0807.5032/qm3.tex
1: \documentclass{article}
2: \usepackage{amsmath}
3: \usepackage{graphicx}
4: \usepackage{amsfonts}
5: \usepackage{amssymb}
6: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
7: 
8: 
9: \begin{document}
10: 
11: \title{Anharmonic oscillator, negative dimensions and inverse factorial convergence
12: of large orders to the asymptotic form}
13: \author{P.V. Pobylitsa\\\emph{Petersburg Nuclear Physics Institute,}\\\emph{Gatchina, St.~Petersburg, 188300, Russia}}
14: \date{}
15: \maketitle
16: 
17: \begin{abstract}
18: The spectral problem for $O(D)$ symmetric polynomial potentials allows for a
19: partial algebraic solution after analytical continuation to negative even
20: dimensions $D$. This fact is closely related to the disappearance of the
21: factorial growth of large orders of the perturbation theory at negative even
22: $D$. As a consequence, certain quantities constructed from the perturbative
23: coefficients exhibit fast inverse factorial convergence to the asymptotic
24: values in the limit of large orders. This quantum mechanical construction can
25: be generalized to the case of quantum field theory.
26: 
27: \end{abstract}
28: 
29: \section{Introduction}
30: 
31: \setcounter{equation}{0} 
32: 
33: The $O(D)$ symmetric anharmonic oscillator
34: \begin{equation}
35: H=\frac{1}{2}\sum\limits_{i=1}^{D}\left(  p_{i}^{2}+x_{i}^{2}\right)
36: +V\left(  \sum\limits_{i=1}^{D}x_{i}^{2}\right)  \label{H-general}%
37: \end{equation}
38: with a polynomial potential $V(r^{2})$ is a good theoretical toy for testing
39: various ideas used in quantum field theory. In particular, one should mention:
40: 
41: 1) large orders of perturbation theory
42: \cite{BW-69,BW-73,Lipatov-77b,BLZ-77,SZ-79,DP-79,Zinn-Justin-81a,Zinn-Justin-81b,ZJ-04}%
43: ,
44: 
45: 2) fermion representations for boson theories in negative dimensions
46: \cite{DH-88,Dunne-89},
47: 
48: 3) hidden symmetries, quasi-exactly solvable models
49: \cite{Turbiner-94,Shifman-ITEP,Ushveridze}, correspondence between integrable
50: quantum field theoretical models and ordinary differential equations
51: \cite{DT-99,BLZ-98,BLZ-03,DDT-07,DDGT-07}, correspondence between $O(D)$
52: symmetric and $D=1$ anharmonic oscillator systems
53: \cite{SZ-79,Andrianov-82,DPM-84,BG-93,Andrianov-07}.
54: 
55: 4) Regge theory of complex angular momenta in nonrelativistic quantum
56: mechanics \cite{DeAlfaro-Regge,Collins-book,Newton-book}.
57: 
58: There is a vast literature where the anharmonic oscillator is studied from
59: various points of view. The aim of this paper is to bring some of these ideas
60: together. The paper consists of two parts. The first part is devoted to the
61: spectrum of the anharmonic oscillator at even negative $D$
62: \begin{equation}
63: D=0,-2,-4,-6,\ldots\label{D-even-negative}%
64: \end{equation}
65: understood in the sense of analytical continuation. A remarkable feature of
66: these values of $D$ is that a part of the spectrum can be found by solving
67: algebraic equations. In principle, this fact is known since long ago
68: \cite{DH-88} but as usual with exactly solvable models, one can arrive at the
69: same results using many different ways, and the alternative derivations
70: discussed here may deserve attention.
71: 
72: The second part of the paper deals with the asymptotic behavior of large
73: orders of the perturbation theory. As is well known, the perturbative series
74: for the anharmonic oscillator is divergent because of the factorial growth of
75: the coefficients. However, one can show that in even negative dimensions
76: (\ref{D-even-negative}) the factorial growth disappears and the perturbative
77: series becomes convergent. Taken alone, this fact is quite natural and not too
78: exciting. But this observation gives an interesting solution to another old problem.
79: 
80: Let us return to the case of general $D$ where we have the factorial growth of
81: perturbative coefficients. It is well known that the asymptotic behavior of
82: large-$k$ orders of the perturbation theory is reached very slowly because of
83: the $O(k^{-1})$ corrections that are usually rather large in practically
84: interesting cases. Therefore it would be rather interesting to construct
85: quantities for which the asymptotic regime is reached faster than $O(k^{-1})$.
86: This problem is especially acute in quantum field theoretical models where
87: only several first orders of the perturbation theory can be usually computed.
88: In this paper we construct and study special quantities whose approach to the
89: asymptotic regime is \emph{factorially fast} instead of the traditional slow
90: $O(k^{-1})$ convergence to the asymptotic form. It is interesting that the
91: same construction can be used in quantum field theory. In fact, this paper
92: appeared as an attempt to understand some superfast convergence effects that
93: are seen in quantum field theory \cite{PP-08}.
94: 
95: Before starting with a detailed derivation of the exact solution for the
96: spectrum of the anharmonic oscillator at negative even dimensions $D$, it
97: makes sense to describe the final result. For simplicity let us concentrate on
98: levels with zero angular momentum $l=0$.
99: 
100: At negative even $D$ a part of the discrete spectrum of energy $E$ is given by
101: roots of the algebraic equation
102: \begin{gather}
103: R_{M}(E)=0\,,\label{R-M-g-spectrum-eq}\\
104: M=-\frac{D}{2}=0,1,2,3,\ldots
105: \end{gather}
106: where $R_{M}(E)$ are certain polynomials of $E$ depending on potential $V$. In
107: this paper a rather elegant representation for these polynomials will be
108: derived
109: \begin{equation}
110: R_{M}(E)=\det\left[  J_{+}+V\left(  2J_{-}\right)  -E\right]  \,.
111: \label{R-N-det}%
112: \end{equation}
113: Here
114: \begin{equation}
115: J_{\pm}=J_{1}\pm iJ_{2}%
116: \end{equation}
117: are standard spin matrices for spin
118: \begin{equation}
119: j=\frac{M}{2}=-\frac{D}{4}=0,\frac{1}{2},1,\frac{3}{2},\ldots
120: \end{equation}
121: 
122: Several comments should be made about the history of the exact solution for
123: the anharmonic oscillator. The limit $D\rightarrow0$ was considered by Dolgov
124: and Popov \cite{DP-79}. From the point of view of representation
125: (\ref{R-N-det}) this case corresponds to spin $j=0$ with the polynomial
126: $R_{0}=-\mathrm{\,}E$ leading according to eq. (\ref{R-M-g-spectrum-eq}) to
127: only level controlled by the exact solution $E=0$. Although this case is
128: trivial from the spectral point of view, the work of Dolgov and Popov
129: \cite{DP-79} contains several interesting results. In particular, the wave
130: function corresponding to the level $E=0$ was computed in Ref. \cite{DP-79}.
131: 
132: The general case of arbitrary negative even $D$ was studied by Dunne and
133: Halliday in Ref. \cite{DH-88} where a method was suggested for a calculation
134: of polynomials describing the spectrum. In principle, this solves the problem
135: of the spectrum. However, some issues were not clarified. Testing some special
136: cases, the authors of \cite{DH-88} found an unexpected factorization of their
137: polynomials into ``elementary polynomials''. In the framework of the current
138: paper the ``elementary polynomials'' of Dunne and Halliday are nothing else
139: but polynomials $R_{M}(E,g)$ given by eq. (\ref{R-N-det}). An important role
140: in the disentanglement of the polynomial structure is played by a ``hidden''
141: $sl(2)$ symmetry which stands behind the spin representation (\ref{R-N-det})
142: and explains the miracles observed in Ref. \cite{DH-88}.
143: 
144: In this paper we discuss several derivations of polynomials (\ref{R-N-det}).
145: One method described in section \ref{sl-2-algebra-section} uses an explicit
146: expression for the Hamiltonian of the anharmonic oscillator in terms of
147: differential operators obeying $sl(2)$ algebra. Another way to polynomials
148: (\ref{R-N-det}) considered in Sec. \ref{Recusrion-relation-subsection} is
149: based on the analysis of a recursion relation for the coefficients of the
150: power series for the wave function. In the third method (Sec.
151: \ref{Analytical-PF-section}) one performs analytical continuation of the $l$
152: decomposition for the partition function. One more method (Sec.
153: \ref{Fermion-section}) uses a fermion representation for the partition
154: function continued analytically to negative dimensions. This approach is very
155: close to the method of original work \cite{DH-88} but we implement this method
156: so that the $sl(2)$ algebra remains explicit.
157: 
158: An important aspect of the problem of the anharmonic oscillator in negative
159: dimensions is the precise definition of negative dimensions. It is interesting
160: that the problem of analytical continuation to negative or complex dimensions
161: is equivalent to the problem of analytical continuation to complex angular
162: momenta which is a corner stone of Regge theory (Sec.
163: \ref{Regge-theory-subsection}).
164: 
165: As was already mentioned, at negative even $D$ we meet interesting phenomena
166: in large orders of the perturbation theory. The energy of the ground state of
167: the anharmonic oscillator has a perturbative expansion
168: \begin{equation}
169: E(g,D)=\sum\limits_{k=0}^{\infty}E^{(k)}(D)g^{k} \label{E-series-0}%
170: \end{equation}
171: with coefficients $E^{(k)}(D)$ which are polynomials of $D$ so that there are
172: no problems with analytical continuation of $E^{(k)}(D)$ to negative
173: dimensions. In the case of the quartic anharmonic oscillator%
174: 
175: \begin{equation}
176: V(r^{2})=\frac{1}{2}r^{2}+gr^{4} \label{H-quartic}%
177: \end{equation}
178: the large-$k$ asymptotic behavior of $E^{(k)}(D)$ is described by formula
179: \cite{BLZ-77}%
180: \begin{equation}
181: E^{(k)}(D)\overset{k\rightarrow\infty}{=}(-1)^{k+1}\Gamma\left(  k+\frac{D}%
182: {2}\right)  3^{k+\frac{D}{2}}\frac{2^{D/2}}{\pi\Gamma\left(  D/2\right)
183: }\left[  1+O(k^{-1})\right]  . \label{E-k-asymptotic-0}%
184: \end{equation}
185: An interesting feature of this formula is that the RHS vanishes at even
186: negative $D$ because of the factor $\left[  \Gamma\left(  D/2\right)  \right]
187: ^{-1}$ on the RHS (\ref{E-k-asymptotic-0}). In section
188: \ref{No-factorials-subsection} we show that the factorial divergence of series
189: (\ref{E-series-0}) disappears at negative even values of $D$ so that the
190: perturbative series has a nonzero convergence radius.
191: 
192: Functions $E^{(k)}(D)$ are polynomials of $D$ of degree $k+1$ and they have
193: $k+1$ roots:%
194: \begin{equation}
195: E^{(k)}(\nu_{k,r})=0\quad(1\leq r\leq k+1).
196: \end{equation}
197: The last part of the paper is devoted to asymptotic properties of these roots
198: in the limit of large $k$. One could wonder why we should care about the roots
199: $\nu_{k,r}$. The answer is that these roots have a very interesting property:
200: in the limit $k\rightarrow\infty$ some subsets of the roots are convergent to
201: points $D=-4,-6,-8,\ldots$ and this convergence is factorially fast. In
202: quantum mechanics this superfast factorial convergence may be an amazing but
203: useless curiosity. However, in the context of quantum field theory analogous
204: effects are important. Indeed, in quantum field theory one is usually limited
205: to a rather small amount of perturbative terms. Four or five orders are
206: usually considered as a great achievement. Therefore the existence of
207: quantities with factorial convergence to the asymptotic form instead of the
208: slow $O(k^{-1})$ convergence of (\ref{E-k-asymptotic-0}) is extremely
209: interesting for field theoretical applications. The case of the anharmonic
210: oscillator allows us to study this phenomenon in detail.
211: 
212: \section{$O(D)$ symmetric Hamiltonians and $sl(2)$ algebra}
213: 
214: \setcounter{equation}{0} 
215: 
216: \label{sl-2-algebra-section}
217: 
218: \subsection{Schr\"{o}dinger equation}
219: 
220: \label{Schroedinger-section}
221: 
222: The $D$-dimensional Hamiltonian with a spherically symmetric potential
223: (\ref{H-general}) becomes in the polar coordinates%
224: \begin{equation}
225: H_{r}=\frac{1}{2}\left(  -\frac{d^{2}}{dr^{2}}-\frac{D-1}{r}\frac{d}%
226: {dr}\right)  +\frac{l(l+D-2)}{2r^{2}}+V(r^{2})\,.
227: \end{equation}
228: In $D\geq2$ dimensions, parameter $l$ runs over nonnegative integer values%
229: \begin{equation}
230: l=0,1,2,3,\ldots\quad(D\geq2)
231: \end{equation}
232: 
233: The degeneracy of the $l$ levels is given by%
234: \begin{equation}
235: m(D,l)=\frac{(2l+D-2)}{l!}\frac{\Gamma(D+l-2)}{\Gamma(D-1)}\,.
236: \label{l-degeneracy-positive}%
237: \end{equation}
238: The cases $D=1$, $D=2$ are somewhat exceptional. But they still can be
239: described by analytical continuation of (\ref{l-degeneracy-positive}) in $D$
240: (at fixed integer nonnegative $l$):
241: 
242: For $D=2$ and nonnegative integer $l$, we obtain%
243: \begin{equation}
244: m(2,l)=2-\delta_{l0}\,,
245: \end{equation}
246: which can interpreted in terms of the $\pm l$ degeneracy for $l\neq0$.
247: 
248: For $D=1$, we have%
249: \begin{equation}
250: m(1,l)=\delta_{l0}+\delta_{l1}%
251: \end{equation}
252: so that the only allowed values are $l=0,1$, and they correspond to the states
253: with positive and negative parity respectively.
254: 
255: After the separation of the factor $r^{l}$ from the wave function one arrives
256: at%
257: \begin{equation}
258: H_{\mathcal{D}}=r^{-l}H_{r}r^{l}=\frac{1}{2}\left(  -\frac{d^{2}}{dr^{2}%
259: }-\frac{\mathcal{D}-1}{r}\frac{d}{dr}\right)  +V(r^{2}) \label{H-D-def}%
260: \end{equation}
261: where%
262: \begin{equation}
263: \mathcal{D}=D+2l\,. \label{Delta-def}%
264: \end{equation}
265: Thus the spectrum depends on $D$ and $l$ only via parameter $\mathcal{D}$:
266: \begin{equation}
267: E_{nl}(D)=E_{n0}(D+2l)=E_{n0}(\mathcal{D})\,. \label{E-D-l}%
268: \end{equation}
269: In the case $D=1$ is the values $l=0$, $1$ correspond the states with positive
270: and negative parity respectively.
271: 
272: Introducing the variable%
273: \begin{equation}
274: \zeta=\frac{r^{2}}{2}%
275: \end{equation}
276: we bring $H_{\mathcal{D}}$ (\ref{H-D-def}) to the form%
277: \begin{equation}
278: H_{\zeta}=\left(  -\zeta\frac{d^{2}}{d\zeta^{2}}-\frac{\mathcal{D}}{2}%
279: \frac{d}{d\zeta}\right)  +V(2\zeta)\,. \label{H-zeta}%
280: \end{equation}
281: The corresponding Schr\"{o}dinger equation is%
282: \begin{equation}
283: \left[  \left(  -\zeta\frac{d^{2}}{d\zeta^{2}}-\frac{\mathcal{D}}{2}%
284: \frac{d}{d\zeta}\right)  +V(2\zeta)-E_{n}(\mathcal{D})\right]  \psi_{n}%
285: (\zeta)=0\,. \label{zeta-Schroedinger}%
286: \end{equation}
287: Below we assume for simplicity that the potential $V(2\zeta)$ is polynomial in
288: $\zeta$.
289: 
290: Equation (\ref{zeta-Schroedinger}) was derived for positive integer dimensions
291: $\mathcal{D}$. However, we can use this equation for analytical continuation
292: of eigenenergies $E_{n}(\mathcal{D})$ to arbitrary $\mathcal{D}$. To this aim
293: we solve equation (\ref{zeta-Schroedinger}) imposing the following boundary
294: conditions on eigenfunctions $\psi_{n}(\zeta)$:
295: 
296: 1) $\psi_{n}(\zeta)$ must be analytical at $\zeta=0$:%
297: \begin{equation}
298: \phi(\zeta)=\sum\limits_{k=0}^{\infty}p_{k}\zeta^{k}\,, \label{phi-series}%
299: \end{equation}
300: 
301: 2) $\psi_{n}(\zeta)$ must decay at $\zeta\rightarrow+\infty$.
302: 
303: \subsection{$sl(2)$ representation for $O(D)$ symmetric Hamiltonians}
304: 
305: Let us define parameter%
306: \begin{equation}
307: j=-\frac{\mathcal{D}}{4} \label{j-j-prime-D}%
308: \end{equation}
309: and operators%
310: \begin{align}
311: T_{+}  &  =-\zeta\frac{d^{2}}{d\zeta^{2}}+2j\frac{d}{d\zeta}%
312: ,\label{T-plus-diff}\\
313: T_{0}  &  =-\zeta\frac{d}{d\zeta}+j,\\
314: T_{-}  &  =\zeta\,. \label{T-minus-diff}%
315: \end{align}
316: These operators provide a representation of $sl(2)$ algebra%
317: \begin{equation}
318: \lbrack T_{+},T_{-}]=2T_{0},\quad\lbrack T_{+},T_{0}]=-T_{+},\quad\lbrack
319: T_{-},T_{0}]=T_{-} \label{sl-2-algebra}%
320: \end{equation}
321: with the Casimir operator%
322: \begin{gather}
323: C_{2}=\frac{1}{2}\left(  T_{+}T_{-}+T_{-}T_{+}\right)  +T_{0}T_{0}\,,\\
324: \left[  C_{2},T_{a}\right]  =0\,.
325: \end{gather}
326: In representation (\ref{T-plus-diff}) -- (\ref{T-minus-diff}) we have%
327: \begin{equation}
328: C_{2}=j(j+1)\,. \label{C2-j}%
329: \end{equation}
330: In terms of operators (\ref{T-plus-diff}) -- (\ref{T-minus-diff}) Hamiltonian
331: (\ref{H-zeta}) becomes%
332: \begin{equation}
333: H_{T}=T_{+}+V(2T_{-})\,. \label{H-T-def}%
334: \end{equation}
335: In the case of the quartic anharmonic oscillator (\ref{H-quartic}) we have%
336: \begin{equation}
337: H_{T}^{\mathrm{QO}}=T_{+}+T_{-}+4gT_{-}^{2}\,.
338: \end{equation}
339: 
340: For integer or half-integer $j$ one could expect a simplification of the
341: problem of the anharmonic oscillator due to the presence of finite dimensional
342: irreducible representations of $sl(2)$. However, we meet a problem on this
343: way: in the case of finite dimensional representations of $sl(2)$ there exists
344: a vector $\psi$ annihilated by $T_{-}$:%
345: \begin{equation}
346: T_{-}\psi=0,\quad\psi\neq0\,,
347: \end{equation}
348: whereas for operator $T_{-}$ (\ref{T-minus-diff}) equation%
349: \begin{equation}
350: T_{-}\psi(\zeta)=\zeta\psi(z)=0
351: \end{equation}
352: leads to $\psi=0$ [or to $\psi(z)=c\delta(z)$ if one allows for generalized
353: functions]. In the next section we show how this problem can be circumvented
354: in negative even dimensions.
355: 
356: \subsection{Effective spin Hamiltonian for even negative dimensions
357: $\mathcal{D}$}
358: 
359: \label{Solution-at-negative-D-section}
360: 
361: Let us consider the case of positive integer or half-integer $j$%
362: \begin{equation}
363: j=0,\frac{1}{2},1,\frac{3}{2},\ldots\, \label{j-special}%
364: \end{equation}
365: According to eq. (\ref{j-j-prime-D}) this corresponds to negative even values
366: (\ref{D-even-negative}) of parameter $\mathcal{D}$. These values are
367: nonphysical. We assume analytical continuation in the sense of solutions of
368: Schr\"{o}dinger equation with boundary conditions described in Sec.
369: \ref{Schroedinger-section}.
370: 
371: Let us consider the subspace $\mathcal{H}_{j}$ of functions $f(\zeta)$ obeying
372: conditions%
373: \begin{equation}
374: f^{(k)}(0)=0\quad(0\leq k\leq2j)\,.
375: \end{equation}
376: This subspace is invariant under the action of operators $T_{a}$
377: (\ref{T-plus-diff}) -- (\ref{T-minus-diff}). Therefore we can reduce the
378: action of operators $T_{a}$ from the space $\mathcal{H}$ of differentiable
379: functions to the factor space $\mathcal{H}/\mathcal{H}_{j}$. Expression
380: (\ref{H-T-def}) for Hamiltonian $H_{T}$ shows that $\mathcal{H}_{j}$ is
381: invariant also under the action of $H_{T}$. Therefore we can also reduce the
382: action of $H_{T}$ from $\mathcal{H}$ to $\mathcal{H}/\mathcal{H}_{j}$. The
383: reduction to $\mathcal{H}/\mathcal{H}_{j}$ is also possible for the spectral
384: problem%
385: \begin{equation}
386: \left(  H_{T}-E\right)  \psi=0\,. \label{H-T-Schroedinger}%
387: \end{equation}
388: Certainly after this reduction only some finite part of the infinite spectrum
389: can survive. Note that the factor space $\mathcal{H}/\mathcal{H}_{j}$ is
390: isomorphic to the space of polynomials $P_{2j}(\zeta)$ of degree $2j$. The
391: action of operators $T_{a}$ (\ref{T-plus-diff}) -- (\ref{T-minus-diff}) on
392: these polynomials is described by the usual differentiation algebra extended
393: by an additional formal rule%
394: \begin{equation}
395: \xi^{2j+k}=0\quad\mathrm{for}\quad k\geq1\,.
396: \end{equation}
397: The action of operators $T_{a}$ in the space of these polynomials is
398: equivalent to the spin-$j$ irreducible $(2j+1)$-dimensional representation of
399: $sl(2)$. This is obvious from the dimension $2j+1$ of this representation and
400: from the eigenvalue $j(j+1)$ of the Casimir operator (\ref{C2-j}). Therefore
401: after the reduction of the spectral problem (\ref{H-T-Schroedinger}) to the
402: factor space $\mathcal{H}/\mathcal{H}_{j}$ we arrive at the matrix eigenvalue
403: problem%
404: \begin{gather}
405: H_{j}=J_{+}+V(2J_{-})\,,\label{H-J-def}\\
406: \left(  H_{j}-E\right)  \phi=0\, \label{det-E-H-J}%
407: \end{gather}
408: where $J_{a}$ are $\left(  2j+1\right)  $-dimensional matrices of the spin-$j$
409: representation of $sl(2)$. These $sl(2)$ matrices $J_{a}$ are connected by relations%
410: 
411: \begin{align}
412: J_{0}  &  =J_{3}\,,\\
413: J_{\pm}  &  =J_{1}\pm iJ_{2} \label{J-pm}%
414: \end{align}
415: with standard spin matrices $J_{1,2,3}$ for the spin-$j$ representation of
416: $su(2)$.
417: 
418: Thus the spectral problem reduces to solving the equation%
419: \begin{equation}
420: \det\left(  H_{j}-E\right)  =0\,. \label{H-E-det}%
421: \end{equation}
422: Obviously%
423: \begin{equation}
424: R_{2j}(E)=\det\left(  H_{j}-E\right)  =\det\left[  J_{+}+V(2J_{-})\,-E\right]
425: \label{R-2j-def}%
426: \end{equation}
427: is a polynomial of degree $2j+1$ so that we deal with an algebraic equation%
428: \begin{equation}
429: R_{2j}(E)=0
430: \end{equation}
431: where $j=-\mathcal{D}/4$ according to (\ref{j-j-prime-D}).
432: 
433: One should keep in mind that matrix Hamiltonian (\ref{H-J-def}) describes only
434: a part of the spectrum. Since $sl(2)$ matrices $J_{\pm}$ are real, polynomials
435: $R_{2j}(E)$ are also real. Nevertheless the eigenvalues of $H_{j} $ may be complex.
436: 
437: In the case of quartic anharmonic oscillator (\ref{H-quartic}) matrix
438: Hamiltonian (\ref{H-J-def}) reduces to%
439: \begin{equation}
440: H_{j}=J_{+}+J_{-}+4gJ_{-}^{2}\,. \label{H-J-quartic}%
441: \end{equation}
442: 
443: \subsection{Connection with quasi-exactly solvable problems}
444: 
445: The Hamiltonians of type (\ref{H-J-quartic}) appear in the context of
446: quasi-exactly solvable (QES) potentials
447: \cite{Turbiner-94,Shifman-ITEP,Ushveridze}. However, there is a certain
448: difference between the $sl(2)$ representation discussed here and in
449: traditional QES problems. Our differential operators $T_{a}$
450: (\ref{T-plus-diff}) -- (\ref{T-minus-diff}) differ from the operators
451: $T_{a}^{\prime}$ used in QES problems:%
452: \begin{align}
453: T_{+}^{\prime}  &  =2j\xi-\xi^{2}\frac{d}{d\xi}\,,\\
454: T_{0}^{\prime}  &  =-j+\xi\frac{d}{d\xi}\,,\\
455: T_{-}^{\prime}  &  =\frac{d}{d\xi}\,.
456: \end{align}
457: Both sets of differential operators obey $sl(2)$ algebra (\ref{sl-2-algebra}).
458: In fact, $T_{a}$ and $T_{a}^{\prime}$ are connected (at least formally) by a
459: Laplace transformation combined with the change of $j\rightarrow-1-j$.
460: Nevertheless the two classes of problems are different. In particular, the
461: methods discussed here allow us to compute a part of the spectrum but not the
462: corresponding wave functions $\psi_{n}(\zeta)$ obeying equation
463: (\ref{zeta-Schroedinger}). It should be emphasized that one has to distinguish
464: two types of wave functions at negative $\mathcal{D}$:
465: 
466: 1) solutions of Schr\"{o}dinger equation (\ref{zeta-Schroedinger}) at negative
467: even $\mathcal{D}$,
468: 
469: 2) eigenvectors of the effective spin Hamiltonian (\ref{H-J-def}).
470: 
471: It is interesting that in the case of the sextic QES potential, Dunne and
472: Halliday \cite{DH-88} could trace the connection between the exact
473: eigenfunctions in physical dimensions $D=1,2,3,\ldots$ and the eigenvectors of
474: the effective finite-dimensional matrix problem corresponding to
475: $D=0,-2,-4\ldots$
476: 
477: \subsection{Example: harmonic oscillator}
478: 
479: In the case of the harmonic oscillator%
480: \begin{align}
481: V^{\mathrm{HO}}(r^{2})  &  =\frac{1}{2}r^{2}\,,\label{V-HO}\\
482: V^{\mathrm{HO}}(2\zeta)  &  =\zeta\,,
483: \end{align}
484: the calculation of determinant (\ref{det-E-H-J}) is trivial:%
485: \begin{equation}
486: \det\left(  J_{+}+J_{-}-E\right)  =\det\left(  2J_{1}-E\right)  =\det\left(
487: 2J_{3}-E\right)  =2^{2j+1}\prod\limits_{n=-j}^{j}\left(  n-\frac{E}{2}\right)
488: \label{det-H-H-HO}%
489: \end{equation}
490: so that solutions of eq. (\ref{H-E-det}) are%
491: \begin{align}
492: E  &  =-2j,-2\left(  j-1\right)  ,\ldots,2\left(  j-1\right)  ,2j\nonumber\\
493: &  =\frac{\mathcal{D}}{2},\frac{\mathcal{D}}{2}+2,\ldots,-\frac{\mathcal{D}%
494: }{2}-2,-\frac{\mathcal{D}}{2}\quad(\mathcal{D}\leq0). \label{E-matrix-EV-HO}%
495: \end{align}
496: It is instructive to compare this algebraic solution with the full spectrum of
497: the harmonic oscillator. Labeling levels with $n=0,1,2,3,\ldots$ in each $l$
498: sector, we can write%
499: \begin{equation}
500: E_{nl}(0,D)=\frac{D}{2}+\left(  2n+l\right)  \,. \label{E-nl-HO}%
501: \end{equation}
502: Combining $D$ and $l$ into parameter $\mathcal{D}$ (\ref{Delta-def}) we arrive
503: at the expression%
504: \begin{equation}
505: E=\frac{\mathcal{D}}{2}+2n,\quad n=0,1,2,\ldots\label{HO-all-levels}%
506: \end{equation}
507: that has a trivial analytical continuation in $\mathcal{D}$. For negative even
508: $\mathcal{D}=-4j$ the lowest $2j+1$ levels (\ref{HO-all-levels}) obviously
509: coincide with the solution of the spin problem (\ref{E-matrix-EV-HO}).
510: 
511: In Fig. \ref{trajectories-HO-fig} we show the ``Regge trajectories'' (the
512: relation between the analytical continuation in $D$ and Regge theory is
513: discussed in Sec. \ref{Regge-theory-subsection}) connecting physical values
514: $D=1,2,3,\ldots$ and with the special points $D=0,-2,-4,\ldots$ It is
515: remarkable that any trajectory corresponding to a physical level after
516: analytical continuation to negative $D$ sooner or later enters into the domain
517: described by the roots of polynomial (\ref{det-H-H-HO}).
518: 
519: \begin{figure}[ptb]
520: \begin{center}
521: \includegraphics[
522: height=2.1015in,
523: width=3.2958in
524: ]{trajectories_HO2.eps}
525: \end{center}
526: \caption{Trajectories continuing the levels of the harmonic oscillator
527: analytically in $\mathcal{D}$ from physical values $\mathcal{D}=1,2,3,\ldots$
528: (marked with crosses) to arbitrary values of $\mathcal{D}$. The levels at
529: $\mathcal{D}=0,-2,-4,\ldots$ that are described by roots of polynomials
530: $R_{2j}$ ($j=-D/2$) are marked with circles.}%
531: \label{trajectories-HO-fig}%
532: \end{figure}
533: 
534: \subsection{Power series and recursion relations}
535: 
536: \label{Recusrion-relation-subsection}
537: 
538: The analytical continuation of levels $E_{n}(\mathcal{D})$ to negative
539: $\mathcal{D}$ uses solutions $\phi(\zeta)$ (\ref{phi-series}) of
540: Schr\"{o}dinger equation (\ref{zeta-Schroedinger}) that are analytical at
541: $\zeta=0$. For polynomial potentials%
542: \begin{align}
543: V(2\zeta)  &  =\sum\limits_{k}w_{k}\zeta^{k}\,,\label{V-w}\\
544: w_{k}  &  =0\quad\mathrm{if}\quad k\leq0\quad\mathrm{or}\quad k>L
545: \end{align}
546: obeying condition%
547: \begin{equation}
548: V(0)=0
549: \end{equation}
550: let us define%
551: \begin{equation}
552: u_{k}=\left\{
553: \begin{array}
554: [c]{ll}%
555: w_{k} & \mathrm{if}\quad\,1\leq k\leq L,\\
556: -E & \mathrm{if}\quad k=0,\\
557: 0 & \mathrm{otherwise\,.}%
558: \end{array}
559: \right.  \label{u-k-def}%
560: \end{equation}
561: Then%
562: \begin{equation}
563: V(2\zeta)-E=\sum\limits_{k=-\infty}^{\infty}u_{k}\zeta^{k}\,.
564: \end{equation}
565: Inserting this series and expansion (\ref{phi-series}) into eq.
566: (\ref{zeta-Schroedinger}) one obtains the recursion relation for the
567: coefficients $p_{k}$ of series (\ref{phi-series}):%
568: \begin{align}
569: a_{k+1}(\mathcal{D})p_{k+1}  &  =\sum\limits_{m=0}^{L}u_{m}p_{k-m}%
570: \,\,,\label{recursion-0}\\
571: a_{k}(\mathcal{D})  &  =k\left(  k-1+\frac{1}{2}\mathcal{D}\right)  \,.
572: \label{a-k-D-def}%
573: \end{align}
574: If $D\neq0,-2,-4,\ldots$ then we can solve this equation iteratively starting
575: from some $p_{0}\neq0$ and obtain%
576: \begin{equation}
577: p_{k+1}=\frac{p_{0}s_{k+1}\left(  \{u_{m}\},\mathcal{D}\right)  }%
578: {a_{k+1}(\mathcal{D})} \label{p-s}%
579: \end{equation}
580: where $s_{k+1}\left(  \{u_{m}\},\mathcal{D}\right)  $ are polynomials of
581: $u_{m}$. Once the series (\ref{phi-series}) is constructed, one has impose the
582: boundary condition on $\phi(\zeta)$ at $\zeta\rightarrow+\infty$ and this will
583: fix the spectrum.
584: 
585: The case $\mathcal{D}\equiv-4j=0,-2,-4,\ldots$ is exceptional because in this
586: case at $k=-\mathcal{D}/2=2j$ we have a zero in the denominator on the RHS of
587: (\ref{p-s}). Therefore instead of the determination of $p_{2j}$ we will arrive
588: at the condition%
589: \begin{equation}
590: s_{2j+1}\left(  \{u_{m}\},-4j\right)  =0\,. \label{s-condition}%
591: \end{equation}
592: Since $s_{2j+1}\left(  \{u_{m}\},\mathcal{D}\right)  $ is a nonzero polynomial
593: in $E$ (via $u_{0}=-E$), this condition does not hold for the most of
594: energies. In fact, in the case $\mathcal{D}=0,-2,-4,\ldots$ the majority of
595: eigenfunctions have power expansion (\ref{phi-series}) with%
596: \begin{equation}
597: p_{0}=p_{1}\ldots=p_{2j+1}=0,\quad p_{2j+2}\neq0\,. \label{p-type-2}%
598: \end{equation}
599: 
600: However, if condition (\ref{s-condition}) holds then we can proceed with
601: iterations and construct the series (\ref{phi-series}). Thus in the special
602: case (\ref{s-condition}) we have two solutions (\ref{phi-series}) of
603: Schr\"{o}dinger equation (\ref{zeta-Schroedinger}) analytical at $\zeta=0$:
604: one solution with $p_{0}\neq1$ and the second solution (\ref{p-type-2}). Thus
605: all solutions of Schr\"{o}dinger equation (\ref{phi-series}) are regular in
606: the case (\ref{s-condition}). This means that taking an appropriate linear
607: combination we can satisfy boundary conditions at infinity so that values of
608: $E$ obeying equation (\ref{s-condition}) automatically belong to the spectrum.
609: 
610: It is easy to see that this construction is equivalent to the spin problem
611: (\ref{H-E-det}). Indeed, recursion relation (\ref{recursion-0}) is nothing
612: else but vector equation (\ref{H-E-det}) written in the basis diagonalizing
613: matrix $J_{0}$. Now we understand that polynomials $s_{2j+1}\left(
614: \{w_{m}\},E,-4j\right)  $ appearing in eq. (\ref{s-condition}) must coincide
615: up to a constant with polynomials $R_{2j}(E)$ defined by eq. (\ref{R-2j-def})%
616: \begin{equation}
617: s_{2j+1}\left(  \{u_{m}\},-4j\right)  =c_{j}\det\left[  J_{+}+V\left(
618: 2J_{-}\right)  -E\right]  \label{s-det-equality}%
619: \end{equation}
620: with some coefficients $c_{j}$.
621: 
622: Let us derive equation (\ref{s-det-equality}) more carefully and compute
623: coefficients $c_{j}$. First we define a $\left(  N+1\right)  \times\left(
624: N+1\right)  $ matrix%
625: \begin{equation}
626: C_{kn}^{(N)}=u_{k-n}-a_{k}\delta_{k,n-1}\quad1\leq k,n\leq N+1\,.
627: \label{C-N-def}%
628: \end{equation}
629: According to eq. (\ref{u-k-def}) we assume that $u_{k}=0$ if $k<0$.
630: 
631: Then recursion relation (\ref{recursion-0}) takes the form%
632: \begin{equation}
633: \sum\limits_{n=1}^{N+1}C_{kn}^{(N)}p_{n-1}=\delta_{k,N+1}a_{N+1}p_{N+1}%
634: \quad(1\leq k\leq N+1)\,. \label{C-recursion}%
635: \end{equation}
636: Matrix (\ref{C-N-def}) is quasitriangular. Its determinant can be computed by
637: making with its rows the same linear manipulations as in the recursive
638: calculation of $s_{k+1}\left(  \{u_{m}\},\mathcal{D}\right)  $ in eq.
639: (\ref{p-s}). As a result, one obtains%
640: \begin{equation}
641: \det\left[  C^{(N)}\left(  \{u_{m}\},\mathcal{D}\right)  \right]
642: =s_{N+1}\left(  \{u_{m}\},\mathcal{D}\right)  \prod\limits_{k=1}^{N}%
643: a_{k}(\mathcal{D})\,.
644: \end{equation}
645: Now we set $\mathcal{D}=-4j,$ $N=2j$. Using eq. (\ref{a-k-D-def}), we find%
646: \begin{equation}
647: \prod\limits_{k=1}^{2j}a_{k}(-2j)=(-1)^{2j}\left[  (2j)!\right]  ^{2}\,.
648: \end{equation}
649: Thus%
650: \begin{equation}
651: \det\left[  C^{(2j)}\left(  \{u_{m}\},-4j\right)  \right]  =(-1)^{2j}\left[
652: (2j)!\right]  s_{2j+1}\left(  \{u_{m}\},-4j\right)  \,. \label{det-C-s}%
653: \end{equation}
654: Therefore algebraic equation (\ref{s-condition}) takes the form%
655: \begin{equation}
656: \det\left[  C^{(2j)}\left(  \{u_{m}\},-4j\right)  \right]  =0\,.
657: \label{det-C-0}%
658: \end{equation}
659: In practical calculations one can compute det$C^{(2j)}$ either directly using
660: definition (\ref{C-N-def})%
661: \begin{equation}
662: \det C^{(2j)}=\det_{1\leq k,n\leq2j+1}\left[  w_{k-n}-k\left(  k-1-2j\right)
663: \delta_{k,n-1}-E\delta_{kn}\right]  \label{det-C-via-w}%
664: \end{equation}
665: or solving recursion relations iteratively and finding in this way
666: $s_{2j+1}\left(  \{u_{m}\},-4j\right)  $. Both methods lead to the same
667: results:%
668: \begin{align}
669: \det C^{(0)}  &  =-E\,,\\
670: \det C^{(1)}  &  =E^{2}-w_{1}\,,\\
671: \det C^{(2)}  &  =-E^{3}+4w_{2}+4Ew_{1}\,,\\
672: \det C^{(3)}  &  =E^{4}-10E^{2}w_{1}+9w_{1}^{2}-24w_{2}E-36w_{3}\,,\\
673: \det C^{(4)}  &  =-E^{5}+20E^{3}w_{1}-64Ew_{1}^{2}+84w_{2}E^{2}-192w_{1}%
674: w_{2}+288Ew_{3}+576w_{4}\,.
675: \end{align}
676: 
677: Note that recursion relation (\ref{C-recursion}) for $N=2j$ is nothing else
678: but the equation%
679: \begin{equation}
680: \left[  T_{+}+V\left(  2T_{-}\right)  -E\right]  \phi(\zeta)=0
681: \end{equation}
682: written in the basis $\zeta^{k}$, $k=0,1,2,\ldots2j$. This basis coincides
683: with the standard basis for spin-$j$ representation up to the subtleties of
684: the normalization and enumeration. Obviously the determinant of matrix
685: $C^{(2j)}$ is invariant with the respect to this trivial change of the basis.
686: Therefore%
687: \begin{equation}
688: \det C^{(2j)}=\det\left[  J_{+}+V\left(  2J_{-}\right)  -E\right]
689: \label{det-C-det-H-J}%
690: \end{equation}
691: where the spin $j$ representation is assumed on the RHS. Combining this result
692: with eq. (\ref{det-C-0}) we see that we have obtained old equation
693: (\ref{H-J-def}).
694: 
695: Certainly this ``new derivation'' of eq. (\ref{H-J-def}) is nothing else but
696: an explicit detailed version of the compact arguments of Sec.
697: \ref{Solution-at-negative-D-section} presented there in terms of the $sl(2)$
698: algebra acting in the factor space $\mathcal{H}/\mathcal{H}_{j}$.
699: 
700: \subsection{Properties of polynomials $R_{2j}(E)$}
701: 
702: Comparing eqs. (\ref{R-2j-def}), (\ref{det-C-via-w}), (\ref{det-C-det-H-J}) we
703: see that we have three equivalent representations for polynomials $R_{2j}(E)$%
704: \begin{align}
705: R_{2j}(E)  &  =\det\left[  J_{+}+V(2J_{-})\,-E\right] \label{R-repr-1}\\
706: &  =\det_{1\leq k,n\leq2j+1}\left[  w_{k-n}-k\left(  k-1-2j\right)
707: \delta_{k,n-1}-E\delta_{kn}\right] \label{R-repr-2}\\
708: &  =(-1)^{2j}\left[  (2j)!\right]  s_{2j+1}\left(  \{u_{m}\},-4j\right)  \,.
709: \label{R-repr-3}%
710: \end{align}
711: Representation (\ref{R-repr-1}) is written in terms of spin-$j$ matrices
712: $J_{\pm}=J_{1}\pm iJ_{2}$ and explicitly expresses the $sl(2)$ symmetry
713: standing behind this construction. Representation (\ref{R-repr-2}) written in
714: terms of coefficients $w_{k}$ of the polynomial potential $V$ (\ref{V-w}) is
715: nothing else but eq. (\ref{R-repr-1}) transformed to a basis with a
716: nonstandard normalization which allows us to get rid of cumbersome square
717: roots appearing in the traditional expressions for the matrix elements of
718: $J_{\pm}$. Representation (\ref{R-repr-3}) is based on the iterative solution
719: (\ref{p-s}) of recursion relations (\ref{recursion-0}) and is useful for
720: computer calculations of polynomials $R_{2j}(E)$.
721: 
722: The higher coefficients of polynomials $R_{2j}(E)$ can be computed using the
723: large-$E$ expansion:%
724: \begin{gather}
725: \det\left[  E-J_{+}-V(2J_{-})\right]  =E^{2j+1}\exp\left\{  \mathrm{Tr}%
726: \ln\left[  1-\frac{J_{+}+V(2J_{-})}{E}\right]  \right\} \nonumber\\
727: =E^{2j+1}\exp\left\{  -\sum\limits_{n=1}^{\infty}\frac{1}{n}E^{-n}%
728: \mathrm{Tr}\left[  J_{+}+V(2J_{-})\right]  ^{n}\right\}  \,.
729: \end{gather}
730: We have%
731: \begin{align}
732: \mathrm{Tr}\left[  J_{+}+V(2J_{-})\right]   &  =0\,,\nonumber\\
733: \mathrm{Tr}\left\{  \left[  J_{+}+V(2J_{-})\right]  ^{2}\right\}   &
734: =2\mathrm{Tr}\left[  J_{+}V(2J_{-})\right]  =2w_{1}\mathrm{Tr}\left(
735: J_{+}J_{-}\right)  \,,\nonumber\\
736: \mathrm{Tr}\left[  J_{+}+V(2J_{-})\right]  ^{3}  &  =3\mathrm{Tr}\left[
737: \left(  J_{+}\right)  ^{2}V(2J_{-})\right]  =3w_{2}\mathrm{Tr}\left[  \left(
738: J_{+}\right)  ^{2}\left(  J_{-}\right)  ^{2}\right]
739: \end{align}
740: where $w_{k}$ are coefficients of the polynomial potential (\ref{V-w}). A
741: straightforward spin algebra gives%
742: \begin{align}
743: \mathrm{Tr}\left(  J_{+}J_{-}\right)   &  =\frac{2}{3}j(j+1)(2j+1)\,,\\
744: \mathrm{Tr}\left[  \left(  J_{+}\right)  ^{2}\left(  J_{-}\right)
745: ^{2}\right]   &  =\frac{2}{15}j(j+1)(2j+1)(2j-1)(2j+3)\,.
746: \end{align}
747: Finally we obtain%
748: 
749: \begin{align}
750: (-1)^{2j+1}R_{2j}(E)  &  =E^{2j+1}-\frac{2}{3}j(j+1)(2j+1)w_{1}E^{2j-1}%
751: \nonumber\\
752: &  -\frac{2}{15}j(j+1)(2j+1)(2j-1)(2j+3)w_{2}E^{2j-2}+O(E^{2j-3}),
753: \end{align}
754: 
755: In the case of harmonic oscillator (\ref{V-HO}) we have according to
756: (\ref{det-H-H-HO})%
757: \begin{equation}
758: R_{2j}^{\mathrm{HO}}(E)=\det\left(  J_{+}+J_{-}-E\right)  =2^{2j+1}%
759: \prod\limits_{n=-j}^{j}\left(  n-\frac{E}{2}\right)  \,.
760: \end{equation}
761: 
762: \subsection{Symmetry $D\rightarrow4-D$}
763: 
764: \label{D-4-D-symmetry-section}
765: 
766: Relation%
767: 
768: \begin{equation}
769: \zeta^{-1+(\mathcal{D}/2)}\left(  -\zeta\frac{d^{2}}{d\zeta^{2}}%
770: -\frac{\mathcal{D}}{2}\frac{d}{d\zeta}\right)  \zeta^{1-(\mathcal{D}%
771: /2)}=-\zeta\frac{d^{2}}{d\zeta^{2}}-\frac{4-\mathcal{D}}{2}\frac{d}{d\zeta}
772: \label{D-4-D-symmetry}%
773: \end{equation}
774: shows that equation (\ref{zeta-Schroedinger}) has a symmetry%
775: \begin{equation}
776: \phi_{4-\mathcal{D}}(\zeta)=\zeta^{-1+(\mathcal{D}/2)}\phi_{\mathcal{D}}%
777: (\zeta)\,, \label{D-4-D-symmetry-phi}%
778: \end{equation}%
779: \begin{equation}
780: \mathcal{D}\rightarrow4-\mathcal{D\,}\,.
781: \end{equation}
782: 
783: However, one should be careful about the boundary conditions. Indeed, for
784: $\mathcal{D}<2$ the factor $\zeta^{-1+(\mathcal{D}/2)}$ is singular so that
785: regular solutions $\phi_{\mathcal{D}}(\zeta)$ may correspond to singular
786: solutions of $\phi_{4-\mathcal{D}}(\zeta)$. In other words, some part of the
787: $\mathcal{D}<2$ spectrum may be lost when one turns from $\mathcal{D}<2$ to
788: $4-\mathcal{D}>2$.
789: 
790: According to results of Sec. \ref{Recusrion-relation-subsection} at
791: $\mathcal{D}=2M=0,-2,-4,\ldots$ there are two types of regular solutions
792: (\ref{phi-series})
793: 
794: 1) solutions with $p_{0}\neq0$,
795: 
796: 2) solutions with $p_{0}=\ldots=p_{M}=0$, $p_{M+1}\neq0$.
797: 
798: Obviously solutions $\phi_{\mathcal{D}}(\zeta)$ of the first type with
799: $p_{0}\neq0$ generate singular functions $\phi_{4-\mathcal{D}}(\zeta)$
800: (\ref{D-4-D-symmetry-phi}). Therefore the spectrum of the $4-\mathcal{D}$
801: problem does not contain that part of the $\mathcal{D}=-0,-2,-4$ spectrum
802: which corresponds to wave functions (\ref{phi-series}) with $p_{0}=0$. We see
803: that under the change $\mathcal{D}\rightarrow4-\mathcal{D}$ we lose exactly
804: those states which are described by the roots of polynomials $R_{M}(E)$.
805: 
806: This symmetry $\mathcal{D}\rightarrow4-\mathcal{D}$ is illustrated in Fig.
807: \ref{trajectories-AO-fig}. It should be stressed that the symmetry
808: $\mathcal{D}\rightarrow4-\mathcal{D}$ is relevant only for integer even values
809: of $\mathcal{D}$.
810: 
811: \section{Quartic anharmonic oscillator}
812: 
813: \setcounter{equation}{0} 
814: 
815: \subsection{Polynomials $R_{2j}(E,g)$}
816: 
817: In the case of the quartic anharmonic oscillator (\ref{H-quartic}) we have%
818: \begin{align}
819: V(2\zeta)  &  =\zeta+4g\zeta^{2}=w_{1}+w_{2}\zeta^{2}\,,\\
820: w_{1}  &  =1\,,\quad w_{2}=4g\,.
821: \end{align}
822: Now one can use one of representations (\ref{R-repr-1}) -- (\ref{R-repr-3})
823: and compute $R_{2j}(E)$. In order to avoid large coefficients it is convenient
824: to define%
825: 
826: \begin{equation}
827: \tilde{R}_{2j}(E,g)=(-2)^{-2j}\mathrm{\,}R_{2j}(E)=(-2)^{-2j}\mathrm{\,}%
828: \det\left(  J_{+}+J_{-}+4gJ_{-}^{2}-E\right)  \,. \label{R-tilde-2j-def}%
829: \end{equation}
830: At $g=0$ we have according to (\ref{det-H-H-HO})%
831: \begin{equation}
832: \tilde{R}_{2j}(E,0)=-2\prod_{k=0}^{2j}\left(  \frac{E-2j}{2}+k\right)  \,.
833: \label{R-HO-res}%
834: \end{equation}
835: Let us define%
836: \begin{equation}
837: \tilde{R}_{2j}^{\prime}(E,g)=\tilde{R}_{2j}(E,g)-\tilde{R}_{2j}(E,0)\,.
838: \label{R-prime-def}%
839: \end{equation}
840: The first polynomials $\tilde{R}_{2j}^{\prime}(E,g)$ are%
841: \begin{align}
842: \tilde{R}_{1}^{\prime}(E,g)  &  =0\,,\\
843: \tilde{R}_{2}^{\prime}(E,g)  &  =4g\,,\\
844: \tilde{R}_{3}^{\prime}(E,g)  &  =12Eg\,,\\
845: \tilde{R}_{4}^{\prime}(E,g)  &  =3(-16+7E^{2})g\,,\\
846: \tilde{R}_{5}^{\prime}(E,g)  &  =4(7E^{3}-55E-200g)g\,,\\
847: \tilde{R}_{6}^{\prime}(E,g)  &  =\frac{9}{2}(7E^{4}-124E^{2}+192-1000Eg)g\,,\\
848: \tilde{R}_{7}^{\prime}(E,g)  &  =\frac{9}{2}(7E^{5}-230E^{3}+1183E-3056E^{2}%
849: g+10976g)g\,.
850: \end{align}
851: Now we find using eqs. (\ref{R-HO-res}) and (\ref{R-prime-def})%
852: \begin{align}
853: \tilde{R}_{0}(E,g)  &  =-E\,,\label{R-0}\\
854: \tilde{R}_{1}(E,g)  &  =-\frac{1}{2}(E^{2}-1)\,,\label{R-1}\\
855: \tilde{R}_{2}(E,g)  &  =-\frac{1}{4}E(E^{2}-4)+4g\,,\label{R-2}\\
856: \tilde{R}_{3}(E,g)  &  =-\frac{1}{8}(E^{2}-9)(E^{2}-1)+12gE. \label{R-3}%
857: \end{align}
858: 
859: \subsection{Ground state at $D=-4$}
860: 
861: The ground state has $l=0$ so that the case of the ground state in $D=-4$
862: dimensions corresponds to $\mathcal{D}=D=-4$. The spectrum is described by the
863: roots of polynomial $\tilde{R}_{2}(E,g)$ (\ref{R-2})%
864: \begin{equation}
865: \tilde{R}_{2}(E,g)=0
866: \end{equation}
867: i.e.%
868: \[
869: E^{3}-4E-16g=0\,.
870: \]
871: Let us define%
872: \begin{align}
873: E  &  =\frac{2}{\sqrt{3}}\varepsilon\,,\\
874: h  &  =3^{3/2}g\,.
875: \end{align}
876: Then%
877: \begin{equation}
878: \varepsilon^{3}-3\varepsilon-2h=0\,.
879: \end{equation}
880: The solutions can be found using Cardano formula:%
881: \begin{equation}
882: \varepsilon=\sqrt[3]{h+\sqrt{h^{2}-1}}+\sqrt[3]{h-\sqrt{h^{2}-1}}\,.
883: \end{equation}
884: Keeping in mind applications to the case of the perturbation theory in small
885: $g$ it is convenient to transform this expression to the form%
886: \begin{align}
887: \varepsilon &  =e^{5\pi i/6}\sqrt[3]{\sqrt{1-h^{2}}-ih}+e^{-5\pi i/6}%
888: \sqrt[3]{\sqrt{1-h^{2}}-ih}\nonumber\\
889: &  =2\mathrm{Re}\left[  \left(  \frac{-\sqrt{3}+i}{2}\right)  \sqrt[3]%
890: {\sqrt{1-h^{2}}-ih}\right]
891: \end{align}
892: which allows for the perturbative expansion%
893: \begin{equation}
894: \varepsilon=-\sqrt{3}+\frac{h}{3}+\ldots
895: \end{equation}
896: This leads to the perturbative expansion for the energy of the ground state in
897: $D=-4$ dimensions%
898: \begin{equation}
899: E(g,-4)=-2+2g+3g^{2}+8g^{3}+\frac{105}{4}g^{4}+96g^{5}+\frac{3003}{16}%
900: g^{6}+1536g^{7}+\ldots\label{E-D-4-series}%
901: \end{equation}
902: In Sec. \ref{Higher-orders-section} we will compute the asymptotic behavior
903: (\ref{E-k-4-asymptotic}) of this series.
904: 
905: \section{Analytical continuation to negative $D$}
906: 
907: \setcounter{equation}{0} 
908: 
909: \subsection{Negative $\mathcal{D}$ and Regge theory}
910: 
911: \label{Regge-theory-subsection}
912: 
913: We have already made some comments about the role of Schr\"{o}dinger equation
914: (\ref{zeta-Schroedinger}) for the analytical continuation in $\mathcal{D}$.
915: Remember that parameter $\mathcal{D}$ (\ref{Delta-def}) is built of $D$ and
916: $l$ so that the problem of analytical continuation in $\mathcal{D}$ has two
917: equivalent formulations:
918: 
919: 1) analytical continuation in $D$ at fixed $l$,
920: 
921: 2) analytical continuation in $l$ at fixed $D$.
922: 
923: The second approach allows us to use some results from Regge theory of complex
924: angular momenta \cite{DeAlfaro-Regge,Collins-book,Newton-book}. Although the
925: traditional Regge theory deals with the scattering problem whereas we are
926: interested in polynomial potentials, the analysis of the behavior of the
927: solutions of Schr\"{o}dinger equation (\ref{zeta-Schroedinger}) at
928: $\zeta\rightarrow0$ is the same in Regge theory of the potential scattering
929: and in our case. In particular, in Regge theory for potentials regular at
930: $r=0$, the Regge poles are at points%
931: \begin{equation}
932: l=-\frac{3}{2},-\frac{5}{2},-\frac{7}{2}\ldots\label{l-Regge}%
933: \end{equation}
934: Combining this with the general formula (\ref{Delta-def}) applied to the case $D=3$%
935: 
936: \begin{equation}
937: \mathcal{D}=3+2l\quad(D=3) \label{D-l-dependence-D3}%
938: \end{equation}
939: we see that $l$-points (\ref{l-Regge}) correspond to negative even values
940: $\mathcal{D}=0,-2,-4,-6,\ldots$.
941: 
942: From the point of view of the historical perspective it is also interesting
943: that analytical continuation in $l$ for potentials growing at infinity, e.g.
944: for the oscillator energies (\ref{E-nl-HO}), was discussed in the context of
945: naive models for quark confinement \cite{Collins-book,Newton-book}.
946: 
947: Although the discussion of analytical continuation in terms of $l$ is
948: preferable from the point of view of explicit connections with Regge theory,
949: we choose the $D$ representation. The $D$ language is natural for the analysis
950: of analytical properties of the partition function (Sec.
951: \ref{Analytical-PF-section}) and for the fermion representation at negative
952: $D$ (\ref{Fermion-section}).
953: 
954: \subsection{$D$ dependence of levels}
955: 
956: We have already discussed the dependence of energy $E_{n}(\mathcal{D})$ on
957: $\mathcal{D}$ for the harmonic oscillator (see Fig. \ref{trajectories-HO-fig}%
958: ). If we switch on a small anharmonicity then the trajectories will be
959: deformed but qualitative features will be preserved. However, in the case of a
960: strong deviation from the harmonic regime (or at larger values of
961: $|\mathcal{D}|$) some new phenomena may occur. In Fig.
962: \ref{trajectories-AO-fig} we show the $\mathcal{D}$ dependence of levels for
963: the quartic anharmonic oscillator (\ref{H-quartic}) with $g=1$. We see that
964: the trajectories of the ground state and of the first excited state meet
965: together at $\mathcal{D}\approx-2.6$ . Polynomials $R_{2j}$ have $2j+1$ roots
966: but some of these roots may be complex. In Fig. \ref{trajectories-AO-fig} we
967: see that this happens for polynomials $R_{2}$, $R_{3}$, $R_{4}$.
968: 
969: \begin{figure}[ptb]
970: \begin{center}
971: \includegraphics[
972: height=2.1041in,
973: width=3.3235in
974: ]{trajectories_AO2.eps}
975: \end{center}
976: \caption{Trajectories continuing the levels of the quartic anharmonic
977: oscillator with $g=1$ analytically in $\mathcal{D}$ from physical values
978: $\mathcal{D}=1,2,3,\ldots$ (marked with crosses) to arbitrary values of
979: $\mathcal{D}$. The levels at $\mathcal{D}=0,-2,-4,\ldots$ that are described
980: by roots of polynomials $R_{2j}$ ($j=-D/2$) are marked with circles. Note that
981: the trajectories of the ground state and of the first excited state meet
982: together at $D\approx-2.6$ . Some polynomials $R_{2j}$ have complex roots so
983: that the number of real levels associated with these polynomials is smaller
984: than $2j+1$. The dashed horizontal lines connect levels related by the
985: symmetry $D\rightarrow4-D$ (for even $D$).}%
986: \label{trajectories-AO-fig}%
987: \end{figure}
988: %EndExpansion
989: 
990: \subsection{Various aspects of the problem of analytical continuation in
991: $\mathcal{D}$}
992: 
993: \label{Various-analytical-aspects}
994: 
995: The problem of analytical continuation in $\mathcal{D}$ can be studied for
996: various quantities:
997: 
998: 1) analytical continuation of separate energy levels
999: \begin{equation}
1000: E_{nl}(D)=E_{n0}(D+2l)=E_{n0}(\mathcal{D})\,,
1001: \end{equation}
1002: .
1003: 
1004: 2) analytical continuation of coefficients $E_{n}^{(k)}(\mathcal{D})$ of the
1005: perturbative expansion%
1006: \begin{equation}
1007: E_{n0}(\mathcal{D},g)=\sum\limits_{k=1}^{\infty}E_{n0}^{(k)}(\mathcal{D})g^{k}
1008: \label{E-n-g-epansion}%
1009: \end{equation}
1010: for anharmonic oscillator,
1011: 
1012: 3) analytical continuation of the partition function%
1013: \begin{equation}
1014: Z(\beta,D)=\mathrm{Tr}\,\exp\left(  -\beta H\right)  \,. \label{Z-beta-def}%
1015: \end{equation}
1016: Let us comment briefly on the analytical properties of these quantities.
1017: 
1018: The situation with analytical continuation of the perturbative coefficients
1019: $E_{n0}^{(k)}(\mathcal{D})$ in expansion (\ref{E-n-g-epansion}) is simple
1020: because $E_{n0}^{(k)}(\mathcal{D})$ are polynomials of $\mathcal{D}$. However,
1021: the asymptotic nature of the perturbative series (\ref{E-n-g-epansion}) does
1022: not allow us to draw simple conclusions about the analyticity in $\mathcal{D}$
1023: of exact levels $E_{n0}(\mathcal{D},g)$ from the trivial analyticity of
1024: $E_{n0}^{(k)}(\mathcal{D}).$
1025: 
1026: The simplest example of analytical continuation in $D$ provides the harmonic
1027: oscillator. Its partition function trivially factorizes into $D$ components%
1028: \begin{equation}
1029: Z^{(0)}(\beta,D,0)=\left[  Z^{(0)}(\beta,1,0)\right]  ^{D}%
1030: \end{equation}
1031: with an obvious analyticity in $D$. The case of anharmonic potentials will be
1032: considered in the next section.
1033: 
1034: \subsection{Analytical continuation of the partition function}
1035: 
1036: \label{Analytical-PF-section}
1037: 
1038: Now let us study analytical continuation of the partition function%
1039: \begin{equation}
1040: Z(\beta,D)=\mathrm{Tr}\exp\left(  -\beta H\right)  \quad(D=1,2,3,\ldots)
1041: \end{equation}
1042: in $D$. At physical values of $D=1,2,3,4,\ldots$ one can use the $l$
1043: decomposition%
1044: \begin{equation}
1045: Z(\beta,D)=\sum\limits_{n,l=0}^{\infty}m(D,l)\exp\left[  -\beta E_{nl}%
1046: (D)\right]  \,. \label{Z-D-l-sum-1}%
1047: \end{equation}
1048: Here $m(D,l)$ is the degeneracy of the $l$-levels in the $D$-dimensional space
1049: given by eq. (\ref{l-degeneracy-positive}). Using eq. (\ref{E-D-l}), we find
1050: from eq. (\ref{Z-D-l-sum-1})%
1051: 
1052: \begin{equation}
1053: Z(\beta,D)=\sum\limits_{l=0}^{\infty}m(D,l)\left\{  \sum_{n}\exp\left[  -\beta
1054: E_{n0}(D+2l)\right]  \right\}  \,.
1055: \end{equation}
1056: Let us define%
1057: \begin{equation}
1058: z(\beta,D)=\sum\limits_{n=0}^{\infty}\exp\left[  -\beta E_{n0}(D)\right]  \,.
1059: \end{equation}
1060: Then%
1061: \begin{equation}
1062: Z(\beta,D)=\sum\limits_{l=0}^{\infty}m(D,l)\,z(\beta,D+2l)\,.
1063: \label{Z-D-l-sum-2}%
1064: \end{equation}
1065: We want to use this series for analytical continuation of $Z(\beta,D)$ to
1066: negative $D$. It is easy to check that for positive integer $l=0,1,2,3\ldots$,
1067: function $m(D,l)$ is analytical in $D$ in the whole complex plane. Analytical
1068: continuation of $m(D,l)$ to $D=1$ and $D=2$ was already discussed in Sec.
1069: \ref{Schroedinger-section}. Now let us consider the general case. Using the
1070: relation%
1071: \begin{equation}
1072: \Gamma(z)\Gamma(1-z)=\frac{\pi}{\sin\pi z}\,,
1073: \end{equation}
1074: we can transform eq. (\ref{l-degeneracy-positive}) to the form%
1075: 
1076: \begin{equation}
1077: m(D,l)=(-1)^{l}\frac{(2l+D-2)}{l!}\frac{\Gamma(2-D)}{\Gamma(3-D-l)}\,.
1078: \label{m-D-l-2}%
1079: \end{equation}
1080: Combining representations (\ref{l-degeneracy-positive}) and (\ref{m-D-l-2}),
1081: we see that $m(D,l)$ is regular for all $D$. The regularity of $m(D,l)$ allows
1082: us to use representation (\ref{Z-D-l-sum-2}) for analytical continuation of
1083: $Z(D)$ to negative $D$.
1084: 
1085: At negative even values $D=-2M=0,-2,-4,\ldots$ we have an important
1086: simplification. Indeed, according to eq. (\ref{m-D-l-2})%
1087: \begin{equation}
1088: m(-2M,l)=(-1)^{l+1}\frac{(2l-2M-2)}{l!}\frac{\Gamma(2+2M)}{\Gamma(3+2M-l)}\,.
1089: \end{equation}
1090: We see that%
1091: \begin{equation}
1092: m(-2M,l)=0\quad\mathrm{if}\quad l\geq2M+3\,.
1093: \end{equation}
1094: Therefore series (\ref{Z-D-l-sum-2}) reduces to a finite sum:%
1095: \begin{equation}
1096: Z(\beta,-2M)=\sum\limits_{l=0}^{2(M+1)}(-1)^{l}\frac{2(-l+M+1)}{l!}%
1097: \frac{(1+2M)!}{(2+2M-l)!}\,\,z(\beta,-2M+2l)\,.
1098: \end{equation}
1099: Now we change the summation variable from $l$ to%
1100: \begin{equation}
1101: k=l-M-1\,.
1102: \end{equation}
1103: Then%
1104: \begin{equation}
1105: Z(\beta,-2M)=\sum\limits_{k=-M-1}^{M+1}\frac{(-1)^{k-M}2k\left[
1106: (1+2M)!\right]  }{(M+1-k)!(1+M+k)!}\,\,z(\beta,2+2k)\,.
1107: \end{equation}
1108: Note that $k=0$ does not contribute to this sum. Combining the contributions
1109: of $k$ and $-k$ we arrive at%
1110: \begin{align}
1111: Z(\beta,-2M)  &  =\sum\limits_{k=1}^{M+1}\frac{(-1)^{k-M}2k\left[
1112: (1+2M)!\right]  }{(M+1-k)!(1+M+k)!}\nonumber\\
1113: &  \times\,\,\left[  z(\beta,2+2k)-z(\beta,2-2k)\right]  \,.
1114: \label{Z-D-l-sum-3}%
1115: \end{align}
1116: Now we use the $D\rightarrow4-D$ symmetry (\ref{D-4-D-symmetry-phi})
1117: connecting the levels of $z(\beta,2+2k)$ and $z(\beta,2-2k)$. As was explained
1118: in Sec. \ref{D-4-D-symmetry-section}, this symmetry is partial: when one
1119: passes from $D=2-2k$ to $D=2+2k$ one loses the states described by the
1120: solutions of the algebraic equation $R_{k-1}(E)=0$. Therefore%
1121: \begin{equation}
1122: z(\beta,2-2k)=z(\beta,2+2k)+\left[  \mathrm{Tr}\exp\left(  -\beta
1123: H_{j=-(k-1)/2}\right)  \right]
1124: \end{equation}
1125: where $H_{j}$ is the matrix Hamiltonian (\ref{H-J-def}) corresponding to the
1126: spin%
1127: \begin{equation}
1128: j=\frac{k-1}{2}\,. \label{j-via-n}%
1129: \end{equation}
1130: Now eq. (\ref{Z-D-l-sum-3}) takes the form%
1131: \begin{align}
1132: Z(\beta,-2M)  &  =\sum\limits_{k=1}^{M+1}\frac{(-1)^{k-M+1}2k\left[
1133: (1+2M)!\right]  }{(M+1-k)!(1+M+k)!}\\
1134: &  \times\mathrm{Tr}\exp\left(  -\beta H_{j=(k-1)/2}\right)  \,.
1135: \end{align}
1136: Let us replace the summation variable $k$ to $j$ (\ref{j-via-n})%
1137: \begin{equation}
1138: Z(\beta,-2M)=\sum\limits_{j=0}^{M/2}(-1)^{2j-M}\frac{2(2j+1)\left[
1139: (1+2M)!\right]  }{(M-2j)!(2+M+2j)!}\,\mathrm{Tr}\exp\left(  -\beta
1140: H_{j}\right)  \,. \label{Z-m2M-res}%
1141: \end{equation}
1142: Here the summation runs over integer and half-integer $j$.
1143: 
1144: \section{Fermion representation}
1145: 
1146: \label{Fermion-section}
1147: 
1148: \setcounter{equation}{0} 
1149: 
1150: \subsection{Derivation}
1151: 
1152: Let us start from the case of the quartic anharmonic oscillator
1153: (\ref{H-quartic}). We can write the path integral for the partition function%
1154: 
1155: \begin{equation}
1156: Z^{\mathrm{QO}}(\beta,D,g)=\int\limits_{q(0)=q(\beta)}Dx\exp\left\{  -\int
1157: _{0}^{\beta}dt\left[  \frac{1}{2}\sum\limits_{i=1}^{D}\left(  \dot{x}_{i}%
1158: ^{2}+x_{i}^{2}\right)  +g\left(  \sum\limits_{i=1}^{D}x_{i}^{2}\right)
1159: ^{2}\right]  \right\}  \,. \label{path-int-1}%
1160: \end{equation}
1161: One can use the standard trick with the auxiliary field $\sigma$%
1162: \begin{gather}
1163: Z^{\mathrm{QO}}(\beta,D,g)=\int\limits_{q(0)=q(T)}Dx\int D\sigma\nonumber\\
1164: \times\exp\left\{  -\int_{0}^{\beta}dt\left[  \frac{1}{2}\sum\limits_{i=1}%
1165: ^{D}x_{i}\left(  -\partial_{t}^{2}+1+i\sqrt{8g}\sigma\right)  x_{i}%
1166: +\frac{\sigma^{2}}{2}\right]  \right\} \nonumber\\
1167: =\int D\sigma\left[  \mathrm{Det}\left(  -\partial_{t}^{2}+1+i\sqrt{8g}%
1168: \sigma\right)  \right]  ^{-D/2}\exp\left(  -\int_{0}^{\beta}dt\frac{\sigma
1169: ^{2}}{2}\right)  \,.
1170: \end{gather}
1171: For negative even $D$%
1172: \begin{equation}
1173: D=-2M \label{D-M}%
1174: \end{equation}
1175: we can use the fermion representation%
1176: \begin{align}
1177: &  \left[  \mathrm{Det}\left(  -\partial_{t}^{2}+1+i\sqrt{8g}\sigma\right)
1178: \right]  ^{-D/2}\nonumber\\
1179: &  =\int D\psi^{+}D\psi\exp\left\{  -\int_{0}^{\beta}dt\left[  \sum
1180: \limits_{i=1}^{M}\psi_{i}^{+}\left(  -\partial_{t}^{2}+1+i\sqrt{8g}%
1181: \sigma\right)  \psi_{i}\right]  \right\}  \,. \label{Det-int-psi}%
1182: \end{align}
1183: Integrating out the field $\sigma$, we obtain%
1184: \begin{align}
1185: &  Z^{\mathrm{QO}}(\beta,D,g)\overset{D\rightarrow-2M}{=}\int D\psi^{+}%
1186: D\psi\nonumber\\
1187: &  \times\exp\left\{  -\int_{0}^{\beta}dt\left[  \sum\limits_{i=1}^{M}\left[
1188: \left(  \partial_{t}\psi_{i}^{+}\right)  \left(  \partial_{t}\psi_{i}\right)
1189: +\psi_{i}^{+}\psi_{i}\right]  +4g\left(  \sum\limits_{i=1}^{M}\psi_{i}^{+}%
1190: \psi_{i}\right)  ^{2}\right]  \right\}  \,.
1191: \end{align}
1192: 
1193: The same trick can be repeated (at least formally) for an arbitrary polynomial
1194: central potential $V(r^{2})$ if one uses more intricate integral
1195: representations involving several auxiliary boson fields $\sigma$. In this way
1196: one arrives at the fermion representation for partition function
1197: (\ref{Z-beta-def}) with an arbitrary polynomial $V$:
1198: \begin{align}
1199: &  Z(\beta,D)\overset{D\rightarrow-2M}{=}\int D\psi^{+}D\psi\nonumber\\
1200: &  \times\exp\left\{  -\int_{0}^{\beta}dt\left[  \sum\limits_{i=1}^{M}\left[
1201: \left(  \partial_{t}\psi_{i}^{+}\right)  \left(  \partial_{t}\psi_{i}\right)
1202: \right]  +V\left(  2\sum\limits_{i=1}^{M}\psi_{i}^{+}\psi_{i}\right)  \right]
1203: \right\}  \,. \label{Z-fermion-general}%
1204: \end{align}
1205: 
1206: Now we introduce the auxiliary fermion field $\chi_{i}$:%
1207: \begin{align}
1208: &  \int D\chi^{+}D\chi\exp\left\{  \int_{0}^{\beta}dt\sum\limits_{i=1}%
1209: ^{M}\left[  \left(  \partial_{t}\psi_{i}^{+}\right)  \chi_{i}-\chi_{i}%
1210: ^{+}\left(  \partial_{t}\psi_{i}\right)  -\chi_{i}^{+}\chi_{i}\right]
1211: \right\} \nonumber\\
1212: &  =\exp\left[  -\int_{0}^{\beta}dt\sum\limits_{i=1}^{M}\left(  \partial
1213: _{t}\psi_{i}^{+}\right)  \left(  \partial_{t}\psi_{i}\right)  \right]  \,.
1214: \end{align}
1215: Then%
1216: \begin{align}
1217: &  Z(\beta,D)\overset{D\rightarrow-2M}{=}\int D\psi^{+}D\psi D\chi^{+}%
1218: D\chi\nonumber\\
1219: &  \times\exp\left\{  \int_{0}^{\beta}dt\left[  \sum\limits_{i=1}^{M}\left[
1220: \left(  \partial_{t}\psi_{i}^{+}\right)  \chi_{i}-\chi_{i}^{+}\left(
1221: \partial_{t}\psi_{i}\right)  -\chi_{i}^{+}\chi_{i}\right]  \right.  \right.
1222: \nonumber\\
1223: &  \left.  \left.  -V\left(  2\sum\limits_{i=1}^{M}\psi_{i}^{+}\psi
1224: _{i}\right)  \right]  \right\}  \,.
1225: \end{align}
1226: Next we change the notation for the integration variables:%
1227: \begin{align}
1228: \psi_{i}  &  \rightarrow a_{i1}\,,\\
1229: \psi_{i}^{+}  &  \rightarrow a_{i2}^{+}\,,\\
1230: \chi_{i}  &  \rightarrow a_{i2}\,,\\
1231: \chi_{i}^{+}  &  \rightarrow a_{i1}^{+}%
1232: \end{align}
1233: so that%
1234: \begin{align}
1235: &  Z(\beta,D)\overset{D\rightarrow-2M}{=}\int Da^{+}Da\nonumber\\
1236: &  \times\exp\left\{  -\int_{0}^{\beta}dt\left[  \sum\limits_{i=1}^{M}%
1237: \sum\limits_{\alpha,\beta=1}^{2}a_{i\alpha}^{+}\left(  \delta_{\alpha\beta
1238: }\partial_{t}+\frac{1}{2}\left(  \sigma_{1}+i\sigma_{2}\right)  _{\alpha\beta
1239: }\right)  a_{i\beta}\right.  \right. \nonumber\\
1240: &  \left.  \left.  +V\left(  \sum\limits_{i=1}^{M}\sum\limits_{\alpha,\beta
1241: =1}^{2}a_{i\alpha}^{+}\left(  \sigma_{1}-i\sigma_{2}\right)  _{\alpha\beta
1242: }a_{i\beta}\right)  \right]  \right\}  \,. \label{Z-Grassmann}%
1243: \end{align}
1244: where $\sigma_{a}$ are standard Pauli matrices.
1245: 
1246: The integral on the RHS corresponds to the effective fermion Hamiltonian%
1247: \begin{equation}
1248: H_{F}=\sum\limits_{i=1}^{M}\sum\limits_{\alpha,\beta=1}^{2}\frac{1}%
1249: {2}a_{i\alpha}^{+}\left(  \sigma_{1}+i\sigma_{2}\right)  _{\alpha\beta
1250: }a_{i\beta}+V\left(  \sum\limits_{i=1}^{M}\sum\limits_{\alpha,\beta=1}%
1251: ^{2}a_{i\alpha}^{+}\left(  \sigma_{1}-i\sigma_{2}\right)  _{\alpha\beta
1252: }a_{i\beta}\right)  \label{H-F-1}%
1253: \end{equation}
1254: with anticommutation relations%
1255: \begin{align}
1256: \left\{  a_{m\alpha},a_{n\beta}^{+}\right\}   &  =\delta_{mn}\delta
1257: _{\alpha\beta}\,,\\
1258: \left\{  a_{m\alpha},a_{n\beta}\right\}   &  =\left\{  a_{m\alpha}%
1259: ^{+},a_{n\beta}^{+}\right\}  =0\,.
1260: \end{align}
1261: 
1262: Thus we have derived the formula%
1263: \begin{equation}
1264: Z(\beta,D)\overset{D\rightarrow-2M}{=}\mathrm{Tr}\,\left[  e^{-\beta H_{F}%
1265: }P_{F}\right]  \label{Tr-continuation}%
1266: \end{equation}
1267: which should be understood in the sense of analytical continuation of the
1268: bosonic partition function $\mathrm{Tr}\,e^{-\beta H}$ from positive integer
1269: dimensions to negative even dimensions. On the RHS we have a trace in a
1270: $2^{2M}$ dimensional vector space. $P_{F}$ is the operator counting the
1271: fermion parity of the state:%
1272: \begin{align}
1273: P_{F}  &  =(-1)^{N_{F}+M}\,,\\
1274: N_{F}  &  =\sum\limits_{i,\alpha}a_{i\alpha}^{+}a_{i\alpha}\,.
1275: \end{align}
1276: $P_{F}$ appears in eq. (\ref{Tr-continuation}) because the integral over
1277: Grassmann variables (\ref{Z-Grassmann}) inherits periodic boundary conditions
1278: of the original boson integral (\ref{path-int-1}).
1279: 
1280: Now we define%
1281: \begin{equation}
1282: \tilde{T}_{a}=\frac{1}{2}\sum\limits_{i=1}^{M}\sum\limits_{\alpha,\beta=1}%
1283: ^{2}a_{i\alpha}^{+}\left(  \sigma_{a}\right)  _{\alpha\beta}a_{i\beta}\,.
1284: \end{equation}
1285: Obviously%
1286: \begin{equation}
1287: \left[  \tilde{T}_{a},\tilde{T}_{b}\right]  =i\varepsilon_{abc}\tilde{T}_{c}%
1288: \end{equation}
1289: Introducing $sl(2)$ generators%
1290: \begin{align}
1291: \tilde{T}_{\pm}  &  =\tilde{T}_{1}\pm i\tilde{T}_{2}\,,\\
1292: \tilde{T}_{0}  &  =\tilde{T}_{3}%
1293: \end{align}
1294: we can write Hamiltonian (\ref{H-F-1}) as%
1295: \begin{equation}
1296: H_{F}=\tilde{T}_{+}+V(2\tilde{T}_{-})\,. \label{H-F-def}%
1297: \end{equation}
1298: This effective Hamiltonian has a form similar to form (\ref{H-T-def}).
1299: 
1300: \subsection{Degeneracy factors}
1301: 
1302: Note that in the first derivation of the spin Hamiltonian $H_{j}$
1303: (\ref{H-J-def}) via the $sl(2)$ expression $H_{T}$ (\ref{H-T-def}) we worked
1304: with the ``effective dimension'' $\mathcal{D}$ given by (\ref{Delta-def})
1305: keeping in mind that the spectrum depends on $D$ and $l$ only via
1306: $\mathcal{D}$. On the contrary, in the derivation based on the fermion
1307: Hamiltonian $H_{F}$ (\ref{H-F-def}) we can trace the contributions of levels
1308: with different $l$. Strictly speaking, $H_{F}$ (\ref{H-F-def}) is not
1309: completely equivalent to $H_{j}$ (\ref{H-J-def}) because $H_{F}$ acts in the
1310: $2^{2M}$ dimensional space corresponding to the reducible representation of
1311: $sl(2)$ (or $su(2)$ if one prefers a more familiar physical interpretation)%
1312: \begin{equation}
1313: \bigotimes\limits_{i=1}^{M}\left(  [0]%
1314: {\textstyle\bigoplus}
1315: [0]%
1316: {\textstyle\bigoplus}
1317: \left[  \frac{1}{2}\right]  \right)  \,. \label{tensor-product-1}%
1318: \end{equation}
1319: Here we use notation $[j]$ for irreducible representations corresponding to
1320: spin $j$. Each factor $[0]%
1321: {\textstyle\bigoplus}
1322: [0]%
1323: {\textstyle\bigoplus}
1324: \left[  \frac{1}{2}\right]  $ in (\ref{tensor-product-1}) is associated with
1325: the states in the $i$-th sector:%
1326: \begin{align}
1327: a_{i\beta}^{+}|0\rangle &  \rightarrow\left[  \frac{1}{2}\right]
1328: \,,\label{one-particle-states}\\
1329: |0\rangle &  \rightarrow\lbrack0]\,,\\
1330: a_{i1}^{+}a_{i2}^{+}|0\rangle &  \rightarrow\lbrack0]\,.
1331: \end{align}
1332: 
1333: Tensor product (\ref{tensor-product-1}) can be decomposed in irreducible
1334: representations $[j]$%
1335: \begin{equation}
1336: \bigotimes\limits_{i=1}^{M}\left(  [0]%
1337: {\textstyle\bigoplus}
1338: [0]%
1339: {\textstyle\bigoplus}
1340: \left[  \frac{1}{2}\right]  \right)  =\bigoplus_{j=0}^{M/2}n(j,M)[j]
1341: \label{N-product}%
1342: \end{equation}
1343: with degeneracies%
1344: \begin{equation}
1345: n(j,M)=\frac{2\left(  2j+1\right)  \Gamma(2M+2)}{\Gamma\left(  M-2j+1\right)
1346: \Gamma(3+M+2j)} \label{n-j-M}%
1347: \end{equation}
1348: which can be computed using the recursion relation%
1349: \begin{equation}
1350: n(j,M+1)=2n(j,M)+n\left(  j-\frac{1}{2},M\right)  +n\left(  j+\frac{1}%
1351: {2},M\right)  \,.
1352: \end{equation}
1353: 
1354: Now we apply decomposition (\ref{N-product}) to the calculation of the
1355: partition function (\ref{Tr-continuation}):%
1356: \begin{equation}
1357: Z(\beta,D)\overset{D\rightarrow-2M}{=}\sum\limits_{j=0}^{M/2}(-1)^{2j-M}%
1358: n(j,M)\,\,\left[  \mathrm{Tr}\exp\left(  -\beta H_{j}\right)  \right]  \,.
1359: \end{equation}
1360: Inserting $n(j,M)$ from eq. (\ref{n-j-M}), we arrive at the result coinciding
1361: with (\ref{Z-m2M-res}). Thus fermion representation (\ref{Z-fermion-general})
1362: for the partition function leads to the same result as the direct calculation
1363: of Sec. \ref{Analytical-PF-section}.
1364: 
1365: \subsection{Connection with the results of Dunne and Halliday}
1366: 
1367: If one uses representation%
1368: \begin{align}
1369: a_{i2}^{+}  &  =\theta_{2i-1}\,,\\
1370: a_{i2}  &  =\frac{\partial}{\partial\theta_{2i-1}}\,,\\
1371: a_{i1}  &  =\theta_{2i}\,,\\
1372: a_{i1}^{+}  &  =\frac{\partial}{\partial\theta_{2i}}%
1373: \end{align}
1374: then Hamiltonian $H_{F}$ takes the form
1375: \begin{equation}
1376: H_{F}=\sum\limits_{i=1}^{M}\frac{\partial}{\partial\theta_{2i}}\frac{\partial
1377: }{\partial\theta_{2i-1}}+V\left(  \sum\limits_{i=1}^{M}\theta_{2i-1}%
1378: \theta_{2i}\right)
1379: \end{equation}%
1380: \begin{equation}
1381: H_{F}=\sum\limits_{i=1}^{M}\sum\limits_{\alpha,\beta=1}^{2}\frac{1}%
1382: {2}a_{i\alpha}^{+}\left(  \sigma_{1}+i\sigma_{2}\right)  _{\alpha\beta
1383: }a_{i\beta}+V\left(  \sum\limits_{i=1}^{M}\sum\limits_{\alpha,\beta=1}%
1384: ^{2}a_{i\alpha}^{+}\left(  \sigma_{1}-i\sigma_{2}\right)  _{\alpha\beta
1385: }a_{i\beta}\right)
1386: \end{equation}
1387: used by Dunne and Halliday \cite{DH-88}. The advantage of this form is the
1388: possibility to work with the pseudo-boson variables%
1389: \begin{align}
1390: \xi_{i}  &  =\theta_{2i-1}\theta_{2i}\,,\\
1391: \xi_{i}\xi_{k}  &  =-\xi_{k}\xi_{i}\,,\\
1392: \left(  \xi_{i}\right)  ^{2}  &  =0
1393: \end{align}
1394: and to look for the eigenstates of $H_{F}$ in the form of polynomials
1395: $\Psi(\{\xi_{i}\})$ in $\xi_{i}$.
1396: 
1397: In our representation these polynomial wave functions correspond to the states
1398: of the form%
1399: \begin{equation}
1400: P(\left\{  a_{i2}^{+}a_{i1}\right\}  )|\Omega\rangle\, \label{P-Omega-states}%
1401: \end{equation}
1402: where $P$ are polynomials of $a_{i2}^{+}a_{i1}$, and $|\Omega\rangle$ is fixed
1403: by the conditions%
1404: \begin{equation}
1405: a_{i1}^{+}|\Omega\rangle=a_{i2}|\Omega\rangle=0\quad(1\leq i\leq M)\,.
1406: \end{equation}
1407: This corresponds to working in the subspace of states obeying the constraint%
1408: \begin{equation}
1409: \sum_{\beta=1}^{2}a_{i\beta}^{+}a_{i\beta}\psi=\psi\quad(1\leq i\leq M)\,.
1410: \end{equation}
1411: In terms of decomposition (\ref{N-product}) this constraint selects states
1412: (\ref{one-particle-states}) in each $i$-sector. Therefore states
1413: (\ref{P-Omega-states}) belong to the subspace:%
1414: \begin{equation}
1415: \bigotimes\limits_{i=1}^{M}\left[  \frac{1}{2}\right]  =\bigoplus
1416: _{j=\{M/2\}}^{M/2}\tilde{n}_{j}(j,M)[j]\,, \label{DH-decomposition}%
1417: \end{equation}
1418: where $\{x\}$ stands for the fractional part of $x$, and the degeneracy
1419: factors%
1420: \begin{equation}
1421: \tilde{n}(j,M)=\frac{(-1)^{M+2j}+1}{2}\frac{(2j+1)M!}{\left(  \frac{M}%
1422: {2}+j+1\right)  !\left(  \frac{M}{2}-j\right)  !}%
1423: \end{equation}
1424: follow from the recursion relation%
1425: 
1426: \begin{equation}
1427: \tilde{n}(j,M+1)=\tilde{n}\left(  j-\frac{1}{2},M\right)  +\tilde{n}\left(
1428: j+\frac{1}{2},M\right)  \,.
1429: \end{equation}
1430: Since Dunne and Halliday \cite{DH-88} (working with coupling constant
1431: $\lambda=8g$) compute their characteristic polynomials $P_{M}(E,\lambda)$ in
1432: the subspace associated with decomposition (\ref{DH-decomposition}), their
1433: polynomials $P_{M}(E,\lambda)$ factorize into our polynomials $R_{M}%
1434: (E,\lambda)$ (\ref{R-2j-def}):%
1435: \begin{equation}
1436: \left[  P_{M}(E,\lambda)\right]  _{\lambda=8g}=\mathrm{const}\,\prod
1437: \limits_{j=\{M/2\}}^{M/2}\left[  R_{2j}(E,g)\right]  ^{\tilde{n}_{j}}\,.
1438: \end{equation}
1439: For example,%
1440: \begin{align*}
1441: P_{1}  &  =\mathrm{const}\,R_{1}\,,\\
1442: P_{2}  &  =\mathrm{const}\,R_{0}R_{2}\,,\\
1443: P_{3}  &  =\mathrm{const}\,R_{1}^{2}R_{3}\,,\\
1444: P_{4}  &  =\mathrm{const}\,R_{0}^{2}R_{2}^{3}R_{4}\,.
1445: \end{align*}
1446: In this way the $sl(2)$ algebra explains the ``miracles'' observed in
1447: \cite{DH-88}.
1448: 
1449: \section{Large orders of the perturbation theory}
1450: 
1451: \setcounter{equation}{0} 
1452: 
1453: \label{Higher-orders-section}
1454: 
1455: \subsection{Disappearance of the factorial divergence at negative even $D$}
1456: 
1457: \label{No-factorials-subsection}
1458: 
1459: In a general case the perturbative series for the energy of the ground state
1460: of the $D$-dimensional quartic anharmonic oscillator (\ref{H-quartic})%
1461: 
1462: \begin{equation}
1463: E(g,D)=\sum\limits_{k=0}^{\infty}E^{(k)}(D)g^{k} \label{E-g-expansion}%
1464: \end{equation}
1465: has factorially growing coefficients \cite{Zinn-Justin-81a}:%
1466: \begin{align}
1467: &  E^{(k)}(D)\overset{k\rightarrow\infty}{=}(-1)^{k+1}\Gamma\left(
1468: k+\frac{D}{2}\right)  3^{k+\frac{D}{2}}\frac{2^{D/2}}{\pi\Gamma\left(
1469: D/2\right)  }\nonumber\\
1470: &  \times\left[  1-\frac{1}{6k}\left(  \frac{5}{3}+\frac{9}{2}D+\frac{7}%
1471: {4}D^{2}\right)  +O(k^{-2})\right]  \,. \label{E-k-N-asymptotic}%
1472: \end{align}
1473: However, at negative even values of $D$ function $\left[  \Gamma\left(
1474: D/2\right)  \right]  ^{-1}$ vanishes. Note that $\left[  \Gamma\left(
1475: D/2\right)  \right]  ^{-1}$ appears on the RHS as a common factor so that at
1476: negative even values of $D$ all terms of the systematic expansion in $1/k$
1477: become zero. This hints that at $D=0,-2,-4$ the factorial growth of
1478: perturbative coefficients disappears. The same happens in for the potential%
1479: \begin{equation}
1480: V(r^{2})=\frac{1}{2}r^{2}+gr^{2N}\,.
1481: \end{equation}
1482: In this case the factorial structure is different \cite{BLZ-77}%
1483: \begin{align}
1484: &  E^{(k)}(D)\overset{k\rightarrow\infty}{=}\left[  k(N-1)\right]  !\left[
1485: -\frac{1}{\pi}\frac{(N-1)^{D/2}}{\Gamma(D/2)}\right]  \left[  \frac{\Gamma
1486: \left(  \frac{2N}{N-1}\right)  }{\Gamma^{2}\left(  \frac{N}{N-1}\right)
1487: }\right]  ^{D/2}\nonumber\\
1488: &  \times k^{(D/2)-1}\left\{  \left[  -\frac{1}{2}\frac{\Gamma\left(
1489: \frac{2N}{N-1}\right)  }{\Gamma^{2}\left(  \frac{N}{N-1}\right)  }\right]
1490: ^{N-1}\right\}  ^{k}\left[  1+O(k)\right]  \,.
1491: \end{align}
1492: but we still have the same factor of $\left[  \Gamma\left(  D/2\right)
1493: \right]  ^{-1}$.
1494: 
1495: This factor of $\left[  \Gamma\left(  D/2\right)  \right]  ^{-1}$ also appears
1496: in the asymptotic expression for $E^{(k)}(D)$ in the $O(D)$ symmetric model
1497: studied in Ref. \cite{RS-95}. This model describes a particle on a
1498: $D$-dimensional sphere with a potential breaking the $O(D+1)$ symmetry of the
1499: sphere to $O(D)$.
1500: 
1501: One can easily trace the origin of this universal factor $\left[
1502: \Gamma\left(  D/2\right)  \right]  ^{-1}$. In the semiclassical path integral
1503: derivation of large-order asymptotic formulas, the common factor of $\left[
1504: \Gamma\left(  D/2\right)  \right]  ^{-1}$ comes from the integration over
1505: rotational collective coordinates of the classical solution breaking the
1506: $O(D)$ symmetry of the problem to $O(D-1)$. As a result of this integration,
1507: the final result is proportional to the surface area of $(D-1$)-dimensional
1508: sphere in the $D$-dimensional space:%
1509: \begin{equation}
1510: S_{D}=\frac{2\,\pi^{D/2}}{\Gamma(D/2)}\,.
1511: \end{equation}
1512: 
1513: Let us study the asymptotic behavior of $E^{(k)}(D)$ at negative even $D=-2M$.
1514: We will consider the case when the potential has the form%
1515: \begin{equation}
1516: V(r)=\frac{1}{2}r^{2}+gU(r^{2})
1517: \end{equation}
1518: where $U(r^{2})$ is some polynomial. We know that at $D=-2M$ the energy of the
1519: ground state (understood as usual in the sense of analytical continuation) is
1520: given (at least for small $g$) by one of the roots of the algebraic equation
1521: (\ref{R-M-g-spectrum-eq}). Since $R_{M}(E,g)$ is a polynomial in both $g$ and
1522: $E$, the resulting dependence of $E(g,-2M)$ is regular for most values of $g$.
1523: The radius of convergence of the perturbative expansion of $E(g,-2M)$ is
1524: controlled by the singularity of $E(g,-2M)$ closest to $g=0$. These
1525: singularities appear at values of $g$ corresponding to degenerate roots of
1526: polynomials $R_{M}(E,g)$ which are described by equations%
1527: \begin{equation}
1528: R_{M}(E_{0},g_{0})=\left[  \frac{\partial}{\partial E}R_{M}(E,g_{0})\right]
1529: _{E=E_{0}}=0\,. \label{R-M-degenerate-root}%
1530: \end{equation}
1531: Note that at $g=0$ polynomial $R_{M}(E,g)$ is given by expression
1532: (\ref{det-H-H-HO}) which has no degenerate roots so that the point $g=0$ is
1533: regular and we have a nonzero convergence radius. Equations
1534: (\ref{R-M-degenerate-root}) may have several solutions. One has to choose the
1535: solution corresponding to the singularity closest to the point $g=0$ (on the
1536: main Riemann sheet). In the vicinity of the degenerate point $g_{0},E_{0}$ we
1537: can use the Taylor expansion%
1538: \begin{equation}
1539: R_{M}(E,g)=\frac{1}{2}(E-E_{0})^{2}\left[  \frac{\partial^{2}}{\partial E^{2}%
1540: }R_{M}(E,g_{0})\right]  _{E=E_{0}}+(g-g_{0})\left[  \frac{\partial}{\partial
1541: g}R_{M}(E_{0},g)\right]  _{g=g_{0}}+\ldots
1542: \end{equation}
1543: Now equation (\ref{R-M-g-spectrum-eq}) leads to the following expression valid
1544: in the vicinity of the singularity of $E(g)$ at $g=g_{0}$ is described by
1545: equation%
1546: \begin{equation}
1547: E(g)\overset{g\rightarrow g_{0}}{=}E_{0}-\sqrt{c\left(  1-\frac{g}{g_{0}%
1548: }\right)  } \label{E-root-singularity}%
1549: \end{equation}
1550: where%
1551: \begin{equation}
1552: c=-2g_{0}\left[  \frac{\partial}{\partial g}R_{M}(E_{0},g)\right]  _{g=g_{0}%
1553: }\left\{  \left[  \frac{\partial^{2}}{\partial E^{2}}R_{M}(E,g_{0})\right]
1554: _{E=E_{0}}\right\}  ^{-1}\,. \label{c-via-R-M}%
1555: \end{equation}
1556: Expanding%
1557: \begin{equation}
1558: \sqrt{1-\frac{g}{g_{0}}}=-\frac{1}{2\sqrt{\pi}}\sum\limits_{k=0}^{\infty
1559: }\frac{\Gamma\left(  k-\frac{1}{2}\right)  }{\Gamma\left(  k+1\right)
1560: }\left(  \frac{g}{g_{0}}\right)  ^{k}%
1561: \end{equation}
1562: and using the limit $k\rightarrow\infty$%
1563: \begin{equation}
1564: \frac{\Gamma\left(  k-\frac{1}{2}\right)  }{\Gamma\left(  k+1\right)
1565: }\rightarrow k^{-3/2}\,,
1566: \end{equation}
1567: we find from eq. (\ref{E-root-singularity})%
1568: \begin{equation}
1569: E^{(k)}(D)\overset{k\rightarrow\infty}{\longrightarrow}\frac{1}{2}%
1570: \sqrt{\frac{c(D)}{\pi}}k^{-3/2}\left[  g_{0}(D)\right]  ^{-k}\,\quad(D=-2M).
1571: \label{E-k-D-asymptotic-D-special}%
1572: \end{equation}
1573: 
1574: \subsection{Cases $D=0,-2,-4,-6$}
1575: 
1576: Now we want to concentrate on the case of the quartic anharmonic oscillator
1577: (\ref{H-quartic}) and to consider cases $D=0,-2,-4,-6$.
1578: 
1579: The values $D=0$ and $D=-2$ are trivial because the corresponding polynomials
1580: $\tilde{R}_{0}$ (\ref{R-0}) and $\tilde{R}_{1}$ (\ref{R-1}) do not depend on
1581: $g$ so that the energy is given by%
1582: \begin{align}
1583: E(0)  &  =0\,,\\
1584: E(-2)  &  =-1
1585: \end{align}
1586: and almost all perturbative coefficients $E^{(k)}$ vanish:%
1587: \begin{align}
1588: E^{(k)}(0)  &  =0\label{E-k-D0}\\
1589: E^{(k)}(-2)  &  =-\delta_{k0} \label{E-k-Dm2}%
1590: \end{align}
1591: 
1592: In the case $D=-4$ we insert polynomial $\tilde{R}_{2}(E,g)$ (\ref{R-2}) into
1593: eq. (\ref{R-M-degenerate-root}) and find two solutions%
1594: 
1595: \begin{equation}
1596: E_{0}(-4)=\mp\frac{2}{\sqrt{3}}\,,\quad g_{0}(-4)=\pm3^{-3/2}\,.
1597: \label{E0-g0-m4-uncertain}%
1598: \end{equation}
1599: Taking into account that we are interested in the energy of the ground state
1600: that corresponds to%
1601: 
1602: \begin{equation}
1603: \left[  E(g,D)\right]  _{g=0}=\frac{D}{2}\,,
1604: \end{equation}%
1605: \begin{equation}
1606: E(0,-4)=-2\,,
1607: \end{equation}
1608: one can check that the relevant singularity corresponds to the upper signs in
1609: eq. (\ref{E0-g0-m4-uncertain})%
1610: \begin{equation}
1611: E_{0}(-4)=-\frac{2}{\sqrt{3}}\,,\quad g_{0}(-4)=3^{-3/2}\,.
1612: \label{g0-solution}%
1613: \end{equation}
1614: Now we find from eq. (\ref{c-via-R-M})%
1615: \begin{equation}
1616: c(-4)=\frac{8}{9}\, \label{c-m4}%
1617: \end{equation}
1618: and insert this into eq. (\ref{E-k-D-asymptotic-D-special})%
1619: \begin{equation}
1620: E^{(k)}(-4)\overset{k\rightarrow\infty}{\longrightarrow}\frac{1}{3}%
1621: \sqrt{\frac{2}{\pi}}\sum\limits_{k\gg1}^{\infty}k^{-3/2}3^{3k/2}g^{k}
1622: \label{E-k-4-asymptotic}%
1623: \end{equation}
1624: for the coefficients $E^{(k)}(D)$ of expansion (\ref{E-series-0}) at $D=-4$.
1625: 
1626: Other cases of negative even $D$ can be analyzed in the same way. For example,
1627: solving equation (\ref{R-M-degenerate-root}) for the polynomial $R_{3}(E,g)$
1628: given by eq. (\ref{R-3}) we find for $D=-6$%
1629: \begin{align}
1630: g_{0}(-6)  &  =\frac{5\sqrt{13}-1}{36\sqrt{3\left(  5+2\sqrt{13}\right)  }%
1631: }\,,\label{g0-minus-6}\\
1632: E_{0}(-6)  &  =-\sqrt{\frac{1}{3}\left(  5+2\sqrt{13}\right)  }\,,\\
1633: c(-6)  &  =\frac{2}{9}\left(  5-\frac{1}{\sqrt{13}}\right)  \label{c-m-6}%
1634: \end{align}
1635: which results in the asymptotic behavior%
1636: \begin{equation}
1637: E^{(k)}(-6)\overset{k\rightarrow\infty}{\longrightarrow}\frac{1}{6}%
1638: \sqrt{\frac{2}{\pi}\left(  5-\frac{1}{\sqrt{13}}\right)  }k^{-3/2}\left[
1639: g_{0}(-6)\right]  ^{-k}\,. \label{E-k-6-asymptotic}%
1640: \end{equation}
1641: 
1642: \subsection{Roots of polynomials $E^{(k)}(D)$}
1643: 
1644: Coefficients $E^{(k)}(D)$ of the perturbative expansion (\ref{E-g-expansion})
1645: are polynomials of degree $k+1$. Let us denote the $k+1$ roots of these
1646: polynomials $\nu_{k,r}$:%
1647: \begin{equation}
1648: E^{(k)}(\nu_{k,r})=0\,\quad(1\leq r\leq k+1) \label{P-nu-root-eq}%
1649: \end{equation}
1650: According to eqs. (\ref{E-k-D0}), (\ref{E-k-Dm2}), at $k\geq1$ polynomials
1651: $E^{(k)}(D)$ have a common factor of $D(D+2)$, i.e.%
1652: 
1653: \begin{equation}
1654: E^{(k)}(D)=D(D+2)P_{k}(D)\,\, \label{E-via-P}%
1655: \end{equation}
1656: where $P_{k}(D)$ is a polynomial of degree $k-1$ with $k-1$ roots $\nu_{k,r}$.
1657: This means that the roots $\nu_{k,r}$ include values $0$ and $-2$:%
1658: \begin{equation}
1659: \left\{  \nu_{k,r}\right\}  =\{0,-2,\ldots\}
1660: \end{equation}
1661: 
1662: Note that the function $\left[  \Gamma\left(  D/2\right)  \right]  ^{-1}$ on
1663: the RHS of eq. (\ref{E-k-N-asymptotic}) has zeros at%
1664: \begin{equation}
1665: D=0,-2,-4,-6,\ldots
1666: \end{equation}
1667: The roots of the asymptotic expression (\ref{E-k-N-asymptotic}) at $D=0,-2$
1668: appear explicitly in the factorized expression on the RHS of eq.
1669: (\ref{E-via-P}). The other roots should appear asymptotically at large $k$. In
1670: other words, in the full set of roots $\nu_{k,r}$ there must be subsets
1671: converging to values $D=-4,-6,\ldots$
1672: 
1673: Let us study the first subset converging to
1674: \[
1675: D=-2M\,\quad(M=-2,-3,\ldots)
1676: \]
1677: We label this subset of roots with the index $r=r(k,-2M)$:%
1678: \begin{align}
1679: E^{(k)}(\nu_{k,r(k,-2M)})  &  =0\,,\nonumber\\
1680: \lim_{k\rightarrow\infty}\nu_{k,r(k,-2M)}  &  =-2M\,.
1681: \end{align}
1682: 
1683: We want to derive a large-$k$ asymptotic formula for $2M+\nu_{k,r(k,-2M)}$. To
1684: this aim we need an asymptotic formula for $E^{(k)}(D)$ in the double limit
1685: $D\rightarrow-2M$ and $k\rightarrow\infty$. This asymptotic formula has the
1686: form%
1687: \begin{equation}
1688: E^{(k)}(D)\overset{k\rightarrow\infty,\,D\rightarrow-2M}{\longrightarrow
1689: }(-1)^{k+1}\Gamma\left(  k+\frac{D}{2}\right)  3^{k+\frac{D}{2}}\frac{2^{D/2}%
1690: }{\pi\Gamma\left(  \frac{D}{2}\right)  }+E^{(k)}(-2M)\,.
1691: \label{E-k-asymptotic-modified}%
1692: \end{equation}
1693: Note that we dropped the $1/k$ correction present in eq.
1694: (\ref{E-k-N-asymptotic}) but added the extra term $E^{(k)}(-2M)$ which becomes
1695: essential at $D\rightarrow-2M$ because of the suppressing factor $\left[
1696: \Gamma\left(  \frac{D}{2}\right)  \right]  ^{-1}$ in the leading term. At
1697: $D\rightarrow-2M$ we can simplify eq. (\ref{E-k-asymptotic-modified})%
1698: \begin{equation}
1699: E^{(k)}(D)\overset{k\rightarrow\infty,\,D\rightarrow-2M}{\longrightarrow
1700: }(-1)^{k+1}k!k^{-M-1}3^{k-M}\frac{2^{-M}}{\pi\Gamma\left(  \frac{D}{2}\right)
1701: }+E^{(k)}(-2M)\,. \label{E-k-asymptotic-modified-2}%
1702: \end{equation}
1703: Here%
1704: \begin{equation}
1705: \Gamma\left(  \frac{D}{2}\right)  \overset{D\rightarrow-2M}{\longrightarrow
1706: }\frac{(-1)^{M}}{\frac{D}{2}+M}\frac{1}{M!}\,.
1707: \end{equation}
1708: Inserting this and eq. (\ref{E-k-D-asymptotic-D-special}) into eq.
1709: (\ref{E-k-asymptotic-modified-2}), we obtain%
1710: \begin{align}
1711: &  E^{(k)}(D)\overset{k\rightarrow\infty,\,D\rightarrow-2M}{\longrightarrow
1712: }(-1)^{k+1}k!k^{-M-1}3^{k-M}\frac{(-1)^{M}2^{-M}M!}{\pi}\left(  \frac{D}%
1713: {2}+M\right) \nonumber\\
1714: &  +\frac{1}{2}\sqrt{\frac{c(-2M)}{\pi}}k^{-3/2}\left[  g_{0}(-2M)\right]
1715: ^{-k}\,.
1716: \end{align}
1717: Using this asymptotic formula, we solve equation%
1718: \begin{equation}
1719: E^{(k)}(\nu_{k,r(k,-2M)})\overset{\nu_{k,r(k,-2M)}\rightarrow-2M}{=}0\,
1720: \end{equation}
1721: and find%
1722: \begin{equation}
1723: \nu_{k,r(k,-2M)}+2M\overset{k\rightarrow\infty}{\longrightarrow}\frac{\left(
1724: -6\right)  ^{M}}{M!}\sqrt{\pi c(-2M)}\frac{(-1)^{k}}{k!}k^{M-(1/2)}\left[
1725: 3g_{0}(-2M)\right]  ^{-k}\,.
1726: \end{equation}
1727: Using above expressions (\ref{E0-g0-m4-uncertain}), (\ref{c-m4}),
1728: (\ref{g0-minus-6}), (\ref{c-m-6}), for $c(-2M)$ and $g_{0}(-2M)$ at $M=2,3$,
1729: we obtain%
1730: \begin{equation}
1731: \nu_{k,r(k,-4)}+4\overset{k\rightarrow\infty}{\longrightarrow}12\sqrt{2\pi
1732: }\frac{(-1)^{k}}{k!}k^{3/2}3^{k/2}\,\,, \label{nu-k-asymptotic-4}%
1733: \end{equation}%
1734: \begin{equation}
1735: \nu_{k,r(k,-6)}+6\overset{k\rightarrow\infty}{\longrightarrow}12\sqrt
1736: {2\pi\left(  5-\frac{1}{\sqrt{13}}\right)  }\frac{(-1)^{k+1}}{k!}%
1737: k^{5/2}\left[  \frac{5\sqrt{13}-1}{12\sqrt{3\left(  5+2\sqrt{13}\right)  }%
1738: }\right]  ^{-k}\,. \label{nu-k-asymptotic-6}%
1739: \end{equation}
1740: 
1741: Asymptotic formulas (\ref{nu-k-asymptotic-4}), (\ref{nu-k-asymptotic-6}) agree
1742: with results of the direct numerical calculation using methods of Refs.
1743: \cite{BW-69,BW-73,SZ-79,Zinn-Justin-81a}. The roots approaching $-4$ appear
1744: rather early. They are listed in Table \ref{table-1}. For small odd values
1745: $k=5,7,9$ these roots have small imaginary parts. Starting from $k=10$, the
1746: roots $\nu_{k,r(k,4)}$ close to $-4$ become real and exhibit a very fast
1747: factorial convergence to $-4$ which is described by the asymptotic formula
1748: (\ref{nu-k-asymptotic-4}). Already at $k=11$ this formula works with accuracy
1749: better than $7\%$.
1750: 
1751: \begin{table}[ptb]%
1752: \begin{tabular}
1753: [c]{||l|l||l|l||}\hline\hline
1754: $k$ & $\nu_{k,r(k,-4)} +4 $ & $k$ & $\nu_{k,r(k,-4)} +4 $\\\hline
1755: $5$ & $-3.22834\pm i0.426293 $ & $33$ & $-4.79523448 \times10^{-26} $\\\hline
1756: $6$ & $-3.44545 $ & $34$ & $+2.55611474 \times10^{-27} $\\\hline
1757: $7$ & $-3.63083\pm i0.34226 $ & $35$ & $-1.32186040 \times10^{-28} $\\\hline
1758: $8$ & $-3.76443 $ & $36$ & $+6.63762841 \times10^{-30} $\\\hline
1759: $9$ & $-3.9583\pm i0.226557 $ & $37$ & $-3.23912410 \times10^{-31} $\\\hline
1760: $10$ & $+0.04231592827 $ & $38$ & $+1.53735607 \times10^{-32} $\\\hline
1761: $11$ & $-0.01231265412 $ & $39$ & $-7.10198670 \times10^{-34} $\\\hline
1762: $12$ & $+0.00178433080 $ & $40$ & $+3.19560318 \times10^{-35} $\\\hline
1763: $13$ & $-0.00027422590 $ & $41$ & $-1.40147930 \times10^{-36} $\\\hline
1764: $14$ & $+0.00003787462 $ & $42$ & $+5.99459657 \times10^{-38} $\\\hline
1765: $15$ & $-4.86252995 \times10^{-6} $ & $43$ & $-2.50229008 \times10^{-39}
1766: $\\\hline
1767: $16$ & $+5.80950053 \times10^{-7} $ & $44$ & $+1.01993415 \times10^{-40}
1768: $\\\hline
1769: $17$ & $-6.49387664 \times10^{-8} $ & $45$ & $-4.06166432 \times10^{-42}
1770: $\\\hline
1771: $18$ & $+6.81906230 \times10^{-9} $ & $46$ & $+1.58111340 \times10^{-43}
1772: $\\\hline
1773: $19$ & $-6.75145346 \times10^{-10}$ & $47$ & $-6.01961315 \times10^{-45}
1774: $\\\hline
1775: $20$ & $+6.32321355 \times10^{-11}$ & $48$ & $+2.24249068 \times10^{-46}
1776: $\\\hline
1777: $21$ & $-5.61842521 \times10^{-12}$ & $49$ & $-8.17805838 \times10^{-48}
1778: $\\\hline
1779: $22$ & $+4.74864227 \times10^{-13}$ & $50$ & $+2.92092144 \times10^{-49}
1780: $\\\hline
1781: $23$ & $-3.82680164 \times10^{-14}$ & $51$ & $-1.02217301 \times10^{-50}
1782: $\\\hline
1783: $24$ & $+2.94682238 \times10^{-15}$ & $52$ & $+3.50623616 \times10^{-52}
1784: $\\\hline
1785: $25$ & $-2.17261017 \times10^{-16}$ & $53$ & $-1.17934411 \times10^{-53}
1786: $\\\hline
1787: $26$ & $+1.53640755 \times10^{-17}$ & $54$ & $+3.89122545 \times10^{-55}
1788: $\\\hline
1789: $27$ & $-1.04388321 \times10^{-18}$ & $55$ & $-1.25990087 \times10^{-56}
1790: $\\\hline
1791: $28$ & $+6.82474923 \times10^{-20}$ & $56$ & $+4.00444478 \times10^{-58}
1792: $\\\hline
1793: $29$ & $-4.29961693 \times10^{-21}$ & $57$ & $-1.24982870 \times10^{-59}
1794: $\\\hline
1795: $30$ & $+2.61370107 \times10^{-22}$ & $58$ & $+3.83179265 \times10^{-61}
1796: $\\\hline
1797: $31$ & $-1.53497157 \times10^{-23}$ & $59$ & $-1.15433788 \times10^{-62}
1798: $\\\hline
1799: $32$ & $+8.71891707 \times10^{-25}$ & $60$ & $+3.41802128 \times10^{-64}
1800: $\\\hline\hline
1801: \end{tabular}
1802: \caption{Roots $\nu_{k,r(k,-4)}$ of polynomials $P_{D}(N)$ approaching the
1803: value $-4$.}%
1804: \label{table-1}%
1805: \end{table}
1806: 
1807: At large values of $k$ there appear roots converging to other integer values
1808: $D=-6,-8,\ldots$ The roots approaching the value $-6$ are listed in Table
1809: \ref{table-2}. At large $k$ they are well described by asymptotic formula
1810: (\ref{nu-k-asymptotic-6}).
1811: 
1812: \begin{table}[ptb]%
1813: \begin{tabular}
1814: [c]{||l|l||l|l||}\hline\hline
1815: $k$ & $\nu_{k,r(k,-6)} +6 $ & $k$ & $\nu_{k,r(k,-6)} +6 $\\\hline
1816: $21$ & $+0.03432898 $ & $41$ & $+1.44684 \times10^{-18} $\\\hline
1817: $22$ & $-0.011016 $ & $42$ & $-1.55969 \times10^{-19} $\\\hline
1818: $23$ & $+0.0020276 $ & $43$ & $+1.63996 \times10^{-20} $\\\hline
1819: $24$ & $-0.00040899 $ & $44$ & $-1.68294 \times10^{-21} $\\\hline
1820: $25$ & $+0.000076759 $ & $45$ & $+1.68654 \times10^{-22} $\\\hline
1821: $26$ & $-0.000013876 $ & $46$ & $-1.65141 \times10^{-23} $\\\hline
1822: $27$ & $+2.40442 \times10^{-6} $ & $47$ & $+1.58077 \times10^{-24} $\\\hline
1823: $28$ & $-4.00519 \times10^{-7} $ & $48$ & $-1.47999 \times10^{-25} $\\\hline
1824: $29$ & $+6.42173 \times10^{-8} $ & $49$ & $+1.35591 \times10^{-26} $\\\hline
1825: $30$ & $-9.92472 \times10^{-9} $ & $50$ & $-1.21614 \times10^{-27} $\\\hline
1826: $31$ & $+1.48040 \times10^{-9} $ & $51$ & $+1.06834 \times10^{-28} $\\\hline
1827: $32$ & $-2.13385 \times10^{-10} $ & $52$ & $-9.19592 \times10^{-30} $\\\hline
1828: $33$ & $+2.97550 \times10^{-11} $ & $53$ & $+7.75910 \times10^{-31} $\\\hline
1829: $34$ & $-4.01814 \times10^{-12} $ & $54$ & $-6.41992 \times10^{-32} $\\\hline
1830: $35$ & $+5.26006 \times10^{-13} $ & $55$ & $+5.21090 \times10^{-33} $\\\hline
1831: $36$ & $-6.68131 \times10^{-14} $ & $56$ & $-4.15065 \times10^{-34} $\\\hline
1832: $37$ & $+8.24173 \times10^{-15} $ & $57$ & $+3.24558 \times10^{-35} $\\\hline
1833: $38$ & $-9.88147 \times10^{-16} $ & $58$ & $-2.49222 \times10^{-36} $\\\hline
1834: $39$ & $+1.15242 \times10^{-16} $ & $59$ & $+1.87991 \times10^{-37} $\\\hline
1835: $40$ & $-1.30830 \times10^{-17} $ & $60$ & $-1.39342 \times10^{-38}
1836: $\\\hline\hline
1837: \end{tabular}
1838: \caption{Roots $\nu_{k,r(k,-6)}$ of polynomials $P_{k}(D)$ approaching the
1839: value $-6$.}%
1840: \label{table-2}%
1841: \end{table}
1842: 
1843: \subsection{Distribution of roots in the complex plane}
1844: 
1845: One should keep in mind that apart from the stable roots at $D=0,-2$ and the
1846: roots asymptotically approaching points $D=-4,-6,\ldots$, there are many other
1847: roots $\nu_{k,r}$. Most of them are complex. In order to get an impression
1848: about the general distribution of roots in the complex plane, we show them for
1849: the cases $k=7$ (Fig. \ref{roots-7}), $k=8$ (Fig. \ref{roots-8}), and $k=30$
1850: (Fig. \ref{roots-30}) and $k=60$ (Fig. \ref{roots-60}). From the plots for
1851: $k=30$ and $k=60$, it is clearly seen that at large $k$ there appears a
1852: certain regular structure in the distribution of complex roots. It is an
1853: interesting problem to give a complete asymptotic description of roots in the
1854: complex plane. Here we make only preliminary comments about the distribution
1855: of roots.
1856: 
1857: It is convenient to work with polynomials $P_{k}(D)$ (\ref{E-via-P}). Using
1858: eq. (\ref{E-k-asymptotic-0}), we find:%
1859: \begin{equation}
1860: P_{k}(D)\overset{k\rightarrow\infty}{\longrightarrow}(-1)^{k+1}\Gamma\left(
1861: k+\frac{D}{2}\right)  3^{k+\frac{D}{2}}\frac{2^{(D/2)-2}}{\pi\Gamma\left(
1862: D/2+2\right)  }\,. \label{P-asymptotic}%
1863: \end{equation}
1864: The polynomial $P_{k}(D)$ has roots $\tilde{\nu}_{k,r}$ which coincide with
1865: the roots of $E^{(k)}(D)$ up to the two stable roots $0$, $-2$:%
1866: \[
1867: \{\nu_{k,1},\nu_{k,2},\ldots\nu_{k,k+1}\}=\{\tilde{\nu}_{k,1},\tilde{\nu
1868: }_{k,2},\ldots\tilde{\nu}_{k,k-1},0,-2\}\,.
1869: \]
1870: We can represent $P_{k}(D)$ in the form%
1871: \begin{equation}
1872: P_{k}(D)=\beta_{k}\prod\limits_{r=1}^{k-1}\left(  D-\tilde{\nu}_{k,r}\right)
1873: \,. \label{P-root-product}%
1874: \end{equation}
1875: The coefficient $\beta_{k}$ was computed in Ref. \cite{DP-79}:%
1876: \begin{equation}
1877: \beta_{k}=(-1)^{k+1}2^{k-2}\frac{\Gamma(\frac{3k-1}{2})}{(k+1)!\Gamma\left(
1878: \frac{k+1}{2}\right)  }\,. \label{beta-k-DP}%
1879: \end{equation}
1880: We find from eqs. (\ref{P-asymptotic}), (\ref{P-root-product})%
1881: 
1882: \begin{equation}
1883: \frac{P_{k}(D)}{P_{k}(0)}=\prod\limits_{r=1}^{k-1}\left(  1-\frac{D}%
1884: {\tilde{\nu}_{k,r}}\right)  \overset{k\rightarrow\infty}{\longrightarrow
1885: }\frac{\left(  6k\right)  ^{D/2}}{\Gamma\left(  D/2+2\right)  }\,.
1886: \end{equation}
1887: Differentiating the logarithm of this expression with respect to $D$, we
1888: obtain%
1889: \begin{equation}
1890: \sum\limits_{r=1}^{k-1}\frac{1}{D-\tilde{\nu}_{k,r}}\overset{k\rightarrow
1891: \infty}{\longrightarrow}\frac{1}{2}\left[  \ln\left(  6k\right)  -\psi\left(
1892: \frac{D}{2}+2\right)  \right]
1893: \end{equation}
1894: where $\psi(z)$ is the logarithmic derivative of the $\Gamma$ function%
1895: \begin{equation}
1896: \psi(z)=\frac{\Gamma^{\prime}(z)}{\Gamma(z)}=-\gamma-\sum\limits_{k=0}%
1897: ^{\infty}\left(  \frac{1}{z+k}-\frac{1}{k+1}\right)
1898: \end{equation}
1899: and $\gamma$ is Euler's constant. Thus%
1900: \begin{equation}
1901: \prod\limits_{r=1}^{k-1}\frac{1}{D-\tilde{\nu}_{k,r}}\overset{k\rightarrow
1902: \infty}{\longrightarrow}\frac{1}{2}\ln\left(  6k\right)  +\frac{1}{2}\left[
1903: \gamma+\sum\limits_{k=0}^{\infty}\left(  \frac{2}{D+4+2k}-\frac{1}%
1904: {k+1}\right)  \right]  \,.
1905: \end{equation}
1906: This relation requires the appearance of roots $\nu_{k,r}$ converging to
1907: $-4,-6,$\ldots\ but does not forbid the existence of additional complex roots
1908: going to infinity or making a quasicontinuous distribution in the limit
1909: $k\rightarrow\infty$. Numerical calculations show a large amount of complex
1910: roots at large $k$ (see Figs. \ref{roots-30} and \ref{roots-60}).
1911: 
1912: In the limit of large $k$, one can compute the product of all roots $\nu
1913: _{k,r}$ of the polynomials $P_{k}(D)$ setting $D=0$ in eqs.
1914: (\ref{P-asymptotic}), (\ref{P-root-product}) and using expression
1915: (\ref{beta-k-DP}) for $\beta_{k}$:%
1916: \begin{equation}
1917: \prod\limits_{r=1}^{k-1}\nu_{k,r}\overset{k\rightarrow\infty}{\longrightarrow
1918: }\frac{1}{\beta_{k}}\frac{3^{k}}{4\pi}\Gamma\left(  k\right)  \overset
1919: {k\rightarrow\infty}{\longrightarrow}(-1)^{k+1}3k^{2}\left(  \frac{k}%
1920: {e\sqrt{3}}\right)  ^{k}\,.
1921: \end{equation}
1922: This gives us an asymptotic estimate for the geometric average of all nonzero
1923: $k$ roots $\nu_{k,r}$ of $E_{k}(D)$ (including the root $\nu_{k,r}=-2$)%
1924: \begin{equation}
1925: \left\langle |\nu_{k,r}|\right\rangle _{\nu_{k,r}\neq0}\equiv\left[  2\left(
1926: \prod\limits_{r=1}^{k-1}|\tilde{\nu}_{k,r}|\right)  \right]  ^{1/k}%
1927: \overset{k\rightarrow\infty}{\longrightarrow}\frac{k}{e\sqrt{3}}\left(
1928: 6k^{2}\right)  ^{1/k}\,.
1929: \end{equation}
1930: The growth of this quantity with $k$ shows that most of the roots go to infinity.%
1931: 
1932: \begin{figure}
1933: [ptb]
1934: \begin{center}
1935: \includegraphics[
1936: height=3.4221in,
1937: width=3.3486in
1938: ]%
1939: {roots7.eps}%
1940: \caption{8 roots $\nu_{8,r}$ of the polynomial $E^{(7)}(D)$. Two complex
1941: conjugate roots $-3.63\pm i0.34$ are in the vicinity of the negative even
1942: value $D=-4$.}%
1943: \label{roots-7}%
1944: \end{center}
1945: \end{figure}
1946: 
1947: \begin{figure}
1948: [ptb]
1949: \begin{center}
1950: \includegraphics[
1951: height=3.2785in,
1952: width=3.4221in
1953: ]%
1954: {roots8.eps}%
1955: \caption{9 roots $\nu_{9,r}$ of the polynomial $E^{(8)}(D)$. One root at
1956: $-3.76443$ is close to the negative even value $-4$.}%
1957: \label{roots-8}%
1958: \end{center}
1959: \end{figure}
1960: 
1961: \begin{figure}
1962: [ptb]
1963: \begin{center}
1964: \includegraphics[
1965: height=3.2785in,
1966: width=3.3944in
1967: ]%
1968: {roots30.eps}%
1969: \caption{31 roots $\nu_{30,r}$ of the polynomial $E^{(30)}(D)$. One can see a
1970: discrete set of roots close to even negative values up to $D=-8$. Most of the
1971: roots belong to the quasicontinuous set formed in the complex plane.}%
1972: \label{roots-30}%
1973: \end{center}
1974: \end{figure}
1975: 
1976: \begin{figure}
1977: [ptb]
1978: \begin{center}
1979: \includegraphics[
1980: height=3.2379in,
1981: width=3.3797in
1982: ]%
1983: {roots60.eps}%
1984: \caption{61 roots $\nu_{60,r}$ of the polynomial $E^{(60)}(D)$. One can see a
1985: discrete set of roots close to even negative values up to $D=-12$. There is
1986: also a quasicontinuous set formed in the complex plane.}%
1987: \label{roots-60}%
1988: \end{center}
1989: \end{figure}
1990: 
1991: \section{Conclusions}
1992: 
1993: \setcounter{equation}{0} 
1994: 
1995: Thus we have established a close connection between
1996: 
1997: 1) exactly solvable features of the anharmonic oscillator in negative even dimensions,
1998: 
1999: 2) disappearance of the factorial growth of perturbative coefficients in
2000: negative even dimensions,
2001: 
2002: 3)\ fast inverse factorial convergence of roots of polynomials $E^{(k)}(D)$ to
2003: the negative even points.
2004: 
2005: Although the content of this paper was restricted to quantum mechanics, the
2006: main motivation comes from quantum field theory. In contrast to quantum
2007: mechanics where the large-order behavior of the perturbation theory can be
2008: easily tested using direct numerical calculations in cases when the analytical
2009: methods fail or raise doubts, in quantum field theory the power of both
2010: analytical and numerical tools is rather limited. The slow $O(k^{-1})$
2011: convergence of perturbative coefficients to the asymptotic form [see e.g. eq.
2012: (\ref{E-k-N-asymptotic})] is rather disturbing in quantum mechanics but one
2013: still can reach the asymptotic regime in large orders. In quantum field theory
2014: calculations rarely go beyond four or five loops. In this situation the
2015: construction of quantities whose perturbative expansion has a fast factorial
2016: convergence to the asymptotic form is very important.
2017: 
2018: Polynomials $E^{(k)}(D)$ and their roots have analogs in quantum field theory.
2019: In many field theoretical models we have dependence on various parameters
2020: (e.g. number of colors, flavors etc.) and Feynman diagrams in any order lead
2021: to a polynomial (or fractional polynomial) dependence on these parameters. The
2022: control of this polynomial dependence is usually trivial compared to the hard
2023: work needed for the calculation of loop integrals. Therefore available
2024: multiloop results in quantum field theory provide many opportunities for
2025: testing the roots of these polynomials. There is some evidence that the
2026: inverse factorial convergence of roots found in the case of the anharmonic
2027: oscillator has analogs in some field theoretical models, e.g. in the
2028: $N$-component $O(N)$ symmetric model \cite{PP-08}. An indirect theoretical
2029: argument comes from the presence of an inverse $\Gamma$ function in field
2030: theoretical asymptotic expressions. This inverse $\Gamma$ function may
2031: generate zeros similar to the zeros at $D=0,-2,-4\ldots$ coming from the
2032: factor $\left[  \Gamma(D)\right]  ^{-1}$ in eq. (\ref{E-k-asymptotic-0}) in
2033: the case of the anharmonic oscillator. Since the convergence is very fast, it
2034: may happen that one can detect it (or its signals) in available multi-loop
2035: perturbative expressions.
2036: 
2037: \textbf{Acknowledgments.} I appreciate discussions with G.V. Dunne, L.N.
2038: Lipatov, V.Yu. Petrov and A. Turbiner.
2039: 
2040: \begin{thebibliography}{9}                                                                                                %
2041: 
2042: \bibitem {BW-69}C.M. Bender, T.T. Wu, Phys. Rev. 184 (1969) 1231.
2043: 
2044: \bibitem {BW-73}C.M. Bender, T.T. Wu, Phys. Rev. D7 (1973) 1620.
2045: 
2046: \bibitem {Lipatov-77b}L.N. Lipatov, Sov. Phys. JETP 45 (1977) 216.
2047: 
2048: \bibitem {BLZ-77}E. Brezin, J. C. Le Guillou and J. Zinn-Justin, Phys. Rev.
2049: D15 (1977) 1544.
2050: 
2051: \bibitem {SZ-79}R. Seznec, J. Zinn-Justin, J. Math. Phys. 20 (1979) 1398.
2052: 
2053: \bibitem {DP-79}A. D. Dolgov and V. S. Popov, Phys. Lett. B86 (1979) 185.
2054: 
2055: \bibitem {Zinn-Justin-81a}J. Zinn-Justin, J. Math. Phys. 22 (1981) 511.
2056: 
2057: \bibitem {Zinn-Justin-81b}J. Zinn-Justin, Phys. Rep. 70 (1981) 109.
2058: 
2059: \bibitem {ZJ-04}J. Zinn-Justin, U.D. Jentschura, Ann. of Phys. 313 (2004) 197, 269.
2060: 
2061: \bibitem {DH-88}G.V. Dunne and I.G. Halliday, Nucl. Phys. B308 (1988) 589.
2062: 
2063: \bibitem {Dunne-89}G.V. Dunne, J. Phys. A22 (1989) 1719.
2064: 
2065: \bibitem {Turbiner-94}A. Turbiner, Quasi-exactly Solvable Differential
2066: Equations, in CRC Handbook of Lie Group Analysis of Differential Equations,
2067: Vol. 3 : New Trends in Theoretical Developments and Computational Methods, ed.
2068: N. H. Ibragimov, CRC Press, 1995 [hep-th/9409068].
2069: 
2070: \bibitem {Shifman-ITEP}M.A. Shifman, ITEP Lectures on Particle Physics and
2071: Field Theory, vol. II, chapter VII, World Scientific (1999).
2072: 
2073: \bibitem {Ushveridze}A. Ushveridze, Quasi-Exactly Solvable Models in Quantum
2074: Mechanics, IOP Publishing, Bristol (1994).
2075: 
2076: \bibitem {DT-99}P. Dorey and R. Tateo, J. Phys. A32 (1999) L419.
2077: 
2078: \bibitem {BLZ-98}V.V. Bazhanov, S.L. Lukyanov and A.B. Zamolodchikov, J.
2079: Statist. Phys 102 (2001) 567.
2080: 
2081: \bibitem {BLZ-03}V.V. Bazhanov, S.L. Lukyanov and A.B. Zamolodchikov, Adv.
2082: Theor. Math. Phys. 7 (2004) 711.
2083: 
2084: \bibitem {DDT-07}P. Dorey, C. Dunning, R. Tateo, The ODE/IM Correspondence,
2085: arXiv: hep-th/0703066.
2086: 
2087: \bibitem {DDGT-07}P. Dorey, C. Dunning, F. Gliozzi and R. Tateo, On the ODE/IM
2088: correspondence for minimal models, arXiv: 0712.2010 [hep-th].
2089: 
2090: \bibitem {Andrianov-82}A.A. Andrianov, Ann. of Phys. 140 (1982) 82.
2091: 
2092: \bibitem {DPM-84}R. Damburg, R. Propin and V. Martyshchenko, J. Phys. A17
2093: (1984) 3493.
2094: 
2095: \bibitem {BG-93}V. Buslaev and V. Grecchi, J. Phys. A26 (1993) 5541.
2096: 
2097: \bibitem {Andrianov-07}A.A. Andrianov, Phys. Rev. D76 (2007) 025003.
2098: 
2099: \bibitem {DeAlfaro-Regge}V. de Alfaro and T. Regge, Potential Scattering,
2100: North-Holland, Amsterdam (1965).
2101: 
2102: \bibitem {Collins-book}P.D.B. Collins. An Introduction to Regge Theory and
2103: High Energy Physics, Cambridge University Press, Cambridge (1977).
2104: 
2105: \bibitem {Newton-book}R.G. Newton, The Complex $J$-Plane, Benjamin, New York (1964).
2106: 
2107: \bibitem {PP-08}P.V. Pobylitsa, arXiv:0807.5136 [hep-th].
2108: 
2109: \bibitem {RS-95}V.A. Rubakov, O.Yu. Shvedov, Nucl. Phys. B434 (1995) 245.
2110: \end{thebibliography}
2111: \end{document}