k\rangle}
189: {\langle\chi_k|\chi_k\rangle}\nonumber\\&\le&\sup_{|\psi\rangle\in D_{k+1}}
190: \frac{\langle\psi|\,H\,|\psi\rangle}{\langle\psi|\psi\rangle}\nonumber\\&\le&
191: \sup_{|\psi\rangle\in D_{k+1}}\frac{\langle\psi|\,{\cal O}\,|\psi\rangle}
192: {\langle\psi|\psi\rangle}\nonumber\\&&\quad\mbox{for all}\ k=0,1,2,\dots\
193: .\label{Eq:EVI}\end{eqnarray}\item All eigenvalues $\widehat E_k$ of ${\cal
194: O}$ are {\em ordered\/} according to$$\widehat E_0\le\widehat E_1\le\widehat
195: E_2\le\cdots\ .$$\item Every $k$-dimensional subspace $D_k$ in the chain of
196: inequalities which constitutes Eq.~(\ref{Eq:EVI}) is spanned by the first $k$
197: eigenvectors of the operator ${\cal O},$ or by precisely those eigenvectors
198: of ${\cal O}$ that correspond to the first $k$ eigenvalues $\widehat
199: E_0,\widehat E_1,\dots,\widehat E_{k-1}$ of our ${\cal O}.$ For this case it
200: is very easy to convince oneself that the supremum of all expectation values
201: of the operator ${\cal O}$ over the $(k+1)$-dimensional subspace $D_{k+1}$ is
202: then identical to the eigenvalue $\widehat E_k$ of ${\cal O}$:
203: $$\sup_{|\psi\rangle\in D_{k+1}}\frac{\langle\psi|\,{\cal O}\,|\psi\rangle}
204: {\langle\psi|\psi\rangle}=\widehat E_k\ .$$\end{itemize}Consequently, an
205: eigenvalue $E_k,$ $k=0,1,2,\dots,$ of the discrete spectrum of $H(\le{\cal
206: O})$ is bounded from above~by the corresponding eigenvalue $\widehat E_k,$
207: $k=0,1,2,\dots,$ of ${\cal O}$:$$E_k\le\widehat E_k\quad\mbox{for all}\
208: k=0,1,2,\dots\ .$$It remains to prove an ``appropriate'' operator inequality.
209: (Summaries of the idea to find bounds by combining the minimum--maximum
210: principle with reasonable operator inequalities may be found, e.~g., in
211: Refs.~\cite{Lucha96rcpaubel,Lucha:Oberwoelz,Lucha:Dubrovnik,Lucha99-1dimsrcp}.)
212:
213: \subsection{Analytical upper bounds}
214:
215: \subsubsection{The trivial nonrelativistic Schr\"odinger
216: bound}\label{Sec:NUB}The inequality $(T-m)^2\ge0$ expressing nothing but~the
217: {\em positivity\/} of the square of the operator $T-m$ may be, for $m>0,$
218: written as an inequality for the kinetic energy~$T$:$$T\le
219: m+\frac{\mbox{\boldmath{$p$}}^2}{2\,m}\ .$$(The right-hand side is the
220: tangent line to the square root in the relativistic kinetic energy $T$ in the
221: point of contact $\mbox{\boldmath{$p$}}^2=0.$) This result proves
222: \cite{Lucha96rcpaubel} that $H$ is bounded from above by a nonrelativistic
223: Schr\"odinger Hamiltonian $H_{\rm S}$:$$H\le H_{\rm
224: S}=m+\frac{\mbox{\boldmath{$p$}}^2}{2\,m}+V\ .$$For a pure Coulomb potential
225: $V(r)=-\alpha/r,$ the energy eigenvalues of the Schr\"odinger Hamiltonian
226: $H_{\rm S}$ depend only on the principal quantum number $n,$ related to both
227: radial and orbital angular-momentum quantum numbers by $n=n_{\rm r}+\ell+1,$
228: with $n_{\rm r}=0,1,2,\dots,$ $\ell=0,1,2,\dots$:$$E_{{\rm
229: S},n}=m\left(1-\frac{\alpha^2}{2\,n^2}\right).$$
230:
231: \subsubsection{A ``squared'' bound}\label{Sec:QUB}A relation between the
232: (semirelativistic) Hamiltonian $H$ and a nonrelativistic Schr\"odinger
233: operator may be found \cite{Lucha96rcpaubel} by considering the square $H^2$
234: of $H$ and by realizing that the anticommutator $T\,V+V\,T$ of relativistic
235: kinetic energy $T$ and potential $V$ generated by the square
236: fulfils$$T\,V+V\,T\le\mbox{\boldmath{$p$}}^2+V^2+2\,m\,V\ ,$$as may be shown
237: \cite{Lucha96rcpaubel} by inspecting some consequences of the positivity of
238: the square of the operator
239: $T-m-V$:\begin{eqnarray*}H^2&=&T^2+V^2+T\,V+V\,T\\&\le&Q\equiv
240: 2\,\mbox{\boldmath{$p$}}^2+m^2+2\,V^2+2\,m\,V\ .\end{eqnarray*}With this
241: inequality, the minimum--maximum principle, recalled in
242: Subsect.~\ref{Sec:MMP}, immediately guarantees that the energy eigenvalues
243: $E_k$ of $H$ are bounded from above by the square root of the corresponding
244: eigenvalues ${\cal E}_{Q,k}$~of the Schr\"odinger operator $Q,$ constructed
245: by squaring $H$:$$E_k\le\sqrt{{\cal E}_{Q,k}}\ ,\quad k=0,1,2,\dots\ .$$For
246: the case of a pure Coulomb potential $V(r)=-\alpha/r,$ the operator $Q$ has
247: the same structure as the Schr\"odinger Hamiltonian $H_{\rm S}$ of
248: Subsect.~\ref{Sec:NUB}, with $\ell$ replaced by an effective orbital angular
249: momentum quantum number $L$ involving both the usual $\ell$ and the Coulomb
250: coupling, $\alpha$:\begin{equation}L\,(L+1)=\ell\,(\ell+1)+\alpha^2\
251: ,\quad\ell=0,1,2,\dots\ .\label{Eq:Leff}\end{equation}The set of eigenvalues
252: ${\cal E}_Q$ of a ``Coulombic'' operator~$Q,$$${\cal
253: E}_{Q,N}=m^2\left(1-\frac{\alpha^2}{2\,N^2}\right),$$is determined by the
254: {\em effective\/} principal quantum number$$N=n_{\rm r}+L+1\ ,\quad n_{\rm
255: r}=0,1,2,\dots\ .$$Unfortunately, in the Coulomb case the squared bounds are
256: above, and thus worse than, the Schr\"odinger bounds.
257:
258: \subsection{Rigorous semianalytical upper bound}\label{Sec:GUB}We regard an
259: energy bound as {\em semianalytical\/} if it can be derived by an (in
260: general, numerical) optimization of an analytically given expression over a
261: single real variable. Taking advantage, as a straightforward generalization
262: of the (simple) line of argument sketched in Subsect.~\ref{Sec:NUB}, of the
263: inequality $(T-\mu)^2\ge0$ requiring an arbitrary real parameter $\mu$ of
264: mass dimension 1 and obviously holding for all self-adjoint $T$
265: \cite{Martin88} implies, for the kinetic
266: energy,$$T\le\frac{\mbox{\boldmath{$p$}}^2+m^2+\mu^2}{2\,\mu}\quad\mbox{for
267: all}\ \mu>0\ .$$This translates \cite{Lucha96rcpaubel} to a set of
268: inequalities for $H,$ each of these involving a Schr\"odinger-like
269: Hamiltonian,~$\widehat H_{\rm S}(\mu)$:$$H\le\widehat H_{\rm
270: S}(\mu)=\frac{\mbox{\boldmath{$p$}}^2+m^2+\mu^2}{2\,\mu}+V\quad\mbox{for
271: all}\ \mu>0\ .$$The best ``Schr\"odinger-like'' upper bound on any energy
272: eigenvalue $E_k$ of $H$ is then provided by the minimum of the
273: $\mu$-dependent energy eigenvalues of $\widehat H_{\rm S}(\mu),$ $\widehat
274: E_{{\rm S},k}(\mu)$:$$E_k\le\displaystyle\min_{\mu>0}\widehat E_{{\rm
275: S},k}(\mu)\ ,\quad k=0,1,2,\dots\ .$$For a pure Coulomb potential
276: $V(r)=-\alpha/r,$ the energy eigenvalues $\widehat E_{{\rm S},n}(\mu)$ of
277: $\widehat H_{\rm S}(\mu)$ read, with $n=n_{\rm r}+\ell+1,$$$\widehat E_{{\rm
278: S},n}(\mu)=\frac{1}{2\,\mu}\left[m^2+\mu^2\left(1-\frac{\alpha^2}{n^2}\right)
279: \right].$$Here, minimizing $\widehat E_{{\rm S},n}(\mu)$ with respect to
280: $\mu$ entails \cite{Lucha96rcpaubel}$$\displaystyle\min_{\mu>0}\widehat
281: E_{{\rm S},n}(\mu)=m\,\sqrt{1-\frac{\alpha^2}{n^2}}\ .$$This (exact) {\em
282: upper\/} bound \cite{Lucha96rcpaubel} to the energy eigenvalues of the
283: so-called ``spinless relativistic Coulomb problem'' holds for all those
284: values of the Coulomb coupling $\alpha$ for which the Hamiltonian $H$ with a
285: Coulomb potential can be regarded as a reasonable operator and arbitrary
286: levels of excitation, and for any value of the principal quantum number $n$
287: it definitely improves the Schr\"odinger bound:$$\min_{\mu>0}\widehat E_{{\rm
288: S},n}(\mu)<E_{{\rm S},n}\quad\mbox{for}\ \alpha\neq 0\ .$$Clearly, fixing
289: $\mu=m$ recovers the Schr\"odinger bounds.
290:
291: \subsection{Exact semianalytical upper and lower bounds from the ``envelope
292: technique''}\label{Sec:EnvULB}Rigorous semianalytical expressions for both
293: upper and lower bounds to the eigenvalues $E_{n\ell}$ of the Hamiltonian $H$
294: are found by a geometrical operator comparison in an approach called
295: ``envelope theory.'' The envelope theory constructs bounds on $E_{n\ell}$ by
296: comparing the spectrum of $H$ with the one of a conveniently formulated
297: ``tangential Hamiltonian'' $\widetilde H$ involving some ``basis potential''
298: $h(r),$$$\widetilde H=\sqrt{m^2+\mbox{\boldmath{$p$}}^2}+c\,h(r)
299: +\mbox{const.}\ ,\quad c>0\ ,$$for which sufficient spectral information
300: (i.~e., either the exact eigenvalues or suitable bounds on these) is known.
301: Let $V(r)$ be a smooth transformation $V=g(h)$ of $h(r),$ with definite
302: convexity of $g(h).$ After optimization with respect to the point of contact
303: of $V(r)$ and the tangential potential, this technique produces bounds on
304: $E_{n\ell}$:~lower bounds for $g(h)$ convex $(g''>0),$ and upper bounds for
305: $g(h)$ concave $(g''<0).$ Suppressing for the moment the quantum numbers
306: $n\ell,$ all these bounds on $E$ may be cast into a common generic form
307: \cite{Lucha00-HO,Lucha01-DMAIa,Lucha01-DMAIb,Lucha02-DMAII,Lucha02-sum,
308: Lucha02-CP} with the individual bounds discriminated by a dimensionless
309: parameter, $P$:\begin{equation}E\approx\min_{r>0}
310: \left[\sqrt{m^2+\frac{1}{r^2}}+V(P\,r)\right].\label{Eq:EnvBd}\end{equation}
311: Here, that cryptic sign of approximate equality indicates that for any
312: definite convexity of $g(h)$ all expressions on the right-hand side represent
313: a lower bound for a convex $g(h)$ and an upper bound for a concave $g(h).$
314: The value of the parameter $P$ used in Eq.~(\ref{Eq:EnvBd}) is determined
315: by~the algebraic structure of the interaction potential $V(r),$ and by its
316: convexity with respect to the basis potential, $h(r)$:\begin{itemize}\item
317: The {\em spinless relativistic Coulomb problem\/} posed by $V(r)=-\alpha/r$
318: is well-defined if its coupling $\alpha$ is constrained to $\alpha<2/\pi$
319: \cite{Herbst}. The bottom of the corresponding spectrum of $H$ (or, its
320: ground-state energy eigenvalue, $E_0$) is bounded from below by$$E_0\ge
321: m\,\sqrt{1-\frac{\alpha^2}{P^2}}\ ,$$with the lower-bound parameter $P$ given
322: either~by $P=2/\pi$ for $\alpha$ fulfilling $0\le\alpha<2/\pi$ \cite{Herbst},
323: or~by$$P\equiv
324: P(\alpha)=\sqrt{\mbox{$\frac{1}{2}$}\left(1+\sqrt{1-4\,\alpha^2}\right)}$$
325: for $0\le\alpha\le\mbox{$\frac{1}{2}$},$ which obviously covers the range
326: $$P(\mbox{$\frac{1}{2}$})=\mbox{$\frac{1}{\sqrt{2}}$}\le P(\alpha)\le P(0)=1\
327: ,$$as derived by weakening
328: \cite{Lucha01-DMAIa,Lucha01-DMAIb,Lucha02-DMAII,Lucha02-CP} an improved lower
329: bound to $E_0$ valid only for $0\le\alpha\le\frac{1}{2}$ \cite{Martin89}.
330:
331: \newpage
332:
333: If $V(r)$ is a convex transform $V=g(h),$ $g''>0,$ of the Coulomb potential
334: $h(r)=-1/r,$ the above envelope approximation generates a {\em lower\/} bound
335: \cite{Lucha01-DMAIa,Lucha01-DMAIb,Lucha02-DMAII,Lucha02-CP} on the
336: ground-state eigenvalue $E_0$ (on the entire spectrum) of the Hamiltonian $H$
337: for any choice of the Coulomb lower-bound parameter $P.$
338:
339: Clearly, the quoted upper bounds on the Coulomb coupling $\alpha$ apply also
340: to any ``effective'' Coulomb coupling in $\widetilde H.$ Consequently, they
341: translate into a constraint on all coupling constants introduced by the
342: interaction potential $V(r)$ under investigation. (An example for these
343: restrictions enforced by the Coulomb menace will be given in
344: Subsect.~\ref{Sec:FP}.)\item For $V(r)$ a concave transform $V=g(h),$
345: $g''<0,$ of the harmonic-oscillator potential $h(r)=r^2,$ a straightforward
346: application of the above envelope approximation yields {\em upper\/} bounds
347: \cite{Lucha01-DMAIa,Lucha01-DMAIb,Lucha02-DMAII,Lucha02-CP} to {\em all\/}
348: the eigenvalues $E_{n\ell}$ of the Hamiltonian $H;$ the parameter $P$ for a
349: given energy level identified by quantum numbers $n\ell$ is, in this case,
350: related to the explicitly algebraically known eigenvalues ${\cal E}_{n\ell}$
351: of the (nonrelativistic) Schr\"odinger operator
352: $\mbox{\boldmath{$p$}}^2+r^2$:$$P\equiv
353: P_{n\ell}(2)=\mbox{$\frac{1}{2}$}\,{\cal
354: E}_{n\ell}=2\,n+\ell-\mbox{$\frac{1}{2}$}\ .$$\item For $V(r)$ a concave
355: transform $V=g(h),$ $g''<0,$ of the linear potential $h(r)=r,$ the
356: application of a ``generalized'' envelope approximation provides {\em
357: upper\/} bounds \cite{Lucha02-DMAII,Lucha02-CP} to {\em all\/} eigenvalues
358: $E_{n\ell}$ of $H$ if the parameters $P$ which characterize the energy levels
359: are given, in terms of the eigenvalues ${\cal E}_{n\ell}$ of the
360: nonrelativistic Schr\"odinger operator $\mbox{\boldmath{$p$}}^2+r,$~by
361: \begin{equation}P\equiv P_{n\ell}(1)=2\left(\mbox{$\frac{1}{3}$}\,{\cal
362: E}_{n\ell}\right)^{3/2}\ ;\label{Eq:P(1)}\end{equation}the parameter values
363: $P_{n\ell}(1)$ corresponding to the lowest-lying energy levels can be found
364: in Table~\ref{Tab:P(1)} (for more details see, for instance,
365: Refs.~\cite{Lucha00-HO,Lucha01-DMAIa,Lucha01-DMAIb,Lucha02-DMAII}).
366: \end{itemize}\begin{table}[ht]\caption{Numerical values of the parameter
367: $P_{n\ell}(1)$ used in the linear-potential-based lower envelope bounds and
368: defined in Eq.~(\ref{Eq:P(1)}) for the lowest-lying energy levels
369: $n\ell.$}\label{Tab:P(1)}\vspace{1ex}
370: \begin{center}\begin{tabular}{ccr}\hline\hline&&\\[-1.5ex]
371: \multicolumn{1}{c}{$n$}&\multicolumn{1}{c}{$\ell$}&
372: \multicolumn{1}{c}{$P_{n\ell}(1)$}\\[1ex]\hline\\[-1.5ex]
373: 1&0&1.37608\\2&0&3.18131\\3&0&4.99255\\4&0&6.80514\\5&0&8.61823\\[.5ex]
374: 1&1&2.37192\\2&1&4.15501\\3&1&5.95300\\4&1&7.75701\\5&1&9.56408\\[.5ex]
375: 1&2&3.37018\\2&2&5.14135\\3&2&6.92911\\4&2&8.72515\\5&2&10.52596\\[1ex]
376: \hline\hline\end{tabular}$\qquad$\begin{tabular}{ccr}\hline\hline&&\\[-1.5ex]
377: \multicolumn{1}{c}{$n$}&\multicolumn{1}{c}{$\ell$}&
378: \multicolumn{1}{c}{$P_{n\ell}(1)$}\\[1ex]\hline\\[-1.5ex]
379: 1&3&4.36923\\2&3&6.13298\\3&3&7.91304\\4&3&9.70236\\5&3&11.49748\\[.5ex]
380: 1&4&5.36863\\2&4&7.12732\\3&4&8.90148\\4&4&10.68521\\5&4&12.47532\\[.5ex]
381: 1&5&6.36822\\2&5&8.12324\\3&5&9.89276\\4&5&11.67183\\5&5&13.45756\\[1ex]
382: \hline\hline\end{tabular}\end{center}\end{table}If the potential $V(r)$ is
383: the sum of several distinct terms,$$V(r)=\sum_iV_i(r)\ ,\quad
384: V_i(r)=c_i\,h_i(r)\ ,$$where every {\em component problem\/} defined by the
385: operator$$\sqrt{m^2+\mbox{\boldmath{$p$}}^2}+c_i\,h_i(r)$$supports, for a
386: sufficiently large $c_i,$ a discrete eigenvalue $E_{i,0}$ at the bottom of
387: its spectrum and information about the lowest energy eigenvalue, $E_{i,0},$
388: is available, all these pieces of information can be combined to a lower
389: bound to $E_0$ \cite{Lucha02-sum}; for sums of pure power-law terms ${\rm
390: sgn}(q)\,r^q,$\begin{equation}V_{\rm PL}(r)=\sum_qa(q)\,{\rm sgn}(q)\,r^q\
391: ,\label{Eq:PLP}\end{equation}where the coefficients $a(q)$ of the pure
392: power-law terms, ${\rm sgn}(q)\,r^q,$ in the potential are positive, that is,
393: $a(q)\ge0,$ and do not vanish {\em all\/}, this yields the ``sum lower
394: bound''$$E_0\ge\min_{r>0}\left[\sqrt{m^2+\frac{1}{r^2}}+\sum_qa(q)\,{\rm
395: sgn}(q)\,(\underline{P}(q)\,r)^q\right]$${\em provided\/} that some set of
396: lower-bound parameters $\underline{P}(q)$ can be derived such that, whenever
397: $V(r)$ consists of just one single component, the above inequality yields
398: either the corresponding exact ground-state energy eigenvalue or, at least, a
399: rigorous lower bound to this latter quantity:\begin{itemize}\item For Coulomb
400: components, that is, $h_i(r)=-1/r,$ $\underline{P}(-1)$ is the Coulomb
401: lower-bound parameter $P.$\item For linear components, that is, $h_i(r)=r,$
402: $\underline{P}(1)$ is derived from the lowest eigenvalue ${\cal E}_0$ of
403: $\sqrt{\mbox{\boldmath{$p$}}^2}+r,$$$\underline{P}(1)=\mbox{$\frac{1}{4}$}\,
404: {\cal E}_0^2=1.2457\ .$$\end{itemize}It is straightforward to (try to)
405: generalize these envelope techniques from the simpler one-body case
406: summarized in this review to systems composed of arbitrary numbers of
407: relativistically moving interacting particles described by a semirelativistic
408: spinless Salpeter equation \cite{Lucha01-NHO,Lucha03-NHO-m=0,Lucha04-NV(r2)}.
409: At least for the particular case of all \mbox{harmonic-oscillator} potentials
410: $V(r)=c\,r^2$ with $c>0$ the generalized upper bounds presented in
411: Subsect.~\ref{Sec:GUB} and the envelope upper bounds can be shown to be
412: equivalent to each other \cite{Lucha02-DMAII}.
413:
414: \subsection{Rayleigh--Ritz (variational) technique}\label{Sec:RRVT}An
415: immediate consequence of the minimum--maximum principle is the
416: ``Rayleigh--Ritz (variational) technique:''\begin{itemize}\item Introduce the
417: {\em restriction\/} $\widehat H$ of some operator $H$ to a subspace $D_d$ by
418: orthogonal projection $P$ to~$D_d$:$$\left.\widehat H\equiv
419: H\right|_{D_d}:=P\,H\,P\ .$$\item Identify all $d$ {\em eigenvalues\/}
420: $\widehat E_k,$ $k=0,1,\dots,d-1,$ of the restricted operator $\widehat H$ as
421: the solutions of the eigenvalue equation of $\widehat H$ for the eigenstates
422: $|\widehat\chi_k\rangle$:$$\widehat H\,|\widehat\chi_k\rangle=\widehat
423: E_k\,|\widehat\chi_k\rangle\ ,\quad k=0,1,\dots,d-1\ .$$\item Let these
424: eigenvalues $\widehat E_k$ be {\em ordered}, according~to$$\widehat
425: E_0\le\widehat E_1\le\cdots\le\widehat E_{d-1}\ .$$\end{itemize}Then every
426: (discrete) eigenvalue $E_k$ of $H$ --- if counting the multiplicity of
427: degenerate levels --- is bounded from above by the eigenvalue $\widehat E_k$
428: of the restricted operator~$\widehat H$:$$E_k\le\widehat E_k\quad\mbox{for
429: all}\ k=0,1,\dots,d-1\ .$$If that $d$-dimensional subspace $D_d$ is spanned
430: by any set of $d$ (of course, linearly independent) basis vectors
431: $|\psi_k\rangle,$ $k=0,1,\dots,d-1,$ the eigenvalues $\widehat E_k$ can
432: immediately be determined, by the diagonalization of the $d\times d$
433: matrix$$\left(\langle\psi_i|\,\widehat H\,|\psi_j\rangle\right),\quad
434: i,j=0,1,\dots,d-1\ ,$$that is, as the $d$ roots of the characteristic
435: equation of~$\widehat H,$\begin{eqnarray*}\det\left(\langle\psi_i|\,\widehat
436: H\,|\psi_j\rangle-\widehat E\,\langle\psi_i|\psi_j\rangle\right)=0\ ,\\
437: i,j=0,1,\dots,d-1\ .\,\end{eqnarray*}To establish this, expand any
438: eigenvector $|\widehat\chi_k\rangle$ of $\widehat H$ over the basis
439: $\{|\psi_i\rangle,\ i=0,1,\dots,d-1\}$ of the subspace $D_d.$
440:
441: \subsection{Variational upper bounds}\label{Sec:VUB}The {\em quality\/}
442: achieved by the variational solution of some eigenvalue problem depends
443: decisively on the definition of the trial subspace $D_d$ employed by the
444: Rayleigh--Ritz technique briefly sketched in Subsect.~\ref{Sec:RRVT}:
445: enlarging $D_d$ to higher dimensions $d$ or choosing a more sophisticated
446: basis $\{|\psi_i\rangle,\ i=0,1,\dots,d-1\}$ which spans $D_d$ will, in
447: general, increase the {\em accuracy\/} of the obtained solutions.
448:
449: For spherically symmetric (central) potentials $V(r),$ that is, for all
450: potentials which depend only on the radial coordinate
451: $r\equiv|\mbox{\boldmath{$x$}}|,$ a convenient and thus rather popular choice
452: for the basis vectors $\{|\psi_i\rangle,\ i=0,1,\dots,d-1\}$ is that one the
453: configuration-space representation of which involves the complete orthogonal
454: system of generalized Laguerre polynomials
455: \cite{Jacobs86,Lucha:LagB,Lucha:Oberwoelz,Lucha:Dubrovnik} ---
456: cf.~Appendix~\ref{App:Lag}.
457:
458: \newpage\noindent In the one-dimensional case \cite{Lucha94:VA-SRCP} realized
459: in the notation of Appendix~\ref{App:Lag} if all quantum numbers
460: $k=\ell=m=0,$ the Laguerre basis collapses to just a single basis vector:
461: $$\psi(\mbox{\boldmath{$x$}})\equiv\psi_{0,00}(\mbox{\boldmath{$x$}})
462: =\sqrt{\frac{\mu^3}{\pi}}\,\exp(-\mu\,r)\ ,\quad\mu>0\ .$$With a trial state
463: $|\psi\rangle$ represented by this exponential and the trivial (nevertheless
464: fundamental) general inequality$$\frac{|\langle\psi|\,{\cal
465: O}\,|\psi\rangle|}{\langle\psi|\psi\rangle}\le\sqrt{\frac{\langle\psi|\,{\cal
466: O}^2\,|\psi\rangle}{\langle\psi|\psi\rangle}}\ ,$$which holds for any
467: self-adjoint, but otherwise arbitrary, operator ${\cal O}$ (${\cal
468: O}^\dagger={\cal O}$), Rayleigh's principle entails,~after optimization with
469: respect to the variational parameter $\mu,$ for a Coulomb potential
470: $V(r)=-\alpha/r$ the upper bound$$E_0\le m\,\sqrt{1-\alpha^2}\ ;$$this is
471: identical to the ``generalized'' upper energy bound on the ground-state or
472: $n=1$ eigenvalue of the Coulomb operator $H$ found by different reasoning in
473: Subsect.~\ref{Sec:GUB}.
474:
475: \subsection{Application to illustrative interactions}Let us appreciate the
476: above bounds' beauty at examples.
477:
478: \subsubsection{Trivial ``testing ground:'' Coulomb potential}\label{Sect:CP}
479: Our first example clearly must be the Coulomb
480: potential$$V(r)=-\frac{\alpha}{r}\ ,\quad\alpha>0\ ;$$this potential arises
481: from the exchange of some massless boson between the interacting objects.
482: Therefore it is of particular interest in many areas of physics. Its
483: effective interaction strength is given by a coupling $\alpha,$ identical~to
484: the fine structure constant in electrodynamics. We study\begin{itemize}\item
485: the somewhat naive nonrelativistic (Schr\"odinger) upper bound given in
486: Subsect.~\ref{Sec:NUB}, equivalent to a tangent line to the relativistic
487: kinetic operator $T,$\item the upper bound of Subsect.~\ref{Sec:QUB},
488: constructed by considering just the square of the Hamiltonian $H,$\item the
489: semianalytical upper bound of Subsect.~\ref{Sec:GUB}, as derived by
490: generalizing the idea of Subsect.~\ref{Sec:NUB},\item all three envelope
491: bounds of Subsect.~\ref{Sec:EnvULB}, namely,\begin{itemize}\item the
492: harmonic-oscillator-based upper bound,\item the upper bound involving a
493: linear potential,\item the lower bound obtained by ``loosening'' an absolute
494: lower bound on the spectrum of the ``semirelativistic Coulomb operator'' $H,$
495: and\end{itemize}\item the ``Rayleigh--Ritz'' upper bound of
496: Subsect.~\ref{Sec:VUB}.\end{itemize}
497:
498: \newpage\noindent For the Coulomb potential under study, the optimization
499: required by the envelope bounds (\ref{Eq:EnvBd}) may be performed
500: analytically, yielding a result of precisely the same form as the generalized
501: upper bounds derived in Subsect.~\ref{Sec:GUB}, or as the squared upper
502: bounds proved in Subsect.~\ref{Sec:QUB}:\begin{equation}E_0(P)\approx
503: m\,\sqrt{1-\frac{\alpha^2}{P^2}}\ ,\label{Eq:CPB}\end{equation}where for the
504: ground state characterized by the quantum numbers $n=1,$ $\ell=0$ the
505: (single) parameter $P$ is given,\begin{itemize}\item for the ``Coulomb lower
506: bound'' (Subsect.~\ref{Sec:EnvULB}), by$$P\equiv P_{\rm
507: C}=P(\alpha)=\sqrt{\mbox{$\frac{1}{2}$}\left(1+\sqrt{1-4\,\alpha^2}\right)}\
508: ,$$\item for the generalized upper bound
509: (Subsect.~\ref{Sec:GUB}),~by$$P\equiv P_{\rm G}=n=1\ ,$$\item for the
510: ``linear upper bound'' (Subsect.~\ref{Sec:EnvULB}), as can be simply read off
511: from the first row in Table~\ref{Tab:P(1)},~by$$P\equiv P_{\rm
512: L}=P_{10}(1)=1.37608\ ,$$\item for the ``squared upper bound''
513: (Subsect.~\ref{Sec:QUB}), in accordance with the solution of
514: Eq.~(\ref{Eq:Leff}) for $L,$~by$$P\equiv P_{\rm
515: Q}=\sqrt{2}\,N=\frac{1+\sqrt{1+4\,\alpha^2}}{\sqrt{2}}\ ,$$\item and, in the
516: case of the ``harmonic-oscillator upper bound'' (Subsect.~\ref{Sec:EnvULB}),
517: from the $P_{n\ell}(2)$ results,~by$$P\equiv P_{\rm
518: H}=P_{10}(2)=\mbox{$\frac{3}{2}$}\ .$$\end{itemize}It is a very trivial
519: observation that, for fixed values of the Coulomb coupling, the ground-state
520: energy bounds (\ref{Eq:CPB}) are (monotone) increasing with increasing
521: parameter $P$:$$\frac{\partial\,E_0(P)}{\partial P}\ge0\ .$$Thus it is
522: straightforward to convince oneself that all the Coulomb ($E_{\rm C}$),
523: generalized ($E_{\rm G}$), nonrelativistic ($E_{\rm N}$), linear ($E_{\rm
524: L}$), squared ($E_{\rm Q}$) and harmonic-oscillator ($E_{\rm H}$) bounds on
525: the ground-state energy eigenvalue $E_0$ of the semirelativistic Coulomb
526: Hamiltonian $H$ have to satisfy\begin{eqnarray*}E_{\rm C}\le E_0\le E_{\rm
527: G}\le E_{\rm N}\le E_{\rm L}\le E_{\rm Q}\le E_{\rm H}\\\mbox{for}\
528: \alpha\le\alpha_0\equiv\sqrt{\mbox{$\frac{3}{8}$}\left(3-2\,\sqrt{2}\right)}\
529: ,\\E_{\rm C}\le E_0\le E_{\rm G}\le E_{\rm N}\le E_{\rm L}\le E_{\rm H}\le
530: E_{\rm Q}\\\mbox{for}\
531: \alpha\ge\alpha_0\equiv\sqrt{\mbox{$\frac{3}{8}$}\left(3-2\,\sqrt{2}\right)}\
532: ,\end{eqnarray*}taking into account the crossing of the upper bounds $E_{\rm
533: H}$ and $E_{\rm Q}$ at $\alpha_0^2=\frac{3}{8}\,(3-2\,\sqrt{2}),$ i.~e.,
534: $E_{\rm Q}(\alpha_0)=E_{\rm H}(\alpha_0).$
535:
536: \newpage
537:
538: For Coulomb-like interactions the only dimensional quantity among the
539: parameters of this theory is the mass $m$ of the interacting particles.
540: Consequently, in this case all energy eigenvalues are proportional to $m$:
541: the energy scale is set by $m.$ The ratio $E_k/m$ is a universal function of
542: the coupling $\alpha;$ w.~l.~o.~g.\ it thus suffices to fix $m=1.$
543:
544: Figure~\ref{Fig:CP} compares for the ground state ($n_{\rm r}=\ell=0$) of the
545: spinless relativistic Coulomb problem the various bounds to the lowest energy
546: eigenvalue, $E_0,$ listed at the beginning of this subsection. Inspecting
547: Fig.~\ref{Fig:CP}, we note:\begin{itemize}\item the squared,
548: harmonic-oscillator, and linear upper bounds are numerically comparable to
549: each other;\item likewise the nonrelativistic and generalized upper bounds
550: are close to each other for all couplings~$\alpha;$\item using a Laguerre
551: trial space of dimension $d=25,$ the Rayleigh--Ritz variational upper bound
552: can be expected to come already pretty close to the exact eigenvalue $E_0$
553: --- which, in turn, clearly indicates that it is highly desirable to find
554: improvements for the lower bounds, in particular for large couplings $\alpha$
555: (this stimulated, e.~g., the analysis of Ref.~\cite{Lucha96:CCC}).
556: \end{itemize}
557:
558: \begin{figure}[h]\begin{center}\psfig{figure=coulomba.ps,scale=0.76789}
559: \caption{Both upper ({\sl full lines\/}) and lower ({\sl dashed line\/})
560: bounds on the ground-state energy eigenvalue ({\bf E}) of the
561: semirelativistic Hamiltonian $H$ with Coulomb potential $V(r)=-a/r$ as a
562: function of the Coulomb coupling,~$a,$ for the s{\bf q}uared ({\bf Q}), {\bf
563: h}armonic-oscillator ({\bf H}), {\bf l}inear ({\bf L}), {\bf n}onrelativistic
564: ({\bf N}), {\bf g}eneralized ({\bf G}), {\bf v}ariational ({\bf V}) and {\bf
565: C}oulomb ({\bf C}) [using the ``optimized'' $P(a)$]
566: approaches.}\label{Fig:CP}\end{center}\end{figure}
567:
568: \subsubsection{Coulomb-plus-linear (or ``funnel'') potential}\label{Sec:FP}
569: Within the field of elementary particle physics, quantum chromodynamics (QCD)
570: is generally accepted to be that relativistic quantum field theory that
571: describes all strong interactions between quarks and gluons by assigning the
572: so-called ``colour'' degrees of freedom to these particles. In the
573: instantaneous approximation inherent to all of the QCD-inspired quark
574: potential models developed for the purely phenomenological description of
575: experimentally observed hadrons, as bound states of {\em constituent\/}
576: quarks, the strong forces are assumed to derive from an effective potential
577: generating the bound states (this description of hadrons within the framework
578: of quark potential models involving either nonrelativistic or relativistic
579: kinematics is reviewed, for instance, in
580: Refs.~\cite{Lucha91:BSQ,Lucha92:QAQBS}.) The prototype of all ``realistic,''
581: that is, phenomenologically acceptable (static) interquark potentials $V(r)$
582: consists of the sum of\begin{itemize}\item a Coulomb contribution generated
583: by a one-gluon exchange between quark bound-state constituents (dominating
584: the potential at short distances $r$) and\item a linear term including all
585: nonperturbative effects (that dominates the potential at large distances
586: $r$).\end{itemize}The resulting interaction potential $V(r)$ is characterized
587: by a ``funnel-type'' Coulomb-plus-linear form; therefore it is called the
588: Coulomb-plus-linear, or funnel, potential. Upon factorizing off a constant
589: $v,$ which spans the range $0<v\le1$ in order to parametrize an overall
590: interaction strength, we (prefer to) analyze this potential in the form
591: \begin{equation}V(r)=-\frac{c_1}{r}+c_2\,r=v\left(-\frac{a}{r}+b\,r\right).
592: \label{Eq:FP}\end{equation}Clearly, given the overall coupling strength $v,$
593: the actual shape of this potential is fixed by the {\em ratio\/} of the
594: positive parameters $a>0$ and $b>0;$ the coupling constants that enter, on
595: the one hand, in the general expression (\ref{Eq:PLP}) for sums of pure
596: power-law terms and, on the other hand, in our funnel potential (\ref{Eq:FP})
597: must be identified according~to\begin{eqnarray*}a(-1)&\equiv&c_1\equiv
598: a\,v>0\ ,\\a(1)&\equiv&c_2\equiv b\,v>0\ .\end{eqnarray*}In view of the lack
599: of fully analytical bounds we explore\begin{itemize}\item the three ``basic''
600: envelope bounds of Subsect.~\ref{Sec:EnvULB}, distinguished by the adopted
601: basis potential, viz.,\begin{itemize}\item the upper bound from a harmonic
602: oscillator,\item the upper bound involving a linear potential,\item the lower
603: bound due to a Coulomb potential,\end{itemize}\item the envelope {\em sum
604: lower bound\/}, derived in the sum approximation recalled by
605: Subsect.~\ref{Sec:EnvULB}, as well as\item the ``Rayleigh--Ritz'' upper bound
606: of Subsect.~\ref{Sec:VUB}.\end{itemize}
607:
608: \newpage\noindent For definiteness, let us fix the potential parameters $a$
609: and $b$ to $a=0.2,$ $b=0.5.$ As done in the Coulomb-potential example (in
610: Subsect.~\ref{Sect:CP}) in order to take advantage of upper and lower bounds,
611: we investigate the ground-state energy $E_0.$ The basic envelope bounds are
612: computed by application of Eq.~(\ref{Eq:EnvBd}), for the appropriate
613: parameter~$P$:\begin{itemize}\item for the ``harmonic-oscillator upper
614: bound'' we use$$P\equiv P_{\rm H}=P_{10}(2)=\mbox{$\frac{3}{2}$}\ ;$$\item
615: for the ``linear upper bound'' we find from Table~\ref{Tab:P(1)}$$P\equiv
616: P_{\rm L}=P_{10}(1)=1.37608\ ;$$\item for the ``Coulomb lower bound,'' in
617: order to derive the maximum value $P$ consistent with $0<v\le1,$ we are
618: forced to evaluate that ``Coulomb coupling constant constraint'' mentioned in
619: Subsect.~\ref{Sec:EnvULB}, in its form \cite{Lucha01-DMAIb,{Lucha02-DMAII}}
620: fixed by our funnel potential (\ref{Eq:FP}),
621: $$c_1+\frac{P^4}{1-P^2}\frac{c_2}{m^2}\le P\,\sqrt{1-P^2}\ ,$$for the maximum
622: values of $c_1$ and $c_2,$ which gives$$P\equiv P_{\rm
623: C}=0.728112397\quad\mbox{for}\ m=1\ .$$\end{itemize}The ``sum lower bound''
624: is extracted from the expression given explicitly in
625: Subsect.~\ref{Sec:EnvULB} for power-law potentials by insertion of the
626: lower-bound parameters $\underline{P}(q=\pm1)$:\begin{itemize}\item the
627: Coulomb lower-bound parameter $P(\alpha)$ leads, for the relevant maximum
628: coupling $\alpha=a=0.2,$ in the Coulomb term of the sum approximation
629: to$$\underline{P}(-1)=P(\alpha)=P(a)=0.9789063\ ;$$\item the lower-bound
630: parameter required for any linear part of sum potentials is copied from
631: Subsect.~\ref{Sec:EnvULB},$$\underline{P}(1)=1.2457\ .$$\end{itemize}As
632: before, the Rayleigh--Ritz or variational upper bound is found in a trial
633: space of dimension $d=25$ spanned by the generalized Laguerre basis
634: (summarized in App.~\ref{App:Lag}).
635:
636: Figure~\ref{Fig:FP} depicts the bounds to the lowest eigenvalue $E_0$ of $H$
637: as function of the overall coupling strength $v$~in the funnel potential
638: (\ref{Eq:FP}). Remarkably, variational upper and sum lower bounds now
639: restrict $E_0$ to a narrow band.
640:
641: \begin{figure}[ht]\begin{center}\psfig{figure=funnel.ps,scale=0.76789}
642: \caption{Three upper ({\sl full lines\/}) and two lower ({\sl dashed
643: lines\/}) bounds on the ground-state energy eigenvalue ({\bf E}) of the
644: semirelativistic Hamiltonian $H$ with the so-called funnel potential
645: $V(r)=v\,(-a/r+b\,r),$ where $a=0.2,$ $b=0.5,$ $m=1.$ These include: the
646: harmonic-oscillator ({\bf H}), linear ({\bf L}) and variational ({\bf V})
647: upper bounds and the sum-approximation ({\bf S}) and Coulomb ({\bf C}) lower
648: bounds.}\label{Fig:FP}\end{center}\end{figure}
649:
650: \section{Approximate Solutions: Quality}Having determined --- for instance,
651: by application of the Rayleigh--Ritz technique sketched in
652: Subsect.~\ref{Sec:RRVT} --- for some $k=0,1,2,\dots$ the state
653: $|\widehat\chi_k\rangle\in D_d$ corresponding to any upper bound $\widehat
654: E_k$ on the exact eigenvalue $E_k$ of $H,$ one question immediately arises:
655: how closely resembles the approximate solution $|\widehat\chi_k\rangle$ the
656: exact eigenstate $|\chi_k\rangle?$
657:
658: \newpage\noindent Standard criteria, such as the (relative) distance between
659: $\widehat E_k$ and true $E_k,$ or the overlap of approximate and exact
660: eigenstates, require the knowledge of the exact solution. In contrast to
661: this, the virial theorem (Sect.~\ref{Sec:RVT}) represents an indicator for
662: the accuracy of approximate eigenstates that merely uses information provided
663: by the variational approach: Since all eigenstates of $H$ satisfy any
664: relation of the form (\ref{Eq:EV(C)}), a significant imbalance in
665: Eq.~(\ref{Eq:RVT}) reveals that this approximation is far from optimum
666: \cite{Lucha:Q/A,Lucha:A/Q,Lucha02-DMAII}. Of course, because of the
667: involvement (\ref{Eq:EV(C)}) of the dilation generator (\ref{Eq:DilGen}) in
668: the derivation of Eq.~(\ref{Eq:RVT}), any variational solution found by
669: minimization of expectation values of $H$ with respect to the scale
670: transformations, or dilations, (\ref{Eq:scale}) will necessarily satisfy our
671: master virial theorem (\ref{Eq:RVT}).
672:
673: \section{Summary, Concluding Remarks}The various efficient approaches
674: presented here allow to analyze the semirelativistic Hamiltonians of the
675: spinless Salpeter equation analytically; this is crucial for general
676: considerations that aim to answer questions of principle, like operator
677: boundedness. For numerical methods, see, for instance,
678: Refs.~\cite{Lucha92:MAP,Lucha:SAMM,{Lucha94:Como}} and the references
679: therein.
680:
681: \appendix\section{The Generalized Laguerre Basis}\label{App:Lag}Assume every
682: basis function of $L_2(R^3)$ to factorize into a function of the radial
683: variable and the angular term. Its configuration-space representation has the
684: general form$$\psi_{k,\ell m}(\mbox{\boldmath{$x$}})=\Phi_{k,\ell}(r)\,{\cal
685: Y}_{\ell m}(\Omega_r)\ ,\quad r\equiv|\mbox{\boldmath{$x$}}|\ ;$$the
686: spherical harmonics ${\cal Y}_{\ell m}(\Omega)$ for angular momentum $\ell$
687: and projection $m$ depend on the solid angle $\Omega\equiv(\theta,\phi)$ and
688: satisfy a well-known orthonormalization condition:$$\int{\rm d}\Omega\,{\cal
689: Y}^\ast_{\ell m}(\Omega)\,{\cal
690: Y}_{\ell'm'}(\Omega)=\delta_{\ell\ell'}\,\delta_{mm'}\ .$$
691:
692: The most popular choice
693: \cite{Jacobs86,Lucha:LagB,Lucha:Oberwoelz,Lucha:Dubrovnik} for the basis
694: states which span the Hilbert space $L_2(R^+)$ of [with the weight
695: $w(x)=x^2$] square-integrable functions $f(x)$ on the positive real line
696: $R^+$ --- which is the Hilbert space of radial trial functions
697: $\Phi_{k,\ell}(r)$ --- involves the generalized Laguerre polynomials
698: $L_k^{(\gamma)}(x),$ for parameter $\gamma$ \cite{Abramowitz,Bateman}:
699: $$\Phi_{k,\ell}(r)=N_{k,\ell}^{(\mu,\beta)}\,r^{\ell+\beta-1}\exp(-\mu\,r)\,
700: L_k^{(2\,\ell+2\,\beta)}(2\,\mu\,r)\ ;$$these generalized Laguerre
701: polynomials for parameter $\gamma$ are orthogonal polynomials, defined by the
702: power series$$L_k^{(\gamma)}(x)=\sum_{t=0}^k\,(-1)^t\left(\begin{array}{c}
703: k+\gamma\\k-t\end{array}\right)\frac{x^t}{t!}\ ,\quad k=0,1,\dots\ ,$$and
704: orthonormalized with weight function
705: $x^\gamma\exp(-x)$:\begin{eqnarray*}&&\int\limits_0^\infty{\rm
706: d}x\,x^\gamma\exp(-x)\,L_k^{(\gamma)}(x)\,L_{k'}^{(\gamma)}(x)\\
707: &&=\frac{\Gamma(\gamma+k+1)}{k!}\,\delta_{kk'}\ ,\quad k,k'=0,1,\dots\
708: .\end{eqnarray*}
709:
710: The basis states defined by the generalized-Laguerre choice for the radial
711: basis functions $\Phi_{k,\ell}(r)$ involve two parameters, both of which may
712: be subsequently adopted for variational purposes: $\mu$ (with the dimension
713: of mass) and $\beta$ (dimensionless); requirements of normalizability of our
714: basis states constrain the parameters to the ranges$$0<\mu<\infty\ ,\quad
715: -1<2\,\beta<\infty\ .$$Therein, the orthonormality of the generalized
716: Laguerre polynomials, inherent to their definition, is equivalent to the
717: orthonormality of the radial basis functions $\Phi_{k,\ell}(r)$:
718: $$\int\limits_0^\infty{\rm d}r\,r^2\,\Phi_{k,\ell}(r)\,\Phi_{k',\ell}(r)
719: =\delta_{kk'}\ ,\quad k,k'=0,1,\dots\ ;$$this condition fixes the
720: normalization constant $N_{k,\ell}^{(\mu,\beta)}$~to
721: $$N_{k,\ell}^{(\mu,\beta)}=\sqrt{\frac{(2\,\mu)^{2\,\ell+2\,\beta+1}\,k!}
722: {\Gamma(2\,\ell+2\,\beta+k+1)}}\ .$$
723:
724: \newpage
725:
726: Fortunately the assumed factorization of every basis function persists in its
727: momentum-space representation:$$\widetilde\psi_{k,\ell
728: m}(\mbox{\boldmath{$p$}})=\widetilde\Phi_{k,\ell}(p)\,{\cal Y}_{\ell
729: m}(\Omega_p)\ ,\quad p\equiv|\mbox{\boldmath{$p$}}|\ .$$Analytical statements
730: about Hamiltonians that involve a kinetic-energy operator nonlocal in
731: configuration space, such as a relativistic square root (\ref{Eq:RKE}), are
732: facilitated by an explicit knowledge of the momentum-space basis states. One
733: of the great advantages of the generalized-Laguerre basis is the availability
734: of its {\em analytic\/} Fourier transform.
735:
736: For all factorizations into radial and angular parts, as consequence of the
737: Fourier transformation acting on the Hilbert space $L_2(R^3)$ of the
738: square-integrable functions on the three-dimensional space $R^3,$ the radial
739: parts of all basis functions that represent the chosen basis vectors in
740: configuration space and momentum space, respectively, are related by
741: so-called Fourier--Bessel
742: transformations:\begin{eqnarray*}\Phi_{k,\ell}(r)&=&{\rm
743: i}^\ell\,\sqrt{\frac{2}{\pi}}\int\limits_0^\infty{\rm
744: d}p\,p^2\,j_\ell(p\,r)\,\widetilde\Phi_{k,\ell}(p)\
745: ,\\\widetilde\Phi_{k,\ell}(p)&=&(-{\rm
746: i})^\ell\,\sqrt{\frac{2}{\pi}}\int\limits_0^\infty{\rm
747: d}r\,r^2\,j_\ell(p\,r)\,\Phi_{k,\ell}(r)\ ,\\&&\quad\mbox{for all}\
748: k=0,1,\dots,\ \ell=0,1,\dots\ ;\end{eqnarray*}the angular-integration
749: remnants $j_n(z)$ ($n=0,\pm 1,\dots$) label the spherical Bessel functions of
750: the first kind \cite{Abramowitz}. For the generalized-Laguerre basis under
751: consideration, these radial basis functions become in momentum space
752: \begin{eqnarray*}\widetilde\Phi_{k,\ell}(p)&=&
753: N_{k,\ell}^{(\mu,\beta)}\,\frac{(-{\rm i})^\ell\,p^\ell}
754: {2^{\ell+1/2}\,\Gamma\left(\ell+\frac{3}{2}\right)}\\
755: &\times&\sum_{t=0}^k\,\frac{(-1)^t}{t!}
756: \left(\begin{array}{c}k+2\,\ell+2\,\beta\\k-t\end{array}\right)\\
757: &\times&\frac{\Gamma(a_{t,\ell;\beta})\,(2\,\mu)^t}
758: {(p^2+\mu^2)^{a_{t,\ell;\beta}/2}}\\
759: &\times&F\left(\frac{a_{t,\ell;\beta}}{2},-\frac{\beta+t}{2};\ell+\frac{3}{2};
760: \frac{p^2}{p^2+\mu^2}\right),\end{eqnarray*}with the hypergeometric series
761: $F(u,v;w;z)$, defined,~in terms of the gamma function $\Gamma,$ by the power
762: series \cite{Abramowitz}\begin{eqnarray*}&&F(u,v;w;z)\\
763: &&=\frac{\Gamma(w)}{\Gamma(u)\,\Gamma(v)}\,\sum_{n=0}^\infty\,
764: \frac{\Gamma(u+n)\,\Gamma(v+n)}{\Gamma(w+n)}\,\frac{z^n}{n!}\
765: ,\end{eqnarray*}and the simplifying abbreviation
766: $a_{t,\ell;\beta}\equiv2\,\ell+\beta+t+2.$
767:
768: Clearly, the momentum-space radial basis functions
769: $\widetilde\Phi_{k,\ell}(p)$ have to satisfy the orthonormalization
770: condition$$\int\limits_0^\infty{\rm
771: d}p\,p^2\,\widetilde\Phi_{k,\ell}^\ast(p)\,\widetilde\Phi_{k',\ell}(p)
772: =\delta_{kk'}\ ,\quad k,k'=0,1,\dots\ .$$
773:
774: \begin{thebibliography}{30}
775: \bibitem{BSE}E.~E.~Salpeter and H.~A.~Bethe, Phys.~Rev.~{\bf 84} (1951) 1232.
776: \bibitem{SE}E.~E.~Salpeter, Phys.~Rev.~{\bf 87} (1952) 328.
777: \bibitem{Lucha91:BSQ}W.~Lucha, F.~F.~Sch\"oberl, and D.~Gromes, Phys.\
778: Rep.~{\bf 200} (1991) 127.
779: \bibitem{Resag94}J.~Resag {\it et al.}, Nucl.~Phys.~A {\bf 578} (1994) 397
780: [nucl-th/9307026].
781: \bibitem{Kopaleishvili01}T.~Kopaleishvili, Phys.~Part.~Nucl.\ {\bf 32} (2001)
782: 560 [hep-ph/0101271].
783: \bibitem{Lucha:RVT}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~Lett.~{\bf 64}
784: (1990) 2733.
785: \bibitem{Lucha:RVTs}W.~Lucha and F.~F.~Sch\"oberl, Mod.~Phys.~Lett.~A~{\bf 5}
786: (1990) 2473.
787: \bibitem{Fock30}V.~Fock, Z.~Phys.~{\bf 63} (1930) 855.
788: \bibitem{Brack83}M.~Brack, Phys.~Rev.~D {\bf 27} (1983) 1950.
789: \bibitem{Lucha94:Como}W.~Lucha and F.~F.~Sch\"oberl, in: Proc.~Int.~Conf.\ on
790: {\it Quark Confinement and the Hadron Spectrum}, edited by N.~Brambilla and
791: G.~M.~Prosperi (World Scientific, River Edge (N.~J.), 1995) p.~100
792: [hep-ph/9410221].
793: \bibitem{Lucha98:Dubna}W.~Lucha and F.~F.~Sch\"oberl, in: Proc.~XI$^{\rm th}$
794: Int.\ Conf.~{\it Problems of Quantum Field Theory}, editors: B.~M.~Barbashov,
795: G.~V.~Efimov, and A.~V.~Efremov (Joint Institute f.~Nuclear Research, Dubna,
796: 1999) p.~482 [hep-ph/9807342].
797: \bibitem{Lucha:Oberwoelz}W.~Lucha and F.~F.~Sch\"oberl, Int.~J.~Mod.~Phys.~A
798: {\bf 14} (1999) 2309 [hep-ph/9812368].
799: \bibitem{Lucha:Dubrovnik}W.~Lucha and F.~F.~Sch\"oberl, Fizika B {\bf 8}
800: (1999) 193 [hep-ph/9812526].
801: \bibitem{Reed78}M.~Reed and B.~Simon, {\em Methods of Modern Mathematical
802: Physics~IV: Analysis~of Operators\/} (Academic Press, New York, 1978)
803: Section~XIII.1.
804: \bibitem{Weinstein72}A.~Weinstein and W.~Stenger, {\em Methods of
805: Intermediate Problems for Eigenvalues -- Theory and Ramifications\/}
806: (Academic Press, New York, 1972) Chapters 1 and 2.
807: \bibitem{Thirring90}W.~Thirring, {\em A Course in Mathematical Physics~3:
808: Quantum Mechanics of Atoms and Molecules\/} (Springer, New York/Wien, 1990)
809: Section~3.5.
810: \bibitem{Lucha96rcpaubel}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~A {\bf 54}
811: (1996) 3790 [hep-ph/9603429].
812: \bibitem{Lucha99-1dimsrcp}W.~Lucha and F.~F.~Sch\"oberl, J.~Math.~Phys.~{\bf
813: 41} (2000) 1778 [hep-ph/9905556].
814: \bibitem{Martin88}A.~Martin, Phys.~Lett.~B {\bf 214} (1988) 561.
815: \bibitem{Lucha00-HO}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, J.~Phys.~A
816: {\bf 34} (2001) 5059 [hep-th/0012127].
817: \bibitem{Lucha01-DMAIa}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, J.~Math.\
818: Phys.~{\bf 42} (2001) 5228 [hep-th/0101223].
819: \bibitem{Lucha01-DMAIb}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, Int.~J.\
820: Mod.~Phys.~A {\bf 17} (2002) 1931 [hep-th/0110220].
821: \bibitem{Lucha02-DMAII}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, Int.~J.\
822: Mod.~Phys.~A {\bf 18} (2003) 2657 [hep-th/0210149].
823: \bibitem{Lucha02-sum}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, J.~Math.\
824: Phys.~{\bf 43} (2002) 5913 [math-ph/0208042].
825: \bibitem{Lucha02-CP}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, in: Proc.\
826: Int.~Conf.~on {\it Quark Confinement and the Hadron Spectrum V},
827: eds.~N.~Brambilla and G.~M.~Prosperi (World Scientific, Singapore, 2003)
828: p.~500.
829: \bibitem{Herbst}I.~W.~Herbst, Commun.~Math.~Phys.~{\bf 53} (1977) 285; {\it
830: ibid}.~{\bf 55} (1977) 316 (addendum).
831: \bibitem{Martin89}A.~Martin and S.~M.~Roy, Phys.~Lett.~B {\bf 233} (1989)
832: 407.
833: \bibitem{Lucha01-NHO}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, J.~Math.\
834: Phys.~{\bf 43} (2002) 1237; {\it ibid}.~{\bf 44} (2003) 2724 (E)
835: [math-ph/0110015].
836: \bibitem{Lucha03-NHO-m=0}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, Phys.\
837: Lett.~A {\bf 320} (2003) 127 [math-ph/0311032].
838: \bibitem{Lucha04-NV(r2)}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, J.~Math.\
839: Phys.~{\bf 45} (2004) 3086 [math-ph/0405025].
840: \bibitem{Jacobs86}S.~Jacobs, M.~G.~Olsson, and C.~Suchyta III, Phys.\ Rev.~D
841: {\bf 33} (1986) 3338; {\it ibid}.~{\bf 34} (1986) 3536 (E).
842: \bibitem{Lucha:LagB}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~A {\bf 56}
843: (1997) 139 [hep-ph/9609322].
844: \bibitem{Lucha94:VA-SRCP}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~D {\bf 50}
845: (1994) 5443 [hep-ph/9406312].
846: \bibitem{Lucha96:CCC}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Lett.~B {\bf 387}
847: (1996) 573 [hep-ph/9607249].
848: \bibitem{Lucha92:QAQBS}W.~Lucha and F.~F.~Sch\"oberl,
849: Int.~J.~Mod.~Phys.~A~{\bf 7} (1992) 6431.
850: \bibitem{Lucha:Q/A}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~A {\bf 60}
851: (1999) 5091 [hep-ph/9904391].
852: \bibitem{Lucha:A/Q}W.~Lucha and F.~F.~Sch\"oberl, Int.~J.~Mod.~Phys.~A {\bf
853: 15} (2000) 3221 [hep-ph/9909451].
854:
855: \newpage
856:
857: \bibitem{Lucha92:MAP}W.~Lucha, H.~Rupprecht, and F.~F.~Sch\"oberl, Phys.\
858: Rev.~D {\bf 45} (1992) 1233.
859: \bibitem{Lucha:SAMM}W.~Lucha and F.~F.~Sch\"oberl, Int.~J.~Mod.~Phys.~C {\bf
860: 11} (2000) 485 [hep-ph/0002139].
861: \bibitem{Abramowitz}{\it Handbook of Mathematical Functions}, eds.~M.\
862: Abramowitz and I.~A.~Stegun (Dover, New York, 1964).
863: \bibitem{Bateman}Bateman Manuscript Project, A.~Erd\'elyi {\it et al}., {\em
864: Higher Transcendental Functions} (McGraw--Hill, New York, 1953) Volume~II.
865: \end{thebibliography}\end{document}
866: