1: %\documentstyle[preprint,revtex]{aps}
2: %\documentstyle[preprint,eqsecnum,psfig]{aps}
3: \documentstyle[multicol,aps,psfig]{revtex}
4: %\setlength{\topmargin}{-0.25in}
5: \begin{document}
6:
7: \draft
8:
9:
10: \title{Arbitrarily Accurate Eigenvalues for One-Dimensional Polynomial
11: Potentials}
12:
13: \author{Y. Meurice \\
14: {\it Department of Physics and Astronomy, The University of Iowa,
15: Iowa City, Iowa 52242, USA}}
16:
17: \maketitle
18: \begin{abstract}
19: We show that the Riccati form of the one-dimensional Schr\"odinger
20: equation can be reformulated in
21: terms of two linear equations depending on an arbitrary function $G$.
22: When $G$ and the potential (as for anharmonic oscillators)
23: are polynomials
24: the solutions of these two equations are entire
25: functions ($L$ and $K$) and
26: the zeroes of $K$ are identical to those of the wave function.
27: Requiring such a zero at a large but finite value of the argument
28: yields the low energy eigenstates with exponentially small errors.
29: Approximate formulas for these errors are provided.
30: We explain how to chose $G$ in order
31: improve dramatically the numerical treatment.
32: The method yields many significant
33: digits with modest computer means. We discuss the extension of this
34: method in the case of several variables.
35:
36:
37: \end{abstract}
38: \pacs{PACS: 03.65.Ge, 03.65.-w, 02.30.Em, 02.30.Mv, 10.10.St, 33.20.-t}
39:
40: \begin{multicols}{2}\global\columnwidth20.5pc
41: \multicolsep=8pt plus 4pt minus 3pt
42:
43: \section{Introduction}
44: Quantum anharmonic oscillators appear
45: in a wide variety of problems in molecular,
46: nuclear or condensed matter physics.
47: Typically,
48: anharmonic terms appear in expansions about a minimum of a potential, when
49: ones tries to incorporate the non-linear features of the
50: forces responsible for this
51: equilibrium.
52: The most celebrated example is the quartic anharmonic oscillator
53: \cite{bender69}
54: where a $\lambda x^4$ term is added to the usual harmonic Hamiltonian.
55: Introducing bilinear couplings among a set of such oscillators leads to
56: a rich spectrum, for instance, multiphonon bound states in one-dimensional
57: lattice models \cite{wang96}.
58: More generally, one can think about the $\lambda \phi^4$ (or
59: higher powers of $\phi$) field theories in
60: various dimensions as systems of coupled anharmonic oscillators.
61:
62: Anharmonic terms can be treated perturbatively and the
63: perturbative series can be represented by Feynman diagrams. Unfortunately,
64: the coefficients of the series\cite{bender69,leguillou90}
65: have a factorial growth
66: and the numerical values obtained from the
67: truncated series have an
68: accuracy which is subject to limitations. At fixed coupling, there is an
69: order at which an optimal accuracy is reached. At fixed order, there is
70: a value of the coupling beyond which the numerical values are meaningless
71: even as an order of magnitude. In the case of the
72: single-well quartic potential, Pad\'e approximants can be used for the
73: series or its Borel transform. Rigorous proofs of convergence can
74: be established in particular cases \cite{loeffel69}.
75: Unfortunately, such a method does not apply to the case of the double-well
76: potential\cite{brezin77}
77: where instanton effects \cite{coleman,zj} need to be taken into
78: account. It should also be noted that even when Pad\'e approximants converge,
79: the convergence rate may be slow. Strong coupling
80: expansions \cite{weninger96} or variational interpolations \cite{kleinert}
81: sometimes provide more accurate results.
82:
83: The above discussion shows that
84: finding an expansion which can be used {\it indiscriminately} for
85: most quantum mechanical problems with polynomial potentials
86: remains a challenging problem.
87: Alternatively, one can use numerical methods. Variational methods
88: are often used to obtain upper and lower bounds on energy levels
89: \cite{bazley,payne}. These methods are based on rigorous inequalities
90: and are considered superior to methods based on numerical integration
91: \cite{payne}. However, the difference between the bounds widens rapidly
92: with the anharmonic coupling and the energy level.
93: Methods based on series expansions in the position variable
94: \cite{biswas,kbeck81,fernandez,bacus} appear to produce more significant digits
95: more easily. However, our understanding of
96: the convergence and numerical stability of these methods seems to be limited
97: to empirical observations. The methods based on series expansions fall into
98: two categories: methods based on the
99: evaluations of determinants \cite{biswas,fernandez} and methods based on
100: boundary conditions at large but finite values of the position
101: \cite{kbeck81,bacus}. The main goal of this article is to provide
102: a systematic discussion of the errors associated with this second category
103: of methods and to show how to make these errors
104: arbitrarily small in the most
105: efficient way. With the exception of Section \ref{sec:multi}, we
106: only consider one-dimensional problems. We discuss two types of errors.
107: First, the numerical errors made in calculating the energy which makes the
108: wave function vanish at some large value of the position $x_{max}$.
109: Second, the intrinsic error due to the finiteness of $x_{max}$.
110:
111: The basic elements the numerical method used hereafter
112: were sketched in Ref.\cite{bacus}
113: and applied to the quartic anharmonic oscillator.
114: We wrote the logarithmic
115: derivative of the wave function which appears in the Riccati equation
116: as $L/K$ and showed that these functions were entire.
117: The values of
118: the first ten eigenvalues with 30 significant digits
119: provided for a particular coupling
120: have been used to test new theoretical methods\cite{antonsen}.
121: Two issues were left open in this formulation:
122: first, the basic equations
123: had an interesting invariance which was not undestood but could
124: be used to improve the numerical efficiency; second,
125: the use of the method for
126: parity non-invariant potentials appeared to be unduly complicated \cite{oktay}.
127:
128: In Section \ref{sec:basic},
129: we present a new formulation where these two
130: issues are settled.
131: The basic equations presented
132: depend on an arbitrary {\it function}
133: denoted $G(x)$.
134: This freedom
135: can be interpreted as a local gauge invariance associated with
136: the fact that only $L/K$ is physical.
137: The wave function is invariant under
138: these local transformations.
139: In section \ref{sec:sol}, we show
140: how to construct power series for $L$ and $K$.
141: The complications in the case of parity non-invariant potentials (such as
142: asymmetric double-wells) are
143: minimal. When the potential and the gauge function are polynomials,
144: these series define {\it entire} function.
145: In other words, it is always possible to
146: construct arbitrarily accurate solutions of the Schr\"odinger equation
147: for arbitrary $E$ within a given range of the position variable,
148: by calculating enough terms in the expansions of $L$ and $K$. This allows
149: us to reproduce the asymptotic behavior of the wave function and determine
150: the energy eigenvalues. In section \ref{sec:sens}, we use the global properties
151: of the flows of the Riccati equation to recall of some basic results related
152: to the WKB approximation and the Sturm-Liouville theorem.
153: We explain how bifurcations in the asymptotic behavior of the functions
154: $K$ and $L$ can be exploited to determine the eigenvalues.
155:
156: It should be noted that the importance of reproducing the proper
157: asymptotic behavior has been emphasized in variational approaches
158: \cite{turbiner84}.
159: It should also be noted that Pad\'e approximants have been used in
160: conjunction with the Riccati equation in Ref. \cite{fernandez}, where
161: the quantization condition used
162: was that the approximants give one
163: additional coefficient in the Taylor expansion.
164: This procedure depends only on the coefficients of the expansions used and
165: there is no reference to any particular value of $x$
166: (as our $x_{max}$). Consequently, there is no
167: obvious connection between the two approaches.
168:
169: In the next two sections,
170: we show how to turn the gauge invariance to our advantage.
171: In Section \ref{sec:bif}, the quantitative aspects of
172: the bifurcation are discussed with an exponential parametrization
173: similar to the one
174: used to determine Lyapounov exponents in the study of chaotic dynamical
175: system. The exponents are $G$-dependent.
176: We provide an approximate way to determine the exponents and
177: the energy resolution.
178: We explain how our freedom in chosing $G$ can be used to
179: make the bifurcation more violent and improve the energy
180: resolution. However, the choice of $G$ also affects
181: the convergence of $L$ and $K$ and consequently the numerical
182: accuracy of the solution of the Schr\"odinger equation.
183: In Section \ref{sec:opt},
184: we show in a particular example that for an expansion of $L$ and $K$ at
185: a given order, a judicious choice of gauge can improve tremendously
186: the numerical accuracy of an energy level. We
187: discuss the two principles which allow to make optimal choices of $G$
188: and provide practical methods to determine approximately this optimal choice
189: for the general case.
190: We use these methods to explain some empirical results found in \cite{kbeck81}.
191:
192: In Section \ref{sec:app}, we discuss the
193: the error $\delta E$ on the energy levels due to the finiteness of $x_{max}$.
194: We propose two approximate formulas valid, respectively, for intermediate
195: and large values of $x_{max}$
196: and compatible in overlapping ranges.
197: Note that one can reinterpret the condition
198: that the wavefunction vanishes at $x_{max}$ as coming from
199: a slightly different problem where the potential becomes infinite
200: at $x_{max}$.
201: In the path-integral formulation (which can be extended immediately
202: to field theory problems), the fact that the potential becomes infinite
203: at $x_{max}$ means that
204: paths with values of $x$ larger than $x_{max}$ are not taken into account.
205: It has been argued \cite{pernice98,convpert}
206: that these configurations
207: are responsible for the asymptotic behavior of the
208: regular perturbative series. In Ref. \cite{convpert}, we showed that
209: the perturbative series of several modified problem were convergent.
210: The error formula sets the accuracy limitations
211: of this approach. Some of the methods used in this section could be used
212: for quantum field theory problems.
213:
214: The anharmonic oscillator can be considered as a
215: field theory with one time and zero space dimensions.
216: It can be used to test
217: approximate methods such as perturbative expansions or
218: semi-classical procedures. An illustrative
219: example is given in Ref. \cite{jentschura}
220: where multi-instanton effects were considered and where the splitting of
221: the two lowest levels of a double-well problem were estimated with more
222: than hundred digits. In Section \ref{sec:chall}, we show that our method
223: can be used to reproduce all these digits. Finally, we discuss the
224: generalization of the method to problems with several variables
225: in Section \ref{sec:multi}. For these problems, our ability to reduce
226: the degree of expansion by using optimal gauge functions may be crucial.
227:
228: \section{Basic equations and their gauge-invariance}
229: \label{sec:basic}
230:
231: We consider a one-dimensional, time-independent
232: Schr\"odinger equation $H\Psi=E\Psi$, for an Hamiltonian
233: \begin{equation}
234: H={p^2\over{2m}}+\sum_{l=1}^{2l}V_jx^j\ .
235: \label{eq:ham}
236: \end{equation}
237: As is well-known,
238: one can reexpress the wave function in terms of its logarithmic
239: derivative
240: \begin{equation}
241: \Psi(x)\propto{\rm e}^{-{1\over \hbar}\int_{x_0}^{x}dy \phi(y)\ ,}
242: \label{eq:repa}
243: \end{equation}
244: and obtain the Riccati form of the equation:
245: \begin{equation}
246: \hbar \phi '=\phi^2+2m(E-V)\ .
247: \label{eq:ric}
248: \end{equation}
249: It is assumed that $m>0$ and that the leading power
250: of $V$ is even with a positive
251: coefficient ($V_{2l}>0$).
252:
253: Writing $\phi=L/K$, we obtain a solution of Eq. (\ref{eq:ric}) provided
254: that we solve the system of equations:
255: \begin{eqnarray}
256: \label{eq:basic1}
257: \hbar L'&+&2m(V-E)K+GL=0 \\
258: \label{eq:basic2}
259: \hbar K'&+&L+GK=0
260: \end{eqnarray}
261: where $G(x)$ is an unspecified function.
262: This can be seen by multiplying (\ref{eq:basic1}) by
263: $K$, (\ref{eq:basic2}) by $L$ and
264: eliminating $GKL$ by taking the difference.
265: One then obtains the Riccati equation (\ref{eq:ric})
266: multiplied by $K^2$. Near a zero of $K$, one can check that Eqs.
267: (\ref{eq:basic1}-\ref{eq:basic2}) remain valid,
268: namely they impose that $\phi$ has a simple pole
269: with residue $-\hbar$. This allows the wave function to
270: become zero and change
271: sign as the contour goes around the pole on either side.
272:
273: Eqs. (\ref{eq:basic1}-\ref{eq:basic2})
274: are invariant under the {\it local} transformation
275: \begin{eqnarray}
276: \nonumber
277: L(x)&\rightarrow & Q(x)L(x) \\
278: K(x)&\rightarrow & Q(x)K(x) \\
279: \nonumber
280: G(x)&\rightarrow & G(x)-\hbar Q'(x)/Q(x) \ ,
281: \label{eq:gt}
282: \end{eqnarray}
283: where $Q(x)$ is an arbitrary function.
284: It is clear that this transformation leaves $\phi$ and the
285: wave function unchanged.
286: If we choose $G=0$ and eliminate $L$ using Eq.
287: (\ref{eq:basic2}), we recover the Schr\"odinger equation for $K$.
288: Starting from this gauge and making an arbitrary transformation, we find
289: that in general
290: \begin{equation}
291: K(x)\propto\Psi(x){\rm e}^{-{1\over \hbar}\int_{x_0}^{x}dy G(y)}
292: \label{eq:kgen}
293: \end{equation}
294: This shows that when $G$ is polynomial,
295: $K$ is simply $\Psi$ multiplied
296: by an entire function {\it with no zeroes} \cite{knopp}.
297: This means that the zeroes of
298: $K$ and $\Psi$ are identical. In other words, there are no spurious zeroes
299: when $G$ is polynomial.
300:
301: By taking the derivative of Eqs. (\ref{eq:basic1}) and (\ref{eq:basic2})
302: and choosing $G(x)$
303: appropriately,
304: one can obtain the basic Equations used in \cite{bacus}.
305: The explicit form of $G(x)$
306: is reached by comparing the two sets of equations and integrating
307: one of the differences. The two possibilities are compatible. The resulting
308: integral expression can be worked out easily by the interested reader.
309: The only important point is that the $G$ found that way is in general
310: not polynomial, justifying the spurious zeroes found with the original
311: formulation.
312:
313: \section{Solutions in terms of entire functions}
314: \label{sec:sol}
315:
316: The function $G$ can be chosen at our convenience.
317: For instance, we could impose the condition $K=1$ by
318: taking $G=-L$ and recover
319: the Riccati equation for $L$. However, the main advantage of
320: Eqs. (\ref{eq:basic1}-\ref{eq:basic2}) is that they are linear
321: first order differential equations with variables coefficients.
322: It is well-known \cite{coddington} that if we consider these equations for
323: complex $x$, the solutions inherit the domain of analyticity of the
324: coefficients (provided that this domain is simply connected).
325: If the coefficients are entire functions,
326: there exists a unique entire solution corresponding to a particular
327: set of initial values. In the following, we restrict ourselves to
328: the case where $V$ and $G$ are
329: polynomials.
330:
331: One can construct the unique solution corresponding to a particular choice
332: of initial values $L(0)$ and $K(0)$ by series expansions.
333: Using $K(x)=\sum_{n=0}^{\infty}K_nx^n$ and similar notations
334: for the other functions, one obtains the simple recursion
335: \begin{eqnarray}
336: \nonumber
337: L_{n+1}&=&{-1\over{\hbar (n+1)}}(\sum_{l+p=n}(2mV_lK_p+L_lG_p)
338: -2mEK_n)
339: \\
340: K_{n+1}&=&{-1\over{\hbar (n+1)}}(L_n+\sum_{l+p=n}K_lG_p)
341: \label{eq:iter}
342: \end{eqnarray}
343: Given $L_0$ and $K_0$, these equations allow to determine
344: all the other coefficients.
345: For potentials which are parity invariant, and
346: if $G$ is
347: an odd function, $L$ and $K$ can be assigned definite and opposite
348: parities. In this case, we can impose the initial conditions $K_0=1$ and
349: $L_0=0$ for even wave functions and $K_0=0$ and
350: $L_0=1$ for odd wave functions. If the Hamiltonian has no special symmetry,
351: as for instance in the case of an asymmetric double-well,
352: one could leave $L_0$
353: indeterminate and fix it at the same time as $E$
354: using conditions on the wave function or its
355: derivative at two different points. These two conditions translate
356: (in good approximation) into
357: two polynomial equations in $L_0$ and $E$ and can be solved by Newton's
358: method.
359:
360: The fact that Eqs. (\ref{eq:iter}) determines entire functions provided that
361: $V$ and $G$ are polynomials can be inferred directly from the fact that
362: the coefficients will decrease as $(n!)^{-\kappa }$ for some positive
363: power $\kappa$ to be determined and in general
364: depending on the choice of $G$.
365: As we will explain in more detail in Section \ref{sec:sens},
366: if the leading term in $V$ is $V_{2l}x^{2l}$,
367: one expects from Eq. (\ref{eq:ric}) that for $x$ large enough,
368: \begin{equation}
369: \phi(x)\simeq\pm \sqrt{2mV_{2l}}x^{l}\ ,
370: \label{eq:aswkb}
371: \end{equation}
372: and asymptotically,
373: \begin{equation}
374: \Psi(x)\propto {\rm e}^{- {\pm 1\over{(l+1)\hbar}} \sqrt{2mV_{2l}}x^{l+1}}\ .
375: \end{equation}
376: Looking at the general expression for $K$ given in Eq. (\ref{eq:kgen}), one
377: sees that $K$ will have the same asymptotic behavior provided that the
378: integral of $G$ grows not faster than $x^{l+1}$. If this is the case,
379: then $\kappa =1/(l+1)$. This behavior is well observed in empirical series.
380:
381: Note that if $G$ grows faster than $x^l$, the coefficients decay more slowly
382: and the procedure seem to be less efficient.
383: In the following, we will mostly discuss the case $l=2$. If we require that
384: $G$ is an odd polynomial growing not faster than $x^2$, this means that
385: $G$ is homogeneous of degree 1.
386:
387: \section{Quantization from global flow properties}
388: \label{sec:sens}
389: In this section, we use
390: the global properties of the flows associated with the Riccati equation
391: to rephrase some implications of Sturm-Liouville
392: theorem and to justify the asymptotic behavior given in Eq. (\ref{eq:aswkb}).
393: The main goal of this section is to provide a simple
394: and intuitive picture of the
395: bifurcation which occurs when the value of $E$ is varied by a small amount
396: above or below an energy eigenvalue. The main results of this section
397: are summarized
398: in Figs. \ref{fig:bif1} and \ref{fig:kas}.
399:
400: We consider the solutions of Eq. (\ref{eq:ric})
401: obtained by varying $E$ with fixed initial values. It is
402: convenient
403: to introduce an additional parameter $s$ and to rewrite
404: the original equation as a 2-dimensional ODE with an $s$-independent
405: r.h.s .
406: \begin{eqnarray}
407: \label{eq:ode1}
408: \hbar \dot{\phi}&=&\phi^2+2m(E-V(x))\ \\
409: \label{eq:ode2}
410: \dot{x}&=&1\ ,
411: \end{eqnarray}
412: where the dot denotes the derivative with respect to $s$.
413:
414: The flows in the
415: $(x,\phi)$ plane have some simple global properties that we now proceed to
416: describe. We consider a solution (phase curve) with
417: initial condition $x=x_0$ and
418: $\phi=\phi_0$ at $s=0$.
419: We assume that for these values the r.h.s of Eq. (\ref{eq:ode1}) is $>0$.
420: It will become clear later that if such a choice is impossible, a
421: normalizable wave function cannot be constructed.
422: With this assumption, the phase curve starts moving up and right as
423: $s$ increases,
424: possibly going through simple poles with residues $-h$. This situation persists
425: unless the r.h.s. of (\ref{eq:ode1}) becomes zero.
426: We call the separating curves defined by a
427: zero for the r.h.s of Eq. (\ref{eq:ode1}),
428: $\phi=\pm \sqrt{2m(V(x)-E})$, ``$WKB$ curves''. After a phase curve crosses
429: (horizontally) a $WKB$ curve, it moves right and down. If it crosses the
430: $WKB$ curve again, we can repeat the discussion as at the beginning.
431:
432: At some point, we reach the ``last'' $WKB$ curve
433: (i.e., the farthest right). For $x$ large enough, the potential is dominated
434: by its largest power and
435: the upper (lower) part of this last
436: WKB curve has a strictly positive (negative) slope.
437: For such values of $x$, if a phase curve crosses the WKB curve,
438: it will do so
439: horizontally and move {\it inside} the region where the r.h.s. of
440: Eq. (\ref{eq:ode1}) is negative. As $s$ further increases, $\phi$ decreases,
441: but the phase curve cannot cross horizontally
442: the lower part of the WKB curve which has a strictly negative slope.
443: In the same region, if $\phi$
444: has a pole, the curve
445: reappears below the lower part of the WKB curve and will
446: never take positive values again.
447:
448: In summary, if in the region described above,
449: a phase curve crosses the WKB curve or develops a pole, then it cannot
450: develop a pole again. The other logical possibility is that the phase
451: curve does none of the above.
452: It is thus clear that for fixed $E$, we can always find
453: a $X$ such that if $x>X$, $\phi(x)$ has no pole. Consequently the two
454: terms involving $\phi$ in Eq. (\ref{eq:ric}) cannot grow faster than $2m(E-V)$.
455: Otherwise, $2m(E-V)$ would become negligible and a pole would be necessary.
456: At least one of these two terms needs to match $2m(E-V)$. Inspection of the
457: two possibilities leads to Eq. (\ref{eq:aswkb}). Only the positive solution
458: which follows asymptotically the upper WKB curve
459: leads to a normalizable wave function.
460:
461: If we compare two phase curves with identical initial conditions
462: but different $E$, the one with larger $E$ initially lays
463: above the other one. If the one with lower $E$ has a first pole at $x_1$,
464: then the one with larger $E$ has a first pole at some $x<x_1$. Remembering
465: that the poles of $\phi$ produce zeroes of $\Psi$, this
466: rephrases the main
467: idea behind the Sturm-Liouville theorem.
468: An exact energy eigenstate $E_n$
469: is obtained when the wave function has its last zero
470: at infinity. When $E$ is fine-tuned to that value, $\phi$ follows
471: closely the upper branch of the WKB curve. This trajectory in unstable under
472: small changes in $E$.
473: If the energy is slightly increased with respect to
474: $E_n$, $\phi$ develops a
475: pole and reappears on the lower part of the WKB curve.
476: If the energy is slightly decreased with respect to $E_n$, $\phi$ crosses
477: the upper part of the WKB curve and reaches the lower part of the WKB curve.
478: This is illustrated in
479: Fig. \ref{fig:bif1} in the case of the ground state of the quartic single-well
480: anharmonic oscillator with
481: $ m=1/2,\ \hbar=1, V_2=1$ and $V_4=0.1$. All the figures in this section and
482: the next two sections have been done with this particular example.
483: \begin{figure}
484: \centerline{\psfig{figure=bifu2.EPS,width=3.in}}
485: \caption{Bifurcations of $\phi(x)$ from the upper part of the WKB curve
486: associated with the ground state energy $E_0$ for
487: energies $E_0\pm 10^{-5}$, $E_0\pm 10^{-10}$, $E_0\pm 10^{-15}$,
488: $E_0\pm 10^{-20}$ and $E_0\pm 10^{-25}$ (from left to right).}
489: \label{fig:bif1}
490: \end{figure}
491: The sensitive dependence on $E$ is also present in the asymptotic
492: behavior of $K$.
493: If the energy is slightly increased with respect to $E_n$, $K$
494: reaches zero at a finite value of $x$.
495: If the energy is slightly decreased with respect to $E_n$, $K$ increases
496: rapidly. This is illustrated in Fig. \ref{fig:kas} for the same example as
497: in Fig. \ref{fig:bif1}.
498: \begin{figure}
499: \centerline{\psfig{figure=bifk3.EPS,width=3.in}}
500: \caption{Bifurcations of $K(x)$ from its trajectory for $E=E_0$.
501: The changes in $E$ are $\pm 10^{-30}$, $\pm 2\times 10^{-30}, \dots$,
502: $\pm 10^{-29}$ }
503: \label{fig:kas}
504: \end{figure}
505: We now discuss the initial value $\phi_0$.
506: For parity invariant potentials,
507: one only needs to consider the cases $\phi_0=0$ (even $\Psi$) or $\phi_0=
508: -\infty$ (odd $\Psi$) at $x_0=0$. For potentials with no reflection symmetry,
509: one needs to
510: insure that the appropriate behavior is reached when $x\rightarrow -\infty$.
511: This can be implemented in good approximation
512: by requiring that the wave function
513: has also a zero at some large negative value $x_{min}$.
514: For potentials with a reflection symmetry about another point $x_1$ than the
515: origin, one can impose that the wave function ($K(x_1)=0$) or its derivative
516: ($L(x_1)=0$) vanish at that point.
517: In all cases, we have an independent condition
518: which allows to determine $\phi_0$.
519:
520: In summary, for $x_{max}$ large enough, the condition
521: \begin{equation}
522: K(x_{max})=0
523: \end{equation}
524: provides sharp upper bound on the energy levels. The lower part of
525: Fig. \ref{fig:kas} makes clear that as $x_{max}$ increases, sharper
526: bounds are reached. For potentials that are not parity invariant, an
527: additional condition has to be imposed. In all cases, one obtains
528: polynomial equations which can be solved for the energy levels given
529: the potential or vice-versa using Newton's method. Note also that
530: a sharp lower bound can be found by solving $L(x_{max})=0$. The
531: fact that in Fig. \ref{fig:kas}, a zero of $K$ at $E_0+\delta$
532: corresponds to a zero of $L$ at $E_0-\delta$, suggests
533: that the exact value should be very close to the average of the two bounds.
534:
535:
536: \section{$G$-dependence of the bifurcation}
537: \label{sec:bif}
538:
539: The strength of the
540: bifurcation in $K$ illustrated in Fig. \ref{fig:kas}
541: can be approximately characterized by local exponents.
542: If we consider the departure $\delta K(x)$ from $K(x)$ calculated at some
543: exact energy level $E_n$, we expect the approximate behavior:
544: \begin{equation}
545: \delta K(x) \simeq C (E-E_n){\rm e}^{xB}\ .
546: \label{eq:exp}
547: \end{equation}
548: In other words ln($|\delta K(x)|$) is linear with a slope $B$
549: independent of the choice of $E$ and an intercept that varies like
550: ln$(|E-E_n|)$. This situation is approximately realized in the
551: example considered before as shown in Fig. \ref{fig:fitexp}.
552: We have checked in the same example that the sign of the energy difference
553: plays no role. In other words, the same values of $C$ and $B$ can be
554: used above and below $E_n$.
555: \begin{figure}
556: \centerline{\psfig{figure=lndelk.EPS,width=3.in}}
557: \caption{Natural logarithm of $\delta K(x)$ for $E-E_0=10^{-30}$ (lower set
558: of point) and $E-E_0=10^{-28}$ (upper set
559: of point). Lines are linear fits. }
560: \label{fig:fitexp}
561: \end{figure}
562: The exponent $B$ is not uniform.
563: It increases with $x$ and is $G-$dependent
564: as shown in Fig. \ref{fig:expdep}. The local values of $B$ have been
565: calculated by fits in regions of width 0.2 with central value displayed in
566: the horizontal label of Fig. \ref{fig:expdep}.
567: \begin{figure}
568: \centerline{\psfig{figure=lyagl4.EPS,width=3.in}}
569: \caption{Value of $B$ for various $x$
570: and for $G=-3x$ (empty hexagons), $G=-2x$ (filled squares) $G=-x$ (crosses)
571: and $G=0$ (triangles). The continuous lines have been drawn using
572: Eq. (\ref{eq:bap}).}
573: \label{fig:expdep}
574: \end{figure}
575: The change in $B$ can be understood as follows. If $E$ is
576: changed from $E_n$ to $E_n+\delta E$, then at some point we have a sudden
577: transition from the upper to the lower WKB curve and
578: asymptotically
579: \begin{equation}
580: \delta \Psi (x)\propto \delta E \
581: {\rm e}^{+{1\over{(l+1)\hbar}} \sqrt{2mV_{2l}}x^{l+1}}\ .
582: \end{equation}
583: Using Eq. (\ref{eq:kgen}) and expanding about $x_{max}$, we obtain that,
584: in good approximation,
585: \begin{equation}
586: B\simeq {1\over \hbar}\bigl(\sqrt{2mV_{2l}}x_{max}^{l}-G(x_{max})\bigr)\ .
587: \label{eq:bap}
588: \end{equation}
589: As shown in Fig. \ref{fig:expdep},
590: this simple expression provides reasonable estimates of $B$. The
591: slight underestimation comes in part
592: from the fact that Eq. (\ref{eq:bap}) does not take into account the
593: harmonic term in $V$.
594: Eq. (\ref{eq:bap}) shows that we can increase the strength of the
595: bifurcation near $x_{max}$
596: by increasing $x_{max}$ or $-G(x_{max})$.
597: This allows us to ``resolve'' the energy more accurately. However, at the
598: same time our numerical resolution of $K(x_{max})$ is affected
599: and we need to take this effect into account.
600: This question is treated in
601: the next Section. In general, if we can establish that $K(x_{max})$
602: at an energy $E$ very close to $E_n$,
603: can be calculated with some numerical accuracy $\delta K^{num.}$, we
604: have the approximate numerical energy resolution
605: \begin{equation}
606: \delta E^{num.}\propto \delta K^{num.}
607: {\rm e}^{+{1\over {\hbar}}
608: \bigl({-1\over{l+1}}\sqrt{2mV_{2l}}x_{max}^{l+1}+\int^{x_{max}}_0
609: dxG(x)\bigr)}\ .
610: \label{eq:eresol}
611: \end{equation}
612:
613: \section{An optimal choice of $G$}
614: \label{sec:opt}
615:
616: In this Section, we show that
617: from a numerical point of view, the choice of $G$ is important.
618: We discuss the question of an optimal choice, first with an example
619: and then in general.
620: We start with the calculation
621: of the ground state in the case
622: $ m=1/2,\ \hbar=1, V_2=1$ and $V_4=0.1$.
623: We discuss the estimation of the ground
624: state energy
625: using the equation $K(x_{max})=0$ with $x_{max}=6$. The fact that we
626: use this finite value for $x_{max}$ creates an error in the 25-th digit
627: (see Section \ref{sec:app}).
628:
629: From the discussion of Section \ref{sec:sol}, it is reasonable to
630: limit the discussion to a gauge function of the form
631: \begin{equation}
632: G(x)=-ax \ ,
633: \label{eq:gchoice}
634: \end{equation}
635: which using Eq. (\ref{eq:kgen}) implies that
636: \begin{equation}
637: K(x)\propto \Psi(x){\rm e}^{{1\over{2\hbar}}a x^2}\ .
638: \label{eq:kpart}
639: \end{equation}
640: With this restriction, the optimization problem is reduced to the
641: determination of $a$. As $a$ increases through
642: positive values, the features
643: of $\Psi$ are exponentially amplified, making the bifurcation
644: displayed in Fig. \ref{fig:kas} more violent.
645: Ideally, we would like to take $a$ as large as possible. However, if $a$ is
646: too large, we may need too many coefficients $K_n$ to get a good
647: approximation. If we consider
648: the problem at a given order, the two requirements of sensitivity and
649: accuracy result in a compromise which determines the optimal value of $a$.
650:
651: As explained in Section \ref{sec:sol},
652: the choice of Eq. (\ref{eq:gchoice}) guarantees a suppression
653: of the form $(n!)^{-1\over{3}}$ for the coefficients of $L$ and $K$.
654: However, the choice of $a$ still affects significantly the behavior
655: of these coefficients as shown in Fig. \ref{fig:k6}.
656: \begin{figure}
657: \centerline{\psfig{figure=coeff2.EPS,width=3.in}}
658: \caption{{\rm ln}($|K_n6^n|$) versus $n$,
659: for $G=0$ (triangles), $G=-x$ (filled squares), $G=-2x$ (crosses) and
660: $G=-3x$ (empty squares). }
661: \label{fig:k6}
662: \end{figure}
663: The quantity $K_nx_{max}^n$ is relevant to decide at which order we need to
664: truncate the series in order to get a good estimate of $K(x_{max})$.
665: For instance, if we require knowing
666: $K(x_{max})$ with errors of order 1, we
667: need about 100 coefficients for $a=2$ but more than 150 for $a=0$. The
668: corresponding values
669: for $a=1$ and 3 fall between these two values, indicating that $a=2$ is close
670: to optimal.
671: This estimate is confirmed by an analysis of the dependence of $K_n$ on $a$.
672: Sample values are shown in Fig. \ref{fig:lncoe}. We observe rapid
673: oscillations (that we will not attempt to explain) and
674: slowly varying amplitudes which have a minimum slightly below 2.
675: Note that on the logarithmic scale of Fig. \ref{fig:lncoe}, the
676: zeroes of $K_n$ give $-\infty$, however due to the discrete sampling of $a$,
677: it just generates isolated dots on the graphs. Note also that in Figs.
678: \ref{fig:k6} and \ref{fig:lncoe}, the coefficients have been calculated for an
679: an accurate value of the ground state energy.
680: \begin{figure}
681: \centerline{\psfig{figure=lncoe.EPS,width=3.in}}
682: \caption{{\rm ln}($|K_{n}|$) versus $a$, for $n$=60 (upper set),
683: 70 (next set), 80 (next set) and 90 (lower set). }
684: \label{fig:lncoe}
685: \end{figure}
686: The behavior of the $K_n$ calculated at
687: value of $E$ sufficiently close to an eigenvalue,
688: can be understood by using the asymptotic form
689: \begin{equation}
690: K(x_{max})\propto
691: {\rm e}^{{1\over \hbar}\bigl(-{1\over{l+1}} \sqrt{2mV_{2l}}x_{max}^{l+1}-
692: \int^{x_{max}}_0dxG(x)\bigr) }\ .
693: \label{eq:kasy}
694: \end{equation}
695: We emphasize that the relative sign between the two terms
696: in the exponential is opposite than
697: in Eq. (\ref{eq:eresol}), because we are now on the upper WKB curve.
698: For $a=0$, Eq. (\ref{eq:kasy}) provides a rough estimate of $K_nx_{max}^n$.
699: Remembering the minus sign in the parametrization of $G$
700: (Eq. (\ref{eq:gchoice})), we see that if
701: $a$ is given a small positive value, the argument of the exponential
702: in Eq. (\ref{eq:kasy}) decreases and we can obtain comparable accuracy
703: with less terms in the expansion. Naively, our optimum choice is
704: obtained when the two terms in the exponential cancel.
705: In the general case, this amounts to having
706: \begin{equation}
707: \sqrt{2mV_{2l}}x_{max}^{l+1}\simeq -(l+1)\int^{x_{max}}_0
708: dxG(x) \ .
709: \label{eq:optim}
710: \end{equation}
711: For the particular example considered here, this cancellation is
712: obtained for $a=(2/3)\sqrt{0.1}x_{max}\simeq 1.27$. It is clear that
713: when the two terms cancel, subleading terms neglected in Eq. (\ref{eq:aswkb})
714: should be taken into account. However, in several examples, we found that
715: this simple procedure gives results close to what is found empirically.
716:
717: We now address the more general question of determining
718: the $G$-dependence of the
719: number of
720: significant digits that can be obtained from the condition
721: $K(x_{max})=0$ using an expansion of $K$
722: truncated at a given order. For the example considered before in this section,
723: we see from Fig. \ref{fig:sd} that, for instance
724: for a truncation at order 100, the
725: most accurate answer is obtained for $a\simeq 1.6$.
726: It is worth noting that for this value of $a$, one gains more than 15
727: significant digits compared to the $G=0$ case!
728: This figure also indicates,
729: that as expected,
730: the best possible answer (in the present case, 25 significant digits)
731: can always be achieved by calculating enough coefficients.
732: \begin{figure}
733: \centerline{\psfig{figure=gaugedep4.EPS,width=3.in}}
734: \caption{Number of significant digits for $E_0$ versus $a$ using the
735: condition $K(6)=0$ with expansions of order 50 (empty diamonds), 75 (filled
736: squares), 100 (crosses), 125 (empty squares) and 150 (stars).}
737: \label{fig:sd}
738: \end{figure}
739: Using Eq.
740: (\ref{eq:eresol}) and Fig. (\ref{fig:lncoe}), we were able to reproduce
741: approximately the left part of Fig. \ref{fig:sd} ($0<a<1$). To give a
742: specific example, at order 100, when one changes $a$ from 0 to 1,
743: $\delta K^{num}$ becomes 4 orders of magnitude smaller and the factor
744: ${\rm e}^{-({a\over{2\hbar}})x_{max}^2}$ improves the resolution by
745: almost 8 orders of
746: magnitude. This approximately accounts for the gain of 11 significant
747: digits observed in Fig. \ref{fig:sd}. A detailed understanding of the
748: figure in the region $1<a<2$ is beyond what can be accomplished using
749: the asymptotic form of the wave function. However, the naive estimate
750: of Eq. (\ref{eq:optim}) provides a reasonable estimate of the location
751: of the optimal $a$.
752:
753: It should be noted that an ansatz of the form of Eq. (\ref{eq:kpart})
754: with $a=1$ has been used in Ref. \cite{biswas} and that the fact that
755: varying $a$ could improve the numerical efficiency was found empirically
756: in Ref. \cite{kbeck81}. Eq. (\ref{eq:optim}) can be used to understand
757: these results. For instance, for $H=p^2+x^2+x^8$, we can obtain a very
758: accurate result with $x_{max}=2.8$ (see Section \ref{sec:app}).
759: According to Eq. (\ref{eq:optim}) the optimal value of $a$ in this case is
760: $a=(2/5)x_{max}^3\simeq 8.8$ which is slightly below the value $(\approx 10)$
761: suggested in \cite{kbeck81}. Note also that equivalently good results
762: can be obtained using $G=-bx^3$.
763: \section{Approximate error formulas}
764: \label{sec:app}
765: In this Section,we discuss the intrinsic error
766: $\delta E =E(x_{max})-E(\infty)$
767: where $E(x_{max})$ is defined by $\psi(x_{max},E(x_{max}))=0$,
768: for a given energy level. We emphasize that $\delta E$ is the error due to
769: the finiteness of $x_{max}$ independently of practical considerations
770: regarding the numerical estimation of $E(x_{max})$ which is assumed to be
771: known with an error much smaller than $\delta E$ in this section.
772: We use the familiar parametrization of the
773: quadratic term of the potential, $V_2={1\over 2}m\omega^2$ and we restore the
774: dependence on $\hbar$ and $m$.
775: The error for the ground state of the harmonic oscillator has been
776: estimated in Eq. (4) of Ref. \cite{convpert}.
777: Using the asymptotic form of the integral in this equation, we obtain
778: \begin{equation}
779: \delta E_0^{harm.} \simeq 2 \bigl({S_0\over{\pi \hbar}}\bigr)^{1\over2}{\rm e}^
780: {-{S_0/{\hbar}}}\ ,
781: \label{eq:harmerr}
782: \end{equation}
783: with
784: \begin{equation}
785: S_0=\int_{-\infty}^{+\infty}dt{1\over 2}m((\dot{x}_c(t))^2+\omega^2(x_c(t))^2)
786: =m\omega x_{max}^2
787: \end{equation}
788: and $x_c(t)=x_{max}{\rm e}^{-\omega |t-t_0|}$.
789: This corresponds to semi-classical approximation where the contribution of
790: the large field configurations
791: are obtained by calculating the quadratic fluctuations with respect
792: to $x_c(t)$.
793: The anharmonic corrections can
794: be approximated to lowest order in the the anharmonic couplings
795: by adding a term $S_{anh}$ to $S_0$ in the exponent of Eq. (\ref{eq:harmerr})
796: with
797: \begin{equation}
798: S_{anh}=\int_{-\infty}^{+\infty} dt V_{anh}(x_c(t))\ ,
799: \end{equation}
800: and $V_{anh}$ is the anharmonic part of the potential.
801: Our final perturbative estimate is thus
802: \begin{equation}
803: \delta E_0 \simeq \delta E_0^{harm.}
804: {\rm e}^{-\sum_{j=2}^{l}({1\over{j\hbar }})V_{2j} x_{max}^{2j}}
805: \label{eq:dele}
806: \end{equation}
807: This estimate is accurate if the $V_{2j}$ and $x_{max}$
808: are small enough. We expect that for the excited states, approximate formulas
809: of the form of Eq. (\ref{eq:dele}) multiplied by a polynomial should hold.
810:
811: When $\lambda$ or $x_{max}$ become too large, Eq. (\ref{eq:dele})
812: is not adequate. To obtain a better approximation, we use
813: \begin{equation}
814: {\partial\over{\partial x_{max}}}\psi(x_{max},E(x_{max}))=0\ ,
815: \end{equation}
816: and the asymptotic behavior of $\Psi$.
817: We estimate that $\partial \Psi/\partial E$ is of the
818: order of the non-normalizable WKB solution and as a consequence,
819: $\delta E$
820: has the asymptotic form
821: \begin{equation}
822: \delta E\simeq P(x_{max})(\psi(x_{max}))^2 \ ,
823: \end{equation}
824: where $P$ is a polynomial. This form is correct for the
825: ground state of the harmonic
826: oscillator.
827: In the case where the leading term of $V$ is $V_{2l}x^{2l}$,
828: this implies the asymptotic order of magnitude
829: estimate
830: \begin{equation}
831: \delta E\approx {\rm e}^{-{2\over{(l+1)\hbar}}\sqrt{2mV_{2l}}x_{max}^{l+1}}\ ,
832: \label{eq:cufit}
833: \end{equation}
834:
835: We have tested the two approximate errors formulas given
836: above (Eqs. (\ref{eq:dele}) and (\ref{eq:cufit})) for
837: the ground state corresponding to $V_{anh}=\lambda x^4$.
838: We used the numerical values $\hbar=m=\omega=1$ and $\lambda=0.1$
839: The results are shown in
840: Fig. \ref{fig:varx}.
841: We see that for small values of $x_{max}$, the perturbative estimate of
842: Eq. (\ref{eq:dele}) corrects properly the harmonic result.
843: However when $x_{max}$ increases, the Eq. (\ref{eq:cufit}) gives better
844: results. If the left part of the graph is displayed with a log-log scale,
845: it is approximately linear with a slope close to 3. In Fig. (\ref{fig:varx}),
846: the proportionality constant not given by Eq. (\ref{eq:cufit}) has been
847: determined by fitting the 5 last data points on the right of the figure.
848: We conclude that by combining the two approximations it is possible to
849: get a reasonable estimate of the errors on $E$ over a wide range of $x_{max}$.
850: \begin{figure}
851: \centerline{\psfig{figure=xmaxerr3.EPS,width=3.in}}
852: \caption{{\rm ln}( $\delta E_0$) as a function of $x_{max}$
853: for $\lambda$=0.1 (black dots).
854: %(lower set of points).
855: The continuous lines are
856: from top to bottom on the left of the figure:
857: the harmonic case (Eq. (\ref{eq:harmerr})),
858: Eq. (\ref{eq:dele}) with $V_4=0.1$ (fits the dots well on the left of the
859: figure), Eq. (\ref{eq:cufit}) (fits the dots well on the right of the
860: figure).}
861: \label{fig:varx}
862: \end{figure}
863: We have tested Eq. (\ref{eq:cufit}) for other potentials. For instance,
864: for $H=p^2+x^2+x^8$, in order to get 30 significant digit,
865: we estimated that $x_{max}\simeq 2.8$. We found that the difference
866: between the ground state energy found from the condition $K=0$ (upper
867: bound) and $L=0$ (lower bound) differed in the 30th significant digits.
868: \section{A challenging test}
869: \label{sec:chall}
870: The only practical limitation of the method proposed here is that
871: in some cases the relevant details of the potential appear in widely separated
872: regions, forcing us to calculate a huge number of coefficients with many
873: significant digits.
874: A simple example where such problem may occur is the symmetric double-well
875: with a small quartic coupling where the separation between the
876: wells goes like the inverse square root of the quartic coupling.
877:
878: In Ref. \cite{jentschura}, the lowest even and
879: odd energies were calculated for a potential with $m=1,\ \hbar=1,
880: \ V_2=-1/4,\ V_4=1/2000$ with 180 significant digits. Remarkably, the
881: authors were able to reproduce the 110 significant digits of the splitting
882: between these two states
883: by calculating instanton effects.
884: We have reproduced the 180 digits of both states using an expansion
885: of order 1700 for $K$ and a value of $x_{max}=46$. The calculations were
886: performed with 700 digit arithmetic. The calculation of one level with
887: such a procedure takes less than two hours
888: with MATHEMATICA on an unexpensive laptop using Pentium3.
889: The computation time increases
890: with the accuracy required. In order to fix the ideas,
891: it takes less than 2 minutes
892: minutes to reproduce the
893: first 120 digits in the above calculation.
894: \section{The multivariable case}
895: \label{sec:multi}
896:
897: The basic equations presented in Section \ref{sec:basic} can be extended
898: when the single variable $x$ is replaced by a $N$-dimensional vector $\vec{x}$.
899: In Eq. (\ref{eq:repa}), $\phi$ becomes a vector $\vec{ \phi}$ and the integral
900: a line integral. In order to guarantee that the wave function is independent
901: of the choice of the line, we require that the curl of $\vec{\phi}$ vanishes.
902: Eq. (\ref{eq:ric}) becomes:
903: \begin{equation}
904: \hbar \vec{\nabla} \vec{\phi} =\vec{\phi}\cdot\vec{\phi}+2m(E-V)\ .
905: \label{eq:ricmany}
906: \end{equation}
907: Using $\vec{ \phi}=\vec{ L}/K$, we write as previously
908: \begin{eqnarray}
909: \hbar \vec{\nabla}\vec{ L}&+&2m(V-E)K+\vec{G}\cdot\vec{L}=0
910: \label{eq:v1} \\
911: \hbar \vec{\nabla} K&+&\vec{L}+\vec{G}K=0 \ ,
912: \label{eq:v2}
913: \end{eqnarray}
914: with $\vec{G}(\vec{x})$ unspecified at this point.
915: These equations imply the multivariable Riccati equation (\ref{eq:ricmany})
916: multiplied by $K^2$. Near a zero of $K$, these equations imply the
917: same singularity as Eq. (\ref{eq:ricmany}).
918: After using Eq. (\ref{eq:v2}), the condition that
919: $\phi$ has no curl reads
920: \begin{equation}
921: \nabla_i L_{(j)}+G_{(i)}L_{(j)}=\nabla_jL_{i}+G_{(j)}L_{(i)}\ .
922: \end{equation}
923: The parenthesis for the vector indices are used in order to distinguish
924: these indices from the order in a power series expansion used later.
925:
926: The transformation Eqs. (\ref{eq:gt}) can vectorized trivially with $Q$ treated
927: as a scalar. In the expression of $K$ given by Eq. (\ref{eq:kgen}),
928: the integral becomes a line integral and we require
929: that $\vec{G}(\vec{x})$ has a vanishing curl. This condition is also necessary
930: to establish that different derivatives acting on $K$ commute.
931:
932: The choice of coordinates to be used depends on the choice of boundary
933: conditions imposed. If we require $\Psi $ to vanish on a large
934: hypersphere, hyperspherical harmonics should be used. If we require $\Psi$
935: to vanish on hypercubes (as suggested for lattice problems in Ref.
936: \cite{convpert}) cartesian coordinates should be used. To fix the ideas,
937: let us consider the case of cartesian coordinates for two variables
938: $x_1$ and $x_2$ with boundary conditions on a rectangle.
939: We expand $K(x_1,x_2)=\sum_{m,n\geq 0}K_{m,n}x_1^mx_2^n$
940: and similar
941: expansions for the two components of $\vec{L}$.
942: The coefficients can be constructed order by order, with the order of
943: $K_{m,n}$ defined as $m+n$. The terms with one derivative yield the
944: higher order terms. For instance, for $K$,
945: we obtain equations providing $K_{m+1,n}$ and $K_{m,n+1}$ in terms
946: of coefficients of higher order just as in Eq. (\ref{eq:iter}).
947: A detailed construction shows that if
948: $V(x_1,x_2)$ has no special symmetry, we can determine all the
949: coefficient up to a given order $l$
950: provided that we supply the values of two coefficients at
951: each intermediate order (for instance $\vec{L}_{m,0}$ for $m\leq l$).
952: These coefficients together with $E$ are fixed by the boundary
953: conditions $K(x_{1min},x_2)=K(x_{1max},x_2)=K(x_1,x_{2min})=K(x_1,x_{2max})=0$.
954: Taking derivatives with respect to the free variables $x_1$ and $x_2$,
955: and setting these variables to 0, we
956: obtain an infinite set of conditions. The truncation of this set,
957: together with the truncation of the expansion in the other variable
958: must be studied carefully. If we consider the special case where the
959: problem can be solved by separation of variables, we see that it is
960: important to maintain a uniform accuracy for all the conditions.
961: If all the coefficients have been calculated up to order $l$, this can
962: achieved in the following way.
963: We retain of the order of $l/2$ derivatives of the four conditions in such
964: a way that we get exactly $2l+3$ conditions which can be expanded up to
965: an order close to
966: $l/2$ in the remaining variable. A practical implementation of this program
967: is in progress.
968:
969: \section{Conclusions}
970:
971: In conclusion, we have shown that accurate estimates of the energy levels
972: of arbitrary polynomial potentials
973: bounded from below can be obtained by solving
974: polynomial equations. The fact that the function $L$ and $K$ are entire
975: guarantees that if we calculate enough terms we will gain proper control
976: of the asymptotic behavior of the wave function. Reaching this goal is
977: in general a difficult task which often requires guesswork and analytical
978: continuations (see e.g., Ref. \cite{bender96}). Here, the convergence of
979: the procedure is guaranteed and the order at which we can terminate the
980: expansion in order to reach a given accuracy can be estimated.
981: In addition, a systematic understanding
982: and control of the errors due to the finite value of $x_{max}$ has been
983: achieved.
984:
985: The understanding of the gauge invariance of the basic equations proposed
986: here completely resolves the issues raised from our initial proposal
987: \cite{bacus}. By varying $G$, from 0 to -$\phi$, we can interpolate
988: between a situation where $K$ is the wave function to another situation
989: where $K=1$ and $L =\phi$. However, for every other choice of $G$, only
990: the ratio $L/K$ has a direct physical meaning.
991: By properly chosing $G$, we can at the same time improve the convergence
992: of $K$ and amplify the bifurcation toward the the non-normalizable
993: behavior.
994:
995: The extreme accuracy obtained for two widely separated wells indicates that
996: for reasonably complicated potential, the number of terms that
997: needs to be calculated is not prohibitive. We intend to use this method
998: to test analytical results regarding the role of large configurations
999: in the path-integral and to test semi-classical treatment of potentials
1000: with asymmetric wells \cite{coleman,zj}.
1001:
1002: The method can be extended in the case of several variables. It remains to be
1003: determined if the simultaneous solution of many polynomial equations
1004: can be accomplished with a reasonable accuracy. For these problems, the
1005: fact that a judicious choice of the arbitrary functions $\vec{G}$ allows
1006: to decrease the order of the expansions may be crucial.
1007:
1008: \begin{acknowledgments}
1009: We thank B. Oktay for communicating his work regarding the treatment of
1010: parity non invariant potentials with the method of Ref. \cite{bacus}.
1011: We thank P. Kleiber, W. Klink, L. Li,
1012: G. Payne, W. Polyzou, M.H. Reno and V.G.J. Rodgers for valuable conversations.
1013: We thank F. Fernandez for pointing out Ref. \cite{kbeck81} to us.
1014: This research was supported in part by the Department of Energy
1015: under Contract No. FG02-91ER40664,
1016: and in part by the CIFRE of the University of Iowa.
1017:
1018: \end{acknowledgments}
1019: %\bibliography{da}
1020: %\bibliographystyle{unsrt}
1021: %\bibliographystyle{prsty}
1022: %\end{multicols}
1023: %\end{document}
1024: \begin{thebibliography}{10}
1025: \bibitem{bender69}
1026: C. Bender and T.~T. Wu, Phys. Rev. {\bf 184}, 1231 (1969).
1027: \bibitem{wang96}
1028: W. Z. Wang, J. Tinka Gammel, A. Bishop and M. Salkola, Phys. Rev. Lett,
1029: {\bf 76}, 3598 (1996).
1030: \bibitem{leguillou90}
1031: J.~C. Le Guillou and J. Zinn-Justin, {\em Large-Order Behavior of Perturbation
1032: Theory} (North Holland, Amsterdam, 1990) ands Refs. therein.
1033: \bibitem{loeffel69}
1034: J. Loeffel, A. Martin, B. Simon, and A. Wightman, Phys. Lett. B {\bf 30}, 656
1035: (1969); B. Simon, Ann. of Phys. {\bf 58}, 76 (1970).
1036: \bibitem{brezin77}
1037: E. Brezin, G. Parisi and J. Zinn-Justin, Phys. Rev. D {\bf16}, 408 (1977).
1038: \bibitem{coleman}
1039: S. Coleman, {\em Aspects of Symmetry} (Cambridge University Press, Cambridge,
1040: 1985).
1041: \bibitem{zj}
1042: J. Zinn-Justin, {\em Quantum Field Theory and Critical Phenomena}, (Oxford,
1043: 1989).
1044: \bibitem{weninger96}
1045: E. Weniger, Phys. Rev. Lett,
1046: {\bf 77}, 2859 (1996);
1047: L. Skala, J. Cizek, E. Weniger and J. Zamastil, Phys. Rev. A
1048: {\bf 59} 102 (1999).
1049: \bibitem{kleinert}
1050: H. Kleinert, Phys. Lett. {A207}, 133 (1995).
1051: \bibitem{bazley}
1052: N. Bazley and D. Fox, Phys. Rev. {\bf 124}, 483 (1961);
1053: C. Reid, J. Chem. Phys. {\bf 43}, S186 (1965).
1054: \bibitem{payne}
1055: G. Payne and L. Schlessinger, J. Comp. Phys. {\bf 13}, 266 (1973).
1056: \bibitem{biswas}
1057: S. Biswas et al. Jour. Math. Phys. {\bf 14}, 1190 (1973)
1058: \bibitem{kbeck81}
1059: J. Killingbeck, Phys. Lett. {\bf 84A}, 95 (1981).
1060: \bibitem{fernandez}
1061: F. Fernandez, Q. Ma and R. Tipping, Phys. Rev. A {\bf 39}, 1605 (1989).
1062: \bibitem{bacus}
1063: B. Bacus, Y. Meurice, and A. Soemadi, J. Phys. A {\bf 28}, L381 (1995).
1064: \bibitem{antonsen}
1065: F. Antonsen, Phys. Rev. A {\bf 60}, 812 (1999); W. Bietenholz and
1066: T. Struckmann, Int. Jour. Mod. Phys. C {\bf 10}, 531 (1999).
1067: \bibitem{oktay}
1068: B. Oktay, unpublished.
1069: \bibitem{turbiner84}
1070: A. Turbiner, Sov. Phys. Ups. {\bf 27}, 668 (1984).
1071: \bibitem{pernice98}
1072: S. Pernice and G. Oleaga, Phys. Rev. D {\bf 57}, 1144 (1998).
1073: \bibitem{convpert}
1074: Y. Meurice, Phys. Rev. Lett. {\bf 88}, 141601 (2002); Y. Meurice,
1075: Nucl. Phys. B (Proc. Suppl.) {\bf 106}, 908 (2002).
1076: \bibitem{knopp}
1077: K. Knopp, {\em The Theory of Functions, Part II} (Dover, New York, 1975), Ch. 1
1078: section 1.
1079: \bibitem{coddington}
1080: E. Coddington and N. Levinson, {\em Theory of Ordinary Differential Equations}
1081: (Mc Graw-Hill, New York, 1955), Ch. 3 section 7.
1082: \bibitem{jentschura}
1083: U. Jentschura and J. Zinn-Justin, J. Phys. A {\bf 34}, L253 (2001).
1084: \bibitem{bender96}
1085: C. Bender and L. Bettencourt, Phys. Rev. Lett {\bf 77}, 4114 (1996).
1086: \end{thebibliography}
1087: \end{multicols}
1088: \end{document}