1: \documentstyle[12pt,epsfig,graphics]{article}\pagestyle{myheadings}
2: \textheight=23.5cm\topmargin=-1.2cm\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: HEPHY-PUB 740/01\\UWThPh-2001-27\\CUQM-86\\hep-th/0110220\\July
9: 2001\end{tabular}\\[.3cm]\Large\begin{center}{\bf DISCRETE SPECTRA OF
10: SEMIRELATIVISTIC HAMILTONIANS FROM ENVELOPE THEORY}\\\vspace{0.5cm}\large{\bf
11: Richard L. HALL\footnote[3]{\normalsize\ {\em E-mail address\/}:
12: rhall@mathstat.concordia.ca}}\\[.1cm]\normalsize Department of Mathematics
13: and Statistics, Concordia University,\\1455 de Maisonneuve Boulevard West,
14: Montr\'eal, Qu\'ebec, Canada H3G 1M8\\[0.4cm]\large{\bf Wolfgang
15: LUCHA\footnote[1]{\normalsize\ {\em E-mail address\/}:
16: wolfgang.lucha@oeaw.ac.at}}\\[.1cm]\normalsize Institut f\"ur
17: Hochenergiephysik, \"Osterreichische Akademie der
18: Wissenschaften,\\Nikolsdorfergasse 18, A-1050 Wien,
19: Austria\\[0.4cm]\large{\bf Franz F.~SCH\"OBERL\footnote[2]{\normalsize\ {\em
20: E-mail address\/}: franz.schoeberl@univie.ac.at}}\\[.1cm]\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: analyze the (discrete) spectrum of the semirelativistic ``spinless-Salpeter''
24: Hamiltonian$$H=\beta\sqrt{m^2+{\bf p}^2}+V(r)\ ,\quad\beta>0\ ,$$where $V(r)$
25: is an attractive, spherically symmetric potential in three dimensions. In
26: order~to locate the eigenvalues of $H,$ we extend the ``envelope theory,''
27: originally formulated only~for nonrelativistic Schr\"odinger operators, to
28: the case of Hamiltonians involving the relativistic kinetic-energy operator.
29: If $V(r)$ is a convex transformation of the Coulomb potential $-1/r$ and a
30: concave transformation of the harmonic-oscillator potential $r^2$, both upper
31: and lower bounds on the discrete eigenvalues of $H$ can be constructed, which
32: may all be expressed~in the
33: form$$E=\min_{r>0}\left[\beta\sqrt{m^2+\frac{P^2}{r^2}}+V(r)\right]$$for
34: suitable values of the numbers $P$ here provided. At the critical point, the
35: relative growth to the Coulomb potential $h(r)=-1/r$ must be bounded by ${\rm
36: d}V/{\rm d}h<2\beta/\pi.$\\[3ex]{\em PACS numbers\/}: 03.65.Ge, 03.65.Pm,
37: 11.10.St\renewcommand{\thefootnote}{\arabic{footnote}}\end{titlepage}
38:
39: \normalsize
40:
41: \section{Introduction}We study the semirelativistic (so-called
42: ``spinless-Salpeter'') Hamiltonian\begin{equation}H=\beta\sqrt{m^2+{\bf
43: p}^2}+V(r)\ ,\quad\beta>0\ ,\label{Eq:SH}\end{equation}in which $V(r)$ is a
44: central potential in three dimensions. The eigenvalue equation~of this
45: operator is called the ``spinless Salpeter equation.'' This equation of
46: motion arises as~a well-defined standard approximation to the Bethe--Salpeter
47: formalism \cite{Salpeter51} for the description of bound states within a
48: (relativistic) quantum field theory and is arrived at by the following
49: simplifying steps:\begin{enumerate}\item Eliminate all timelike variables by
50: assuming the Bethe--Salpeter kernel that describes the interactions between
51: the bound-state constituents to be static, i.e., instantaneous; the result of
52: this reduction step is called the ``instantaneous Bethe--Salpeter equation''
53: or the ``Salpeter equation'' \cite{Salpeter52}.\item Neglect the spin of the
54: bound-state constituents, assume the Bethe--Salpeter kernel~to be of
55: convolution type (as is frequently the case), and consider merely
56: positive-energy solutions $\psi,$ in order to arrive at the so-called
57: ``spinless Salpeter equation'' $H\psi=E\psi,$ with a Hamiltonian $H$ of the
58: form (\ref{Eq:SH}). (For two particles, this form of the Hamiltonian $H$
59: holds only for equal masses $m$ of the bound-state
60: constituents.)\end{enumerate}(For a more detailed account of the reduction of
61: the Bethe--Salpeter equation to the spinless Salpeter equation, consult,
62: e.g., the introductory sections of Refs.~\cite{Lucha98O,Lucha98D}.) This wave
63: equation describes the bound states of spin-zero particles (scalar bosons) as
64: well as the spin-averaged spectra of the bound states of fermions.
65:
66: In this paper we consider potentials which are at the same time convex
67: transformations $V(r)=g(h(r))$ of the Coulomb potential $h(r)=-1/r$ and
68: concave transformations of the harmonic-oscillator potential $h(r)=r^2.$ The
69: reason for this is that spectral information is known for these two ``basis''
70: potentials $h(r)$. Thus the class of potentials is those $V(r)$ that have a
71: dual representation$$V(r)=g^{(1)}\left(-\frac{1}{r}\right)=g^{(2)}(r^{2})\
72: ,$$in which $g^{(1)}$ is convex (${g^{(1)}}''>0$) and $g^{(2)}$ is concave
73: (${g^{(2)}}''<0$). An example of a potential in this class
74: is\begin{equation}V(r)=-\frac{c_1}{r}+c_2\ln r+c_3r+c_4r^2\
75: ,\label{Eq:pot-class}\end{equation}where the coefficients $\{c_{i}\}$ are not
76: negative and are not all zero. Thus tangent lines to the transformation
77: function $g(h)$ are of the form $ah+b$ and are either Coulomb
78: potentials~lying below $V$, or harmonic-oscillator potentials lying above
79: $V.$ This geometrical idea is the basis for our approach to the spectral
80: problem posed by $H.$ We shall consider applications of this idea to the
81: (nonrelativistic) Schr\"odinger problem, the relativistic kinetic-energy
82: operator, and the full Salpeter Hamiltonian in Secs.~\ref{Sec:NRET},
83: \ref{Sec:RKE}, and \ref{Sec:RET}, respectively. We shall show that all our
84: upper and lower bounds on the eigenvalues of the semirelativistic Salpeter
85: Hamiltonian $H$ of Eq.~(\ref{Eq:SH}) can be expressed in the compact form
86: $$E\approx\min_{r>0}\left[\beta\sqrt{m^2+\frac{P^2}{r^2}}+V(r)\right],$$where
87: $P$ is a constant for each bound, and a sign of approximate equality is used
88: to indicate that, for definite convexity of $g(h),$ the envelope theory
89: yields lower bounds for convex $g(h)$ and upper bounds for concave $g(h).$
90: The main purpose of the present considerations is to establish the general
91: envelope formalism in terms of which such bounds can be proved, and to
92: determine the appropriate values of $P.$
93:
94: It is fundamental to our method that we first know something about the
95: spectrum~of~$H$ in those cases where $V(r)$ is one of the basis potentials,
96: i.e., the Coulomb and the~harmonic oscillator. These two spectra are
97: discussed in Sec.~\ref{Sec:CHOP} below. In Sec.~\ref{Sec:FP} we look at the
98: example of the Coulomb-plus-linear potential.
99:
100: \section{The Coulomb and harmonic-oscillator potentials}\label{Sec:CHOP}
101: \subsection{Scaling behaviour}Since the two basis potentials are both pure
102: powers, it is helpful first to determine what~can be learnt about the
103: corresponding eigenvalues by the use of standard scaling arguments. By
104: employing a wave function $\phi(cr)$ depending on a scale variable $c>0,$ we
105: find the following scaling rule for the eigenvalues corresponding to
106: attractive pure power potentials~$v\,{\rm sgn}(q)r^q.$ The
107: Hamiltonian$$H=\beta\sqrt{m^2+{\bf p}^2}+v\,{\rm sgn}(q)r^q$$has the (energy)
108: eigenvalues $E(v,\beta,m),$ where$$E(v,\beta,m)=\beta
109: mE\!\left(\frac{v}{\beta m^{1+q}},1,1\right),\quad q\ge -1\ .$$The scaling
110: behaviour described by the above formula allows us to consider the
111: one-particle, unit-mass special case $m=\beta=1$ initially, that is to say,
112: to work w.l.o.g.\ with the operator$$H=\sqrt{1+{\bf p}^2}+v\,{\rm
113: sgn}(q)r^{q}\ .$$
114:
115: \subsection{Coulomb potential}\label{Subsec:CP}In the case of the Coulomb
116: potential $V(r)=-v/r$ it is well known \cite{Herbst77} that the Hamiltonian
117: $H$ has a Friedrichs extension provided the coupling constant $v$ is not too
118: large. Specifically, it is necessary in this case that $v$ is smaller than a
119: critical value $v_{\rm c}$ of the coupling constant:$$v<v_{\rm
120: c}=\frac{2}{\pi}\ .$$With this restriction, a lower bound to the bottom of
121: the spectrum is provided by Herbst's formula\begin{equation}
122: E_0\ge\sqrt{1-(\sigma v)^{2}}\ ,\quad\sigma\equiv\frac{\pi}{2}\
123: .\label{Eq:Herbst-bound}\end{equation}By comparing the spinless Salpeter
124: problem to the corresponding Klein--Gordon equation, Martin and Roy
125: \cite{Martin89} have shown that if the coupling constant is further
126: restricted by $v<\frac{1}{2},$ then an improved lower bound is provided by
127: the expression\begin{equation}E_0\ge\sqrt{\frac{1+\sqrt{1-4v^{2}}}{2}}\
128: ,\quad v<\frac{1}{2}\ .\label{Eq:MR-bound}\end{equation}It turns out that our
129: lower-bound theory has a simpler form when the Coulomb eigenvalue bound has
130: the form of Eq.~(\ref{Eq:Herbst-bound}) rather than that of
131: Eq.~(\ref{Eq:MR-bound}). For this reason, we have derived from
132: Eq.~(\ref{Eq:MR-bound}), by rather elementary methods, a new family of
133: Coulomb bounds. To this~end, we begin with the ansatz
134: $$\sqrt{\frac{1+\sqrt{1-4v^{2}}}{2}}\ge\sqrt{1-(\sigma v)^{2}}$$and look for
135: conditions under which it becomes true. Since both sides are positive, we may
136: square the ansatz and rearrange to yield$$v^2\le\frac{\sigma^2-1}{\sigma^4}\
137: .$$Meanwhile from (\ref{Eq:MR-bound}) we must always satisfy $v<\frac{1}{2}.$
138: This establishes the inequality we'll~need,
139: namely,\begin{equation}E_0\ge\sqrt{1-(\sigma v)^2}\ ,\quad
140: v\le\frac{\sqrt{\sigma^2-1}}{\sigma^2}<\frac{1}{2}\
141: .\label{Eq:NLB}\end{equation}Examples are\begin{eqnarray*}&&\sigma^2=2\
142: ,\quad v\le\frac{1}{2}\ ;\\[1ex]
143: &&\sigma^2=\frac{3}{2}\left(3-\sqrt{5}\right)\approx 1.145898\ ,\quad
144: v\le\frac{1}{3}\ ;\\[1ex]&&\sigma^2=8-4\sqrt{3}\approx 1.071797\ ,\quad
145: v\le\frac{1}{4}\ .\end{eqnarray*}All these (lower) bounds are slightly weaker
146: than the Martin--Roy bound (\ref{Eq:MR-bound}) but above the Herbst bound
147: (\ref{Eq:Herbst-bound}). We note that these functions of the coupling
148: constant $v$ are all monotone and {\it concave\/}.
149:
150: \subsection{Harmonic-oscillator potential}\label{Subsec:HOP}In the case of
151: the harmonic-oscillator potential, i.e., $V(r)=vr^2,$ much more is
152: known~\cite{Lucha99Q,Lucha99A}. In momentum-space representation the operator
153: ${\bf p}$ becomes a $c$-variable and thus, from the spectral point of view,
154: the Hamiltonian$$H=\sqrt{1+{\bf p}^2}+vr^2$$is equivalent to the
155: Schr\"odinger operator\begin{equation}H=-v\Delta+\sqrt{1+r^2}\
156: .\label{Eq:SHam-HO}\end{equation}Since the potential in this operator
157: increases without bound, we know \cite{Reed78} that the spectrum of this
158: operator is entirely discrete. We call its eigenvalues ${\cal E}_{n\ell}(v),$
159: $n=1,2,\dots,$ $\ell=0,1,\dots,$ where $n$ counts the radial states in each
160: angular-momentum subspace labelled by $\ell.$ In what follows we shall either
161: approximate the eigenvalues ${\cal E}_{n\ell}(v)$ analytically or presume
162: that they are known numerically. The concavity of nonrelativistic
163: Schr\"odinger energy eigenvalues has been discussed in
164: Refs.~\cite{Narnhofer75,Thirring90}. Theorem~2~of Ref.~\cite{Hall83}
165: establishes concavity for the ground state; the same proof can be applied to
166: states which are (1) in the subspace corresponding~to angular momentum $\ell$
167: and (2) orthogonal to the first $n-1$ exact energy eigenstates in this
168: subspace. This establishes concavity also for all the higher Schr\"odinger
169: energy eigenvalues. Thus the eigenvalues ${\cal E}_{n\ell}(v),$ regarded as
170: functions of the coupling parameter $v,$ are {\it concave\/}.
171:
172: \subsection{The spectral comparison theorem}\label{Subsec:SCT}For the class
173: of interaction potentials given by (\ref{Eq:pot-class}) with the coefficient
174: of the Coulombic~term satisfying the constraint$$\lim_{r\to
175: 0}r^2V'(r)<\frac{2\beta}{\pi}\ ,$$the semirelativistic Salpeter Hamiltonian
176: $H$ is bounded below and is essentially self-adjoint \cite{Herbst77}.
177: Consequently, the discrete spectrum of $H$ is characterized variationally
178: \cite{Reed78} and it follows immediately from this that, if we compare two
179: such Hamiltonians $H$ having the potentials $V^{(1)}(r)$ and $V^{(2)}(r),$
180: respectively, and we know that $V^{(1)}(r)<V^{(2)}(r),$ then we may conclude
181: that the corresponding discrete eigenvalues $E_{n\ell}$ satisfy the
182: inequalities $E_{n\ell}^{(1)}<E_{n\ell}^{(2)}.$ We shall refer to this
183: fundamental result as the ``spectral comparison theorem.'' In the more~common
184: case of nonrelativistic dynamics, i.e., for a (nonrelativistic) kinetic term
185: of the form $\beta{\bf p}^2/2m$ in the Hamiltonian $H$, a constraint similar
186: to the above would hold for the coefficient of a possible additional
187: (attractive) $-1/r^2$ term in the potential $V(r).$
188:
189: \section{Envelope representations for Schr\"odinger
190: operators}\label{Sec:NRET}We distinguish a potential $V(r)=vf(r)$ from its
191: shape $f(r),$ where the positive parameter~$v$ is often called the ``coupling
192: constant.'' The idea behind envelope representations \cite{Hall83,Hall84} is
193: suggested by the question: if one potential $f(r)$ can be written as a smooth
194: transformation $f(r)=g(h(r))$ of another potential $h(r),$ what spectral
195: relationship might this induce?~We consider potential shapes that support at
196: least one discrete eigenvalue for sufficiently large values of the coupling
197: $v$ and suppose for the sake of definiteness that the lowest eigenvalue~of
198: $-\Delta+vh(r)$ is given by $H(v)$ and that of $-\Delta+vf(r)$ by $F(v).$ If
199: the transformation function $g(h)$ is smooth, then each tangent to $g$ is an
200: affine transformation of the ``envelope basis''~$h$ of the form $f^{({\rm
201: t})}(r)=a(t)h(r)+b(t),$ where $r=t$ is the point of contact. The
202: coefficients~$a(t)$ and $b(t)$ are obtained by demanding that the
203: ``tangential potential'' $f^{({\rm t})}(r)$ and its derivative agree with
204: $f(r)$ at the point of contact $r=t.$ Thus we have$$a(t)=\frac{f'(t)}{h'(t)}\
205: ,\quad b(t)=f(t)-a(t)h(t)\ .$$The corresponding geometrical configuration is
206: illustrated in Fig.~\ref{Fig:shape} in which the potential~$f$ is chosen to
207: be the Coulomb-plus-linear potential, $f(r)=-1/r+r,$ and the envelope
208: basis~$h$ is, for the upper family, the harmonic-oscillator potential
209: $h(r)=r^2$ and, for the lower family, the Coulomb potential $h(r)=-1/r.$ The
210: spectral function $F^{({\rm t})}$ for the tangential potential $f^{({\rm
211: t})}(r)=a(t)h(r)+b(t)$ is given by $F^{({\rm t})}(v)=H(va(t))+vb(t).$ If the
212: transformation $g(h)$ has definite convexity, say $g''(h)>0,$ then each
213: tangential potential~$f^{({\rm t})}(r)$ lies beneath~$f(r)$ and, as a
214: consequence of the spectral comparison theorem, we know that each
215: corresponding tangential spectral function $F^{({\rm t})}(v),$ and the
216: envelope of this set, lie beneath $F(v).$ Similarly, in the case where $g$ is
217: concave, i.e., $g''(h)<0$, we obtain upper bounds to $F(v).$ These purely
218: geometrical arguments, depending on the spectral comparison theorem, extend
219: easily to~the excited states of the problem under consideration. The spectral
220: curves corresponding to~the envelope representations for the potential in
221: Fig.~\ref{Fig:shape} are shown in Fig.~\ref{Fig:bounds} for the excited state
222: $(n,\ell)=(2,4).$ For comparison the exact curve $E=F(v)$ is also shown in
223: Fig.~\ref{Fig:bounds}; this curve will be close to the Coulomb envelope for
224: large $v$ and to the oscillator envelope for~small~$v.$ Of course, the
225: envelopes of which we speak still have to be determined explicitly.
226: Extensions of this idea to completely new problems, such as simultaneous
227: transformations of each of~a number of potential terms \cite{Hall98}, the
228: Dirac equation \cite{Hall86,Hall99}, or the spinless-Salpeter problem of the
229: present paper, are best formulated initially with the basic argument outlined
230: above.
231:
232: \begin{figure}[ht]\vspace*{-2cm}\begin{center}
233: \psfig{figure=wsfig1.ps,scale=.75}\vspace*{-1cm}\caption{The
234: ``Coulomb-plus-linear'' potential shape $f(r)=-a/r+br$ represented as the
235: envelope curve of two distinct families of tangential potentials of the form
236: $\alpha h(r)+\beta.$ In the upper family $h(r)=r^2$ is the
237: harmonic-oscillator potential; in the lower family $h(r)=-1/r$ is~the Coulomb
238: potential. The adopted values of the relevant physical parameters $a$
239: and~$b$~are $a=0.2$ and $b=0.5.$}\label{Fig:shape}\end{center}\end{figure}
240:
241: \begin{figure}[ht]\vspace*{-2cm}\begin{center}
242: \psfig{figure=wsfig2.ps,scale=.75}\vspace*{-1cm}\caption{The spectral
243: approximation corresponding to Fig.~\ref{Fig:shape}. Each ``tangential''
244: potential $f^{({\rm t})}(r)=\alpha h(r)+\beta$ generates a corresponding
245: tangential energy curve $F^{({\rm t})}(v)=H(\alpha v)+\beta v.$ The envelopes
246: of these spectral families generate upper and lower bounds to the exact curve
247: $E=F(v),$ shown here for the case $(n,\ell)=(2,4).$}
248: \label{Fig:bounds}\end{center}\end{figure}
249:
250: In the 1-term case the question arises as to whether there is a simple way of
251: determining the envelopes of the families of upper and lower spectral
252: functions. One effective solution of this problem is by the use of ``kinetic
253: potentials'' which were introduced \cite{Hall83,Hall84} precisely for this
254: purpose. The idea is as follows. To each spectral function $F_{n\ell}(v)$
255: there is a corresponding ``kinetic potential'' (that is, a {\it minimum mean
256: iso-kinetic potential\/}) $\bar f_{n\ell}(s).$ The relationship between
257: $F_{n\ell}$ and $\bar f_{n\ell}$ is invertible and is essentially that of a
258: Legendre transformation \cite{Gelfand}:~we can prove in general that $F$ is
259: concave, $\bar f(s)$ is convex, and$$F''(v)\bar f''(s)=-\frac{1}{v^3}\ .$$The
260: explicit transformation formulas are as follows:\begin{equation}\bar
261: f_{n\ell}(s)=F'_{n\ell}(v)\ ,\quad s=F_{n\ell}(v)-vF'_{n\ell}(v)
262: \label{eq:TF1}\end{equation}and\begin{equation}\frac{F_{n\ell}(v)}{v}=\bar
263: f_{n\ell}(s)-s\bar f'_{n\ell}(s)\ ,\quad\frac{1}{v}=-\bar f'_{n\ell}(s)\
264: .\label{eq:TF2}\end{equation}An {\it a-priori\/} definition of the
265: ground-state kinetic potential $\bar f_{10}(s)=\bar f(s)$ is given by$$\bar
266: f(s)=\inf_{{{\scriptstyle\psi\in{\cal
267: D}(H)}\atop{\scriptstyle(\psi,\psi)=1}}\atop{\scriptstyle(\psi,-\Delta\psi)=s}}
268: (\psi,f\psi)\ ,$$where ${\cal D}\subset L^{2}(\mbox{\vrule R}^3)$ is the
269: domain of the Hamiltonian. The definition for the excited states~is a little
270: more complicated \cite{Hall84} and, in view of (\ref{eq:TF1}) and
271: (\ref{eq:TF2}), will not be needed in what follows.
272:
273: What is crucial is that the spectral functions, either exact or approximate,
274: are recovered from the corresponding kinetic potentials by a minimization
275: over the kinetic-energy variable $s.$ In this way the total minimization
276: required by the minimum--maximum principle \cite{Reed78,Thirring90} has been
277: divided into two steps: the first is constrained by $(\psi,-\Delta\psi)=s$
278: and the second~is~a minimization over $s.$ We have in all
279: cases:$$F_{n\ell}(v)=\min_{s>0}\left[s+v\bar f_{n\ell}(s)\right].$$Another
280: form of this expression is possible for the kinetic potential is monotone and
281: allows us to change variables $(s\rightarrow r)$ by $f(r)=\bar f_{n\ell}(s).$
282: Thus we have\begin{equation}
283: F_{n\ell}(v)=\min_{r>0}\left[K^{(f)}_{n\ell}(r)+vf(r)\right],\quad
284: K^{(f)}_{n\ell}=\bar f_{n\ell}^{-1}\circ f\ .\label{Eq:CV}\end{equation}The
285: two corresponding expressions of the envelope approximation then become$$\bar
286: f_{n\ell}(s)\approx g(\bar h_{n\ell}(s))\ ,\quad{\rm or}\quad
287: K^{(f)}_{n\ell}\approx K^{(h)}_{n\ell}\ .$$The second form (\ref{Eq:CV}) of
288: the explicit expression for $F_{n\ell}(v)$ isolates the potential
289: shape~$f$~itself and leads to an inversion sequence \cite{Hall00a} which
290: reconstructs the potential from a single given spectral function; but this is
291: another story \cite{Hall95,Hall99a}.
292:
293: It is useful here to provide the formulas for the kinetic potentials $\bar f$
294: corresponding~to~pure power-law potentials $V(r)=v\,{\rm sgn}(q)r^q.$
295: Elementary scaling arguments for the Hamiltonian$$H=-\Delta+v\,{\rm
296: sgn}(q)r^q$$show that the eigenvalues satisfy
297: $$F_{n\ell}(v)=v^{\frac{2}{2+q}}F_{n\ell}(1)\ ,$$where
298: $F_{n\ell}(1)=E_{n\ell}(q)$ are the eigenvalues of $H$ with coupling $v=1,$
299: i.e., of $-\Delta+{\rm sgn}(q)r^q.$ From Eq.~(\ref{eq:TF1}) we immediately
300: find that the kinetic potentials $\bar f_{n\ell}(s)$ for these
301: potentials~$V(r)$ are~given by\begin{equation}\bar f_{n\ell}(s)=\frac{2}{q}
302: \left|\frac{qE_{n\ell}(q)}{2+q}\right|^{\frac{2+q}{2}}s^{-\frac{q}{2}}\ .
303: \label{Eq:PL-KP}\end{equation}Meanwhile the corresponding $K$-functions all
304: have the same simple form \cite{Hall93}$$K_{n\ell}^{(q)}(r)=
305: \left(\frac{P_{n\ell}(q)}{r}\right)^2\ ,$$where the $P$ numbers are given by
306: $$P_{n\ell}(q)=\left|E_{n\ell}(q)\right|^{\frac{2+q}{2q}}
307: \left(\frac{2}{2+q}\right)^{\frac{1}{q}}
308: \left|\frac{q}{2+q}\right|^{\frac{1}{2}}\ ,\quad q\ne 0\ .$$Consequently, the
309: power-law kinetic potentials (\ref{Eq:PL-KP}) may be expressed in the simple
310: form$$\bar f_{n\ell}(s)={\rm
311: sgn}(q)\left(\frac{P_{n\ell}(q)}{\sqrt{s}}\right)^q\ ,\quad q\ne 0\ .$$Some
312: of the eigenvalues $E,$ and thus the corresponding $P$ numbers, are known
313: exactly from elementary quantum mechanics. From the known eigenvalues $E$ for
314: the Coulomb potential, $E_{n\ell}(-1)=-[2(n+\ell)]^{-2},$ and the
315: harmonic-oscillator potential, $E_{n\ell}(2)=4n+2\ell-1,$ we immediately
316: obtain the corresponding $P$ numbers:\begin{eqnarray*}P_{n\ell}(-1)&=&n+\ell\
317: ,\\[1ex]P_{n\ell}(2)&=&2n+\ell-\frac{1}{2}\ .\end{eqnarray*}The case $q=0$
318: corresponds {\it exactly\/} to the logarithmic potential
319: \cite{Quigg79,Hall00b}; the $P$ numbers~for the lowest states of the
320: logarithmic and the linear potentials may be found in
321: Table~\ref{Tab:P-numbers}.
322:
323: \begin{table}[ht]\caption{Numerical values for the $P$ numbers for the
324: logarithmic potential ($q=0$) and the linear potential ($q=1$) used in the
325: Schr\"odinger eigenvalue formula (\ref{Eq:EVs}).}\label{Tab:P-numbers}\small
326: \begin{center}\begin{tabular}{ccrr}\hline\hline&&\\[-1.5ex]
327: \multicolumn{1}{c}{$n$}&\multicolumn{1}{c}{$\ell$}&
328: \multicolumn{1}{c}{$P_{n\ell}(0)$}&\multicolumn{1}{c}{$P_{n\ell}(1)$}\\[1ex]
329: \hline\\[-1.5ex]
330: 1&0&1.21867&1.37608\\2&0&2.72065&3.18131\\3&0&4.23356&4.99255\\
331: 4&0&5.74962&6.80514\\5&0&7.26708&8.61823\\[.5ex]
332: 1&1&2.21348&2.37192\\2&1&3.68538&4.15501\\3&1&5.17774&5.95300\\
333: 4&1&6.67936&7.75701\\5&1&8.18607&9.56408\\[.5ex]
334: 1&2&3.21149&3.37018\\2&2&4.66860&5.14135\\3&2&6.14672&6.92911\\
335: 4&2&7.63639&8.72515\\5&2&9.13319&10.52596\\[.5ex]
336: 1&3&4.21044&4.36923\\2&3&5.65879&6.13298\\3&3&7.12686&7.91304\\
337: 4&3&8.60714&9.70236\\5&3&10.09555&11.49748\\[.5ex]
338: 1&4&5.20980&5.36863\\2&4&6.65235&7.12732\\3&4&8.11305&8.90148\\
339: 4&4&9.58587&10.68521\\5&4&11.06725&12.47532\\[.5ex]
340: 1&5&6.20936&6.36822\\2&5&7.64780&8.12324\\3&5&9.10288&9.89276\\
341: 4&5&10.56970&11.67183\\5&5&12.04517&13.45756\\[1ex]
342: \hline\hline\end{tabular}\end{center}\end{table}\normalsize
343:
344: In summary, if the potential $V(r)$ is a smooth transformation $V(r)=g({\rm
345: sgn}(q)r^q)$ of~the pure power-law potential ${\rm sgn}(q)r^q$, then the
346: eigenvalues of$$H=-\Delta+V(r)$$are given approximately by the
347: expression\begin{equation}E_{n\ell}\approx\min_{r>0}
348: \left[\frac{P_{n\ell}^2(q)}{r^2}+V(r)\right].\label{Eq:EVs}\end{equation}
349: Here, a sign of approximate equality is used to indicate that, for a definite
350: convexity~of~$g(h),$ Eq.~(\ref{Eq:EVs}) yields lower bounds for convex $g$
351: ($g''>0$) and upper bounds for concave $g$ ($g''<0$). The numbers
352: $P_{n\ell}(q)$ can be derived from the eigenvalues of the operator
353: $-\Delta+{\rm sgn}(q)r^q.$
354:
355: The lower bounds derived in the framework of envelope theory can be improved
356: by~use~of the refined comparison theorems of Ref.~\cite{Hall92}, which allow
357: comparison potentials~to intersect; a detailed study of the latter approach
358: is, however, beyond the scope of the present~analysis.
359:
360: As an immediate application we consider the (nonrelativistic) Schr\"odinger
361: Hamiltonian (\ref{Eq:SHam-HO}) for the semirelativistic spinless-Salpeter
362: harmonic-oscillator problem (\ref{Eq:SH}). Here we have$$H=-v\Delta+V(r)\
363: ,$$with$$V(r)=\beta\sqrt{m^2+r^2}\ ;$$this potential is a convex
364: transformation of a linear potential and a concave transformation of a
365: harmonic-oscillator potential. We conclude therefore from Eq.~(\ref{Eq:EVs})
366: (see also Ref.~\cite{Lucha00-HO}):\begin{equation}
367: \min_{r>0}\left[v\frac{P_{n\ell}^2(1)}{r^2}+\beta\sqrt{m^2+r^2}\right]\leq
368: {\cal E}_{n\ell}(v)\leq
369: \min_{r>0}\left[v\frac{P_{n\ell}^2(2)}{r^2}+\beta\sqrt{m^2+r^2}\right],
370: \label{Eq:SEVF-HO}\end{equation}where the numbers $P_{n\ell}(1)$ are given in
371: Table~\ref{Tab:P-numbers} and $P_{n\ell}(2)=2n+\ell-\frac{1}{2}$. By a simple
372: change of variables, $r\rightarrow r'=P/r,$ we are able to recast the
373: inequalities (\ref{Eq:SEVF-HO}) into the ``preferred'' form$${\cal
374: E}_{n\ell}(v)\approx
375: \min_{r>0}\left[\beta\sqrt{m^2+\frac{P^2_{n\ell}}{r^2}}+V(r)\right],\quad
376: V(r)=vr^2\ ,$$in which the function to be minimized is simply the
377: spinless-Salpeter Hamiltonian with~the momentum operator $|{\bf p}|$ replaced
378: by $P/r;$ the $P$ numbers yielding upper and lower bounds are as in
379: Eq.~(\ref{Eq:SEVF-HO}). Interestingly, the upper and lower bounds in
380: Eq.~(\ref{Eq:SEVF-HO}) are equivalent~to~the corresponding bounds
381: obtained~in~Ref.~\cite{Lucha99A}; however, these earlier specific bounds were
382: not derived as part of the general envelope theory. If we approximate the
383: square root from~above {\it again\/}, by using the elementary
384: inequality$$\sqrt{m^2+r^2}\le m+\frac{r^2}{2m}\ ,$$we obtain from
385: (\ref{Eq:SEVF-HO}) the weaker upper bound$${\cal E}_{n\ell}(v)\le\beta
386: m+\sqrt{\frac{\beta v}{2m}}(4n+2\ell-1)\ ,$$which is identical to that given
387: by the general ``Schr\"odinger upper~bound''
388: \cite{Lucha96a,Lucha98O,Lucha98D} obtained by initially approximating the
389: relativistic kinetic-energy operator above by$$\beta\sqrt{m^2+{\bf
390: p}^2}\le\beta\left(m+\frac{{\bf p}^2}{2m}\right).$$The latter upper bound,
391: and an improvement on it, will be discussed in the next section.
392:
393: \section{Envelope approximations for the relativistic kinetic
394: energy}\label{Sec:RKE}The relativistic kinetic-energy operator
395: $T=\beta\sqrt{m^2+{\bf p}^2}$ is a concave transformation of ${\bf p}^2.$
396: Thus ``tangent lines'' to this operator all are of the form $a{\bf p}^2+b$
397: and each one generates a Schr\"odinger operator that provides an upper bound
398: to $T.$ In a given application with given parameter values, one can search
399: for the best such upper bound. By elementary analysis~we can establish the
400: operator inequality\begin{equation}H=\beta\sqrt{m^2+{\bf p}^2}+V(r)\le
401: \frac{\beta}{2}\left(-\frac{\Delta}{\mu}+\mu+\frac{m^2}{\mu}\right)+V(r)\
402: ,\label{Eq:OI}\end{equation}where $\mu=\sqrt{m^2+p_1^2},$ and $|{\bf p}|=p_1$
403: is the ``point of contact'' of the tangent line with the square-root
404: function. The inequality (\ref{Eq:OI}) is identical to that obtained
405: \cite{Lucha96a} by employing the inequality $(T-\mu)^2\ge 0.$ Optimization
406: over $\mu$ for the Coulomb case $V(r)=-v/r$ recovers the explicit upper-bound
407: formula of Ref.~\cite{Lucha96a}:$$E_{n\ell}(v)\le
408: m\beta\sqrt{1-\left(\frac{v}{\beta(n+\ell)}\right)^2}\ .$$In the case of the
409: harmonic-oscillator potential, $V(r)=vr^2,$ we obtain the upper
410: bounds~\cite{Lucha99A}$$E_{n\ell}(v)\le\min_{\mu>0}\left[\sqrt{\frac{\beta
411: v}{2\mu}}(4n+2\ell-1)+\frac{\beta}{2}\left(\mu+\frac{m^2}{\mu}\right)\right].$$
412: (Brief reviews of analytical upper bounds on the energy eigenvalues of the
413: spinless Salpeter equation derived by combining operator inequalities with
414: the minimum--maximum principle may be found in
415: Refs.~\cite{Lucha98O,Lucha98D,Lucha98R}.)
416:
417: The strategy of the present section is to regard the relativistic
418: kinetic-energy operator~$T$ as a concave function of ${\bf p}^2,$ so that
419: ``tangent lines'' generate ``upper'' Schr\"odinger operators. This general
420: approach leads to the same upper bounds as those of Martin \cite{Martin88}
421: who used~the particular square-root form of the relativistic kinetic energy
422: to construct an operator whose positivity yields the bounds.
423:
424: \section{Envelope approximations for Salpeter Hamiltonians}\label{Sec:RET}
425: \subsection{The principal envelope formula}Let us now turn to our main topic
426: and consider the spinless-Salpeter Hamiltonian of Eq.~(\ref{Eq:SH}),
427: $$H=\beta\sqrt{m^2+{\bf p}^2}+V(r)\ ,$$and its eigenvalues $E.$ We shall
428: assume that the potential $V(r)$ is a smooth transformation $V(r)=g(h(r))$ of
429: another potential $h(r)$ and that $g$ has definite convexity so that
430: we~obtain bounds to the energy eigenvalues $E$. We suppose that the ``basis''
431: potential $h(r)$ generates~a ``tangential'' Salpeter problem$${\cal
432: H}=\beta\sqrt{m^2+{\bf p}^2}+vh(r)\ ,$$for which the eigenvalues $e(v),$ or
433: bounds to them, are known. We shall follow here as~closely as possible the
434: development in Sec.~\ref{Sec:NRET} for the corresponding Schr\"odinger
435: problem.
436:
437: First of all, we recall that the approximations or bounds to the energy
438: eigenvalues of~the relativistic Coulomb and harmonic-oscillator problems we
439: shall eventually use from Secs.~\ref{Sec:CHOP} and \ref{Sec:NRET}, when
440: regarded as functions of the coupling~$v,$ are all {\it concave\/}.
441: Furthermore, it is~easy to convince oneself that all the (unknown) energy
442: functions $e(v)$ of the ``tangential'' Salpeter problem~are {\it concave\/},
443: that is, $e''(v)<0.$ Suppose that the exact eigenvalue and~(normalized)
444: eigenvector for the problem posed by the ``tangential'' Hamiltonian$${\cal
445: H}=\beta\sqrt{m^2+{\bf p}^2}+vh(r)$$are $e(v)$ and $\psi(v,r).$ Then, by
446: differentiating the expectation value $(\psi,{\cal H}\psi)$ with respect~to
447: the coupling $v,$ we find$$e'(v)=(\psi,h\psi)\ .$$If we now apply $\psi(v,r)$
448: as a trial vector to estimate the energy of the operator$$\beta\sqrt{m^2+{\bf
449: p}^2}+uh(r)\ ,$$in which $v$ has been replaced by $u,$ we obtain an upper
450: bound to the corresponding energy function $e(u)$ which may be written in the
451: form$$e(u)\leq e(v)+(u-v)e'(v)\ .$$This inequality tells~us that the function
452: $e(u)$ lies beneath its tangents; that is to say,~$e(u)$~is {\it concave\/}.
453: Convexity properties of the energy functions of the corresponding
454: (nonrelativistic) Schr\"odinger problem have been investigated in
455: Refs.~\cite{Narnhofer75,Hall83,Thirring90}.
456:
457: Next, in order to prove the main result of this section, the ``principal
458: envelope formula,'' we begin by using an envelope representation for the
459: potential $V(r)$ in the Hamiltonian~(\ref{Eq:SH}) and then demonstrate that
460: all the spectral formulas that follow possess a certain structure. Finally,
461: as an application, we specialize to the case of pure power-law ``basis''
462: potentials~$h(r)$ and, more particularly, to the Coulomb potential and the
463: harmonic-oscillator potential for which, at this time, we have spectral
464: information [cf.\ the discussions in Secs.~\ref{Subsec:CP} and
465: \ref{Subsec:HOP}, and the exact bounds (\ref{Eq:SEVF-HO}) on the energy
466: levels ${\cal E}_{n\ell}(v)$ of the relativistic harmonic oscillator].
467:
468: The tangential potentials we shall employ have the form $V^{({\rm
469: t})}(r)=a(t)h(r)+b(t),$~where, as in the Schr\"odinger case, the coefficients
470: $a(t)$ and $b(t)$ are given by$$a(t)=\frac{V'(t)}{h'(t)}=g'(h(t))\ ,\quad
471: b(t)=V(t)-a(t)h(t)=g(h(t))-g'(h(t))h(t)\ ,$$and $r=t$ is the point of contact
472: of the potential $V(r)$ and its tangent $V^{({\rm t})}(r)$. If, for~the~sake
473: of definiteness, we assume that $V=g(h)$ with $g$ concave (i.e., $g''<0$), we
474: obtain a family~of upper bounds given by$$E\le\varepsilon(t)=e(a(t))+b(t)\
475: .$$The best of these is given by optimizing over $t$:$$E\le\varepsilon(\hat
476: t)=e(a(\hat t))+b(\hat t)\ ,$$where $\hat t,$ the value of $t$ which
477: optimizes these bounds, is to be determined as the
478: solution~of$$e'(g'(h(\hat{t})))=h(\hat{t})\ .$$In the spirit of the Legendre
479: transformation \cite{Gelfand} we now consider another problem which~has the
480: same solution; this second problem is the one that provides us with our
481: basic~eigenvalue formula. We consider$${\cal
482: E}\equiv\min_{v>0}[e(v)-ve'(v)+g(e'(v))]\ ,$$which is well defined since
483: $e(v)$ is concave. The solution has the critical point$$\hat v=g'(e'(\hat
484: v))\ .$$If we now apply the correspondence $h(\hat t)=e'(\hat{v}),$ it
485: follows that the critical point $\hat v$ becomes$$\hat v=g'(h(\hat t))\
486: ,$$and the tangential-potential coefficients $a$ and $b$
487: become\begin{equation}a(\hat t)=g'(e'(v))=v\ ,\quad b(\hat
488: t)=g(e'(v))-ve'(v)\ ,\quad v=\hat v\ .\label{Eq:TPC}\end{equation}Meanwhile
489: the original critical (energy) value is given by$$\varepsilon(\hat
490: t)=e(a(\hat t))+b(\hat t)=e(v)-ve'(v)+g(e'(v))\ ,\quad v=\hat v\ .$$Thus we
491: conclude that the spectral approximation obtained by envelope methods is
492: given by the following ``principal envelope formula:''\begin{equation}
493: E\approx{\cal E}\equiv\min_{v>0}[e(v)-ve'(v)+g(e'(v))]\
494: .\label{Eq:PEF}\end{equation}If $g$ is concave (that is, $g''<0$), then
495: $E\le{\cal E};$ if $g$ is convex (that is, $g''>0$), then $E\ge{\cal E}.$
496: From the above considerations it follows immediately that, if the {\it
497: exact\/} energy function $e(v)$ corresponding to the basis potential $h$ is
498: not available, then, for $g(h)$ concave, concave~{\it upper\/} approximations
499: $e_{\rm u}(v)>e(v)$ or, for $g(h)$ convex, concave {\it lower\/}
500: approximations $e_{\rm l}(v)<e(v)$ may be used instead of the exact energy
501: function $e(v)$ in the principal envelope formula~(\ref{Eq:PEF}). Then all
502: the lower tangents will lie even lower and all the upper tangents will lie
503: even~higher. If $g$ is convex, we obtain a lower bound; if $g$ is concave, we
504: obtain an upper bound; because~of the concavity of $e(v),$ {\it this\/}
505: extremum is a minimum in {\it both\/} cases. If we wish to use numerical
506: solutions to the ``basis'' problem (generated by $h(r)$), or if a completely
507: new energy-bound expression becomes available, the principal envelope formula
508: (\ref{Eq:PEF}) is the relation that would at first be used.
509:
510: Interestingly, in the formula (\ref{Eq:PEF}) the tangential-potential
511: apparatus is no longer evident; only the correct convexity is required. As in
512: the Schr\"odinger case \cite{Hall83}, once we have the basic result, the
513: reformulation in terms of ``kinetic potentials'' is often useful: the kinetic
514: potential $\bar h(s)$ corresponding to some basis potential $h(r)$ is given
515: by the Legendre transformation~\cite{Gelfand}$$\bar h(s)=e'(v)\ ,\quad
516: s=e(v)-ve'(v)\ .$$Meanwhile the envelope approximation has the
517: kinetic-potential expression $\bar{V}(s)\approx g(\bar h(s)).$
518:
519: For both the Coulomb lower bounds (\ref{Eq:Herbst-bound}) or (\ref{Eq:NLB})
520: and the harmonic-oscillator upper bounds (\ref{Eq:SEVF-HO}) which we have at
521: present, we may express our general results in a special common~form which
522: will now be derived.
523:
524: \subsection{The Coulomb lower bound}We consider first the Coulomb lower bound
525: in which we assume that the potential $V(r)$~is~a convex transformation
526: $V(r)=g(h(r))$ of the Coulomb potential $h(r)=-1/r.$ According~to
527: Sec.~\ref{Subsec:CP}, in this case all the ``lower'' $e_{\rm l}(v)$ have been
528: arranged---with the parameters $\beta$ and~$m$ returned---in the form$$e_{\rm
529: l}(v)=\beta m\sqrt{1-\left(\frac{\sigma v}{\beta}\right)^2}\ .$$From this it
530: follows by elementary algebra that if we define a new optimization
531: variable~$r$~by $e_{\rm l}'(v)=h(r)=-1/r,$ we have$$e_{\rm l}(v)-ve_{\rm
532: l}'(v)=\beta\sqrt{m^2+\frac{P^2}{r^2}}\ ,\quad P\equiv\frac{1}{\sigma}\ .$$
533: Consequently, the lower bound on the energy eigenvalues $E$ of the spinless
534: Salpeter equation becomes\begin{equation}
535: E\ge\min_{r>0}\left[\beta\sqrt{m^2+\frac{P^2}{r^2}}+V(r)\right],\quad v<\beta
536: v_P\ .\label{Eq:SSE-LB}\end{equation}Here the boundary value $v_P$ of the
537: Coulomb coupling $v$ is given, when simply determined by the requirement of
538: boundedness from below of the operator (\ref{Eq:SH}), by the critical
539: coupling~$v_{\rm c}$,$$v_P=v_{\rm c}=\frac{2}{\pi}\ ,$$and, when arising from
540: the region of validity of our Coulomb-like family of lower
541: bounds~(\ref{Eq:NLB}), via $P=1/\sigma,$ by\begin{equation}
542: v_P=P\sqrt{1-P^2}<\frac{1}{2}\ .\label{Eq:CCUBvP}\end{equation}Some
543: $\{P,v_P\}$ pairs may be found in Table~\ref{Tab:PvP-pairs}; others can be
544: easily generated from the upper bound on the coupling $v$ given in
545: Eq.~(\ref{Eq:CCUBvP}). The meaning of the Coulomb-coupling~constraint is
546: $a(\hat t)<\beta v_P,$ where $a$ is the coefficient in the tangential Coulomb
547: potential given by (\ref{Eq:TPC}).
548:
549: \begin{table}[ht]\caption{Explicit values of some $\{P,v_P\}$ pairs, obtained
550: via the equality $P=1/\sigma$ either~from the Herbst lower bound
551: (\ref{Eq:Herbst-bound}) or the expression (\ref{Eq:NLB}) for our new lower
552: bounds on the spectrum of the spinless relativistic Coulomb problem (in three
553: spatial dimensions).}\label{Tab:PvP-pairs}\small
554: \begin{center}\begin{tabular}{cc}\hline\hline\\[-1.5ex]
555: \multicolumn{1}{c}{$P$}&\multicolumn{1}{c}{$v_P$}\\[1ex]\hline\\[-1.5ex]
556: $\displaystyle\frac{2}{\pi}$&$\displaystyle\frac{2}{\pi}$\\[2ex]
557: $\displaystyle\frac{1}{\sqrt{2}}$&$\displaystyle\frac{1}{2}$\\[2ex]
558: $\sqrt{\displaystyle\frac{2}{9-3\sqrt{5}}}$&$\displaystyle\frac{1}{3}$\\[2ex]
559: $\displaystyle\frac{1}{2\sqrt{2-\sqrt{3}}}$&$\displaystyle\frac{1}{4}$\\[3ex]
560: \hline\hline\end{tabular}\end{center}\end{table}\normalsize
561:
562: As a rather trivial consistency check of our formalism, the Coulomb lower
563: energy bound of Eq.~(\ref{Eq:SSE-LB}) may be applied to the Coulomb potential
564: $V(r)=-v/r$ in order to re-derive,~for $P=2/\pi,$ the Herbst formula
565: (\ref{Eq:Herbst-bound})---which is nothing else~but the starting point of the
566: present ``lower-bound'' considerations.
567:
568: \subsection{The harmonic-oscillator upper bounds}Next, let us turn to the
569: harmonic-oscillator upper bounds. Our main assumption is here that
570: $V(r)=g(r^2),$ with $g''<0.$ In this case the only difficulty is that the
571: basis problem~$h(r)=r^2$ is equivalent to a Schr\"odinger problem whose
572: solution ${\cal E}_{n\ell}(v)$ is not known exactly. Following the discussion
573: of suitable bounds after the proof of the principal envelope formula,
574: Eq.~(\ref{Eq:PEF}), let us call the upper bound provided by
575: Eq.~(\ref{Eq:SEVF-HO}) $e_{\rm u}(v),$ and let us introduce the shorthand
576: notation $P_{n\ell}(2)=2n+\ell-\frac{1}{2}=P.$ Then we have~the following
577: parametric equations~for~$e_{\rm u}(v)$:$$e_{\rm
578: u}(v)=v\frac{P^2}{r^2}+\beta\sqrt{m^2+r^2}\ ,\quad v=\frac{\beta
579: r^4}{2P^2\sqrt{m^2+r^2}}\ ,\quad e_{\rm u}'(v)=\frac{P^2}{r^2}\ .$$By
580: substituting these expressions into the fundamental envelope formula
581: (\ref{Eq:PEF}) we obtain~the following upper bound on all the eigenvalues of
582: the spinless-Salpeter problem with potential $V(r)=g(r^2)$ and $g''<0$:
583: \begin{equation}
584: E_{n\ell}\le\min_{r>0}\left[\beta\sqrt{m^2+\frac{P^2}{r^2}}+V(r)\right],\quad
585: P=P_{n\ell}(2)=2n+\ell-\frac{1}{2}\ .\label{Eq:SSE-UB}\end{equation}
586:
587: \section{The Coulomb-plus-linear (or ``funnel'') potential}\label{Sec:FP}In
588: order to illustrate the above results by a physically motivated example, let
589: us apply~these considerations to the Coulomb-plus-linear or (in view of its
590: shape) ``funnel'' potential$$V(r)=-\frac{c_1}{r}+c_2r\ ,\quad c_1\ge 0\
591: ,\quad c_2\ge 0\ .$$(This potential provides a reasonable overall description
592: of the strong interactions of quarks in hadrons. For the phenomenological
593: description of hadrons in terms of both nonrelativistic and semirelativistic
594: potential models, see, e.g., Refs.~\cite{Lucha91:BSQ,Lucha92:QAQBS}.) By
595: choosing as basis potential the Coulomb potential $h(r)=-1/r,$ we may write
596: $V(r)=g(h(r))$ with$$g(h)=c_1h-\frac{c_2}{h}\ ,$$which is clearly a convex
597: function of $h<0$: $g''>0.$ Thus the convexity condition is~satisfied.
598: However, we are not free to choose the coupling constants $c_1$ and $c_2$ as
599: large as we please.~It is immediately obvious that, for a particular $\{P,
600: v_P\}$ pair, we must in any case have~$c_1<\beta v_P.$ For the full problem
601: the coefficient $c_2$ of the linear term will also be involved. The
602: coupling~$v$ we are concerned about is given by (\ref{Eq:TPC}). We
603: have$$v=g'(e'(v))=\frac{\beta
604: P^2}{r\sqrt{m^2+\left(\displaystyle\frac{P}{r}\right)^2}}
605: =c_1+\frac{c_2}{h^2}=c_1+c_2r^2\ .$$From this we obtain, for given values of
606: the parameters $m$ and $\beta$ and for a given $\{P, v_P\}$~pair, as a
607: sufficient condition for $v<\beta v_P$ the ``Coulomb coupling constant
608: constraint'' on the~two coupling strengths $c_1$ and $c_2$ in the funnel
609: potential:\begin{equation}c_1+\frac{P^2}{m^2}\left(\frac{P^2}{v_P^2}-1\right)
610: c_2<\beta v_P\ .\label{Eq:CCCC}\end{equation}In the case $\{P=1/\sqrt{2},
611: v_P=1/2\}$ and $\beta=m=1$ this condition reduces to
612: $c_1+\frac{1}{2}c_2<\frac{1}{2}.$ For Herbst's lower bound
613: (\ref{Eq:Herbst-bound}), i.e., $P=v_P=v_{\rm c}=2/\pi,$ this constraint
614: clearly yields~$c_1<\beta v_P.$ There is no escaping this feature of all
615: energy bounds involving the Coulomb potential: the constraint derives from
616: the fundamental observation that the Coulomb coupling $v$ must~not be too
617: large, so that the (relativistic) kinetic energy is able to counterbalance
618: the Coulomb potential in order to maintain the Hamiltonian (\ref{Eq:SH}) with
619: $V(r)=-v/r$ bounded from below.
620:
621: For example, if we seek the largest allowed value of the parameter $P$ by
622: solving Eqs.~(\ref{Eq:CCUBvP}) and (\ref{Eq:CCCC}) together, we find that
623: this largest $P$ is given by
624: \begin{equation}\frac{c_2\sin^4t}{\cos^2t(\beta\sin t\cos t-c_1)}=m^2\ ,\quad
625: P\equiv\sin t\ .\label{Eq:P(m)}\end{equation}
626:
627: \begin{figure}[ht]\vspace*{-2cm}\begin{center}
628: \psfig{figure=wgfig4.ps,scale=0.792}\vspace*{-0.5cm}\caption{Lower bounds
629: (L), according to (\ref{Eq:SSE-LB}), and upper bounds (U), according to
630: (\ref{Eq:SSE-UB}),~on the energy eigenvalue $E$ of the ground state
631: [$(n,\ell)=(1,0)$] of the spinless Salpeter equation with a
632: Coulomb-plus-linear potential $V(r)=-c_1/r+c_2r,$ for $\beta=1,$ $c_1=0.1,$
633: and $c_2=0.25.$ The lower bound is given by the general result
634: (\ref{Eq:SSE-LB}) with the ``best'' $P(m)$ provided~by~(\ref{Eq:P(m)}). In
635: order to satisfy the Coulomb coupling constraint (\ref{Eq:P(m)}), the mass
636: $m$ must fulfil $m>\sqrt{5}/4.$ For comparison, a (very accurate)
637: Rayleigh--Ritz variational upper bound $E$ is
638: depicted~too.}\label{Fig:Bounds}\end{center}\end{figure}
639:
640: For the Coulomb-plus-linear potential $V(r)=-c_1/r+c_2r$ under consideration,
641: Fig.~\ref{Fig:Bounds} shows the lower and upper bounds on the lowest energy
642: eigenvalue $E$ of the spinless Salpeter equation, given by the envelopes of
643: the lower and upper families of tangential energy curves (\ref{Eq:SSE-LB})
644: and (\ref{Eq:SSE-UB}), as functions $E(m)$ of the mass $m$ entering in the
645: semirelativistic Hamiltonian. In the case of the Coulomb lower bound
646: (\ref{Eq:SSE-LB}), we have employed for each $m$ the best possible $P(m)$
647: provided by (\ref{Eq:P(m)}). As $m\to 0,$ the ``basis'' Coulomb problem
648: $H=\beta\sqrt{m^2+{\bf p}^2}-v/r$~has energy $e(m)\to 0;$ thus the Coulomb
649: lower bound for a non-Coulomb problem becomes very weak for small values of
650: $m.$ Of course, Eq.~(\ref{Eq:SSE-UB}) provides us with rigorous upper bounds
651: for {\it every\/} energy level.
652:
653: In order to get an idea of the location of the exact energy eigenvalues $E$,
654: Fig.~\ref{Fig:Bounds} also shows the ground-state energy curve $E(m)$
655: obtained by the Rayleigh--Ritz variational technique~\cite{Reed78} with the
656: Laguerre basis states for the trial space defined in Ref.~\cite{Lucha97}.
657: Strictly speaking, this energy curve represents only an upper bound to the
658: precise eigenvalue $E$. However,~from~the findings of Ref.~\cite{Lucha97} the
659: deviations of these Laguerre bounds from the exact eigenvalues~may be
660: estimated, for the superposition of 25 basis functions used here, to be of
661: the order of~1\,\%.
662:
663: \begin{figure}[ht]\vspace*{-2cm}\begin{center}
664: \psfig{figure=wgfig3.ps,scale=0.792}\vspace*{-1cm}\caption{Lower bounds (L),
665: according to (\ref{Eq:SSE-LB}), and upper bounds (U), according to
666: (\ref{Eq:SSE-UB}),~on the energy eigenvalue $E$ of the ground state
667: [$(n,\ell)=(1,0)$] of the spinless Salpeter equation with Coulomb-plus-linear
668: potential $V(r)=v(-a/r+br),$ for $a=0.2,$ $b=0.5,$ $m=\beta=1.$ The lower
669: bound is given by the general result (\ref{Eq:SSE-LB}) with $P(m=1)$ computed
670: from Eq.~(\ref{Eq:P(m)}), evaluated for mass $m=1$ and the funnel-potential
671: coupling strengths $c_1=va$ and $c_2=vb.$ For comparison, a (very accurate)
672: Rayleigh--Ritz variational upper bound $E$ is
673: depicted~too.}\label{Fig:SSE-bounds}\end{center}\end{figure}
674:
675: Figure~\ref{Fig:SSE-bounds} shows, for a Coulomb-plus-linear potential of the
676: form $V(r)=v(-a/r+br),$~the lower and upper bounds on the lowest energy
677: eigenvalue $E$ of the spinless Salpeter equation, given by the envelopes of
678: the lower and upper families of tangential energy curves (\ref{Eq:SSE-LB})
679: and (\ref{Eq:SSE-UB}), as functions $E(v)$ of the ``overall'' coupling
680: parameter $v$ which multiplies the potential shape $-a/r+br.$ Again we
681: compare these bounds with the ground-state energy curve $E(v),$ obtained by
682: the Rayleigh--Ritz variational technique \cite{Reed78} with the Laguerre
683: basis states \cite{Lucha97}.
684:
685: \section{Summary and conclusion}In this analysis we have studied the discrete
686: spectrum of semirelativistic ``spinless-Salpeter'' Hamiltonians $H,$ defined
687: in Eq.~(\ref{Eq:SH}), by an approach which is based principally on convexity.
688: We have at our disposal very definite information concerning, on the one
689: hand, the bottom of the spectrum of $H$ for the Coulomb potential,
690: $h(r)=-1/r,$ and, on the other hand, the entire spectrum of $H$ for the
691: harmonic-oscillator potential, $h(r)=r^2.$ The class of potentials that are
692: at the same time a convex transformation of $-1/r$ and a concave
693: transformation~of $r^2$ includes, for example, arbitrary linear combinations
694: of Coulomb, logarithmic, linear,~and harmonic-oscillator terms. In order to
695: obtain information about the eigenvalues $E$ of $H$ for arbitrary members
696: within this class of potentials, we have extended the---for nonrelativistic
697: Schr\"odinger operators well-established---formalism of envelope theory to
698: Hamiltonians with relativistic kinetic energies. The envelope technique
699: applied here takes advantage of the fact that all ``tangent lines'' to the
700: interaction potential $V(r)=g(h(r))$ in $H$ are potentials~of~the form
701: $ah(r)+b,$ and that, by convexity and the comparison theorem recalled in
702: Subsec.~\ref{Subsec:SCT}, the energy eigenvalues corresponding to these
703: ``tangent'' potentials provide rigorous bounds to the unknown exact
704: eigenvalues $E$ of $H.$ If $e(v)$ denotes the energy function---or a suitable
705: bound to it---corresponding to the problem posed by a ``basis'' potential
706: $vh(r),$ where $v$~is~a positive coupling parameter, the envelopes of upper
707: and lower families of energy curves~may be found with the help of the
708: ``principal envelope formula''$$E\approx\min_{v>0}[e(v)-ve'(v)+g(e'(v))]\
709: .$$Here, a sign of approximate equality is used to indicate that, for a
710: definite convexity~of~$g(h),$ the envelope theory yields lower bounds for
711: convex $g(h)$ and~upper bounds for concave $g(h).$ With the above principal
712: envelope formula at hand, all new spectral pairs $\{h(r),e(v)\}$ which may
713: become available at some future time can immediately be used to enrich our
714: collection of energy bounds. If the basis potential~$h(r)$ is a pure power,
715: these bounds can be written~as$$E_{n\ell}\approx\min_{r>0}
716: \left[\beta\sqrt{m^2+\frac{P_{n\ell}^2}{r^2}}+V(r)\right],$$where the numbers
717: $P_{n\ell}$ are obtained from the corresponding underlying basis
718: problems.~The power of this technique is illustrated, in Sec.~\ref{Sec:FP},
719: by our application to the ``funnel'' potential, $V(r)=-c_1/r+c_2r.$ For this
720: problem, we have employed both the semirelativistic Coulomb and
721: harmonic-oscillator problems to calculate, respectively, lower and upper
722: bounds on the energy eigenvalues of the spinless Salpeter equation.
723:
724: We expect that such results would provide bounds on the energy eigenvalues
725: for general theoretical discussions, or be used as guides for more tightly
726: focussed analytic or numerical studies of the spectra of semirelativistic
727: ``spinless-Salpeter'' Hamiltonians.
728:
729: \section*{Acknowledgement}Partial financial support of this work under Grant
730: No.~GP3438 from the Natural Sciences and Engineering Research Council of
731: Canada, and the hospitality of the Erwin Schr\"odinger International
732: Institute for Mathematical Physics in Vienna is gratefully acknowledged by
733: one of us (R.~L.~H.).
734:
735: \small\begin{thebibliography}{30}
736: \bibitem{Salpeter51}E.~E.~Salpeter and H.~A.~Bethe, Phys.~Rev.\ {\bf 84},
737: 1232 (1951).
738: \bibitem{Salpeter52}E.~E.~Salpeter, Phys.~Rev.\ {\bf 87}, 328 (1952).
739: \bibitem{Lucha98O}W.~Lucha and F.~F.~Sch\"oberl, Int.~J.~Mod.~Phys.~A {\bf
740: 14}, 2309 (1999), hep-ph/9812368.
741: \bibitem{Lucha98D}W.~Lucha and F.~F.~Sch\"oberl, Fizika B {\bf 8}, 193
742: (1999), hep-ph/9812526.
743: \bibitem{Herbst77}I.~W.~Herbst, Commun.~Math.~Phys.\ {\bf 53}, 285 (1977);
744: {\bf 55}, 316 (1977) (addendum).
745: \bibitem{Martin89}A.~Martin and S.~M.~Roy, Phys.~Lett.~B {\bf 233}, 407
746: (1989).
747: \bibitem{Lucha99Q}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~A {\bf 60}, 5091
748: (1999), hep-ph/9904391.
749: \bibitem{Lucha99A}W.~Lucha and F.~F.~Sch\"oberl, Int.~J.~Mod.~Phys.~A {\bf
750: 15}, 3221 (2000), hep-ph/9909451.
751: \bibitem{Reed78}M.~Reed and B.~Simon, {\em Methods of Modern Mathematical
752: Physics~IV: Analysis~of Operators\/} (Academic Press, New York, 1978).
753: \bibitem{Narnhofer75}H.~Narnhofer and W.~Thirring, Acta Phys.~Austriaca {\bf
754: 41}, 281 (1975).
755: \bibitem{Thirring90}W.~Thirring, {\em A Course in Mathematical Physics 3:
756: Quantum Mechanics of Atoms and Molecules\/} (Springer, New York/Wien, 1990).
757: \bibitem{Hall83}R.~L.~Hall, J.~Math.~Phys.\ {\bf 24}, 324 (1983).
758: \bibitem{Hall84}R.~L.~Hall, J.~Math.~Phys.\ {\bf 25}, 2708 (1984).
759: \bibitem{Hall98}R.~L.~Hall and N.~Saad, J.~Chem.~Phys.\ {\bf 109}, 2983
760: (1998).
761: \bibitem{Hall86}R.~L.~Hall, J.~Phys.~A {\bf 19}, 2079 (1986).
762: \bibitem{Hall99}R.~L.~Hall, Phys.~Rev.~Lett.\ {\bf 83}, 468 (1999).
763: \bibitem{Gelfand}I.~M.~Gelfand and S.~V.~Fomin, {\em Calculus of
764: Variations\/} (Prentice-Hall, Englewood Cliffs, 1963). Legendre
765: transformations are discussed on p. 72.
766: \bibitem{Hall00a}R.~L.~Hall, Phys.~Lett.~A {\bf 265}, 28 (2000).
767: \bibitem{Hall95}R.~L.~Hall, J.~Phys.~A {\bf 28}, 1771 (1995).
768: \bibitem{Hall99a}R.~L.~Hall, J.~Math.~Phys.\ {\bf 40}, 699 (1999).
769: \bibitem{Hall93}R.~L.~Hall, J.~Math.~Phys.\ {\bf 34}, 2779 (1993).
770: \bibitem{Quigg79}C.~Quigg and J.~L.~Rosner, Phys.~Rep.~{\bf 56}, 167 (1979).
771: \bibitem{Hall00b}R.~L.~Hall, J.~Phys.~G {\bf 26}, 981 (2000); ESI programme
772: {\em Confinement\/} (Vienna, Austria, 2000).
773: \bibitem{Hall92}R.~L.~Hall, J.~Phys.~A {\bf 25}, 4459 (1992).
774: \bibitem{Lucha00-HO}R.~L.~Hall, W.~Lucha, and F.~F.~Sch\"oberl, J.~Phys.~A
775: {\bf 34}, 5059 (2001), hep-th/0012127.
776: \bibitem{Lucha96a}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~A {\bf 54}, 3790
777: (1996), hep-ph/9603429.
778: \bibitem{Lucha98R}W.~Lucha and F.~F.~Sch\"oberl, in: Proceedings of the
779: XI$^{\rm th}$ International Conference ``Problems of Quantum Field Theory,''
780: edited by B.~M.~Barbashov, G.~V.~Efimov, and A.~V.~Efremov, July 13 -- 17,
781: 1998, Dubna, Russia (Joint Institute for Nuclear Research, Dubna, 1999),
782: p.~482, hep-ph/9807342.
783: \bibitem{Martin88}A.~Martin, Phys.~Lett.~B {\bf 214}, 561 (1988).
784: \bibitem{Lucha91:BSQ}W.~Lucha, F.~F.~Sch\"oberl, and D.~Gromes,
785: Phys.~Rep.~{\bf 200}, 127 (1991).
786: \bibitem{Lucha92:QAQBS}W.~Lucha and F.~F.~Sch\"oberl, Int.~J.~Mod.~Phys.~A
787: {\bf 7}, 6431 (1992).
788: \bibitem{Lucha97}W.~Lucha and F.~F.~Sch\"oberl, Phys.~Rev.~A {\bf 56}, 139
789: (1997), hep-ph/9609322.
790: \end{thebibliography}\end{document}
791: