hep-th0507207/semi.tex
1: \documentclass[a4paper,12pt]{article}
2: \usepackage{graphicx}
3: \usepackage{amssymb}
4: \usepackage{a4}
5: 
6: %opening
7: \title{The quantum Neumann model: asymptotic analysis.}
8: \date{July 2005}
9: \author{Marc Bellon and Michel Talon\thanks{LPTHE, CNRS et Universit\'es
10: Paris VI--Paris VII (UMR 7589),
11:  Bo\^{\i}te 126,
12:  4 place Jussieu, F-75252 PARIS CEDEX 05}
13: }
14: 
15: \begin{document}
16: 
17: \maketitle
18: 
19: \begin{abstract}
20: We use semi--classical and perturbation methods to establish the quantum
21: theory of the Neumann model, and explain the features observed in previous
22: numerical computations.
23: \end{abstract}
24: \vfill
25: LPTHE--05--17
26: \eject
27: 
28: 
29: \section{Introduction}
30: 
31: In a previous publication~\cite{BT05}, we have explained how to determine
32: numerically the spectrum of the quantum Neumann model. The Neumann model
33: describes a particle on a sphere subject to a quadratic potential. It has been
34: introduced in the nineteenth century by Neumann~\cite{Neumann}, who showed that
35: it is integrable.
36: More recently, Moser~\cite{Moser} and Mumford~\cite{Mu84} have shown how it fits
37: in the modern approaches to algebraically integrable systems and generalized the model
38: to any number of dimensions.  More to our point, the quantum version of the
39: model is also integrable~\cite{AvTa90,BaTa92}. This means that
40: its solution reduces to the resolution of a one dimensional Schr\"odinger equation,
41: as it is the generic behaviour for algebraically integrable systems. The notable
42: feature compared to the
43: textbook examples is that the different conserved quantities, which all
44: appear in the same equation, must be solved for simultaneously. This is a non
45: trivial problem
46: which has been attacked in~\cite{Gurarie}. We will precise and prove some
47: statements made in this paper. Moreover, the Neumann model is related with
48: St\"ackel systems as noted by Gurarie and fully developed recently
49: in~\cite{DuVa05}.
50: 
51: We will start with a brief presentation of the model and the equations which
52: have to be solved. The physics of the model is driven by a general scale
53: parameter~$v$, which is proportional both to the strength of the interaction
54: and the radius of the sphere. Small values of $v$ correspond to a quantum
55: regime and are next studied by perturbation theory. Then, we introduce
56: the semi--classical analysis which is appropriate for large values of $v$. The 
57: comparison of energy levels obtained by semi--classical methods and by the
58: numerical solution of the Schr\"odinger equation shows that it is indeed the
59: case. However, we find a small constant discrepancy between the two. In a last
60: part, we recover this constant by analytic studies of the model at large $v$
61: according to these two methods.
62: 
63: \section{The model.}
64: 
65: The Hamiltonian of the system is given in terms of angular momenta
66: $J_{kl} = x_k \dot{x}_l - x_l\dot{x}_k$:
67: \begin{equation}
68: H = {1 \over 4} \sum_{k \neq l} J_{kl}^2 ~+ ~ 
69: {1 \over 2} \sum_k a_k x_k^2
70: \label{hamiltonian}
71: \end{equation}
72: with the constraint $\sum x_k^2=r^2$. We assume $a_1<a_2<\cdots<a_N$. One can
73: always shift all the $a_k$ by a
74: constant since this merely adds a constant to $H$. The system is classically
75: and quantum mechanically integrable since the following quantities commute,
76: classically and quantically:
77: \begin{equation}
78: F_k = x_k^2 + \sum_{l \neq k} {J_{kl}^2 \over a_k - a_l}, \quad H = {1\over2}
79: \sum a_k F_k
80: \label{conserved}
81: \end{equation}
82: 
83: 
84: To solve the equations of motions, one introduces the ``separated" variables
85: $t_k$ given
86: as the roots of:
87: $$ \sum_k {x_k^2 \over t-a_k}=0$$
88: Positivity of $x_k^2$ implies that there is exactly one root in each interval
89: $[a_k,a_{k+1}]$.
90: It has been shown in~\cite{BaTa92} that  finding the common eigenvector 
91: for the $F_k$ with eigenvalue $f_k$ reduces to solving the unique
92: one--dimensional Schr\"odinger equation:
93: \begin{equation}
94: \left[{d^2\over dt^2}+{1\over 2}\sum_k{1\over t-a_k}\;{d\over dt}
95: -{1\over 4\hbar^2}\sum_k{f_k\over t-a_k}\right]\,\Psi(t)=0
96: \label{lame}
97: \end{equation}
98: This is a linear differential equation with singularities at the $a_k$ and at
99: infinity. The situation is explained in more details in~\cite{BT05}.
100: 
101: 
102: It can be useful to write the potential term in the form:
103: \begin{equation} \label{produit}
104: \sum _k{f_k\over t-a_k}= v\; {\prod (t-b_l)\over \prod (t-a_k)}
105: \end{equation}
106: with $v = \sum_k f_k = \sum_k x_k^2=r^2$.
107: Classically the action $S$ satisfies the separated
108: Hamilton-Jacobi equation:
109: $$ {1\over 2}\left({dS \over dt}\right)^2 = - v\; {\prod (t-b_l)\over \prod
110: (t-a_k)}$$
111: In Figure~1 we draw the different possibilities for the potential in
112: the Neumann case $N=3$, with oscillator strengths 0, 1 and $y$.
113: 
114: \begin{figure}[t]
115: \includegraphics[width=12cm]{figs}
116: \label{figs}\caption{The four different possibilities for $t_1,t_2$, with
117: $y=2$.}
118: \end{figure}
119: 
120: For $v$ large, we are in case A), $b_1$ is close to 1 and $b_2$ is close to
121: $y$. Classically the allowed region is such that for each $t_k$ one has
122: $P(t)=\prod (t -a_n)\prod(t-b_m) <0$ so that, since $t_k\in [a_k,a_{k+1}]$ 
123: one has $b_1\leq t_1 \leq 1$ and $b_2 \leq t_2 \leq y$. As $v$ is lowered, both
124: $b_1$ and $b_2$ move to the left, and when $v\to 0$, one has $b_1 \to
125: -\infty$ so that one is in case C) or D). One goes from case D) to case
126: A) either through the cases B) or C). In case B), $b_1$ and $b_2$ both belong
127: to the interval $[0,1]$ so that the allowed region corresponds to $b_1\leq t_1
128: \leq b_2$ and $t_2 $ describes the whole interval $[1,y]$. In case C) and D),
129: $b_1$ is negative and $b_2$ is on either side of 1. One of the $t_k$ describes
130: the whole interval and the other is constrained away from 1 by $b_2$.
131: It is easy to check that no other 
132: configuration is compatible with the existence of $t_1$ and $t_2$ satisfying
133: $P(t)<0$ in the appropriate intervals. In the general case, we will show in the
134: sequel that all the $b_i$ are real and belong to the interval
135: $[a_{i-1},a_{i+1}]$, with $a_0$
136: supposed to be $-\infty$. There is therefore a total of $2^{N-1}$ possibilities
137: for the relative positions of the $b_i$ and $a_k$.
138: 
139: 
140: At $v=0$, the system reduces to a free particle on the sphere. The energy
141: levels are given by the usual formula ${1\over 2}j(j+1)$ for spherical
142: harmonics, but with a different
143: choice of basis in the $(2j+1)$-dimensional eigenspace. These basic
144: functions are called spheroidal harmonics. These harmonics diagonalize the
145: perturbation by the potential, so that energy levels separate linearly for
146: small $v$.
147: 
148: For large $v$, the potential term becomes dominant and the point is restricted
149: to the vicinity of the two opposite poles $x_1=\pm r$, $x_2=0$, $x_3=0$. In the
150: next approximation, we can neglect the curvature of the sphere and we have two
151: independent harmonic oscillators in $x_2$ and $x_3$. The configuration around
152: the two poles are linked by an exponentially small tunnelling amplitude, so that
153: all states come in almost degenerate pairs.
154: 
155: \section{Small $v$ perturbation theory.}
156: 
157: In the case $v=0$, the singularity at infinity of the equation~(\ref{lame})
158: becomes regular. Even if the behaviour of the solution at infinity has no
159: direct bearing on the physical wave function, it appears important for the
160: characterization of the solutions and provides the quantization condition for
161: the energy.
162: 
163: Indeed, recall that around a regular singularity at $x=0$, a second order
164: differential equation has two independent solutions of the form $x^\alpha(1+ b x
165: + \cdots)$, with the exponents $\alpha$ roots of a second order equation.
166: At the finite singularities $a_k$, the exponents are 0 and $1/2$ with
167: respective monodromies $+1$ and $-1$. In our previous work~\cite{BT05}, we
168: determined that
169: to produce a single valued wave function on the sphere, the solutions must be
170: of definite monodromy at each of the points $a_k$. Monodromy at infinity is the
171: product of the monodromies at the $a_k$ and is therefore $\pm 1$. Since the
172: exponent at infinity is solution of:
173: \begin{equation}
174: 	\alpha ^2 + \alpha ({N\over 2}-1) - { E \over 2 \hbar^2}=0
175: \end{equation} 
176: and $\alpha$ must be integer or half integer to get the right monodromy, we
177: conclude that the energy $E$ is ${1\over2}\hbar^2 j(j+N-2)$ with any non
178: negative integer $j=2\alpha$. 
179: 
180: For a given energy $E$, there is a second root $\alpha$ which yields a
181: solution vanishing at infinity. However it cannot be the solution we are
182: seeking with diagonal monodromy at the $a_k$, since such a solution would be
183: single valued and regular on a compact Riemann surface. In the case where all
184: monodromies are one, the solution has just one singularity, a pole at infinity
185: and is therefore a polynomial of degree $\alpha$. These functions are known as
186: spheroidal harmonics of first species~\cite{WW}.
187: 
188: In order to find these polynomial solutions, we write them in the form $\prod
189: (t-\lambda_j)$. Inserting in equation~(\ref{lame}), we get:
190: \begin{equation}
191: \sum_{i\neq j} {1\over(t-\lambda_i)(t-\lambda_j)}+{\textstyle 1 \over
192: 2}\sum_{k,j} {1\over(t-a_k)(t-\lambda_j)} - {1\over
193: 4\hbar^2}\sum_k{f_k\over t-a_k} =0
194: \end{equation}
195: As it should be, this equation requires that $\sum_k f_k=0$ and we recover the
196: value for the energy from the dominant term at $t \to \infty$. This equation is
197: equivalent to the condition that its residues at the $a_k$ and $\lambda_j$ all
198: vanish. The residues at the $\lambda_j$ give a system of non-linear equations
199: which determine the $\lambda_j$:
200: \begin{equation}
201: \sum_k{1\over \lambda_j-a_k}+4 \sum_{h\neq j}{1\over \lambda_j- \lambda_h} = 0
202: \label{eqzero}
203: \end{equation}
204: The residues at the $a_k$ then determine the conserved quantities as:
205: \begin{equation}
206: f_k = 2 \hbar^2 \sum_j{1\over a_k - \lambda_j}
207: \end{equation}
208: The system of equations~(\ref{eqzero}) can be seen as the condition of
209: equilibrium of a system of points with logarithmic repulsive potentials, some
210: of which being fixed. In this formulation, it is therefore clear that there is
211: a unique solution with a definite number of $\lambda_j$ in each of the intervals
212: $[a_k,a_{k+1}]$. In the case $N=3$, this gives $d+1$ different solutions of
213: degree $d$. Similar calculations can be done for solutions which are of the
214: form $\prod_{k\in I}\sqrt{t-a_k}\prod_j(t-\lambda_j)$ where $I$ is a subset of
215: the singularities. This gives spheroidal harmonics of the second, third and
216: fourth species~\cite{WW} for $N=3$. The total number of these harmonics with
217: energy ${1\over2}j(j+1)$ is exactly $2j+1$, the known degeneracy of this energy
218: level.
219: 
220: Numerical solutions of the system of equations~(\ref{eqzero}) are easy to get
221: by a multidimensional Newton method and have been used to produce initial
222: solutions in our previous paper.
223: 
224: From the knowledge of this $v=0$ solutions, we will describe the behaviour of
225: the solutions for small $v$. As soon as $v$ is different from zero, the
226: solution gets an essential singularity at infinity. We therefore multiply the
227: spheroidal harmonic by the simplest function singular at infinity, namely
228: $\exp(pt)$. In the vicinity of $v=0$, we expect $p$ to be small and neglect
229: higher powers of $p$. Inserting in equation~(\ref{lame}) gives:
230: \begin{eqnarray*}
231: \sum_{i\neq j} {1\over(t-\lambda_i)(t-\lambda_j)}
232: +{\textstyle 1 \over2}\sum_{k,j} {1\over(t-a_k)(t-\lambda_j)} 
233: - {1\over4\hbar^2}\sum_k{f_k\over t-a_k} &&\\+ 2 p \sum_j{1 \over t-\lambda_j}
234: +{1\over 2}p\sum_k{1\over t - a_k}&=&0
235: \end{eqnarray*} 
236: The term proportional to $1/t$ in this equation immediately gives that $p$ is
237: proportional to $v$, so that the hypothesis that $p$ is small is valid. 
238: \[ 2np + {1\over2}N p -{1\over4\hbar^2}v =0
239: \]
240: Here $n$ is the degree of the polynomial. The roots $\lambda_j$ of the wave
241: function satisfy equations which are small perturbations of
242: equation~(\ref{eqzero}), hence are close to their values for
243: $v=0$. The term in $1/t^2$ then gives the perturbation of the energy:
244: \[ E =  \hbar^2 n(2n+N-2) + v { \sum a_k + 4 \sum \lambda_j \over 2(N + 4n)} \]
245: Similar calculations can be done for the spheroidal harmonics of the other
246: species and this nicely reproduces the slopes of the numerical results at small
247: $v$.
248: 
249: \section{Semiclassical analysis.}
250: 
251: As an integrable system, the Neumann model lends itself to a simple
252: semiclassical analysis writing basically that the Bohr--Sommerfeld conditions
253: quantize the Liouville tori. Equivalently, the separated Schr\"odinger equation
254: can be analyzed by the WKB method. 
255: 
256: \subsection{The WKB treatment.}
257: 
258: The presence of a singular factor in the
259: first derivative term of the equation~(\ref{lame}) seems to be a problem for
260: the application of the WKB method in our case, but we will show that in
261: contrast to what happens in the spherical symmetry case~\cite{Lan38}, this does
262: not cause troubles. In a first approach, we write an expansion in powers of
263: $\hbar$ for the action $S$ such that $\Psi = \exp (iS/\hbar)$:
264: \begin{equation}
265: \label{expans} S = S_0 + \hbar S_1 + \hbar^2 S_2 + \cdots
266: \end{equation} 
267: As usual the dominant term in $\hbar^{-2}$ does not depend on the first
268: derivative term in the equation and gives for $S_0$:
269: \[ S_0 = \int dt \;\sqrt{-v\prod(t-b_i) \over 4 \prod (t-a_k)} \]
270: The following term gives:
271: \[ i S_0'' - 2 S_0'S_1' + {1\over2} i S_1' \sum{1\over t- a_k} = 0 \]
272: Due to the form of $S_0$, the singular terms at $a_k$ cancel and one gets:
273: \[ \exp( iS_1) =  \prod (t-b_j) ^{-{1\over4}} \]
274: At this order, the wave function is regular at the $a_k$ and one can also show
275: that the following correction to $S$ is regular at $a_k$. In contrast, the
276: first correction is mildly singular at the turning point $b_l$ and the
277: following correction $S_2$ has so bad singularities that the expansion breaks
278: down. In the cases where the $b_l$ are away from the $a_k$, the usual
279: connection formulae around the turning point~\cite{Lan38} justify an additional
280: $\pi/4$
281: phase factor in the quantization conditions.
282: 
283: 
284: Another approach is to desingularize the equation~(\ref{lame}) by introducing a
285: new variable $x$ defined by the (hyper-)elliptic integral~\cite{WW}:
286: \begin{equation}\label{ellparam}
287: x = \int^t {dt \over \sqrt{\prod(t-a_k)}}
288: \end{equation}
289:  The application of the chain rule gives:
290: \[ {d^2\Psi\over dt^2}+{1\over 2}\sum_k{1\over t-a_k}\;{d\Psi\over dt} =
291: {1\over\prod(t-a_k)} {d^2\Psi\over dx^2} \]
292: so that the Schr\"odinger equation becomes:
293: \begin{equation}\label{ellequ}
294: {d^2\Psi\over dx^2} - {v\over 4\hbar^2} \prod ( t-b_l) \Psi =0
295: \end{equation}  
296: This equation has neither singularities nor first derivative term, so that the
297: usual WKB formula gives:
298: \[ \Psi ( x) \simeq \left( \prod (t-b_j) \right) ^{-{1\over4}} \exp\left(
299: {1\over\hbar} \int dx \sqrt{{v\over 4}\prod(t-b_i) }\right) \]
300: It should be emphasized that the variation of $x$ is alternatively real and
301: purely imaginary when $t$ crosses the $a_k$, introducing additional $i$
302: factors in the action. In the traditional $N=3$ case, $x$ is an elliptic
303: integral and $t$ can be expressed with Weierstra\ss \ elliptic functions. When
304: $t$ goes from $-\infty$ to $+\infty$, $x$ describes the boundary of the
305: half--periods rectangle. In the cases where there are more singularities, it is
306: possible that the $t$ variable is not well defined as an inverse function of
307: $x$, however this is of no consequence for the study of the Schr\"odinger equation
308: which proceeds separately on each of the intervals $[a_k,a_{k+1}]$.
309: 
310: \subsection{Remarks on the zeroes of the potential.}
311: 
312: The above elliptic formulation allows to demonstrate quickly the claim we 
313: have made that there must be in each interval $[a_j,a_{j+1}]$ a region
314: where the product $P(t)=\prod (t -a_n)\prod(t-b_m)$ takes negative values.
315: Recall from~\cite{BT05} that the wave function is either even or odd under
316: each parity operation $x_k \to -x_k$. In terms of the $t$ variable, this
317: translates into monodromy around $a_k$. Now, the $x$ variable of the elliptic
318: parametrization behaves as $x(a_k)+c_k\sqrt{t-a_k}$ in the vicinity of $a_k$ so
319: that the wave function is even or odd as a function of $x-x(a_k)$, which means
320: that either $\Psi$ or its derivative vanishes at each of the $x(a_k)$.
321: 
322: Multiplying eq.~(\ref{ellequ}) by $\Psi$ and integrating
323: by parts, we then get for each $k$: 
324: \begin{equation} \label{ene}
325: \int_{x(a_k)}^{x(a_{k+1})} [\Psi'^2(x) + {v\over 4\hbar^2}\prod(t(x)-b_l)
326: \Psi^2(x)]
327: dx = 0
328: \end{equation}
329: Under the elliptic parametrization, the image of the real axis is a sequence
330: of alternatively real and purely imaginary intervals, according to the sign
331: of the product $\prod (t-a_k)$. The $\Psi'^2(x)$ term has therefore the same
332: sign as this product. Since $\Psi$ is real, Eq.~(\ref{ene}) cannot be satisfied if the
333: potential remains of the same sign as the kinetic term.   This is
334: equivalent to the condition $P(t)<0$ at some point of the interval
335: $[a_j,a_{j+1}]$.
336: Finally, we prove that this allows to establish that  quantum mechanically the
337: $b_j$ obey the constraints we have discussed in the framework of the classical analysis.
338: 
339: The relative position of the $a_k$ and $b_j$ depends on the signs of the $f_k$.
340: The only one with a fixed sign is $f_N$, which is positive. All combinations of
341: the signs of the other ones are possible, which gives $2^{N-1}$ possibilities.
342: We will show that each of these sign combinations corresponds to a unique set
343: of relative positions of the $b_j$. The fact that we identify $N-1$ necessary
344: real zeroes of the potential shows that all $b_j$ must be real.
345: 
346: More precisely, we will show that $b_j$ belongs to the interval $]a_j,a_{j+1}[$
347: if $f_j$ is positive and to the interval $]a_{j-1},a_{j}[$ otherwise. This
348: means that we have to show that in the interval $]a_j,a_{j+1}[$, there is
349: exactly one zero if $f_j$ and $f_{j+1}$ are of the same sign and two if $f_j$
350: is positive and $f_{j+1}$ is negative. If the $f$ are of the same sign, the
351: potential reaches opposite infinite limits at the ends of the interval, so that
352: it must have a zero in between. In the other case, the potential goes to
353: positive infinity at the two ends of the interval and we proved that the
354: potential must be negative somewhere in the interval. This is only possible if
355: the potential has two zeroes which we identify as $b_j$ and $b_{j+1}$. The only
356: remaining thing to prove is that there is a zero in the interval
357: $]-\infty,a_1[$ when $f_1$ is negative. In this case, we use the fact that
358: $\sum f_k$ is positive, so that the dominant term near $-\infty$ is negative.
359: The potential is positive at $0-$ hence there must be  a zero in between. Since
360: we identified $N-1$ zeroes, we have exhausted the whole set of zeroes for the
361: potential and there cannot be any other, proving in particular that all zeroes
362: are real.
363: 
364: \subsection{The quantization conditions.}
365: 
366: In fact, using any of the two WKB approaches, we get the same semiclassical
367: quantization conditions:
368: \begin{equation}\label{semi}
369: 	\int_{b_j}^{a_{j+1}} dt\;\sqrt{-{v\over4}{\prod (t-b_i) \over \prod
370: (t-a_k)}} = (n_j  + {\textstyle 1\over4})\pi \hbar
371: \end{equation} 
372: in the case where all the monodromies are one and $b_j$ is in the interval
373: $[a_j,a_{j+1}]$. The integer $n_j$ is the number of zeroes of the wave
374: function in the interval $[a_j,a_{j+1}]$. If the monodromy around $a_{j+1}$ is
375: $-1$, we should add one half to $n_j$. In the cases where $b_j$ is near
376: $a_j$ or becomes smaller, the $\pi/4$ phase factor is no more a good
377: approximation and we will postpone to further studies the relevant
378: phase factors. This equation has already appeared in the literature, for
379: example in~\cite{BaTa92,Gurarie}, where the refinement with the $\pi/4$ phase
380: first appears in the last paper. It is introduced there using the machinery of
381: Maslov indices. In this case, the justification through the connection formulae
382: is very direct, it simply uses the linear approximation of the potential around
383: the turning point and consequently the asymptotic analysis of the Airy
384: function:$$
385: Ai(x) \sim_{+\infty} {1\over 2\sqrt\pi x^{1/4} } e^{-{2\over3}x^{3/2}},
386: \qquad Ai(-x) \sim_{+\infty} {1\over \sqrt\pi x^{1/4} } \cos( {2\over3}x^{3/2}
387: - {\pi\over4})
388: $$
389: \begin{figure}[t]
390: \includegraphics[angle=270,width=15cm]{compsemiy2}
391: \caption{Energy levels for $y=2$.\label{ener1}}
392: \end{figure}
393: \begin{figure}[t]
394: \includegraphics[angle=270,width=15cm]{compsemiy5}
395: \caption{Energy levels for $y=5$.\label{ener2}}
396: \end{figure}
397: In order to compare the semiclassical predictions to the exact numerical
398: solutions found in~\cite{BT05}, we solve numerically the quantization
399: conditions~(\ref{semi}) and plot the resulting expression for the energy as a
400: function of $v$ compared with two exact energy levels which are degenerate in
401: the large $v$ limit. We do this for a small number of energy levels in order to
402: keep the graphs uncluttered and plot them for oscillator strengths equal to
403: $(0,1,2)$ in Fig.~2 and $(0,1,5)$ in Fig.~2, both in the
404: case $N=3$. In either figure, the three heavy lines are the WKB predictions. We
405: see that the small $v$ regime is poorly accounted for, which is of no surprise
406: since in this case, we are no longer in the situation A) of Fig.~1 and their
407: should be phase corrections different from $\pi/4$ and moreover, we are in the
408: deep quantum regime, where we do not expect WKB approximation to be good.
409: Nevertheless, we plan to study in further works the different cases of Fig.~1
410: and the transitions between them.
411: 
412: The most notable feature of the curves plotted in Fig. 2 and~3 is that in the
413: large $v$ limit, there is a small constant separation between the exact and WKB
414: curves. This separation is independent both on the energy level and the
415: oscillator strengths. This contradicts the naive expectation that the
416: corrections to the WKB approximation are of order $\hbar^2$ with respect to the
417: order zero solution which has an energy of order $1/\hbar$. This order of the
418: correction is true for the solution away from the turning points, 
419: but the difference from the Airy function may create
420: order $\hbar$ correction to the phase as noted in~\cite{Lan38}. It is however
421: remarkable that this finite correction to the energy does not depend on the
422: level, so that the transition energies have vanishing errors in the small
423: $\hbar$ regime. It is the purpose of the following section to show through
424: perturbative computations the validity of this result.
425: 
426: 
427: 
428: \section{Large $v$ study.}
429: 
430: We first compute perturbatively for large $v$ or equivalently small
431: $\hbar$ the energy levels as  obtained from the ``enhanced''
432: Bohr--Sommerfeld rules, and then as perturbations of quantum oscillators
433: sitting around the pole of the sphere.
434: 
435: \subsection{Semi--classical calculation.}
436: 
437: We limit ourselves to the $N=3$ case and we fix the oscillator strengths to
438: $(0,1,y)$ as in~\cite{BT05}. Since $v$ is large, the point on the sphere is
439: limited to the vicinity of the points $(\pm1,0,0)$, whose
440: separated variables are $t_1=1$ and $t_2=y$. With $b_1$ close to 1 and $b_2$
441: close to $y$, $t_1$ is classically limited to the interval $[b_1,1]$ and $t_2$
442: to the interval $[b_2,y]$, which limits the point to the vicinity of the pole.
443: To do our perturbative calculations, we will write $b_1=1-\alpha$ and $b_2 = y
444: - \beta$, so that $\alpha$ and $\beta$ are small. The integrand in the
445: Bohr--Sommerfeld rules takes the form:
446: \begin{equation}
447: \sqrt {-{\frac { \left( t-1+\alpha \right)  \left( t-y+\beta \right) }
448: {t \left( t-1 \right)  \left( t-y \right) }}}
449: \end{equation} 
450: When integrating in $[b_1,1]$, we cannot expand the factors $t-1+\alpha$ and
451: $t-1$, but all the others can be expanded around $t=1$ and in powers of $\beta$.
452: The resulting integrals are easily computable yielding:
453: $${1\over2}\,\alpha\,\pi +{1\over16}\,{\alpha}^{2}\pi -{1\over4}\,{\frac
454: {\alpha\,\beta }{y-1}}\,\pi + \cdots $$
455: Similarly for the integral in $[b_2,y]$ we keep the factors $t-y+\beta$ and
456: $t-y$ and expand the others around $t=y$ in powers of $\alpha$. We then get the
457: integral:
458: $$
459: {1\over2}\,{\frac{\beta }{\sqrt {y}}}\,\pi+
460: {1\over16}\,{\frac {{\beta}^{2}}{{y}^{3/2}}}\,\pi +
461: {1\over4}\,{\frac {\alpha\,\beta}{\sqrt {y} ( y-1 )}}\,\pi + \cdots $$
462: We write that the first integral is equal to
463: ${\hbar\pi\over\sqrt{v}}(n_1+{1\over2})$ and similarly the second equal to
464: ${\hbar\pi\over\sqrt{v}}(n_2+{1\over2})$ which determines $\alpha$ and $\beta$:
465: \begin{eqnarray*}
466: \alpha&=&  2\hbar(n_1+{1\over2})\left( {1\over\sqrt{v}} -{\hbar\over v}
467: \;\Bigl({1\over 4}(n_1+{1\over2})
468: -(n_2+{1\over2}){\sqrt {y} \over (y-1)}\Bigr)+ \cdots \right) \\
469: \beta&=&2\hbar (n_2+{1\over2})\left( \sqrt{y\over v} -{\hbar\over v}\;
470: \Bigl({1 \over 4}(n_2+{1\over2})+(n_1+{1\over2}){\sqrt {y} \over (y-1)}
471: \Bigr)+ \cdots \right)
472: \end{eqnarray*}
473: Note that this justifies a posteriori the claim we made that $\alpha$
474: and $\beta$ are small, and become smaller when $v$ increases.
475: Finally the energy $E={1\over2}v(1+y-b_1-b_2)={1\over2}v(\alpha+\beta)$ takes
476: the simple form, in
477: which there is no mixing between $n_1$ and $n_2$ excitations:
478: \begin{equation}
479: \label{niveausc}
480: E_{WKB}=\hbar\sqrt{v}\left((n_1+{1\over2})+\sqrt {y}(n_2+{1\over2})\right)
481: -{\hbar^2(n_1+{1\over2})^2\over 4}-{\hbar^2(n_2+{1\over2})^2\over 4}
482: \end{equation}
483: Of course the first term corresponds to the energy of two independent harmonic
484: oscillators of respective excitations $n_1$ and $n_2$, and nicely fits
485: the $\sqrt{v}$ look of the energy levels at large $v$  apparent in the figures
486: of~\cite{BT05}. The second and third term shift these square root terms towards
487: their correct position, but miss it by a small constant.  As we show below this
488: small constant is accessible to a more correct quantum computation.
489: 
490: \subsection{Quantum calculation.}
491: 
492: The principle of this calculation is to work out a perturbation of harmonic
493: oscillators. We detail it in the case $N=3$ and it is then convenient to
494: go to the elliptic parametrization given by eq.~(\ref{ellparam}).  We
495: expand the change of variable in the vicinity of $t=1$ and
496: $t=y$. Taking the origin in $t=y$, we write 
497: \[x= \int { d\tau \over \sqrt\tau} {1\over\sqrt{y(y-1)}}\Bigl( 1 + {\tau\over 2
498: y} + {\tau \over 2(y-1)} + \cdots\Bigr) \]
499: which can be inverted to yield
500: \begin{equation}
501:  \tau = y-t = {1\over4}y(y-1) x^2 - {1\over48} y (y-1) (2y-1) x^4 + \cdots
502: \end{equation} 
503: A similar computation gives the parametrization around $t=1$
504: \begin{equation}
505:  \tau = 1-t = {1\over4}(y-1) x^2 - {1\over48} (y-1) (y-2) x^4 + \cdots
506: \end{equation} 
507: These values of the variable $t$ will be used in the
508: Schr\"odinger equation~(\ref{ellequ}) to express in terms of the variable $x$
509: the potential $\pm(v/4\hbar^2)(t-b_1)(t-b_2)$. The sign can change since in the
510: calculation of $x$, we forgot the imaginary unit $i$ which can appear according
511: to the sign of the term under the square root.
512: 
513: In the large $v$ limit, $b_1$ is near $1$ and $b_2$ is near $y$ and the lowest
514: relevant order for the potential for $t$ around $y$ is:
515: $$ {1\over4\hbar^2} v(y-1)\Bigl( (y-b_2) - {1\over4}y(y-1) x^2 \Bigr) $$
516: which is the potential for a harmonic oscillator with energy
517: $(v/4\hbar^2)(y-1)(y-b_2)$ and oscillator strength $(v/4\hbar^2)y(y-1)^2$.
518: This gives $y-b_2= 2\hbar\sqrt{y/v} ( n_2+ 1/2)$, which is exactly the
519: semiclassical result at this order. Similarly, $1-b_1=2\hbar/\sqrt{v}
520: (n_1+1/2)$. In order to choose the terms which will contribute to the next
521: order, we should remember that for the harmonic oscillator, $x^2$ is of order
522: $v^{-1/2}$. We therefore get corrections stemming from the variation of the
523: oscillator strength, from the factor in front of $y-b_2$ and from 
524: anharmonic terms in $x^4$ which stem in part from the one in $\tau$ and in part
525: from the product of the two $x^2$ terms occurring in both factors $t-b_i$.
526: The order one corrected potential is:
527: \begin{eqnarray*}
528: {1\over4\hbar^2} v  \mkern -20mu &&\left(
529: (y-b_2)\bigl(y-1+{2\hbar\over \sqrt{v}}(n_1+{1\over2})\bigr) -{1\over4}y(y-1)^2
530: x^2\right. \\
531: &&-{\hbar\over2\sqrt{v}} \Bigl( n_1+{1\over2}+
532: (n_2+{1\over2}) \sqrt y \Bigr)y(y-1)  x^2 \left.
533: +{1\over48}y(y-1)^2(5y-1)x^4\right)\\
534: \end{eqnarray*} 
535: With the anharmonic oscillator Schr\"odinger equation written as $ - \Psi''+1/4
536: (x^2+\lambda x^4)\Psi = E \Psi$, first order perturbation theory yields the
537: energy levels $E_n = n + 1/2 +3/4 \lambda (2n^2+2n+1)$. This result is easily
538: obtained writing $x^4$ in terms of creation and annihilation operators.
539: Putting these corrections together, we get:
540: $$y-b_2=2\hbar \sqrt{ y\over v} ( n_2+{1\over 2})-
541: {\hbar^2\over v} \Bigl({2\sqrt{y}\over (y-1)}(n_1+{1\over 2})(n_2+{1\over 2}) +
542: {1\over 2}(n_2+{1\over 2})^2 +{1\over 8}{5y-1\over y-1}\Bigr)$$
543: Note that this only differs from the semiclassical result by the last term,
544: independent on the $n_i$.
545: 
546: A similar computation yields the potential developed around $t=1$ and the first
547: order correction to the energy of the corresponding anharmonic oscillator gives
548: $$ 1-b_1 = 2\hbar { 1\over \sqrt v} ( n_1+{1\over 2})+
549: {\hbar^2\over v} \Bigl({2\sqrt{y}\over (y-1)}(n_1+{1\over 2})(n_2+{1\over 2}) -
550: {1\over 2}(n_1+{1\over 2})^2 -{1\over 8}{y-5\over y-1}\Bigr)$$
551: Whatever happens outside of these two regions is exponentially suppressed for
552: large $v$, since the wave function of the harmonic oscillator are
553: already exponentially small there. Combining the values of $b_1$ and $b_2$, we
554: obtain for the energy at this order in $1/v$ a result which differs from the semiclassical one of eq.~(\ref{niveausc}) only by
555: a constant
556: $$ E_{Q} = E_{WKB} - {3\over 8} \hbar^2 $$
557: This nicely reflects our findings on Figs.\ 2 and~3 where $\hbar = 1$.
558: 
559: 
560: \section{Conclusion.}
561: The interaction of the study of the Neumann model and the generalized Lam\'e
562: equation allowed us to give precise results on the behaviour of the spectrum in
563: some limits. The generalized Lam\'e equation has independent interest since it
564: is the most general linear differential equation of second order with only
565: regular singularities at finite distance. 
566: 
567: We were able to precisely count the different cases for the positions
568: of the zeroes of the potential, showing in particular that they are all real.
569: It is also remarkable that there is a constant error in the WKB approximation.
570: This error is however independent on the level, so that it does not appear in
571: transition energies. In this integrable case, the WKB approximation gives pretty
572: good results even for low excitation numbers and the further study of its
573: behaviour around turning points at low energy could make it even more useful.
574: 
575: \begin{thebibliography}{10}
576: 
577: \bibitem{BT05}
578: M. Bellon and M. Talon,
579: \newblock {Spectrum of the quantum Neumann model.}
580: \newblock Phys. Lett. A 337 (2005), pp. 360--368.
581: 
582: \bibitem{Neumann}
583: C. Neumann, {\em De problemate quodam mechanico, quod ad primam integralium 
584: ultraellipticorum classem revocatur.} 
585: \newblock Crelle Journal 56(1859), pp. 46--63.
586: 
587: \bibitem{Moser}
588: J. Moser, {\em Various aspects of integrable Hamiltonian systems.}
589: \newblock Proc.\ CIME Bressanone, Progress in Mathematics, Birkh\"auser (1978)
590: 233.
591: 
592: \bibitem{Mu84}
593: D.~Mumford.
594: \newblock {\em Tata lectures on Theta II}.
595: \newblock Birkhauser Boston,  (1984).
596: 
597: \bibitem{AvTa90}
598: J.~Avan and M.~Talon, {\em Poisson structure and integrability of the
599: {N}eumann-{M}oser-{U}hlenbeck model}.
600: \newblock Intern. Journ. Mod. Phys. A 5 (1990), pp. 4477--4488.
601: 
602: 
603: \bibitem{BaTa92}
604: O.~Babelon and M.~Talon, {\em Separation of variables for the classical and
605: quantum Neumann model}.
606: \newblock Nucl. Phys. B379 (1992), pp. 321--339.
607: 
608: \bibitem{Gurarie}
609: D.~Gurarie, {\em Quantized Neumann problem, separable potentials on S$^n$ and 
610: the Lam\'e equation.} 
611: \newblock J. Math. Phys. 36 (1995), pp. 5355--5391.
612: 
613: \bibitem{DuVa05}
614: C.~Duval and G.~Valent, {\em
615: Quantum integrability of quadratic Killing tensors.}
616: \newblock J. Math. Phys. 46 (2005), 053516.
617: \newblock (arXiv:math-ph/0412059)
618: 
619: \bibitem{WW}
620: E.T. Whittaker and G.N. Watson.
621: \newblock {\em A course of modern analysis}.
622: \newblock Cambridge University Press,  (1902).
623: 
624: \bibitem{Lan38}
625: Rudolph E.\ Langer, {\em On the Connection Formulas and the
626: Solutions of the Wave Equation.} \newblock Phys.\ Rev.\ {\bf 51}(1938),
627: pp.\ 669--676.
628: 
629: \end{thebibliography}
630: 
631: 
632: \end{document}
633: 
634: 
635: 
636: