hep-th0210149/WC6
1: \documentstyle[12pt,epsfig,graphics]{article}\pagestyle{myheadings}
2: \textheight=25cm\topmargin=-2.5cm\textwidth=16.2cm
3: \oddsidemargin-0.1cm\evensidemargin-0.1cm\sloppy\frenchspacing\flushbottom
4: \renewcommand{\topfraction}{1.0}\renewcommand{\bottomfraction}{1.0}
5: \renewcommand{\textfraction}{0.0}\renewcommand{\floatpagefraction}{0.0}
6: \begin{document}\bibliographystyle{plain}\begin{titlepage}
7: \renewcommand{\thefootnote}{\fnsymbol{footnote}}\hfill\begin{tabular}{l}
8: CUQM-94\\HEPHY-PUB 760/02\\UWThPh-2002-19\\hep-th/0210149\\October
9: 2002\end{tabular}\\[1.5cm]\Large\begin{center}{\bf DISCRETE SPECTRA OF
10: SEMIRELATIVISTIC HAMILTONIANS}\\\vspace{0.8cm}\large{\bf Richard L.
11: HALL\footnote[3]{\normalsize\ {\em E-mail address\/}:
12: rhall@mathstat.concordia.ca}}\\[.3cm]\normalsize Department of Mathematics
13: and Statistics, Concordia University,\\1455 de Maisonneuve Boulevard West,
14: Montr\'eal, Qu\'ebec, Canada H3G 1M8\\[0.7cm]\large{\bf Wolfgang
15: LUCHA\footnote[1]{\normalsize\ {\em E-mail address\/}:
16: wolfgang.lucha@oeaw.ac.at}}\\[.3cm]\normalsize Institut f\"ur
17: Hochenergiephysik,\\\"Osterreichische Akademie der
18: Wissenschaften,\\Nikolsdorfergasse 18, A-1050 Wien,
19: Austria\\[0.7cm]\large{\bf Franz F.~SCH\"OBERL\footnote[2]{\normalsize\ {\em
20: E-mail address\/}: franz.schoeberl@univie.ac.at}}\\[.3cm]\normalsize Institut
21: f\"ur Theoretische Physik, Universit\"at Wien,\\Boltzmanngasse 5, A-1090
22: Wien, Austria\vfill {\normalsize\bf Abstract}\end{center}\normalsize We
23: review various attempts to localize the discrete spectra of semirelativistic
24: Hamiltonians of the form $H=\beta\sqrt{m^2+{\bf p}^2}+V(r)$ (defined, without
25: loss of generality but for definiteness, in three spatial dimensions) as
26: entering, for instance, in the spinless Salpeter equation;~every Hamiltonian
27: in this class of operators consists of the relativistic kinetic energy
28: $\beta\sqrt{m^2+{\bf p}^2},$ where $\beta>0$ allows for the possibility of
29: more than one particles of mass $m,$ and a spherically symmetric attractive
30: potential $V(r),$ $r\equiv|{\bf x}|.$ In general, accurate eigenvalues of a
31: nonlocal Hamiltonian operator can only be found by the use of a numerical
32: approximation procedure. Our main emphasis, however, is on the derivation of
33: rigorous semi-analytical expressions~for both upper and lower bounds to the
34: energy levels of such operators. We compare the~bounds obtained within
35: different approaches and present relationships existing between the bounds.
36: 
37: \vspace{3ex}
38: 
39: \noindent{\em PACS numbers\/}: 03.65.Ge, 03.65.Pm, 11.10.St
40: \renewcommand{\thefootnote}{\arabic{footnote}}\end{titlepage}
41: 
42: \normalsize
43: 
44: \section{Introduction: The ``spinless-Salpeter'' Hamiltonian}We investigate
45: the discrete spectrum---that is, the set of eigenvalues $E_{n\ell}$ below the
46: onset~of the essential spectrum, corresponding to bound states characterized
47: by the radial quantum number $n=1,2,3,\dots$ and the angular-momentum quantum
48: number $\ell=0,1,2,\dots$---of the semirelativistic ``spinless-Salpeter''
49: Hamiltonian for particles of mass $m$ and momentum~${\bf p},$\begin{equation}
50: H=\beta\sqrt{m^2+{\bf p}^2}+V(r)\ ,\quad r\equiv|{\bf x}|\ ,\quad\beta>0\
51: ,\label{Eq:(sr)SSH}\end{equation}where $V(r)$ is an attractive central
52: potential in three spatial dimensions. As is evident~from the presence of the
53: relativistic kinetic-energy operator $\beta\sqrt{m^2+{\bf p}^2}$ (with
54: $\beta>0$ allowing, e.g., for more than one particle), this Hamiltonian may
55: be regarded as the generalization~of the nonrelativistic Schr\"odinger
56: Hamiltonian towards relativistic kinematics. The eigenvalue equation of the
57: operator (\ref{Eq:(sr)SSH}), called the ``spinless Salpeter equation,''
58: arises as a well-defined approximation to the Bethe--Salpeter formalism for
59: the description of bound states within~a relativistic quantum field theory.
60: Our main goal is to discuss bounds on the energy levels~$E.$
61: 
62: \section{Envelope theory for energy bounds}\label{Sec:ET}The envelope theory,
63: developed in Refs.~\cite{Hall83,Hall84}, constructs bounds on the discrete
64: eigenvalues $E_{n\ell}$ of a given operator $H$ by comparing the spectrum of
65: $H$ with the spectrum of a suitably formulated ``tangential problem,'' for
66: which some spectral information is known (or may~be obtained with
67: considerable more ease than for the original problem). More
68: specifically,~when applied to the (semirelativistic) Hamiltonian $H$ defined
69: in Eq.~(\ref{Eq:(sr)SSH}) \cite{Lucha00-HO,Lucha01-DMAI}, the envelope theory
70: compares $H$ with a corresponding ``tangential Hamiltonian''$$\widetilde
71: H=\beta\sqrt{m^2+{\bf p}^2}+vh(r)\ ,\quad v>0\ ,$$involving some ``basis
72: potential'' $h(r).$ Under the assumption that the interaction potential under
73: consideration, $V(r),$ is a smooth transformation $V(r)=g(h(r))$ of the basis
74: potential $h(r),$ where the function $g(h)$ has definite convexity, this
75: comparison yields rigorous bounds on the discrete eigenvalues $E_{n\ell}$ of
76: $H.$ More precisely, if the function $g(h)$ is convex, that~is, $g''(h)>0,$
77: we obtain lower energy bounds; if the function $g(h)$ is concave, that is,
78: $g''(h)<0,$ we obtain upper energy bounds. If the eigenvalues $\widetilde E$
79: of $\widetilde H$ are not known (analytically),~the envelope theory may take
80: advantage of the knowledge of suitable bounds on the eigenvalues, namely, of
81: lower bounds on $\widetilde E_{n\ell}$ for convex $g(h)$ or of upper bounds
82: on $\widetilde E_{n\ell}$ for concave~$g(h).$
83: 
84: Suppressing the quantum numbers $n,$ $\ell$ identifying the bound state under
85: consideration, we denote, for a given value of the coupling $v$ entering into
86: the ``tangential Hamiltonian''~$\widetilde H,$ the eigenvalues of $\widetilde
87: H$ or the appropriate---in the above sense---bounds to the latter by $e(v).$
88: The strict bounds on the discrete eigenvalues $E$ of $H$ predicted within the
89: framework of~the envelope technique may then be summarized in form of the
90: ``principal envelope formula''~\cite{Lucha01-DMAI}\begin{equation}
91: E\approx\min_{v>0}[e(v)-ve'(v)+g(e'(v))]\ ,\label{Eq:PEF}\end{equation}where
92: the sign of approximate equality is used to recall that, for a definite
93: convexity~of~$g(h),$ the envelope theory yields lower bounds for convex
94: $g(h)$ and~upper bounds for concave $g(h).$
95: 
96: For the tangential problem posed by the (semirelativistic) ``tangential
97: Hamiltonian'' $\widetilde H,$ spectral information is, at present, available
98: if the basis potential $h(r)$ is either the Coulomb potential, that is,
99: $h(r)=-1/r,$ or the harmonic-oscillator potential, that is, $h(r)=r^2.$~The
100: energy bounds resulting in these two cases are discussed, in turn, in
101: Sec.~\ref{Sec:CLB} and Sec.~\ref{Sec:HOUB}. The case of the interaction
102: potential $V(r)$ in Eq.~(\ref{Eq:(sr)SSH}) being the harmonic-oscillator
103: potential~may be traced back to a nonrelativistic Schr\"odinger problem; this
104: is studied for its own in Sec.~\ref{Sec:HOP}.
105: 
106: \section{Harmonic-oscillator potential}\label{Sec:HOP}The ``relativistic
107: harmonic-oscillator problem,'' defined by the special case of $V(r)$
108: being~the harmonic-oscillator potential $V(r)=vr^2,$ may be analyzed by
109: observing \cite{Lucha96a,Lucha98O,Lucha98D,Lucha99Q,Lucha99A} that~in
110: momentum-space representation the semirelativistic Hamiltonian
111: $H=\beta\sqrt{m^2+{\bf p}^2}+vr^2$~of Eq.~(\ref{Eq:(sr)SSH}) simplifies to
112: the (nonrelativistic) Schr\"odinger operator ${\cal H}$ given
113: by\begin{equation}{\cal H}=-v\Delta+{\cal V}(r)\ ,\quad{\cal
114: V}(r)=\beta\sqrt{m^2+r^2}\ .\label{Eq:Ham-RHOP}\end{equation}Thus the
115: eigenvalue equation for $H$ reduces to an easier-to-treat nonrelativistic
116: Schr\"odinger problem with an effective interaction potential ${\cal V}(r)$
117: which is reminiscent of the square root.
118: 
119: \subsection{Upper bounds on harmonic-oscillator energy
120: levels}\label{Subsec:HOUB}The potential ${\cal V}$ is a concave transform of
121: a harmonic-oscillator potential. Thus we may~find upper bounds on the
122: eigenvalues $E_{n\ell}(v)$ of ${\cal H}$ (cf.\ Eq.~(2.2) of
123: Ref.~\cite{Lucha00-HO} or Eq.~(12) of
124: Ref.~\cite{Lucha01-DMAI}):\begin{equation}E_{n\ell}(v)\leq
125: \min_{r>0}\left[\beta\sqrt{m^2+\frac{P_{n\ell}^2(2)}{r^2}}+vr^2\right],\quad
126: P_{n\ell}(2)=2n+\ell-\frac{1}{2}\ .\label{Eq:HO-UBs}\end{equation}The
127: (dimensionless) numbers $P_{n\ell}(2)$ introduced here are related to the
128: eigenvalues ${\cal E}_{n\ell}$~of~the Hamiltonian $\widetilde H={\bf
129: p}^2+vr^2$ by ${\cal E}_{n\ell}=2\sqrt{v}P_{n\ell}(2).$
130: 
131: Upper bounds on the eigenvalues of self-adjoint operators bounded from below
132: may~also be found by combining the minimum--maximum principle
133: \cite{Reed78,Thirring90} with appropriate operator inequalities. (For
134: details, see, e.g., Sec.~3.4 of Ref.~\cite{Lucha96a}, Sec.~3.5 of
135: Ref.~\cite{Lucha98O}, Sec.~2.4 of Ref.~\cite{Lucha98D}, or Appendix~A of
136: Ref.~\cite{Lucha00-1DRCP}.) Exploiting the obvious positivity of
137: $(\sqrt{m^2+{\bf p}^2}-\mu)^2,$ where~$\mu$ is an arbitrary real parameter
138: (with the dimension of mass), we obtain a set of inequalities for the
139: square-root operator $\sqrt{m^2+{\bf p}^2}$ entering into the kinetic energy
140: of the Hamiltonian~(\ref{Eq:(sr)SSH}); cf.\ also Ref.~\cite{Martin88}:
141: $$\sqrt{m^2+{\bf p}^2}\le\frac{{\bf p}^2+m^2+\mu^2}{2\mu}\quad\forall\ \mu>0\
142: .$$(This operator inequality may also be found by constructing the tangent
143: line $a({\bf p}_1^2){\bf p}^2+b({\bf p}_1^2)$ to the square root
144: $\sqrt{m^2+{\bf p}^2}$ at the point of contact ${\bf p}_1^2=\mu^2-m^2;$ see
145: Sec.~4 of Ref.~\cite{Lucha01-DMAI}.) Amending this result by the interaction
146: potential $V(r),$ one finds that the spinless-Salpeter Hamiltonian
147: (\ref{Eq:(sr)SSH}) satisfies$$H\le H_{\rm NR}(\mu)\equiv\beta\frac{{\bf
148: p}^2+m^2+\mu^2}{2\mu}+V(r)\quad\forall\ \mu>0\ .$$The min--max principle then
149: tells us that the discrete eigenvalues $E_{n\ell}$ of the semirelativistic
150: Hamiltonian $H$ of Eq.~(\ref{Eq:(sr)SSH}) are bounded from above by the
151: corresponding discrete eigenvalues $E_{{\rm NR},n\ell}(\mu)$ of the
152: nonrelativistic Hamiltonian $H_{\rm NR}(\mu)$ and consequently also by the
153: minimum $\bar E_{n\ell}$ with respect to $\mu$ of all these upper bounds
154: \cite{Lucha96a}:$$E_{n\ell}\le\bar E_{n\ell}\equiv\min_{\mu>0}E_{{\rm
155: NR},n\ell}(\mu)\ .$$Applying these general findings to the case of the
156: harmonic-oscillator potential, one ends~up with the following upper bounds
157: for all energy levels of the relativistic harmonic-oscillator problem (see
158: Appendix~A of Ref.~\cite{Lucha99A}; for the ease of comparison, we express
159: here the bounds of Ref.~\cite{Lucha99A} in terms of $P_{n\ell}(2)$ rather
160: than in terms of $E_{{\rm NR},n\ell}(\mu)$):
161: \begin{equation}E_{n\ell}(v)\le\bar E_{n\ell}\equiv\min_{\mu>0}\left[
162: \beta\frac{m^2+\mu^2}{2\mu}+\sqrt{\frac{2\beta v}{\mu}}P_{n\ell}(2)\right],
163: \quad P_{n\ell}(2)=2n+\ell-\frac{1}{2}\ .\label{Eq:HO-MMOI-UBs}\end{equation}
164: 
165: \newpage Despite their seemingly different form, the upper bounds in
166: Eqs.~(\ref{Eq:HO-UBs}) and (\ref{Eq:HO-MMOI-UBs}) are, in~fact, identical, in
167: the sense that the two functions on the right-hand side of these
168: inequalities~have their minimum at the same critical point and take the same
169: miminal value at this point.
170: 
171: In order to demonstrate this equivalence, let us start from the ``min--max''
172: expression~(\ref{Eq:HO-MMOI-UBs}) for these upper bounds. For simplicity of
173: notation, we suppress for the moment the quantum numbers $n,$ $\ell$ which
174: identify the bound states under consideration. The upper bound $\bar E$~then
175: reads$$\bar E=\min_{\mu>0}
176: \left[\beta\frac{m^2+\mu^2}{2\mu}+\sqrt{\frac{\beta v}{2\mu}}2P\right].$$We
177: introduce the abbreviation $q=p^2,$ define the function $f(q)=\sqrt{m^2+q},$
178: which~satisfies$$f'(q)=\frac{1}{2f(q)}\ ,\quad f^2(q)=m^2+q\ ,$$and identify
179: the (positive but otherwise arbitrary) parameter $\mu$ according to
180: $\mu=f(q).$~Next, we introduce a function $F$ by the definition
181: $F(v)=2P\sqrt{v}.$ A couple of algebraic steps then recasts the upper bound
182: $\bar E$ into the form$$\bar E=\min_{q>0}[F(\beta vf')+\beta(f-qf')]\ .$$The
183: critical points of this latter expression are obtained by differentiation
184: with respect~to~$q.$ Since $f''(q)\ne 0,$ this yields$$\hat q=vF'(\beta
185: vf'(\hat q))$$and$$\bar E=F(\beta vf'(\hat q))-\beta vf'(\hat q)F'(\beta
186: vf'(\hat q))+\beta f(\hat q)\ .$$
187: 
188: Let us define a new problem, related to the above one by a Legendre
189: transformation~\cite {Gelfand}, which has the same minimum and the same
190: critical point, but the function to be minimized is {\em different}. Let
191: $u(q)=\beta vf'(q)>0$ and consider the energy expression$$\hat
192: E=\min_{u>0}[F(u)-uF'(u)+\beta f(vF'(u))]\ .$$Since $F''(u)\ne 0,$ we find
193: that the critical point is given by $\hat u=\beta vf'(vF'(\hat u)).$ But,
194: given the definition of $u,$ this equation is equivalent to the earlier
195: critical equation $\hat q=vF'(\beta vf'(\hat q)).$ Meanwhile, the critical
196: {\em value\/} $\hat E$ is also the same as the critical value $\bar E$
197: because we have$$\hat E=F(\beta vf'(\hat q))-\beta vf'(\hat q)F'(\beta
198: vf'(\hat q))+\beta f(\hat q)=\bar E\ .$$
199: 
200: Having shown that $\bar E$ yields the same results as $\hat E,$ we now
201: transform $\hat E$ into the form~(\ref{Eq:HO-UBs}). First of all, we note
202: that $F(u)-uF'(u)=P\sqrt{u}.$ Next, we define a new dummy~minimization
203: variable $r>0$ by$$\sqrt{u}=\frac{vr^2}{P}\ .$$From this, it follows
204: that$$F'(u)=\frac{P}{\sqrt{u}}=\frac{P^2}{vr^2}\ .$$Finally, remembering the
205: definition $f(q)=\sqrt{m^2+q}$ of $f(q),$ we may transform $\hat E$ into the
206: expression on the right-hand side of the inequality (\ref{Eq:HO-UBs}); this
207: establishes the equivalence~of~the upper bounds (\ref{Eq:HO-UBs}) and
208: (\ref{Eq:HO-MMOI-UBs}).
209: 
210: \subsection{Lower bounds on harmonic-oscillator energy
211: levels}\label{Subsec:HOLB}The potential ${\cal V}$ in Eq.~(\ref{Eq:Ham-RHOP})
212: is a convex transform of a linear potential. Therefore we may~find lower
213: bounds on the eigenvalues $E_{n\ell}(v)$ of ${\cal H}$ (cf.\ Eq.~(2.2) of
214: Ref.~\cite{Lucha00-HO} or Eq.~(12) of
215: Ref.~\cite{Lucha01-DMAI}):\begin{equation}E_{n\ell}(v)\ge
216: \min_{r>0}\left[\beta\sqrt{m^2+\frac{P_{n\ell}^2(1)}{r^2}}+vr^2\right],
217: \label{Eq:HO-LBs}\end{equation}where the (dimensionless) numbers
218: $P_{n\ell}(1)$ introduced here are related to the eigenvalues~${\cal
219: E}_{n\ell}$ of the Hamiltonian $\widetilde H={\bf p}^2+vr$ by$${\cal
220: E}_{n\ell}=3\left[\frac{vP_{n\ell}(1)}{2}\right]^{2/3}\ ;$$for the lowest
221: states ($n=1,2,\dots,5,$ $\ell=0,1,\dots,5$) these numbers may be found
222: in~Table~\ref{Tab:P(1)-numbers}.
223: 
224: \begin{table}[ht]\caption{Numerical values of the energy-related numbers
225: $P_{n\ell}(1)$ for the lowest energy levels.}\label{Tab:P(1)-numbers}
226: \begin{center}\begin{tabular}{ccr}\hline\hline&&\\[-1.5ex]
227: \multicolumn{1}{c}{$n$}&\multicolumn{1}{c}{$\ell$}&
228: \multicolumn{1}{c}{$P_{n\ell}(1)$}\\[1ex]\hline\\[-1.5ex]
229: 1&0&1.37608\\2&0&3.18131\\3&0&4.99255\\4&0&6.80514\\5&0&8.61823\\[.5ex]
230: 1&1&2.37192\\2&1&4.15501\\3&1&5.95300\\4&1&7.75701\\5&1&9.56408\\[1ex]
231: \hline\hline\end{tabular}$\qquad$\begin{tabular}{ccr}\hline\hline&&\\[-1.5ex]
232: \multicolumn{1}{c}{$n$}&\multicolumn{1}{c}{$\ell$}&
233: \multicolumn{1}{c}{$P_{n\ell}(1)$}\\[1ex]\hline\\[-1.5ex]
234: 1&2&3.37018\\2&2&5.14135\\3&2&6.92911\\4&2&8.72515\\5&2&10.52596\\[.5ex]
235: 1&3&4.36923\\2&3&6.13298\\3&3&7.91304\\4&3&9.70236\\5&3&11.49748\\[1ex]
236: \hline\hline\end{tabular}$\qquad$\begin{tabular}{ccr}\hline\hline&&\\[-1.5ex]
237: \multicolumn{1}{c}{$n$}&\multicolumn{1}{c}{$\ell$}&
238: \multicolumn{1}{c}{$P_{n\ell}(1)$}\\[1ex]\hline\\[-1.5ex]
239: 1&4&5.36863\\2&4&7.12732\\3&4&8.90148\\4&4&10.68521\\5&4&12.47532\\[.5ex]
240: 1&5&6.36822\\2&5&8.12324\\3&5&9.89276\\4&5&11.67183\\5&5&13.45756\\[1ex]
241: \hline\hline\end{tabular}\end{center}\end{table}
242: 
243: Lower bounds on the eigenvalues of (self-adjoint) Hamiltonians involving the
244: relativistic kinetic energy $\sqrt{m^2+{\bf p}^2}$ may also be found by
245: taking into account the obvious positivity~of $[{\bf
246: p}^2\xi^2-m^2(1-\xi^2)]^2$ for an arbitrary real parameter $\xi,$ which may
247: be converted into the~set of operator inequalities\begin{equation}
248: \sqrt{m^2+{\bf p}^2}\ge|{\bf p}|\sqrt{1-\xi^2}+m\xi\ ,\quad 0\le\xi\le 1\
249: .\label{Eq:RKE-LB}\end{equation}Applying these general findings to the
250: semirelativistic Hamiltonian $H$ of Eq.~(\ref{Eq:(sr)SSH}), with $V(r)$ being
251: the harmonic-oscillator potential, we obtain {\em lower\/} bounds on {\em
252: all\/} discrete energy levels $E_{n\ell}$ of the relativistic
253: harmonic-oscillator problem (for details, see Appendix~A of
254: Ref.~\cite{Lucha99A}):\begin{equation}E_{n\ell}(v)\ge\underline
255: E_{n\ell}\equiv\max_{0\le\xi\le 1}[\beta\xi
256: m+\beta^{2/3}(1-\xi^2)^{1/3}E_{n\ell}(v;m=0)]\
257: ,\label{Eq:HO-xi-LBs}\end{equation}where $E_{n\ell}(v;m=0)$ denote the
258: corresponding discrete eigenvalues of the operator $|{\bf p}|+vr^2,$ i.e., of
259: the Hamiltonian (\ref{Eq:(sr)SSH}) for $\beta=1$ and a vanishing particle
260: mass $m.$ For states with~$\ell=0,$ for instance, these ``zero-mass'' energy
261: levels are given, in terms of the zeros $z_n$ of the Airy function ${\rm
262: Ai}(z)$ \cite{Abramowitz} ($-z_n=2.33810,\,4.08794,\,5.52055,\,\dots$), by
263: $E_{n0}(v;m=0)=-v^{1/3}z_n$~\cite{Lucha99A}.
264: 
265: As in Sec.~\ref{Subsec:HOUB} we observe that the lower bounds in
266: Eqs.~(\ref{Eq:HO-LBs}) and (\ref{Eq:HO-xi-LBs}) appear, at first~sight, to be
267: rather different. A closer inspection, however, reveals that these bounds are
268: identical.
269: 
270: In order to present an analytical proof of the equivalence of the lower
271: bounds (\ref{Eq:HO-LBs}) and~(\ref{Eq:HO-xi-LBs}) on the energy eigenvalues
272: of the relativistic harmonic-oscillator problem, we recall that the envelope
273: lower bounds in (\ref{Eq:HO-LBs}) are based (in the momentum-space
274: representation~of~$H$)~on~the observation that, since its second derivative
275: with respect to $|{\bf p}|$ is positive, the square-root operator
276: $\sqrt{m^2+{\bf p}^2}$ of the relativistic kinetic energy is bounded from
277: below by tangent~lines $a(t)+b(t)|{\bf p}|;$ this may be expressed by the
278: geometrical inequality $\sqrt{m^2+{\bf p}^2}\ge a(t)+b(t)|{\bf p}|,$ where,
279: by elementary calculus,$$a(t)=\frac{m^2}{\sqrt{m^2+t^2}}\ ,\quad
280: b(t)=\frac{t}{\sqrt{m^2+t^2}}\ ,$$and $|{\bf p}|=t$ is the point of contact
281: between the square root of the relativistic kinetic energy and its
282: straight-line lower approximation. The equivalence of the above two
283: expressions for the harmonic-oscillator lower bound is then shown by a simple
284: change of variable from~$t$~to~$\xi$ given explicitly by$$\xi=\frac{a(t)}{m}\
285: ,$$which, after a little algebra, leads to $b(t)=\sqrt{1-\xi^2}.$ For these
286: coefficient functions $a(t)$~and $b(t)$ the above geometrical inequality
287: becomes exactly the lower bound (\ref{Eq:RKE-LB}) on the relativistic kinetic
288: energy. Consequently, with the basic inequalities being identical, the
289: resulting lower energy bounds have to be identical too.
290: 
291: \section{Convex transform of Coulomb potential}\label{Sec:CLB}For the
292: ``relativistic Coulomb problem'' posed by the spinless-Salpeter
293: Hamiltonian~(\ref{Eq:(sr)SSH})~with $V(r)$ being the Coulomb potential
294: $V(r)=-v/r,$ for coupling constants $v$ smaller than $\beta v_{\rm c},$ that
295: is, $v<\beta v_{\rm c},$ where the critical value $v_{\rm c}$ of the Coulomb
296: coupling constant $v$ is given by$$v_{\rm c}=\frac{2}{\pi}\ ,$$a lower bound
297: to the ground-state energy eigenvalue $E_0$ of $H$ (the bottom of the
298: spectrum of $H$) has been derived by Herbst \cite{Herbst77}:\begin{equation}
299: E_0\ge\beta m\sqrt{1-\left(\frac{\sigma v}{\beta}\right)^{2}}\
300: ,\quad\sigma\equiv\frac{\pi}{2}\ .\label{Eq:RCP-HLB}\end{equation}For some
301: part of the allowed range of the coupling $v,$ this lower bound has been
302: improved~by Martin and Roy \cite{Martin89}:\begin{equation}E_0\ge\beta
303: m\sqrt{\frac{1}{2}\left[1+\sqrt{1-\left(\frac{2v}{\beta}\right)^2}\right]}\
304: ,\quad v<\frac{\beta}{2}\ .\label{Eq:RCP-MRLB}\end{equation}In
305: Ref.~\cite{Lucha01-DMAI}, we recast this Martin--Roy bound into a class of
306: lower bounds on the spectrum~of $H$ all of which are of the form of Herbst's
307: bound (\ref{Eq:RCP-HLB}) [but, of course, slightly weaker than~the
308: Martin--Roy bound (\ref{Eq:RCP-MRLB})]; the new lower bounds are parametrized
309: by a real parameter~$\sigma\ge 1$:\begin{equation}E_0\ge\beta
310: m\sqrt{1-\left(\frac{\sigma v}{\beta}\right)^2}\ ,\quad
311: v\le\beta\frac{\sqrt{\sigma^2-1}}{\sigma^2}<\frac{\beta}{2}\
312: .\label{Eq:RCP-NLB}\end{equation}The constraint on the Coulomb coupling $v$
313: in Eq.~(\ref{Eq:RCP-NLB}) automatically guarantees $v\le\beta/\sigma;$ this
314: latter constraint is a necessary and sufficient condition for the reality of
315: the lower~energy bounds in Eq.~(\ref{Eq:RCP-NLB}).
316: 
317: From the lower bounds (\ref{Eq:RCP-HLB}) or (\ref{Eq:RCP-NLB}) on the
318: spectrum of the relativistic Coulomb problem, for the class of potentials
319: $V(r)$ which are convex transforms $V(r)=g(h(r)),$ $g''>0,$~of~the Coulomb
320: potential $h(r)=-1/r$ an (``absolute'') envelope lower bound on the
321: ground-state eigenvalue $E_0$ (and, hence, on the entire spectrum) of the
322: spinless-Salpeter Hamiltonian~(\ref{Eq:(sr)SSH}) may be found (cf.\ Eq.~(16)
323: of Ref.~\cite{Lucha01-DMAI}):\begin{equation}
324: E_0\ge\min_{r>0}\left[\beta\sqrt{m^2+\frac{P^2}{r^2}}+V(r)\right],\quad
325: v<\beta v_P\ .\label{Eq:SSE-LB}\end{equation}Here we introduced a parameter
326: $P$ by defining $P=1/\sigma.$ Accordingly, the boundary value~$v_P$ of the
327: Coulomb coupling constant $v$ is given, when arising from demanding the
328: Hamiltonian (\ref{Eq:(sr)SSH}) to be bounded from below, by the critical
329: coupling $v_{\rm c},$ i.e., $v_P=v_{\rm c},$ or, when arising~from the region
330: of validity of our new family of Coulomb lower bounds (\ref{Eq:RCP-NLB}),
331: by\begin{equation}v_P=P\sqrt{1-P^2}<\frac{1}{2}\
332: .\label{Eq:BV}\end{equation}It is easy to convince oneself that the Coulomb
333: lower energy bound on the right-hand~side~of the inequality (\ref{Eq:SSE-LB})
334: is a monotone increasing function $\underline E(P)$ of the parameter $P.$ To
335: this~end, we rewrite the inequality (\ref{Eq:SSE-LB}) in the
336: form$$E_0\ge\underline E(P)=\min_{r>0}F(r,P)\
337: ,$$with$$F(r,P)\equiv\beta\sqrt{m^2+\frac{P^2}{r^2}}+V(r)\ .$$Remembering
338: that $V(r)$ is the transform $V(r)=g(h(r))$ of $h(r)=-1/r,$ and assuming~that
339: $g'(h)>0,$ the critical point $\hat r(P)$ (corresponding to the minimum of
340: $F(r,P)$ for a given $P$) may be found, from$$\left.\frac{\partial
341: F(r,P)}{\partial r}\right|_{r=\hat r(P)}=0\ ,$$by solving$$\frac{\beta
342: P^2}{\sqrt{m^2\hat r^2+P^2}}=g'(h(\hat r))\ .$$The substitution of the
343: solution $\hat{r}(P)$ of the latter equation in $F(r,P)$ gives the
344: lower~bound $\underline E(P)=F(\hat r(P),P);$ here we are interested in the
345: total derivative of $\underline E(P)$ with respect~to~$P.$ Formally, this
346: derivative would be given by$$\frac{{\rm d}\underline E(P)}{{\rm
347: d}P}=\frac{\partial F(\hat r(P),P)}{\partial P}+\left.\frac{\partial
348: F(r,P)}{\partial r}\right|_{r=\hat r(P)}\frac{{\rm d}\hat r(P)}{{\rm d}P}\
349: .$$At the critical point, however, the first factor of the second term
350: vanishes and one~is~left~with$$\frac{{\rm d}\underline E(P)}{{\rm
351: d}P}=\frac{\partial F(\hat r(P),P)}{\partial P}=\frac{\beta P}{\hat
352: r\sqrt{m^2\hat r^2+P^2}}>0\ .$$From this we conclude that the lower bound
353: $\underline E(P)$ is a monotone increasing function of $P.$
354: 
355: The existence of the upper bound $v<\beta v_P<\beta/2$ on the Coulomb
356: coupling constant~$v$~is reflected in a corresponding constraint on the
357: parameters of the interaction potential~$V(r).$ The envelope theory
358: constructs a ``tangential potential'' to $V(r)$ at a point of contact~$r=t.$
359: For a potential $V(r)$ which is a convex transform $V(r)=g(h(r))$ of the
360: basis potential~$h(r),$ this tangential potential forms a lower bound to
361: $V(r);$ for $h(r)$ being the Coulomb potential, $h(r)=-1/r,$ this lower bound
362: is a Coulomb potential with the effective coupling constant $v(t)=t^2V'(t),$
363: shifted by the constant $V(t)+tV'(t)$:$$V(r)\ge\widetilde
364: V(r)\equiv-\frac{t^2V'(t)}{r}+V(t)+tV'(t)\ .$$The relativistic Coulomb
365: problem is posed by the ``Coulombic'' Hamiltonian$$H_{\rm
366: C}(v)=\beta\sqrt{m^2+{\bf p}^2}-\frac{v}{r}\ ,\quad v<\beta v_{\rm c}\
367: ,$$which is nothing else but the tangential Hamiltonian $\widetilde H$ with
368: $h(r)=-1/r.$ In the notation~of Sec.~\ref{Sec:ET}, the eigenvalues $E_{\rm
369: C}(v)$ of $H_{\rm C}(v)$ are bounded from below according to $E_{\rm C}(v)\ge
370: e(v),$ with $e(v)$ given by Herbst's bound (\ref{Eq:RCP-HLB}), our new bound
371: (\ref{Eq:RCP-NLB}), or the Martin--Roy~bound~(\ref{Eq:RCP-MRLB}). Expressed
372: in terms of the Coulombic Hamiltonian $H_{\rm C}(v),$ the semirelativistic
373: Hamiltonian (\ref{Eq:(sr)SSH}) thus satisfies the inequality $H\ge H_{\rm
374: C}(v(t))+V(t)+tV'(t);$ its eigenvalues $E$ are bounded from below according
375: to$$E_0\ge E_{{\rm C},0}(v(t))+V(t)+tV'(t)\ge e(v(t))+V(t)+tV'(t)\ .$$The
376: optimized lower bound on the discrete spectrum of $H$ is then found by
377: maximizing~the expression on the right-hand side of this latter inequality
378: with respect to the point of contact $r=t.$ The equation which determines the
379: critical point $\hat t$ has to be solved~by~observing~that, at the critical
380: point $r=\hat t,$ the effective Coulomb coupling $v(t)$ still has to
381: satisfy~$v(\hat t)<\beta v_P;$ this nontrivial requirement imposes the
382: announced ``Coulomb coupling constant constraint'' on the coupling parameters
383: entering in the potential~$V(r).$
384: 
385: It goes without saying that, when applying our ``Herbst-based'' lower energy
386: bounds of Eq.~(\ref{Eq:RCP-NLB}), care has to be taken in order to assure
387: that the chosen values of the parameters~in the potential $V(r)$ under
388: consideration respect this Coulomb coupling constant constraint.
389: 
390: Let us illustrate these general considerations by applying them to a
391: Coulomb-plus-linear or (in view of its shape) ``funnel'' potential
392: \begin{equation}V(r)=-\frac{c_1}{r}+c_2r\ ,\quad c_1\ge 0\ ,\quad c_2\ge 0\
393: .\label{Eq:FP}\end{equation}In this case, the effective Coulomb coupling
394: $v(t)$ explicitly reads $v(t)=c_1+c_2t^2$ and, for~the set of ``Herbst-like''
395: lower energy bounds $e(v)$ given by Eqs.~(\ref{Eq:RCP-HLB}) and
396: (\ref{Eq:RCP-NLB}), the critical point~$\hat t$ is, after some algebra,
397: determined by the relation$$v(\hat t)=\frac{\beta P^2}{\sqrt{m^2\hat
398: t^2+P^2}}\ .$$The above Coulomb coupling constant constraint becomes an upper
399: bound on a particular linear combination of the two involved coupling
400: parameters $c_1$ and $c_2$ (cf.\ Sec.~6 of Ref.~\cite{Lucha01-DMAI}):
401: \begin{equation}c_1+\frac{P^2}{m^2}\left(\frac{P^2}{v_P^2}-1\right)c_2<\beta
402: v_P\ .\label{Eq:CCCC}\end{equation}
403: 
404: The Coulomb lower bounds (\ref{Eq:SSE-LB}) on the spectrum of the
405: semirelativistic Hamiltonian~(\ref{Eq:(sr)SSH}) may be optimized by seeking
406: that value of the parameter $P=1/\sigma$ which, for given values~of the mass
407: $m$ of the bound-state constituent(s) and of the coupling parameters $c_1$
408: and $c_2$ in~the Coulomb-plus-linear potential (\ref{Eq:FP}), solves the
409: Coulomb coupling constant constraint~(\ref{Eq:CCCC}). Taking into account the
410: definition (\ref{Eq:BV}) of the boundary value $v_P$ of the Coulomb
411: coupling~$v,$ this optimum value of $P$ is obtained as the solution of the
412: relation\begin{equation}\frac{c_2\sin^4t}{\cos^2t(\beta\sin t\cos
413: t-c_1)}=m^2\ ,\quad P\equiv\sin t\ .\label{Eq:P(m)}\end{equation}
414: 
415: The set (\ref{Eq:SSE-LB}) of envelope lower bounds distinguished from each
416: other by the parameter~$P$ may be improved somewhat by taking into account
417: the Martin--Roy lower bound (\ref{Eq:RCP-MRLB})~on~the energy spectrum of the
418: relativistic Coulomb problem. To this end, let us denote the energy bound on
419: the right-hand side of the inequality~(\ref{Eq:RCP-MRLB}) by
420: $e(v)$:$$e(v)=\beta
421: m\sqrt{\frac{1}{2}\left[1+\sqrt{1-\left(\frac{2v}{\beta}\right)^2}\right]}\
422: .$$Performing the change of variables$$v=\frac{\beta}{2}\sin 2t\ ,\quad 0\le
423: t<\frac{\pi}{4}\ ,$$we find$$e(v)=\beta m\cos t\ ,\quad e(v)-ve'(v)=\beta
424: m\frac{\cos^3t}{\cos 2t}\ ,$$while from $e'(v)=h(r)$ we get$$r=\frac{\cos
425: 2t}{m\sin t}\ .$$Upon inserting these intermediate results in our principal
426: envelope formula (\ref{Eq:PEF}), we arrive~at the ``Martin--Roy-based''
427: Coulomb lower bound$$E_0\ge\min_{0\le t<\pi/4}\left[\beta
428: m\frac{\cos^3t}{\cos 2t}+V\left(\frac{\cos 2t}{m\sin t}\right)\right].$$
429: 
430: \section{Concave transform of harmonic-oscillator potential}\label{Sec:HOUB}
431: For the class of potentials $V(r)$ which are concave transforms
432: $V(r)=g(h(r)),$ $g''<0,$~of~the harmonic-oscillator potential $h(r)=r^2$
433: envelope upper bounds on the eigenvalues $E_{n\ell}$~of~the spinless-Salpeter
434: Hamiltonian (\ref{Eq:(sr)SSH}) may be found for all energy levels (cf.\
435: Eq.~(18) of Ref.~\cite{Lucha01-DMAI}):\begin{equation}
436: E_{n\ell}\le\min_{r>0}\left[\beta\sqrt{m^2+\frac{P^2}{r^2}}+V(r)\right],\quad
437: P\equiv P_{n\ell}(2)=2n+\ell-\frac{1}{2}\ .\label{Eq:SSE-HOUB}\end{equation}
438: 
439: \section{Concave transform of linear potential}\label{Sec:LUB}For the class
440: of potentials $V(r)$ which are concave transforms $V(r)=g(h(r)),$
441: $g''<0,$~of~the linear potential $h(r)=r$ the ``harmonic-oscillator based''
442: envelope upper bounds (\ref{Eq:SSE-HOUB}) on~the eigenvalues $E_{n\ell}$ of
443: the spinless-Salpeter Hamiltonian (\ref{Eq:(sr)SSH}) may be (significantly)
444: improved by considering the composition of the envelope approximations for
445: the kinetic-energy operator and the interaction-potential operator along the
446: following lines. In the present investigation we focus our attention to
447: attractive potentials $V(r),$ that is, to all potentials which exhibit~a
448: monotone increase with increasing radial coordinate $r$: $g'>0.$ Every
449: function $g(r)$ that~is~a monotone increasing [$g'(r)>0$] and concave
450: [$g''(r)<0$] function of $r$ is a concave function $f(r^2)$ of $r^2$:
451: $g(r)=f(r^2)$ with $f''(r^2)<0.$ Consequently, all potentials $V(r)$ studied
452: in this section also belong to the class of potentials considered in
453: Sec.~\ref{Sec:HOUB}.
454: 
455: The square-root operator $k({\bf p}^2)\equiv\sqrt{m^2+{\bf p}^2}$ of the
456: relativistic kinetic energy, regarded as a function of ${\bf p}^2,$ is
457: monotone increasing (i.e., $k'({\bf p}^2)>0$) and concave (i.e., $k''({\bf
458: p}^2)<0$). Let us assume that the interaction potential $V(r)$ exhibits a
459: similar behaviour, that is,~that the operator $V(r)=g(r)$ is monotone
460: increasing, i.e., $g'(r)>0,$ and concave, i.e., $g''(r)<0.$ These operators
461: are bounded from above by their respective upper-tangent approximations:
462: \begin{eqnarray*}&&k({\bf p}^2)\le a_1(q)+b_1(q){\bf p}^2\ ,\\[1ex]&&g(r)\le
463: a_2(t)+b_2(t)r\ ,\end{eqnarray*}with\begin{eqnarray}&&b_1(q)=k'(q)\ ,\quad
464: a_1(q)=k(q)-qk'(q)\ ,\nonumber\\[1ex]&&b_2(t)=g'(t)\ ,\quad
465: a_2(t)=g(t)-tg'(t)\ ,\label{Eq:comp-coeff}\end{eqnarray}where ${\bf p}^2=q$
466: denotes the point of contact between the kinetic-energy operator $k({\bf
467: p}^2)$ and~its upper-tangent approximation $a_1(q)+b_1(q){\bf p}^2,$ and
468: $r=t$ labels the point of contact between the interaction-potential operator
469: $g(r)$ and {\em its\/} upper-tangent approximation $a_2(t)+b_2(t)r.$
470: Consequently, our semirelativistic Hamiltonian $H$ in Eq.~(\ref{Eq:(sr)SSH})
471: satisfies the operator inequality$$H\equiv\beta k({\bf p}^2)+V(r)\le\bar
472: H(q,t)\equiv\beta\left[a_1(q)+b_1(q){\bf p}^2\right]+a_2(t)+b_2(t)r\ .$$The
473: eigenvalues $\bar E_{n\ell}(q,t)$ of the upper ``tangential Hamiltonian''
474: $\bar H(q,t)$ are given in terms~of the numbers $P_{n\ell}(1)$ listed in
475: Table~\ref{Tab:P(1)-numbers} by the exact formula\begin{equation}\bar
476: E_{n\ell}(q,t)=3\left[\frac{\beta
477: b_1(q)b_2^2(t)P_{n\ell}^2(1)}{4}\right]^{1/3}+\beta a_1(q)+a_2(t)\
478: .\label{Eq:comp-UEB}\end{equation}In order to find the best such upper bound,
479: we must minimize with respect to both $q$ and~$t.$ The critical equations
480: immediately yield the necessary
481: conditions\begin{equation}\frac{1}{2\beta}\frac{g'(\hat t)}{k'(\hat
482: q)}=\frac{\hat q^{3/2}}{P_{n\ell}(1)} =\frac{P_{n\ell}^2(1)}{\hat t^3}\
483: .\label{Eq:nec-cond}\end{equation}From the second of the above equalities we
484: deduce that, for given quantum numbers $n$ and~$\ell,$ the minimizing values
485: of the parameters $q$ and $t$ are constrained to the curve $\hat q\hat
486: t^2=P_{n\ell}^2(1)$~in the $(q,t)$ parameter plane. On this curve, and with
487: the coefficients (\ref{Eq:comp-coeff}) of our upper-tangent approximations
488: and the necessary (critical) conditions (\ref{Eq:nec-cond}), the minimum of
489: the expression on the right-hand side of the upper energy bound
490: (\ref{Eq:comp-UEB}) may be cast into a new, simpler~form:$$\min_{\hat
491: t>0}\left[\beta k\left(\frac{P_{n\ell}^2(1)}{\hat t^2}\right)+g(\hat
492: t)\right].$$By remembering the above definitions of $k({\bf p}^2)$ and
493: $g(r),$ this expression eventually becomes\begin{equation}
494: E_{n\ell}\le\min_{r>0}\left[\beta\sqrt{m^2+\frac{P^2}{r^2}}+V(r)\right],\quad
495: P\equiv P_{n\ell}(1)\ .\label{Eq:LUB}\end{equation}These upper bounds on the
496: entire discrete spectrum $\{E_{n\ell}\}$ are valid for any central potential
497: $V(r)$ which is a monotone increasing and concave function of the linear
498: potential $h(r)=r.$
499: 
500: \section{Variational upper bounds}\label{Sec:VarUB}The standard tool for the
501: derivation of rigorous upper bounds on the eigenvalues $E$ of some
502: self-adjoint operator $H$ is the Rayleigh--Ritz variational technique
503: \cite{Reed78,Thirring90}, which is~based~on the minimum--maximum~principle:
504: If the (discrete) eigenvalues $E_k,$ $k=0,1,2,\dots,$ of $H$~are ordered
505: according to $E_0\le E_1\le E_2\le\cdots,$ then the first $d$ of them are
506: bounded from above by the $d$ eigenvalues $\widehat E_k,$ $k=0,1,\dots,d-1,$
507: ordered according to $\widehat E_0\le\widehat E_1\le\cdots\le\widehat
508: E_{d-1},$~of that operator $\widehat H$ which is obtained by the restriction
509: of $H$ to a $d$-dimensional subspace $D_d$ of the domain of $H,$ that is,
510: $E_k\le\widehat E_k$ for all $k=0,1,\dots,d-1.$ Hence, for the
511: accuracy~of~the obtained results, an appropriate, ``reasonable'' definition
512: of this trial subspace $D_d$ is crucial. If this $d$-dimensional subspace
513: $D_d$ is spanned by a chosen set of $d$ linearly independent~basis vectors
514: $|\psi_k\rangle,$ $k = 0,1,\dots,d-1,$ the set of eigenvalues $\widehat E$
515: may immediately be determined~by diagonalizing the $d\times d$ matrix
516: $(\langle\psi_i|\widehat H|\psi_j\rangle),$ $i,j=0,1,\dots,d-1,$ that is, as
517: the $d$ roots~of~the characteristic equation$$\det(\langle\psi_i|\widehat
518: H|\psi_j\rangle-\widehat E\,\langle\psi_i|\psi_j\rangle)=0\ ,\quad
519: i,j=0,1,\dots,d-1\ .$$
520: 
521: The relativistic virial theorem derived in Ref.~\cite{Lucha89:RVT}---for a
522: very comprehensive review,~see Ref.~\cite{Lucha90:RVTs}---allows to define a
523: precise quantitative measure \cite{Lucha99Q,Lucha99A} for the quality of the
524: results found within the framework of variational techniques: according to
525: the analysis presented~in Refs.~\cite{Lucha99Q,Lucha99A}, for some generic
526: Hamiltonian operator $H$ consisting of a momentum-dependent kinetic-energy
527: term $T({\bf p})$ and a coordinate-dependent interaction-potential term
528: $V({\bf x}),$ i.e.,$$H=T({\bf p})+V({\bf x})\ ,$$the accuracy of trial states
529: $|\varphi\rangle$ which approximate the exact bound state under study~can~be
530: quantitatively estimated by the deviation from zero of the quantity (called
531: $\nu$ in Refs.~\cite{Lucha99Q,Lucha99A})$$Q\equiv1-\frac{\langle\varphi|{\bf
532: p}\cdot\frac{\partial}{\partial{\bf p}}T({\bf p})|\varphi\rangle}
533: {\langle\varphi|{\bf x}\cdot\frac{\partial} {\partial{\bf x}}V({\bf
534: x})|\varphi\rangle}\ (\equiv-\nu)\ .$$The main advantage of this measure for
535: the accuracy of approximate eigenstates $|\varphi\rangle$ is that it does not
536: require any information on the solutions of the investigated eigenvalue
537: problem other than the one provided by the variational approximation
538: technique itself.
539: 
540: In order to get a first idea of the position of the lowest energy level
541: $E_0,$ a one-dimensional trial~space is sufficient; a basis for the latter
542: might be any of the two-parameter (normalized) trial functions$$\phi({\bf
543: x})=\sqrt{\frac{n\alpha^{3/n}}{4\pi\Gamma(\frac{3}{n})}}
544: \exp\left(-\frac{\alpha}{2}r^n\right).$$Here the scale parameter $\alpha>0$
545: is varied. The parameter $n>0$ allows for later optimization. For the sake of
546: definiteness, let us consider an interaction potential $V(r)$ which is the
547: sum~of (attractive) pure power-law terms $a(q)\,{\rm sgn}(q)r^q$ with $q\ne0$
548: and a logarithmic term~$a(0)\ln r$:$$V(r)=\sum_{q\ne0}a(q)\,{\rm
549: sgn}(q)r^q+a(0)\ln r\ .$$We assume, of course, that the potential
550: coefficients $a(q)\ge0$ are not {\em all\/} zero. With the~help of Jensen's
551: inequality \cite{Jensen}, and by a tricky redefinition of the scale variable
552: $\alpha,$~the~variational upper bound on the ground-state eigenvalue $E_0$ of
553: the Hamiltonian (\ref{Eq:(sr)SSH}), resulting from this choice of $\phi({\bf
554: x})$ as basis vector of our one-dimensional trial space, may be cast into the
555: form\begin{equation}E_0\le\min_{r>0}\left[\beta\sqrt{m^2+\frac{1}{r^2}}+
556: \sum_{q\ne0}a(q)\,{\rm sgn}(q)({\cal P}(n,q)r)^q+a(0)\ln({\cal
557: P}(n,0)r)\right],\label{Eq:VUB}\end{equation}with the
558: abbreviations\begin{eqnarray*}{\cal
559: P}(n,q)&=&\frac{n}{2}\left(\frac{\Gamma(2+\frac{1}{n})}
560: {\Gamma(\frac{3}{n})}\right)^{1/2}
561: \left(\frac{\Gamma(\frac{q+3}{n})}{\Gamma(\frac{3}{n})}\right)^{1/q}\ ,\quad
562: q\neq 0\ ,\\[1ex]{\cal
563: P}(n,0)&=&\frac{n}{2}\left({\frac{\Gamma(2+\frac{1}{n})}
564: {\Gamma(\frac{3}{n})}}\right)^{1/2}
565: \exp\left(\frac{1}{n}\psi\left(\frac{3}{n}\right)\right),\end{eqnarray*}where
566: $\psi$ denotes the digamma function
567: $$\psi(z)\equiv\frac{1}{\Gamma(z)}\frac{{\rm d}\Gamma(z)}{{\rm d}z}\ .$$
568: 
569: \section{The ``Laguerre'' trial space}\label{Sec:LagTS}Unless the eigenstate
570: corresponding to some eigenvalue $E_k$ is already an element of the~trial
571: space $D_d,$ any upper bounds on the eigenvalues of an operator bounded from
572: below may~be improved by enlarging $D_d$ to higher $d,$ or by spanning $D_d$
573: by more sophisticated basis states.
574: 
575: For the class of spherically symmetric potentials $V(r)$ studied here, a very
576: popular~choice for the basis states which span the $d$-dimensional trial
577: space $D_d$ of the variational technique are ``Laguerre'' trial states, given
578: in their configuration-space representation by
579: \cite{LTS,Lucha97,Lucha98O,Lucha98D}\begin{equation}\psi_{k,\ell m}({\bf x})=
580: \sqrt{\frac{(2\mu)^{2\ell+2\rho+1}k!}{\Gamma(2\ell+2\rho+k+1)}}r^{\ell+\rho-1}
581: \exp(-\mu r)L_k^{(2\ell+2\rho)}(2\mu r){\cal Y}_{\ell m}(\Omega_{\bf x})\
582: .\label{eq:LagTF}\end{equation}Here $L_k^{(\gamma)}(x)$ are the generalized
583: Laguerre polynomials (for the parameter $\gamma$) \cite{Abramowitz},
584: defined~by the power series$$L_k^{(\gamma)}(x)=\sum_{t=0}^k(-1)^t
585: \left(\begin{array}{c}k+\gamma\\k-t\end{array}\right)\frac{x^t}{t!}$$and
586: orthonormalized, with the weight function $x^\gamma\exp(-x),$ according
587: to$$\int\limits_0^\infty{\rm d}x\,x^\gamma\exp(-x)L_k^{(\gamma)}(x)
588: L_{k'}^{(\gamma)}(x)=\frac{\Gamma(\gamma+k+1)}{k!}\delta_{kk'}\ ;$$${\cal
589: Y}_{\ell m}(\Omega)$ are the spherical harmonics for angular momentum $\ell$
590: and projection~$m,$ depending on the solid angle $\Omega$ and orthonormalized
591: according to$$\int{\rm d}\Omega\,{\cal Y}^\ast_{\ell m}(\Omega){\cal
592: Y}_{\ell'm'}(\Omega)=\delta_{\ell\ell'}\delta_{mm'}\ .$$The trial functions
593: (\ref{eq:LagTF}) involve two variational parameters, $\mu$ (with the
594: dimension of~mass) and $\rho$ (dimensionless), which, by the requirement of
595: normalizability of these functions, are subject to the constraints $\mu>0$
596: and $2\rho>-1.$
597: 
598: One of the main advantages of the choice (\ref{eq:LagTF}) for the variational
599: trial states is the easy availability of an analytic expression for their
600: momentum-space representation, obtained~by Fourier transformation of
601: $\psi_{k,\ell m}({\bf x})$:\begin{eqnarray*}\widetilde\psi_{k,\ell m}({\bf
602: p})&=&\sqrt{\frac{(2\mu)^{2\ell+2\rho+1}k!}{\Gamma(2\ell+2\rho+k+1)}}
603: \frac{(-{\rm i})^\ell|{\bf p}|^\ell}
604: {2^{\ell+1/2}\Gamma\left(\ell+\frac{3}{2}\right)}\sum_{t=0}^k\frac{(-1)^t}{t!}
605: \left(\begin{array}{c}k+2\ell+2\rho\\k-t\end{array}\right)\\[1ex]
606: &\times&\frac{\Gamma(2\ell+\rho+t+2)(2\mu)^t}{({\bf
607: p}^2+\mu^2)^{(2\ell+\rho+t+2)/2}}
608: F\left(\frac{2\ell+\rho+t+2}{2},-\frac{\rho+t}{2};\ell+\frac{3}{2};
609: \frac{{\bf p}^2}{{\bf p}^2+\mu^2}\right)\\[1ex]&\times&{\cal Y}_{\ell
610: m}(\Omega_{\bf p})\ ,\end{eqnarray*}with the hypergeometric series $F,$
611: defined by$$F(u,v;w;z)=\frac{\Gamma(w)}{\Gamma(u)\Gamma(v)}\sum_{n=0}^\infty
612: \frac{\Gamma(u+n)\Gamma(v+n)}{\Gamma(w+n)}\frac{z^n}{n!}\ .$$
613: 
614: The choice of the Laguerre states as the basis of the trial space $D_d$
615: allows to calculate~the matrix elements of (any linear combination of) pure
616: power-law potentials $V(r)=a(q)r^q$~and under certain circumstances (for
617: instance, for $\mu=m$ for {\em either\/} radial {\em or\/} orbital
618: excitations) the matrix elements of the kinetic term on purely analytical
619: grounds, and to arrive therefore at analytical expressions for the $d\times
620: d$ Hamiltonian matrix $(\langle\psi_i|\widehat H|\psi_j\rangle),$
621: $i,j=0,1,\dots,d-1.$ The explicit algebraic expressions of these matrix
622: elements may be found in Refs.~\cite{Lucha97,Lucha98O,Lucha98D}. Up to and
623: including a trial-space dimension $d=4,$ the Hamiltonian matrix
624: $(\langle\psi_i|\widehat H|\psi_j\rangle)$~may be diagonalized algebraically.
625: For the funnel potential (\ref{Eq:FP}), the case $d=1$ yields
626: the~bound$$E_0\le\left(\frac{64\beta}{15\pi}-c_1\right)m+\frac{3c_2}{2m}\ .$$
627: 
628: \section{The ``local-energy'' theorem}\label{Sec:LET}Some information on {\em
629: both\/} upper {\em and\/} lower bounds on the isolated eigenvalues
630: $E_{n\ell}$ of some Hamiltonian $H$ may be gained with the help of the
631: so-called ``local-energy'' theorem~\cite{LET,Thirring90}. Unfortunately, the
632: proof of this fundamental criterion makes use of the ``nodal theorem''~for
633: the eigenstates of $H.$ As a consequence of this, for Hamiltonians $H$
634: defined in more than~one dimension the local-energy theorem can be applied
635: only to the ground state of the spectrum. It has been applied to the
636: (three-dimensional) ``relativistic Coulomb problem'' in Ref.~\cite{Raynal94}.
637: 
638: Upper bounds on the energy levels of Hamiltonians bounded from below can be
639: obtained with considerably more efficiency by the variational technique
640: described in Sec.~\ref{Sec:VarUB}. Therefore we employ here the local-energy
641: theorem for the derivation of {\em lower\/} bounds on the spectrum of $H.$ In
642: order to formulate the local-energy theorem in momentum-space
643: representation,~we introduce the ``local energy''$${\cal E}({\bf p})\equiv
644: \beta\sqrt{m^2+{\bf p}^2}+\frac{\displaystyle\int{\rm d}^3q\,\widetilde
645: V({\bf p}-{\bf q})\phi({\bf q})} {\phi({\bf p})}\ .$$Here, $\widetilde V({\bf
646: p})$ is the Fourier transform of the interaction potential $V({\bf x})$ under
647: consideration:$$\widetilde V({\bf p})\equiv\frac{1}{(2\pi)^3}\int{\rm
648: d}^3x\exp(-{\rm i}{\bf p}\cdot{\bf x})V({\bf x})\ .$$$\phi({\bf p})$ denotes
649: a suitably chosen, positive trial function: $\phi({\bf p})>0.$ The
650: ``lower-bound part''~of the local-energy theorem then states that the
651: lowest-lying eigenvalue $E_0$ of the Hamiltonian $H$ is bounded from below by
652: the infimum of ${\cal E}({\bf p})$ with respect to the momentum
653: variable~${\bf p}$:\begin{equation}E_0\ge\inf_{{\bf p}}{\cal E}({\bf p})\
654: .\label{Eq:LET}\end{equation}
655: 
656: The application of the local-energy theorem to {\em confining\/} potentials
657: demands special~care for the following reason: the Fourier transform of any
658: confining potential is not well-defined; the potential has to be regarded as
659: a {\em distribution\/} and an appropriate regularization must~be applied. For
660: instance, the linear potential $V_{\rm L}(r)=\lambda r,$ where $\lambda$ is a
661: coupling parameter with mass dimension 2, may be regularized by the
662: exponential $\exp(-\varepsilon r)$, $\varepsilon\ge0,$ or by writing~it as
663: the derivative of this exponential or as the second derivative of a
664: Yukawa-type function:$$V_{\rm L}(r)=\lambda
665: r=\lambda\lim_{\varepsilon\downarrow0}r\exp(-\varepsilon
666: r)=-\lambda\lim_{\varepsilon\downarrow0}\frac{\partial}{\partial\varepsilon}
667: \exp(-\varepsilon r)=\lambda\lim_{\varepsilon\downarrow0}
668: \frac{\partial^2}{\partial\varepsilon^2}\frac{\exp(-\varepsilon r)}{r}\ .$$
669: 
670: \newpage\noindent With these regularizations, the Fourier transform
671: $\widetilde V_{\rm L}({\bf p})$ of the linear potential $V_{\rm L}(r)$
672: reads$$\widetilde V_{\rm L}({\bf p})=\widetilde V_{\rm L}(|{\bf
673: p}|)=-\frac{\lambda}{\pi^2}\lim_{\varepsilon\downarrow0}\frac{{\bf
674: p}^2-3\varepsilon^2}{({\bf p}^2+\varepsilon^2)^3}\ .$$Note that (as a
675: consistency check) the integral of the Fourier transform $\widetilde V_{\rm
676: L}({\bf p})$ has to vanish: $\int{\rm d}^3p\,\widetilde V_{\rm L}({\bf
677: p})=0.$ This holds for every power-law potential $V(r)=a_nr^n$ with
678: exponent~$n>0.$ The singularity structure of $\widetilde V_{\rm L}({\bf p})$
679: becomes manifest when writing the derivative in $V_{\rm L}(r)$~as a
680: difference quotient,$$V_{\rm L}(r)=\lambda r
681: =-\lambda\lim_{\varepsilon\downarrow0}\frac{\exp(-\varepsilon
682: r)-1}{\varepsilon}\ ,$$which yields$$\widetilde V_{\rm L}({\bf p})=\widetilde
683: V_{\rm L}(|{\bf p}|)=\lambda\lim_{\varepsilon\downarrow0}
684: \left[\frac{1}{\varepsilon}\delta^{(3)}({\bf p})-\frac{1}{\pi^2({\bf
685: p}^2+\varepsilon^2)^2}\right].$$Thus the Fourier transform $\widetilde V_{\rm
686: L}({\bf p})$ of the linear potential consists of a (negative) regular~part
687: and a singular part, which is a distribution localized at the origin ${\bf
688: p}={\bf 0}.$ However, a detailed inspection shows that for the case of the
689: funnel potential the momentum-space~ground-state eigenfunction is indeed
690: positive. This fact justifies our use of a positive trial function~$\phi({\bf
691: p}).$
692: 
693: The simplest normalizable trial function which comes to one's mind is the
694: exponential:$$\phi(r)=\sqrt{\frac{\mu^3}{\pi}}\exp(-\mu r)\ .$$This choice
695: for the ground-state trial function $\phi(r)$ corresponds to the case
696: $k=\ell=m=0$ and $\rho=1$ of the Laguerre basis functions $\psi_{k,\ell
697: m}({\bf x})$ introduced in Sec.~\ref{Sec:LagTS}. It is~straightforward to
698: calculate the Fourier transform of $\phi(r),$ or to extract it from the
699: general expression given explicitly in Sec.~\ref{Sec:LagTS}:$$\phi({\bf
700: p})=\phi(|{\bf p}|)=\frac{\sqrt{2\mu^3}}{\pi|{\bf p}|({\bf p}^2+\mu^2)}
701: \sin\left(2\arctan\frac{|{\bf p}|}{\mu}\right)=\frac{2\sqrt{2\mu^5}}{\pi({\bf
702: p}^2+\mu^2)^2}\ .$$In order to illustrate the application of the local-energy
703: theorem, let us consider the example of a potential $V(r)$ which is a sum of
704: pure power-law terms:$$V(r)=\sum_qa(q)r^q\ .$$For this power-law potential
705: and the ground-state trial function $\phi(r),$ it is straightforward~to
706: obtain the corresponding local energy ${\cal E}({\bf p})$ by Fourier
707: transformation of the product $r^q\phi(r)$: $${\cal E}({\bf
708: p})=\beta\sqrt{m^2+{\bf p}^2}+
709: \sum_q\frac{a(q)\Gamma(q+2)}{2\mu|{\bf p}|({\bf p}^2+\mu^2)^{q/2-1}}
710: \sin\left[(q+2)\arctan\frac{|{\bf p}|}{\mu}\right].$$The variational
711: parameter $\mu$ entering into our ground-state trial function $\phi(r)$
712: either may~be given a chosen fixed value, or it may be adjusted in order to
713: maximize the lower bound~(\ref{Eq:LET}).
714: 
715: \section{Sums of distinct potential terms}\label{Sec:SLB}For the class of
716: potentials $V(r)$ which consist of two or more distinct terms $V^{(i)}(r)\ne
717: V^{(j)}(r)$ for $i\ne j,$ where, for every single component problem defined
718: by the ``one-term'' Hamiltonian $H_i\equiv\beta\sqrt{m^2+{\bf
719: p}^2}+V^{(i)}(r),$ information about the bottom of its spectrum is available,
720: these individual pieces of information may be combined to form a lower bound
721: on the spectrum~of the Hamiltonian (\ref{Eq:(sr)SSH}) with such a sum
722: potential $V(r).$ The precise requirements for this ``sum approximation'' to
723: work are the following:\begin{enumerate}\item The interaction potential
724: $V(r)$ entering in the semirelativistic Hamiltonian of Eq.~(\ref{Eq:(sr)SSH})
725: is a sum of (more than one) different components $V^{(i)}(r)=c_ih^{(i)}(r)$:
726: $$V(r)=\sum_ic_ih^{(i)}(r)\ .$$\item Every component problem
727: $H_i\equiv\beta\sqrt{m^2+{\bf p}^2}+c_ih^{(i)}(r)$ supports, for
728: sufficiently~large values of its coupling parameter $c_i,$ a discrete
729: eigenvalue at the bottom of its spectrum.\end{enumerate}The sum approximation
730: has already been studied for the case of Schr\"odinger Hamiltonians
731: (involving nonrelativistic kinetic energies) in Refs.~\cite{Hall83,NRSA}. The
732: analysis of the corresponding semirelativistic problem has been presented in
733: full generality in Ref.~\cite{Lucha02-Sum}. Therefore, here~we summarize only
734: the result for the case in which the potential $V(r)$ is a sum of
735: attractive~pure power-law terms:\begin{equation}V(r)=\sum_{q\ne0}a(q)\,{\rm
736: sgn}(q)r^q\ ,\label{Eq:SPPP}\end{equation}where the coupling parameters
737: $a(q)$ are non-negative and not all zero. For this potential, a lower bound
738: to the bottom of the spectrum of the semirelativistic Hamiltonian
739: (\ref{Eq:(sr)SSH}), which~is, by assumption, the lowest eigenvalue $E_0$ of
740: $H,$ is given by the expression
741: \begin{equation}E_0\ge\min_{r>0}\left[\beta\sqrt{m^2+\frac{1}{r^2}}+
742: \sum_{q\ne0}a(q)\,{\rm sgn}(q)(P(q)r)^q\right],\label{Eq:SLB}\end{equation}
743: {\em if}, for the particular potential $V(r)$ under consideration, there
744: exist (or there can be found) $P$ numbers $P(q)$ such that, whenever there is
745: only one term present in the potential (\ref{Eq:SPPP}), the right-hand side
746: of the inequality (\ref{Eq:SLB}) yields either the exact ground-state energy
747: eigenvalue of the corresponding ``single-term'' problem or a lower bound to
748: it.
749: 
750: Consequently, when applying our general sum lower bound (\ref{Eq:SLB}) to a
751: given sum potential $V(r),$ the main task is to derive appropriate $P$
752: numbers for all components in the sum~(\ref{Eq:SPPP}). Let us illustrate this
753: by considering the two examples relevant for the funnel potential
754: (\ref{Eq:FP}).\begin{itemize}\item For a Coulomb component potential
755: $h^{(i)}(r)=-1/r,$ corresponding to the case $q=-1$ in Eq.~(\ref{Eq:SPPP}),
756: we may take advantage from the explicit results presented in
757: Sec.~\ref{Sec:CLB}. By~the change of variables $r\to Pr,$ the Coulomb lower
758: bound (\ref{Eq:SSE-LB}) may be cast into the form$$E_0\ge\min_{r>0}
759: \left[\beta\sqrt{m^2+\frac{1}{r^2}}+V(Pr)\right],\quad v<\beta v_P\
760: .$$Specifying this general result to the Coulomb potential $V(r)=-v/r$ yields
761: the bound
762: $$E_0\ge\min_{r>0}\left[\beta\sqrt{m^2+\frac{1}{r^2}}-\frac{v}{Pr}\right],\quad
763: v<\beta v_P\ .$$This bound is, of course, nothing else but, for $P=2/\pi,$
764: the Herbst lower bound~(\ref{Eq:RCP-HLB})~or, for arbitrary $P,$ the
765: Herbst-like lower bound (\ref{Eq:RCP-NLB}) to the relativistic Coulomb
766: problem. Consequently, for the Coulomb component term we may identify the $P$
767: number $P(-1)$ required for the lower bound (\ref{Eq:SLB}) with the
768: lower-bound parameter $P$ defined in Sec.~\ref{Sec:CLB}: $P(-1)=P.$\item For
769: a linear component potential $h^{(i)}(r)=r,$ that is, for the case $q=1$ in
770: Eq.~(\ref{Eq:SPPP}),~(a possible value of) the $P$ number $P(1)$ required for
771: the lower bound (\ref{Eq:SLB}) turns out to~be related to the ground-state
772: eigenvalue $e_0(v)$ of the Hamiltonian operator $\sqrt{{\bf
773: p}^2}+vr$~by$$e_0(v)=2\sqrt{vP(1)}\ ;$$a numerical determination of this
774: lowest energy eigenvalue $e_0(v)$ yields $P(1)=1.2457.$\end{itemize}
775: 
776: \section{Comparison: The case of the funnel potential}\label{Sec:Comp}In
777: order to demonstrate the power of the different energy bounds derived so far,
778: let us apply the above results to the very illustrative example of the
779: Coulomb-plus-linear potential~(\ref{Eq:FP}). By factorizing off an overall
780: coupling constant $0<v\le1,$ we write this potential in the~form
781: \begin{equation}V(r)=v\left(-\frac{a}{r}+br\right).\label{Eq:CLP}\end{equation}
782: For the coupling parameters entering in this funnel potential $V(r),$
783: different notations have been used in Eqs.~(\ref{Eq:FP}), (\ref{Eq:SPPP}),
784: and (\ref{Eq:CLP}). Evidently, the three sets of potential coefficients~have
785: to be identified according to $a(-1)\equiv c_1\equiv av>0$ and $a(1)\equiv
786: c_2\equiv bv>0.$
787: 
788: Fig.~\ref{Fig:SSE-bounds} compares the various energy bounds derived or
789: investigated in Secs.~\ref{Sec:CLB} through~\ref{Sec:SLB} for the
790: ground-state energy eigenvalue $E_0$ of the Hamiltonian (\ref{Eq:(sr)SSH})
791: with the funnel potential (\ref{Eq:CLP}) as a function of the overall
792: coupling parameter $v.$ These energy bounds necessarily~form a bunch of
793: concave curves $E(v),$ all starting at the (free-energy) value
794: $E_0(v=0)=\beta m=1.$
795: 
796: \begin{figure}[ht]\vspace*{-2cm}\begin{center}
797: \psfig{figure=dma2fig1.ps,scale=.6243}\vspace*{-1cm}\caption{Comparison of
798: bounds to the ground-state [$(n,\ell)=(1,0)$] energy eigenvalue of the
799: semirelativistic Hamiltonian $H=\beta\sqrt{m^2+{\bf p}^2}+V(r)$ with a
800: Coulomb-plus-linear~potential $V(r)=v(-a/r+br),$ for $a=0.2,$ $b=0.5,$
801: $m=\beta=1.$ The bounds are plotted as functions $E(v)$ of the overall
802: coupling parameter $v$ in $V(r).$ The energy bounds compared~here include
803: (from top to bottom) the harmonic-oscillator upper bound (H) of
804: Eq.~(\ref{Eq:SSE-HOUB}), the linear upper bound (L) of Eq.~(\ref{Eq:LUB}),
805: the ``one-dimensional'' variational upper bound (V) of Eq.~(\ref{Eq:VUB})
806: with $n=1.74,$ the variational upper bound (E) based on a 25-dimensional
807: Laguerre trial space, which should come rather close to the exact eigenvalue,
808: the sum lower bound (S) of Eq.~(\ref{Eq:SLB}) with $P$ numbers
809: $P(-1)=0.72811$ and $P(1)=1.2457,$ the lower bound (T) provided by~the
810: local-energy theorem (\ref{Eq:LET}), and two Coulomb lower bounds (C)
811: according to Eq.~(\ref{Eq:SSE-LB}),~where with respect to the parameter $P$
812: the lower curve corresponds to the fixed value $P=0.72811$ whereas the upper
813: curve corresponds to the optimized value of $P$ calculated from
814: Eq.~(\ref{Eq:P(m)}).}\label{Fig:SSE-bounds}\end{center}\end{figure}
815: 
816: The Coulomb lower bound (\ref{Eq:SSE-LB}) may be applied to the funnel
817: potential either with~a~fixed value of the lower-bound parameter $P$ or with
818: the optimized $P(v)$ computed from Eq.~(\ref{Eq:P(m)}). For ``fixed-$P$''
819: bounds, the maximum value of $P$ allowed by the Coulomb coupling constant
820: constraint discussed in Sec.~\ref{Sec:CLB} is found, for given values of the
821: kinetic-term parameters~$\beta,$~$m$ and potential coefficients $c_1,$ $c_2,$
822: as solution of that relation which is obtained when equating the left-hand
823: and right-hand sides of the inequality (\ref{Eq:CCCC}) with the critical
824: coupling $v_P$ given by the definition (\ref{Eq:BV}), that is, by solving,
825: for the {\em maximum\/} values of $c_1$ and $c_2$ occuring~here, the
826: constraint~(\ref{Eq:P(m)}):$$c_1+\frac{P^4}{1-P^2}\frac{c_2}{m^2}=\beta
827: P\sqrt{1-P^2}\ .$$For a single, unit-mass particle, i.e., $\beta=1$ and
828: $m=1,$ and the values of the funnel-potential coupling parameters $a$ and $b$
829: used in Fig.~\ref{Fig:SSE-bounds}, $a=0.2$ and $b=0.5,$ this yields
830: $P=0.728112.$
831: 
832: We try to approach the state we are interested in as closely as possible by
833: the application of the Rayleigh--Ritz variational technique brief\/ly
834: sketched in Sec.~\ref{Sec:VarUB}, with the trial space~$D_d$ spanned by the
835: Laguerre basis defined in Sec.~\ref{Sec:LagTS}. By fixing the variational
836: parameter $\mu$ to~the value $\mu=m,$ we are able to obtain the matrix
837: elements of the Hamiltonian (\ref{Eq:(sr)SSH}) analytically, avoiding thereby
838: the necessity to use a numerical integration procedure for~the evaluation~of
839: these matrix elements; for definiteness, we fix the variational parameter
840: $\rho$ to the value~$\rho=1.$
841: 
842: The accuracy of such an approximation to an eigenstate may be estimated with
843: the~help of the ``virial-theorem-inspired'' quality measure $Q$ discussed in
844: Sec.~\ref{Sec:VarUB}. Fig.~\ref{Fig:accuracy} shows, for~the trial-space
845: dimension $d=25$ chosen for the variational bound (E) in
846: Fig.~\ref{Fig:SSE-bounds}, the dependence of the accuracy $Q$ on the overall
847: coupling $v.$ We obtain $Q=|Q|\le1.1\times 10^{-4}$ over the~whole parameter
848: range $0<v\le1.$ Such high accuracy should be sufficient for the present
849: purpose.
850: 
851: \begin{figure}[ht]\vspace*{-0.395cm}\begin{center}
852: \psfig{figure=accuracy1.eps,scale=1.1416}\caption{Measure $Q$ of the accuracy
853: of the (variational) approximation of the ground-state eigenfunction~of the
854: semirelativistic Hamiltonian of Fig.~\ref{Fig:SSE-bounds} [that is,
855: $H=\beta\sqrt{m^2+{\bf p}^2}+V(r),$ where $V(r)$ is the Coulomb-plus-linear
856: potential $V(r)=v(-a/r+br)$ with $a=0.2,$ $b=0.5,$ $m=\beta=1$] by a
857: superposition of 25 Laguerre basis functions (\ref{eq:LagTF}) with $\mu=m$
858: and $\rho=1.$}\label{Fig:accuracy}\end{center}\end{figure}
859: 
860: For the sake of completeness, we plot, in Fig.~\ref{Fig:Laguerre-approx}, the
861: behaviour of the ground state of~the Hamiltonian (\ref{Eq:(sr)SSH}) with the
862: Coulomb-plus-linear potential (\ref{Eq:CLP}), for the maximum value $v=1$ of
863: the overall coupling parameter $v,$ determined by the variational technique
864: of Sec.~\ref{Sec:VarUB}, with the Laguerre basis of Sec.~\ref{Sec:LagTS}, in
865: both configuration and momentum space. As expected, the ground-state
866: momentum-space eigenfunction possesses no node, which~is in~agreement with
867: our assumption of a positive momentum-space trial function $\phi({\bf p})$
868: (which is also depicted,~as the case $d=1,$ in
869: Fig.~\ref{Fig:Laguerre-approx}) in the derivation of the local-energy lower
870: energy bound in Sec.~\ref{Sec:LET}.
871: 
872: \begin{figure}[ht]\begin{center}\psfig{figure=Lagapp-r.ps,scale=1.1658}
873: \\[1ex](a)\\[1ex]\psfig{figure=Lagapp-p.ps,scale=1.1658}\\[1ex](b)
874: \caption{Variational approximation of the lowest-lying eigenstate in
875: configuration space~(a) and momentum space (b) of the semirelativistic
876: Hamiltonian (\ref{Eq:(sr)SSH}), $H=\beta\sqrt{m^2+{\bf p}^2}+V(r),$ with a
877: Coulomb-plus-linear potential $V(r)=-a/r+br,$ for $a=0.2,$ $b=0.5$ and
878: $m=\beta=1,$ by a superposition of Laguerre basis functions (\ref{eq:LagTF})
879: with the variational parameters fixed~to $\mu=m$ and $\rho=1,$ for the
880: trial-space dimensions $d=25$ (full lines) and $d=1$ (dashed~lines).}
881: \label{Fig:Laguerre-approx}\end{center}\end{figure}
882: 
883: \section{Summary and conclusions}This review of the discrete spectra of
884: semirelativistic Hamiltonians $H$ of the form (\ref{Eq:(sr)SSH})~should be
885: considered as a compilation of rigorous (semi-)analytical results for both
886: upper and~lower bounds on the discrete eigenvalues $E_{n\ell}$ of $H,$
887: presented in a way suitable for immediate~use: the algebraic structure and
888: the convexity properties with respect to suitable basis potentials $h(r)$~of
889: the interaction potential $V(r)$ one wishes to use in one's semirelativistic
890: Hamiltonian (\ref{Eq:(sr)SSH}) unambiguously determine which of all the
891: energy bounds discussed here actually apply. With the explicit analytical
892: expressions for these bounds on the discrete eigenvalues at~one's disposal,
893: one is left with the (numerical) optimization with respect to a single real
894: variable. Clearly, within the respective limitations, the Rayleigh--Ritz
895: variational technique of Secs.~\ref{Sec:VarUB} and \ref{Sec:LagTS} as well as
896: the local-energy theorem of Sec.~\ref{Sec:LET} apply to {\em arbitrary\/}
897: reasonable potentials.
898: 
899: For the particular example considered in Sec.~\ref{Sec:Comp}, viz., the
900: Coulomb-plus-linear potential (\ref{Eq:CLP}), the sum lower bound turns out
901: to be somewhat better than the lower bound provided by the local-energy
902: theorem. Clearly, the sum lower bound cannot always be better than the
903: local-energy bound because one could, in principle, choose in the
904: local-energy theorem~(\ref{Eq:LET}) the exact eigenfunction for the trial
905: function $\phi$ and thus recover the exact energy eigenvalue.
906: 
907: In conclusion, let us stress that all but two of the various energy bounds we
908: have studied in this work may eventually be expressed by essentially the same
909: general semi-classical~form. We can view this in the following way: Since the
910: square of the momentum~${\bf p}^2$ scales like~$1/r^2,$ we expect
911: that$$\langle{\bf p}^2\rangle=\frac{P^2}{r^2}\ ,$$where $P$ is a suitable
912: coefficient; thus the balance between the kinetic and potential energies
913: required by the minimum--maximum principle might lead to a search for the
914: minimum of~an energy expression of the form$$\sqrt{m^2+\frac{P^2}{r^2}}+V(r)\
915: .$$What makes this heuristic argument interesting is the fact that definite
916: values of $P$ can be prescribed so that this minimum provides {\it a priori}
917: a bound on the (unknown) exact energy. In the case where $V(r)$ is a sum of
918: terms, a trivial change of variables is used to move~the~$P$ parameters
919: inside the potential $V(r),$ and then optimal distinct parameters~can be
920: used~for each potential term.
921: 
922: The lower bound by the sum approximation incorporates our best known lower
923: bounds for the sub-problems containing only a single potential term; the
924: final result is the optimized mixture of these contributions. The lower bound
925: by the local-energy theorem (\ref{Eq:LET}) could, in principle, achieve
926: arbitrarily high accuracy. The difficulty here, as with variational methods,
927: is to find a suitable trial function. To the extent that we have searched we
928: did not manage to find a better lower bound with the help of the local-energy
929: theorem (\ref{Eq:LET}) than that~provided immediately by the sum
930: approximation (\ref{Eq:SLB}).
931: 
932: Our most accurate results overall were upper bounds obtained by using a
933: 25-dimensional trial space spanned by Laguerre functions. We estimate the
934: error for these results to be less than $0.01\%$ over the total range of
935: potential parameters studied; such accurate estimates~are invaluable for
936: estimating the effectiveness of the large variety of alternative simpler
937: energy bounds which we have reviewed in this article.
938: 
939: \newpage
940: 
941: \begin{thebibliography}{30}
942: \bibitem{Hall83}R.~L.~Hall, J.~Math.~Phys.\ {\bf 24}, 324 (1983).
943: \bibitem{Hall84}R.~L.~Hall, J.~Math.~Phys.\ {\bf 25}, 2708 (1984).
944: \bibitem{Lucha00-HO}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, J.~Phys.~A
945: {\bf 34}, 5059 (2001), hep-th/0012127.
946: \bibitem{Lucha01-DMAI}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl,
947: Int.~J.~Mod.~Phys.~A {\bf 17}, 1931 (2002), hep-th/0110220.
948: \bibitem{Lucha96a}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~A {\bf 54}, 3790
949: (1996), hep-ph/9603429.
950: \bibitem{Lucha98O}W.~Lucha and F.~F.~Sch\"oberl, Int.~J.~Mod.~Phys.~A {\bf
951: 14}, 2309 (1999), hep-ph/9812368.
952: \bibitem{Lucha98D}W.~Lucha and F.~F.~Sch\"oberl, Fizika B {\bf 8}, 193
953: (1999), hep-ph/9812526.
954: \bibitem{Lucha99Q}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~A {\bf 60}, 5091
955: (1999), hep-ph/9904391.
956: \bibitem{Lucha99A}W.~Lucha and F.~F.~Sch\"oberl, Int.~J.~Mod.~Phys.~A {\bf
957: 15}, 3221 (2000), hep-ph/9909451.
958: \bibitem{Reed78}M.~Reed and B.~Simon, {\em Methods of Modern Mathematical
959: Physics~IV: Analysis~of Operators\/} (Academic Press, New York, 1978).
960: \bibitem{Thirring90}W.~Thirring, {\em A Course in Mathematical Physics 3:
961: Quantum Mechanics of Atoms and Molecules\/} (Springer, New York/Wien, 1990).
962: \bibitem{Lucha00-1DRCP}W.~Lucha and F.~F.~Sch\"oberl, J.~Math.~Phys.\ {\bf
963: 41}, 1778 (2000), hep-ph/9905556.
964: \bibitem{Martin88}A.~Martin, Phys.~Lett.~B {\bf 214}, 561 (1988).
965: \bibitem{Gelfand}I.~M.~Gelfand and S.~V.~Fomin, {\em Calculus of
966: Variations\/} (Prentice-Hall, Englewood Cliffs, 1963).
967: \bibitem{Abramowitz}{\em Handbook of Mathematical Functions}, edited by
968: M.~Abramowitz and I.~A.~Stegun (Dover, New York, 1964).
969: \bibitem{Herbst77}I.~W.~Herbst, Commun.~Math.~Phys. {\bf 53}, 285 (1977);
970: {\bf 55}, 316 (1977) (addendum).
971: \bibitem{Martin89}A.~Martin and S.~M.~Roy, Phys.~Lett.~B {\bf 233}, 407
972: (1989).
973: \bibitem{Lucha89:RVT}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~Lett.\ {\bf
974: 64}, 2733 (1990).
975: \bibitem{Lucha90:RVTs}W.~Lucha and F.~F.~Sch\"oberl, Mod.~Phys.~Lett.~A {\bf
976: 5}, 2473 (1990).
977: \bibitem{Jensen}W. Feller, {\em An introduction to probability theory and its
978: applications, Volume II\/} (John Wiley, New York, 1971).
979: \bibitem{LTS}S.~Jacobs, M.~G.~Olsson, and C.~Suchyta III, Phys.~Rev.~D {\bf
980: 33}, 3338 (1986); {\bf 34}, 3536 (1986) (E).
981: \bibitem{Lucha97}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~A {\bf 56}, 139
982: (1997), hep-ph/9609322.
983: \bibitem{LET}R.~J.~Duffin, Phys.~Rev.\ {\bf 71}, 827 (1947);\\M.~F.~Barnsley,
984: J.~Phys.~A {\bf 11}, 55 (1978);\\B.~Baumgartner, J.~Phys.~A {\bf 12}, 459
985: (1979).
986: \bibitem{Raynal94}J.~C.~Raynal, S.~M.~Roy, V.~Singh, A.~Martin, and
987: J.~Stubbe, Phys.~Lett.~B {\bf 320}, 105 (1994).
988: \bibitem{NRSA}R.~L.~Hall, Phys.~Rev.~D {\bf 37}, 540 (1988); J.~Math.~Phys.\
989: {\bf 33}, 1710 (1992).
990: \bibitem{Lucha02-Sum}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl,
991: math-ph/0208042, J.~Math.~Phys.\ (in print).
992: \end{thebibliography}\end{document}
993: