1: % Article for J. Phys. A, authors Vincenzo Aquilanti and
2: % Robert G. Littlejohn
3: %
4: %\documentclass[12pt]{iopart} % preprint
5: \documentclass{iopart} % journal version
6: \usepackage{iopams}
7: \usepackage{setstack}
8:
9:
10: \begin{document}
11: \title{Semiclassical analysis of Wigner $3j$-symbol}
12: \author{Vincenzo Aquilanti}
13: \address{Dipartimento di Chimica, Universit\`a di Perugia, Perugia,
14: Italy 06100}
15: \author{Hal M. Haggard, Robert G. Littlejohn and Liang Yu}
16: \address{Department of Physics, University of
17: California, Berkeley, California 94720 USA}
18:
19: \ead{robert@wigner.berkeley.edu}
20:
21: \begin{abstract}
22:
23: We analyze the asymptotics of the Wigner $3j$-symbol as a matrix
24: element connecting eigenfunctions of a pair of integrable systems,
25: obtained by lifting the problem of the addition of angular momenta
26: into the space of Schwinger's oscillators. A novel element is the
27: appearance of compact Lagrangian manifolds that are not tori, due to
28: the fact that the observables defining the quantum states are
29: noncommuting. These manifolds can be quantized by generalized
30: Bohr-Sommerfeld rules and yield all the correct quantum numbers. The
31: geometry of the classical angular momentum vectors emerges in a clear
32: manner. Efficient methods for computing amplitude determinants in
33: terms of Poisson brackets are developed and illustrated.
34:
35: \end{abstract}
36:
37: \pacs{03.65.Sq, 02.20.Qs, 02.30.Ik, 02.40.Yy}
38:
39: %\maketitle %use this if you want to have title on separate page
40:
41: \section{Introduction}
42:
43: This article is a study of the asymptotics of the Wigner $3j$-symbol
44: from the standpoint of semiclassical mechanics, that is, essentially
45: multidimensional WKB theory for integrable systems. The principal
46: result itself, the leading asymptotic expression for the $3j$-symbol,
47: has been known since Ponzano and Regge (1968). Nevertheless our
48: analysis presents several novel features. One is the exploration of
49: Lagrangian manifolds in phase space that are not tori (the usual case
50: for eigenstates of integrable systems). Instead, one of the states
51: entering into the $3j$-symbol is supported semiclassically on a
52: Lagrangian manifold that is a nontrivial 3-torus bundle over $SO(3)$.
53: This manifold can be quantized by generalized Bohr-Sommerfeld rules,
54: whereupon it yields the exact eigenvalues required by the quantum
55: $3j$-symbol, as well as the correct amplitude and phase of its
56: asymptotic form. This unusual Lagrangian manifold arises because the
57: quantum state in question is an eigenstate of a set of noncommuting
58: operators. Other novel features include the expression of the asymptotic
59: phase of the $3j$-symbol in terms of the phases of Schwinger's
60: harmonic oscillators and the determination of stationary phase points
61: by geometrically transparent operations on angular momentum vectors in
62: three-dimensional space. Yet another is the representation of
63: multidimensional amplitude determinants as matrices of Poisson
64: brackets. Representations of this type have been known for some time,
65: but they are generalized here to the case of sets of noncommuting
66: operators. The final result is a one-line derivation of the amplitude
67: of the asymptotic form of the $3j$-symbol. Similarly brief
68: derivations are possible for the amplitudes of the $6j$- and
69: $9j$-symbols.
70:
71: In addition our analysis of the $3j$-symbol may prove to be useful for
72: the asymptotic study of the $3nj$-symbols for higher $n$. The leading
73: order asymptotics of the $6j$-symbol were derived by Ponzano and Regge
74: (1968), but the understanding of the asymptotics of the $9j$-symbol is
75: still incomplete. These symbols are important in many applications in
76: atomic, molecular and nuclear physics, for example, the $9j$-symbols
77: are needed in atomic physics to convert from an $LS$-coupled basis to
78: a $jj$-coupled basis. These symbols are all examples of closed spin
79: networks, of which more elaborate examples occur in applications, each
80: of which presents a challenge to asymptotic analysis. Moreover in
81: recent years new interest in this subject has arisen from researches
82: into quantum computing (Marzuoli and Rasetti, 2005) and quantum
83: gravity, where new derivations of the asymptotics of the Wigner
84: $6j$-symbol have been produced as well as generalizations to other
85: groups such as the Lorentz group. The $3nj$-symbols and their
86: asymptotics have also been used recently in algorithms for molecular
87: quantum mechanics (De~Fazio \etal\ 2003 and Anderson and Aquilanti,
88: 2006), which exploit the connections with the theory of discrete
89: orthogonal polynomials (Aquilanti \etal\ 1995, 2001a,b and references
90: therein).
91:
92: The asymptotic formula for the $3j$-symbol is closely related to that
93: for the $6j$-symbol, being a limiting case of the latter. These were
94: first derived by Ponzano and Regge (1968), using intuitive methods and
95: building on Wigner's earlier result for the amplitude of the
96: $6j$-symbol (Wigner, 1959). Later Neville (1971) analyzed the
97: asymptotics of the $3j$- and $6j$- symbols by a discrete version of
98: WKB theory, applied to the recursion relations satisfied by those
99: symbols, without apparently knowing of the work of Ponzano and Regge.
100: His formulas are not presented in a particularly transparent or
101: geometrical manner, but appear to reproduce some of the results of
102: Ponzano and Regge. The formula for the $3j$-symbol (in the form of
103: Clebsch-Gordan coefficients) was later derived again by Miller (1974),
104: who presented it as an example of his general theory of semiclassical
105: matrix elements of integrable systems. Miller called on the fact that
106: the phase of the semiclassical matrix element is a generating function
107: of a canonical transformation, and used the classical transformation
108: that most obviously corresponds to the quantum addition of angular
109: momenta to reconstruct the generating function. The method leads to a
110: difficult integral, which, once done, yields the five terms in the
111: phase of the asymptotic formula for the $3j$-symbol. Somewhat later
112: Schulten and Gordon (1975a,b) presented a rigorous derivation of the
113: Ponzano and Regge results for the $3j$- and $6j$-symbols, using
114: methods similar to those of Neville but carrying them out in a more
115: thorough and elegant manner. Schulten and Gordon also provided
116: uniform approximations for the transition from the classical to
117: nonclassical regimes, work that has recently been reanalyzed (Geronimo
118: \etal, 2004) and extended to non-Euclidean and quantum groups (Taylor
119: and Woodward, 2005). Somewhat later Biedenharn and Louck (1981b)
120: presented a review and commentary of the results of Ponzano and Regge,
121: as well as a proof based on showing that the result satisfies
122: asymptotically a set of defining relations for the $6j$-symbol. More
123: recently the asymptotics of the $3j$-symbol was derived again by
124: Reinsch and Morehead (1999), working with an integral representation
125: constructed out of Wigner's single-index sum for the Clebsch-Gordan
126: coefficients. About the same time, Roberts (1999) derived the Ponzano
127: and Regge results for the $6j$-symbol, using methods of geometric
128: quantization. Finally, Freidel and Louapre (2003) presented a
129: derivation of the asymptotic expression for the square of the
130: $6j$-symbol, based on an analysis of an $SU(2)$ path integral. This
131: work was part of a larger study of generalizations of the $6j$-symbol
132: to other groups (for example, the $10j$-symbol) that are important in
133: quantum gravity. See also Barrett and Steele (2003) and Baez,
134: Christensen and Egan (2002).
135:
136: There are many variations on the calculation of the asymptotic forms
137: of the $3nj$-symbols that have been considered by different authors.
138: There are asymptotic forms inside and outside the classically allowed
139: regions; uniform approximations connecting two or more of these
140: regions; asymptotic forms when only some of the quantum numbers are
141: large and others small; and higher order terms. Ponzano and Regge
142: (1968) covered many of these issues, while Reinsch and Morehead
143: computed some higher order terms.
144:
145: The outline of this paper is as follows. In Sec.~\ref{scwfis} we
146: review the semiclassical mechanics of integrable systems in the
147: generic case that one has sets of commuting observables, drawing
148: attention to an expression for the amplitude determinant in terms of
149: Poisson brackets. In Sec.~\ref{Schwingermodel} we review the
150: Schwinger model for representing angular momentum operators in terms
151: of harmonic oscillators. This model allows us to express angular
152: momentum eigenstates in terms of wave functions on ${\mathbb R}^n$,
153: which we use in Sec.~\ref{3jSchwinger} to express the $3j$-symbols in
154: terms of scalar products of such functions. In Sec.~\ref{cmSchwinger}
155: we study the Schwinger model from a classical standpoint, in which an
156: important element is the reduction of the Schwinger phase
157: space (the ``large phase space'') by the torus group $T^3$, producing
158: the Poisson manifold ${\mathbb R}^3 \times {\mathbb R}^3 \times
159: {\mathbb R}^3$ (``angular momentum space'') and the reduced phase
160: space $S^2 \times S^2 \times S^2$ (the ``small phase space''). In
161: Secs.~\ref{jmtori} and \ref{Wignermanifold} we study the two
162: Lagrangian manifolds that support the states whose scalar product is
163: the $3j$-symbol. One is a conventional invariant torus (the
164: ``$jm$-torus''), but the other, what we call the ``Wigner manifold,''
165: is compact and Lagrangian but not a torus. This manifold supports
166: Wigner's state of zero total angular momentum that enters into the
167: definition of the $3j$-symbols. In Secs.~\ref{intersections} we study
168: the intersections of the $jm$-torus and the Wigner manifold, which are
169: the stationary phase points of the $3j$-symbol, and show how these can
170: be found by elementary geometrical considerations in three-dimensional
171: space (that is, by rotating angular momentum vectors). The
172: intersection of the two manifolds turns out to be a pair of 4-tori.
173: In Sec.~\ref{actionintegrals} we compute the action integrals along
174: the respective Lagrangian manifolds to points on the two 4-tori, whose
175: difference is the Ponzano and Regge phase of the $3j$-symbol. In
176: Sec.~\ref{BSquant} we apply generalized Bohr-Sommerfeld quantization
177: to the $jm$-torus and the Wigner manifold, a standard procedure for
178: the $jm$-torus, although it leads in an interesting way to the extra
179: $1/2$ in the classical values representing the lengths of the angular
180: momentum vectors. This extra $1/2$ was guessed by Ponzano and Regge
181: and Miller and derived systematically by Schulten and Gordon, Reinsch
182: and Morehead and by us, although it is missing from the results of
183: Roberts. In our work it is essentially a Maslov index. In
184: Sec.~\ref{ampdet} we generalize known expressions for the amplitude
185: determinant of semiclassical matrix elements of integrable systems in
186: terms of Poisson brackets to the case of collections of noncommuting
187: observables (whose level sets nevertheless are Lagrangian). The
188: result allows us to compute the amplitude of the $3j$-symbol as a
189: $2\times 2$ matrix of Poisson brackets. We then put all the pieces
190: together to obtain the final asymptotic form. Finally, in
191: Sec.~\ref{conclusions} we present some comments on the work, prospects
192: for further work, and conclusions.
193:
194: \section{Semiclassical wave functions for integrable systems}
195: \label{scwfis}
196:
197: The semiclassical mechanics of integrable systems is well understood
198: (Einstein, 1917; Brillouin, 1926; Keller, 1958; Percival, 1973; Berry
199: and Tabor, 1976; Gutzwiller, 1990; Brack and Bhaduri, 1997; Cargo
200: \etal, 2005a, 2005b). Here we summarize the basic facts, some of
201: which require modification for our application.
202:
203: We consider the quantum mechanics of a particle moving in ${\mathbb
204: R}^n$ (with wave function $\psi(x_1,\ldots,x_n)$ and Hilbert space
205: $L^2({\mathbb R}^n)$). We speak of an integrable system if we have a
206: complete set of commuting observables $\{{\hat A}_1, \ldots, {\hat
207: A_n}\}$ acting on this Hilbert space. We use hats to distinguish
208: quantum operators from classical quantities with a similar meaning.
209: Sometimes the Hamiltonian is one of these operators or a function of
210: them, but in our application there is no Hamiltonian, or, rather all
211: the ${\hat A}_i$'s are Hamiltonians on an equal footing. These
212: operators may be converted into their classical counterparts by the
213: Weyl transform (Weyl 1927, Wigner 1932, Groenewold 1946, Moyal 1949,
214: Voros 1977, Berry 1977, Balazs and Jennings 1984, Hillery \etal\ 1984,
215: Littlejohn 1986, McDonald 1988, Estrada \etal\ 1989, Gracia-Bond\'\i a
216: and V\'arilly 1995 and Ozorio de Almeida 1998). The Weyl transforms
217: (or Weyl ``symbols'') of these operators are functions on the
218: classical phase space ${\mathbb R}^{2n}$, that is, functions of
219: $(x_1,\ldots,x_n; p_1,\ldots,p_n)$. They are normally even power
220: series in $\hbar$, as we assume, of which the leading term is the
221: ``principal symbol.'' We denote the principal symbols of $\{{\hat
222: A}_1, \ldots, {\hat A_n}\}$ by $\{A_1, \ldots, A_n\}$ (without the
223: hats). In view of the Moyal star product representation (Moyal 1949)
224: of the vanishing commutators $[{\hat A}_i, {\hat A}_j]=0$, the
225: principal symbols Poisson commute, $\{A_i, A_j\}=0$, thus defining a
226: classically integrable system (Arnold 1989, Cushman and Bates 1997).
227: (We use curly brackets $\{\;\}$ both to denote a set and for Poisson
228: brackets.) Then according to the Liouville-Arnold theorem (Arnold
229: 1989), the compact level sets of $\{A_1, \ldots, A_n\}$ are
230: generically $n$-tori. The Hamiltonian vector fields generated by the
231: $A_i$ are commuting and linearly independent on the tori; thus the
232: tori are not only the level sets of the $A_i$, they are also the
233: orbits of the Abelian group generated by the corresponding Hamiltonian
234: flows. One can define an action function $S$ on a torus as the
235: integral of $\sum_i p_i \, dx_i$ relative to some initial point; it is
236: multivalued because of the topologically distinct paths going from the
237: initial to the final point, but otherwise is independent of the path.
238:
239: Let $A_i=a_i$ be one of these tori ($A_i$ are the functions, $a_i$ the
240: values). The torus has a projection onto configuration space defining
241: a classically allowed region in that space; the inverse projection is
242: multivalued. The function $S$ may be projected onto configuration
243: space, defining a function we shall denote by $S_k(x,a)$ (where for
244: brevity $x$ and $a$ stand for $(x_1, \ldots,x_n)$ and $(a_1, \ldots,
245: a_n)$, etc). Here $k$ labels the branches of the inverse projection;
246: function $S_k$ has an additional multivaluedness due to the choice of
247: contour connecting initial and final points on the torus. Then as
248: explained by Arnold (1989), $S(x,a)$ is the generating function of the
249: canonical transformation $(x,p) \to (\alpha,A)$, where
250: $\alpha=(\alpha_1, \ldots, \alpha_n)$ is the set of angle variable
251: conjugate to the conserved quantities $(A_1, \ldots, A_n)$. Action
252: variables may be defined in the usual way as $(1/2\pi) \oint p\,dx$
253: around the independent basis contours on the torus; these are
254: functions of the $A_i$ and their conjugate variables are the angles
255: that cover the torus once when varying between $0$ and $2\pi$.
256: Sometimes however it is more convenient to work with the $A_i$ instead
257: of the actions (the $A_i$ are not necessarily actions, and their flows
258: are not necessarily periodic on the torus).
259:
260: Tori are quantized, that is, associated with a consistent solution of
261: the simultaneous Hamiltonian-Jacobi and amplitude transport equations
262: for the operators ${\hat A}_i$, only if they satisfy the
263: Bohr-Sommerfeld or EBK quantization conditions, discussed in
264: Sec.~\ref{BSquant}. Associated with a quantized torus is a
265: semiclassical wave function in configuration space, which in the
266: classically allowed region is given by
267: \begin{equation}
268: \psi(x) = \langle x \vert a \rangle =
269: \frac{1}{\sqrt V} \sum_k |\Omega_k|^{1/2}
270: \exp\{i[S_k(x,a) - \mu_k\pi/2]\}.
271: \label{scpsi}
272: \end{equation}
273: The meaning of this formula is the following. First, here and below
274: we set $\hbar=1$. Next, given the point $x$ in the classically
275: allowed region, its inverse projection onto the quantized torus is a
276: set of points indexed by $k$. We assume the projection is nonsingular
277: at these points (we are not at a caustic). The phase $S_k(x,a)$ is
278: the integral of $p\,dx$ from a given initial point on the torus to the
279: $k$-th point of the inverse projection, and $\mu_k$ is the Maslov
280: index (Maslov 1981, Mishchenko \etal\ 1990, de Gosson 1997) of the
281: same path. The amplitude determinant $\Omega_k$ is given by
282: \begin{equation}
283: \Omega_k= \det\frac{\partial^2 S_k(x,a)}{\partial x_i \partial a_j}
284: =[ \det \{x_i,A_j\} ]^{-1},
285: \label{ampldet}
286: \end{equation}
287: where in the second form the Poisson brackets are evaluated on the
288: $k$-th branch of the inverse mapping from $x$ to the Lagrangian
289: manifold. The amplitude determinant is a density on configuration
290: space (to within the semiclassical approximation, the probability
291: density corresponding to a single branch), which is the projection
292: onto configuration space of a density on the torus. The latter
293: density is required to be invariant under the Hamiltonian flows
294: generated by the $A_i$ (this is the meaning of the amplitude transport
295: equations for the $A_i$); in terms of the variables $\alpha_i$
296: conjugate to the $A_i$ this means that the density is constant (it is
297: the $n$-form $d\alpha_1 \wedge \ldots \wedge d\alpha_n$). Finally,
298: the quantity $V$ in (\ref{scpsi}) is the volume of the torus,
299: measured with respect to this density. If the $A_i$ are action
300: variables, then $V=(2\pi)^n$. The overall phase of the wave function
301: (its phase convention) is determined by the choice of the initial
302: point on the torus.
303:
304: Now let $\{{\hat A}_1, \ldots, {\hat A}_n\}$ and $\{{\hat B}_1,
305: \ldots, {\hat B}_n\}$ be two complete sets of commuting observables,
306: with principal symbols $A_i$ and $B_i$, conjugate angles $\alpha_i$
307: and $\beta_i$ and action functions $S_A(x,a)$ and $S_B(x,b)$, and let
308: $a$ and $b$ refer to two quantized tori (an $A$-torus and a
309: $B$-torus). We assume initially that the two sets of functions $A_i$
310: and $B_i$ are independent. We compute $\langle b \vert a \rangle$
311: as an integral of the wave functions over $x$, evaluated by the
312: stationary phase approximation. The stationary phase points are
313: geometrically the intersections of the $A$-torus with the $B$-torus.
314: Generically the two tori intersect in finite set of isolated points
315: that we index by $k$, denoting the corresponding $\alpha$ and $\beta$
316: values by $\alpha_k$ and $\beta_k$. (For given $k$, $\alpha_k$ and
317: $\beta_k$ refer to the same point in phase space.) Then the result is
318: \begin{equation}
319: \langle b\vert a\rangle =
320: \frac{(2\pi i)^{n/2}}{\sqrt{V_A V_B}}
321: \sum_k |\Omega_k|^{1/2}
322: \exp\{ i[S_A(\alpha_k)-S_B(\beta_k)-\mu_k\pi/2]\}.
323: \label{abmatrixelement}
324: \end{equation}
325: Here $V_A$ and $V_B$ are the volumes of the respective tori, as in
326: (\ref{scpsi}), and the actions $S_A$ and $S_B$ are considered
327: functions of the $\alpha$ or $\beta$ coordinates on the respective
328: tori.
329:
330: As shown by Littlejohn (1990), the amplitude determinant $\Omega_k$
331: can be written in terms of the Poisson brackets of the observables
332: $A_i$ and $B_i$,
333: \begin{equation}
334: \Omega_k = [\det \{A_i, B_j\}]^{-1}.
335: \label{PBdet}
336: \end{equation}
337: The Maslov index $\mu_k$ in (\ref{abmatrixelement}) is not the
338: same as in (\ref{scpsi}).
339:
340: Another case considered by Littlejohn (1990) is the one in which some
341: of the $A_i$ are functionally dependent on some of the $B_i$. For
342: this case it is convenient to assume that the first $r$ of the two
343: sets of variables $A$ and $B$ are functionally independent, while the
344: last $n-r$ are identical, so that $A=\{A_1, \ldots, A_r, A_{r+1},
345: \ldots, A_n\}$ and $B=\{B_1, \ldots, B_r, A_{r+1}, \ldots, A_n\}$.
346: Then the stationary phase points are still the intersections of the
347: two $n$-tori, but now the intersections are generically a finite set
348: of isolated $(n-r)$-tori, upon which linearly independent vector
349: fields are the Hamiltonian vector fields associated with the
350: $(A_{r+1}, \ldots, A_n)$. Such an $(n-r)$-torus is the orbit of the
351: Abelian group action generated by the corresponding Hamiltonian flows.
352: In this case we find
353: \begin{equation}
354: \langle b \vert a \rangle =
355: \frac{(2\pi i)^{r/2}}{\sqrt{V_A V_B}}
356: \sum_k V_k |\Omega_k|^{1/2}
357: \exp\{i[S_A(\alpha_k)-S_B(\beta_k)-\mu_k\pi/2]\},
358: \label{abmatrixelement1}
359: \end{equation}
360: where now $V_k$ is the volume of the $k$-th intersection (an
361: $(n-r)$-torus on which the volume measure is $d\alpha_{r+1} \wedge
362: \ldots \wedge d\alpha_n$), and where now the amplitude determinant
363: $\Omega_k$ is still given by (\ref{PBdet}), except that it is
364: understood that only the first $r$ of the $A$'s and $B$'s enter (thus,
365: it is an $r\times r$ determinant instead of an $n \times n$ one). The
366: phase difference $S_A - S_B$ for branch $k$ can be evaluated at any
367: point on the $(n-r)$-torus which is the intersection, since the
368: integral of $p\,dx$ back and forth along a path lying in the
369: intersection vanishes.
370:
371: \section{The Schwinger model}
372: \label{Schwingermodel}
373:
374: The Schwinger (or $SU(2)$ or boson) model for angular momenta is
375: explained well in Schwinger's original paper (reprinted by Biedenharn
376: and van Dam (1965), the original 1952 paper being unpublished), and
377: reworked in an interesting way by Bargmann (1962). For further
378: perspective see Biedenharn and Louck (1981a) and Smorodinskii and
379: Shelepin (1972). Introductions are given by Sakurai (1994) and
380: Schulman (1981). Here we define the notation for the Schwinger model
381: and emphasize some aspects that will be important for our application.
382:
383: In the Schwinger model each independent angular momentum vector is
384: associated with two harmonic oscillators. We shall refer to the
385: $1j$-, $3j$-, etc models, depending on how many independent angular
386: momenta there are. The number of $j$'s in the model is not
387: necessarily the number of $j$'s in the Wigner symbol; for example,
388: Miller (1974) used a $2j$-model to study the Clebsch-Gordan
389: coefficients, essentially the $3j$-symbols.
390:
391: We start with the $1j$-model, for which there are two harmonic
392: oscillators indexed by Greek indices $\mu, \nu, \ldots = 1,2$. (These
393: are just labels of the two oscillators; sometimes other labels such as
394: $1/2$, $-1/2$ are more suitable.) The wave functions are $\psi(x_1,x_2)$
395: and the Hilbert space is ${\cal H} = L^2({\mathbb R}^2)$. We write
396: ${\hat H}_\mu=(1/2)({\hat x}_\mu^2 + {\hat p}_\mu^2)$ for the two
397: oscillator Hamiltonians, and we define ${\hat H}=\sum_\mu {\hat
398: H}_\mu$. The eigenvalues of ${\hat H}$ are $n+1$, with
399: $n=0,1,\ldots$, and energy level $E_n$ is $(n+1)$-fold degenerate. We
400: introduce usual annihilation and creation operators $a_\mu= ({\hat
401: x}_\mu + i{\hat p}_\mu)/\sqrt{2}$, $a^\dagger_\mu = ({\hat x}_\mu -i
402: {\hat p}_\mu)/\sqrt{2}$, omitting the hats on the $a$'s and
403: $a^\dagger$'s since these will always be understood to be operators.
404: We define operators
405: \begin{equation}
406: {\hat I} = \frac{1}{2} \sum_\mu a^\dagger_\mu a_\mu
407: =\frac{1}{2} ({\hat H} - 1)
408: \label{Imudef}
409: \end{equation}
410: and
411: \begin{equation}
412: {\hat J}_i = \frac{1}{2}
413: \sum_{\mu\nu} a^\dagger_\mu \, \sigma^i_{\mu\nu} \, a_\nu,
414: \label{Jidef}
415: \end{equation}
416: where $\sigma^i$ is the $i$-th Pauli matrix. Here and below we use
417: indices $i,j,\ldots = 1,2,3$ (or $x,y,z$ if that is more clear) to
418: denote the Cartesian components of a 3-vector. Notice that ${\hat I}$
419: and ${\hat J}_i$ are quadratic functions of the $x$'s and $p$'s of the
420: system. The eigenvalues of ${\hat I}$ are $n/2$ for $n=0,1,\ldots$.
421: These operators satisfy the commutation relations $[{\hat I},{\hat
422: J}_i] =0$ and $[{\hat J}_i, {\hat J}_j]= i\sum_k \epsilon_{ijk} {\hat
423: J}_k$. We also define ${\hat{\bf J}}^2 = \sum_i {\hat J}_i^2$, so
424: that $[{\hat I},{\hat{\bf J}}^2]=0$ and $[{\hat J_i}, {\hat{\bf
425: J}}^2]=0$. It avoids some confusion with indices to always denote the
426: square of a vector by a bold face symbol, as we have done here. We
427: note the important operator identity ${\hat{\bf J}}^2 = {\hat I}({\hat
428: I} +1)$, expressing the quartic operator ${\hat{\bf J}}^2$ as a
429: function of the quadratic operator ${\hat I}$.
430:
431: From this identity and the known eigenvalues of ${\hat I}$ it follows
432: that the eigenvalues of ${\hat{\bf J}}^2$ are $(n/2)[(n/2)+1]$, for
433: $n=0,1,\ldots$, which leads us to identify $n/2$ with
434: $j=0,1/2,1,\ldots$, the usual angular momentum quantum number. The
435: $n$-th (or $j$-th) eigenspace of ${\hat H}$ or ${\hat I}$ is
436: $(2j+1)$-dimensional, and so must contain a single copy of the $j$-th
437: irrep of $SU(2)$. Each irrep (both integer and half-integer values of
438: $j$) occurs precisely once in the Hilbert space ${\cal H}$. We denote
439: these subspaces by ${\cal H}_j$, and write ${\cal H} = \sum_j \oplus
440: {\cal H}_j$. The standard basis in ${\cal H}_j$ is the eigenbasis of
441: ${\hat J}_z =\frac{1}{2}({\hat H}_1 - {\hat H}_2)$, with the usual
442: quantum number $m$, so that if $n_\mu$ are the usual quantum numbers
443: of the oscillators ${\hat H}_\mu$, then $n_1=j+m$, $n_2=j-m$. The
444: simultaneous eigenstates of ${\hat{\bf J}}^2$ and ${\hat J}_z$ are
445: $\vert jm\rangle$ or $\vert n_1 n_2
446: \rangle$.
447:
448: In the $Nj$-model we index the angular momenta with indices $r,s,\dots
449: = 1, \ldots, N$. The oscillators are now labelled ${\hat H}_{r\mu}$
450: with coordinates and momenta ${\hat x}_{r\mu}$ and ${\hat p}_{r\mu}$
451: and annihilation and creation operators $a_{r\mu}$ and
452: $a^\dagger_{r\mu}$. The wave functions are now $\psi(x_{11}, x_{12},
453: x_{21}, \ldots, x_{N2})$ and the Hilbert space is $L^2({\mathbb
454: R}^{2N})$. We define operators
455: \begin{eqnarray}
456: {\hat I}_r &= \frac{1}{2} \sum_\mu a^\dagger_{r\mu} a_{r\mu},
457: \label{Irdef} \\
458: {\hat J}_{ri} &= \frac{1}{2} \sum_{\mu\nu}
459: a^\dagger_{r\mu} \, \sigma^i_{\mu\nu} \, a_{r\nu},
460: \label{Jridef} \\
461: {\hat{\bf J}}^2_r &= \sum_i {\hat J}_{ri}^2,
462: \label{Jrsquareddef} \\
463: {\hat J}_i &= \sum_r {\hat J}_{ri},
464: \qquad \hbox{\rm or} \qquad
465: {\hat{\bf J}} = \sum_r {\hat{\bf J}}_r,
466: \label{Jtotalidef} \\
467: {\hat{\bf J}}^2 & = \sum_i {\hat J}_i^2,
468: \label{Jtotalsquareddef}
469: \end{eqnarray}
470: most of which are obvious generalizations from the case $N=1$. These
471: satisfy the identity
472: \begin{equation}
473: {\hat{\bf J}}_r^2 = {\hat I}_r({\hat I}_r+1),
474: \label{JIidentity}
475: \end{equation}
476: and the commutation relations $[{\hat I}_r, {\hat J}_{si}] = [{\hat
477: I}_r, {\hat{\bf J}}^2_s]=0$. Each angular momentum vector ${\hat{\bf
478: J}}_r$ also obeys the standard commutation relations among its
479: components and square, which we omit, as does any sum of these angular
480: momenta (partial or total).
481:
482: The angular momenta generate an action of $[SU(2)]^N$ on the Hilbert
483: space (one copy for each pair of oscillators). Here we discuss only
484: the simultaneous rotation of all oscillator degrees of freedom by the
485: same element of $SU(2)$, which is generated by the total angular
486: momentum, but partial rotation operators can also be defined and are
487: useful. We begin with the commutation relations,
488: \begin{eqnarray}
489: [{\hat J}_i, a_{r\mu}] &= -\frac{1}{2} \sum_\nu \sigma^i_{\mu\nu} \,
490: a_{r\nu},
491: \label{Jacomrel} \\ \relax
492: [{\hat J}_i, a^\dagger_{r\mu}] &=
493: +\frac{1}{2} \sum_\nu a^\dagger_{r\nu} \,\sigma^i_{\nu\mu},
494: \label{Jadaggercomrel}
495: \end{eqnarray}
496: which define the transformation properties of the operators
497: $a_{r\mu}$, $a^\dagger_{r\mu}$ under infinitesimal rotations. We
498: define a finite rotation operator in axis-angle or Euler angle form by
499: \begin{eqnarray}
500: U({\bf n},\theta) &= \exp(-i\theta {\bf n} \cdot {\bf J}),
501: \label{Uaxisangledef} \\
502: U(\alpha,\beta,\gamma) &= U({\bf z},\alpha) U({\bf y},\beta)
503: U({\bf z},\gamma),
504: \label{UEulerangledef}
505: \end{eqnarray}
506: where ${\bf n}$ is a unit vector defining an axis and $\theta$ an
507: angle of rotation about that axis, and where ${\bf x}$, ${\bf y}$ and
508: ${\bf z}$ are respectively the unit vectors along the three coordinate
509: axes. The $U$ operators form a faithful representation of $SU(2)$.
510:
511: We use the symbol $u({\bf n},\theta)$ or $u(\alpha,\beta,\gamma)$
512: for the $2 \times 2$ matrices belonging to $SU(2)$, in axis-angle or
513: Euler angle parameterization (not to be confused with the $U$
514: operators that act on the Hilbert space of the $2N$ oscillators). Thus
515: \begin{equation}
516: u({\bf n},\theta) =
517: \exp(-i\theta{\bf n}\cdot \bsigma/2)=
518: \cos\theta/2 - i{\bf n}\cdot {\bsigma}
519: \sin\theta/2.
520: \label{udef}
521: \end{equation}
522: The exponentiated versions of equations (\ref{Jacomrel}) and
523: (\ref{Jadaggercomrel}) are
524: \begin{equation}
525: \eqalign{
526: U^\dagger a_{r\mu} U &= \sum_\nu u_{\mu\nu} \, a_{r\nu}, \\
527: U^\dagger a^\dagger_{r\mu} U &= \sum_\nu
528: a^\dagger_{r\nu}\, (u^{-1})_{\nu\mu},}
529: \label{rotatea}
530: \end{equation}
531: where both $U$ and $u$ have the same parameterization. In the language
532: of irreducible tensor operators the pair of operators
533: $(a_{r1},a_{r2})$ transforms as a spin-$1/2$ operator.
534:
535: Similarly, vector operators are the angular momenta themselves, which
536: satisfy the conjugation relations,
537: \begin{equation}
538: U^\dagger {\hat J}_{ri} U = \sum_j R_{ij} {\hat J}_{rj},
539: \label{rotateJr}
540: \end{equation}
541: where $R$ is the $3\times 3$ orthogonal rotation matrix with the same
542: axis and angle as $U$. The relation between $R$ and $u$ (with the
543: same axis and angle) is
544: \begin{equation}
545: R_{ij} = \frac{1}{2} \mathop{\rm tr}(U^\dagger \sigma_i U \sigma_j).
546: \label{Ruproj}
547: \end{equation}
548: This is the usual projection from $SU(2)$ to $SO(3)$, in which the
549: inverse image of a given $R \in SO(3)$ is a pair $(u,-u)$ in $SU(2)$.
550:
551: \section{The Wigner $3j$-symbols in the Schwinger Model}
552: \label{3jSchwinger}
553:
554: We now define the $3j$-symbols in the context of the Schwinger
555: model. We take the $3j$-model, $N=3$. One complete set of commuting
556: observables on the Hilbert space ${\cal H} \otimes {\cal H}
557: \otimes {\cal H}$ is $({\hat I}_1, {\hat I}_2, {\hat I}_3, {\hat
558: J}_{1z}, {\hat J}_{2z}, {\hat J}_{3z})$, with corresponding eigenstates
559: $\vert j_1j_2j_3m_1m_2m_3 \rangle = \vert j_1 m_1 \rangle \vert j_2
560: m_2 \rangle \vert j_3 m_3\rangle$. Another complete set arises in the
561: usual problem of addition of three angular momenta, in which we
562: consider the values of $j$ and $m$ (the quantum numbers of ${\hat{\bf
563: J}}^2$ and ${\hat J}_z$) that occur in the product space ${\cal
564: H}_{j_1} \otimes {\cal H}_{j_2} \otimes {\cal H}_{j_3}$ for fixed
565: values of $(j_1,j_2,j_3)$, a subspace of ${\cal H} \otimes {\cal H}
566: \otimes {\cal H}$. The set of the five commuting operators $({\hat
567: I}_1, {\hat I}_2, {\hat I}_3, {\hat J}_3, {\hat{\bf J}}^2)$ that
568: arises in this way is however not complete (the simultaneous
569: eigenstates in general possess degeneracies), so to resolve these we
570: introduce a sixth commuting operator, conventionally taken to be
571: ${\hat{\bf J}}^2_{12} = ({\hat{\bf J}}_1 + {\hat{\bf J}}_2)^2$ with
572: quantum number $j_{12}$ (${\hat{\bf J}}_{23}^2$ or ${\hat{\bf
573: J}}_{13}^2$ will also work).
574:
575: The Wigner $3j$-symbols only involve the case $j=0$, but we mention
576: the others anyway because the foliation of the classical phase space
577: into Lagrangian manifolds involves the other values. The usual rules
578: for the addition of angular momenta show that if $(j_1,j_2,j_3)$
579: satisfy the triangle inequality, then there exists precisely a
580: one-dimensional subspace of ${\cal H}_{j_1} \otimes {\cal H}_{j_2}
581: \otimes {\cal H}_{j_3}$ with $j=0$; if they do not, then no such
582: subspace exists. If we enlarge our point of view to the full Hilbert
583: space ${\cal H} \otimes {\cal H} \otimes {\cal H}$, then there is an
584: infinite dimensional subspace with $j=0$, a basis in which is
585: specified by all triplets $(j_1,j_2,j_3)$ that satisfy the triangle
586: inequalities. If $j=0$, then the quantum number $m$ is superfluous,
587: since $m=0$; the quantum number $j_{12}$ is superfluous as well, since
588: $j_{12}=j_3$.
589:
590: We note that if $\langle \psi \vert {\hat{\bf J}}^2
591: \vert \psi \rangle=0$ for any state $\vert \psi \rangle$, then ${\hat
592: J}_i \vert \psi \rangle=0$ for $i=1,2,3$. Although the components of
593: ${\hat{\bf J}}$ do not commute and so do not possess simultaneous
594: eigenstates in general, the case of a state with $j=0$ is an
595: exception, since it is a simultaneous eigenstate of all three
596: components of ${\hat{\bf J}}$ with eigenvalues 0. With this in mind
597: we denote the basis of states in the subspace of the full Hilbert
598: space with $j=0$ by $\vert j_1j_2j_3 {\bf 0} \rangle$, where the zero
599: vector ${\bf 0}$ indicates the vanishing eigenvalues of ${\hat{\bf
600: J}}$. These basis states are also eigenstates of the operators
601: ${\hat{\bf J}}_{ij}^2$, for example, ${\hat{\bf J}}_{12}^2 \vert
602: j_1j_2j_3 {\bf 0}\rangle = j_3(j_3+1) \vert j_1j_2j_3 {\bf
603: 0}\rangle$.
604:
605: When the phase of the state $\vert j_1j_2j_3 {\bf 0} \rangle$ is
606: chosen to agree with Wigner's convention for the phases of the
607: $3j$-symbols, we have
608: \begin{equation}
609: \left(
610: \begin{array}{ccc}
611: j_1 & j_2 & j_3 \\
612: m_1 & m_2 & m_3
613: \end{array}
614: \right) = \langle j_1j_2j_3 m_1m_2m_3 \vert j_1 j_2 j_3 {\bf 0}
615: \rangle.
616: \label{3jdef}
617: \end{equation}
618: In this manner we have expressed the $3j$-symbol as a matrix element
619: connecting the eigenstates of two sets of observables, $({\hat I}_1,
620: {\hat I}_2, {\hat I}_3, {\hat J}_{1z}, {\hat J}_{2z}, {\hat J}_{3z})$
621: on the left and $({\hat I}_1, {\hat I}_2, {\hat I}_3, {\hat J}_x,
622: {\hat J}_y, {\hat J}_z)$ on the right. Since the second set is
623: noncommuting, we will require a generalization of
624: (\ref{abmatrixelement}) to compute the semiclassical approximation to
625: the $3j$-symbols.
626:
627: \section{Classical mechanics of the Schwinger model}
628: \label{cmSchwinger}
629:
630: The classical mechanics of the Schwinger model must be well understood
631: in order to carry out a semiclassical analysis. A general reference
632: on the classical mechanics of integrable systems from the modern point
633: of view is Cushman and Bates (1997), where harmonic oscillators in
634: particular are treated.
635:
636: \subsection{The $1j$-model}
637:
638: We start with the $1j$-model, defining two classical oscillators
639: $H_\mu = (1/2)(x_\mu^2 + p_\mu^2)$, and $H=\sum_\mu H_\mu$, as in the
640: quantum case. The classical configuration space is ${\mathbb R}^2$
641: and the phase space is ${\mathbb R}^4$. We introduce complex
642: coordinates on phase space $z_\mu = (x_\mu +i p_\mu)/\sqrt{2}$ and
643: ${\bar z}_\mu = (x_\mu - i p_\mu)/\sqrt{2}$, where we use an overbar
644: for complex conjugation. These are the Weyl symbols of the operators
645: $a_\mu$, $a^\dagger_\mu$. The complex coordinates $z_\mu$, ${\bar
646: z}_\mu$ allow us to identify the phase space ${\mathbb R}^4$ with
647: ${\mathbb C}^2$, that is, knowledge of $z_1$ and $z_2$ allows us to
648: find all four real coordinates $(x_1,x_2,p_1,p_2)$, since the ${\bar
649: z}$'s are complex conjugates of the $z$'s. As we shall see,
650: coordinates $(z_1,z_2)$, arranged as a 2-component column vector,
651: transform as a spinor under certain $SU(2)$ transformations.
652: Variables $z_\mu$ and $i{\bar z}_\mu$ are canonically conjugate ($q$'s
653: and $p$'s respectively), so that the Poisson bracket of two functions
654: $f$ and $g$ on phase space can be written,
655: \begin{eqnarray}
656: \{f,g\} &= \sum_\mu \left(
657: \frac{\partial f}{\partial x_\mu}
658: \frac{\partial g}{\partial p_\mu}-
659: \frac{\partial f}{\partial p_\mu}
660: \frac{\partial g}{\partial x_\mu}\right) \nonumber \\
661: &=
662: \sum_\mu \left(
663: \frac{\partial f}{\partial z_\mu}
664: \frac{\partial g}{\partial (i{\bar z}_\mu)}-
665: \frac{\partial f}{\partial (i{\bar z}_\mu)}
666: \frac{\partial g}{\partial z_\mu}\right).
667: \label{PB}
668: \end{eqnarray}
669:
670: The basic building blocks of the classical Schwinger model are the function
671: \begin{equation}
672: I=\frac{1}{2} \sum_\mu {\bar z}_\mu z_\mu =
673: \frac{1}{2}\sum_\mu |z_\mu|^2,
674: \label{1jIdef}
675: \end{equation}
676: and the three functions
677: \begin{equation}
678: J_i = \frac{1}{2} \sum_{\mu\nu} {\bar z}_\mu \, \sigma^i_{\mu\nu} \,
679: z_\nu,
680: \label{classJidef}
681: \end{equation}
682: for $i=1,2,3$, which define a classical angular momentum vector. We
683: also define ${\bf J}^2 = \sum_i J_i^2$. These functions satisfy the
684: identity ${\bf J}^2 = I^2$ and the Poisson bracket relations
685: $\{I,J_i\}=0$, $\{J_i, J_j\} = \sum_k \epsilon_{ijk} J_k$, $\{J_i,
686: {\bf J}^2\}=0$.
687:
688: There are two groups of interest that act on the phase space ${\mathbb
689: R}^4$ or ${\mathbb C}^2$. The first is $U(1)$, generated by $I$.
690: Hamilton's equations for $I$ are
691: \begin{equation}
692: \frac{d z_\mu}{d \psi} = \frac{\partial I}{\partial (i{\bar z}_\mu)} =
693: -\frac{i}{2} z_\mu,
694: \qquad
695: \frac{d (i{\bar z}_\mu)}{d\psi} =
696: -\frac{\partial I}{\partial z_\mu} = -\frac{1}{2} {\bar z}_\mu,
697: \label{Iflowodes}
698: \end{equation}
699: where $\psi$ is the parameter of the orbits. These have the solution
700: \begin{equation}
701: z_\mu(\psi) = \exp(-i\psi/2) \,z_\mu(0), \qquad
702: {\bar z}_\mu(\psi) = \exp(i\psi/2) \,{\bar z}_\mu(0).
703: \label{Iflow}
704: \end{equation}
705: Under the $I$-flow, the two-component spinor $(z_1,z_2)$ just gets
706: multiplied by an overall phase $\exp(-i\psi/2)$. Except for
707: the special initial condition $(z_1,z_2)=(0,0)$ (the origin of phase
708: space ${\mathbb R}^4$ or ${\mathbb C}^2$), the orbits are circles with
709: period $4\pi$ with respect to the variable $\psi$. Henceforth when
710: citing equations such as (\ref{Iflowodes}) or (\ref{Iflow}) we shall
711: omit the second half, when it is simply the complex conjugate of the
712: first half.
713:
714: We denote a value of $I$ by $j\ge0$. This is convenient notation, but
715: in this classical context $j$ is a continuous variable not to be
716: identified with the quantum number of any operator (see
717: Sec.~\ref{BSquant}). Except for the origin $j=0$, the level set $I=j$
718: (or equivalently, ${\bf J}^2 = j^2$) is the sphere $S^3$, which is
719: foliated into circles by the action (\ref{Iflow}). This foliation is
720: precisely the Hopf fibration (Frankel 1997, Nakahara 2003), yielding
721: the quotient space $S^2 = S^3/S^1$.
722:
723: The second group acting on phase space is $SU(2)$, whose action is
724: generated by the $J_i$. Explicitly, if ${\bf n}$ is a unit vector and
725: $\theta$ an angle, then the solutions of Hamilton's equations
726: \begin{equation}
727: \frac{d z_\mu}{d\theta} =
728: \frac{\partial ({\bf n}\cdot {\bf J})}{\partial (i{\bar z}_\mu)}=
729: -\frac{i}{2} \sum_\nu ({\bf n}\cdot\bsigma)_{\mu\nu} \,
730: z_\nu
731: \label{ndotJeqns}
732: \end{equation}
733: and its complex conjugate are
734: \begin{equation}
735: z_\mu(\theta) = \sum_\nu u({\bf n},\theta)_{\mu\nu}\, z_\nu(0)
736: \label{ndotJflow}
737: \end{equation}
738: and its complex conjugate. These are the obvious classical analogs of
739: equations (\ref{rotatea}); notice that the period in $\theta$ is
740: $4\pi$. It is because of this $SU(2)$ action that we say that
741: coordinates $(z_1,z_2)$ form a spinor. This classical action of
742: $SU(2)$ can be understood as a subgroup of the classical group of
743: linear canonical transformations, $Sp(4)$ (Littlejohn 1986); in
744: general, $Sp(2N)$ possesses a subgroup $Sp(2N) \cap O(2N)$ that is
745: isomorphic to $U(N)$, which contains the subgroup $SU(N)$ (in this
746: case, $N=2$). When the symplectic matrices lying in the $SU(2)$
747: subgroup are expressed in the complex basis $(z_\mu, i{\bar z}_\mu)$,
748: they block diagonalize with $u$ multiplying the $z$'s and ${\bar u}$
749: multiplying the ${\bar z}$'s.
750:
751: Equation~(\ref{classJidef}) defines a map (a projection) $\pi:{\mathbb
752: R}^4 \;(\hbox{\rm or}\; {\mathbb C}^2) \to {\mathbb R}^3$, where
753: ${\mathbb R}^3$ is ``angular momentum space,'' the space with
754: coordinates $(J_1,J_2,J_3)$. Here and below we use $\pi$ to denote
755: this map or its generalization to the $Nj$-model. The map $\pi$ maps
756: a larger space onto a smaller one, and so is not one-to-one. The
757: inverse image of a point ${\bf J}$ of angular momentum space is a set
758: of spinors that differ by an overall phase. It is easy to see that
759: the definition (\ref{classJidef}) does not depend on the overall phase
760: of the spinor. Thus, the inverse image is a circle, except in the
761: case ${\bf J}=0$ when it is a single point (the origin of
762: phase space ${\mathbb C}^2$ or ${\mathbb R}^4$).
763:
764: These circles are precisely the orbits of the $I$-flow (\ref{Iflow}).
765: Any function $f$ that is constant on these circles projects onto a
766: well defined function on angular momentum space. But such functions
767: are those that Poisson commute with $I$, $\{f,I\}$=0. This includes
768: $I$ itself as well as the three $J_i$. We can write such a function
769: as $f(z_1,z_2,{\bar z}_1,{\bar z}_2)$ or $f({\bf J})$. Now if $f$ and
770: $g$ are any two such functions, then so is their Poisson bracket
771: $\{f,g\}$, as follows from the Jacobi identity, $\{\{f,g\},I\} =
772: \{f,\{g,I\}\} + \{g,\{I,f\}\}=0$. Thus, this Poisson bracket can be
773: computed directly in angular momentum space without going back to the
774: bracket (\ref{PB}); the result is the Lie-Poisson bracket,
775: \begin{equation}
776: \{f,g\} = {\bf J} \cdot \left(
777: \frac{\partial f}{\partial {\bf J}} \times
778: \frac{\partial g}{\partial {\bf J}} \right).
779: \label{LiePoisson}
780: \end{equation}
781:
782: Interpretations of these spaces may be given in terms of the theory of
783: ``reduction'' (Marsden and Ratiu, 1999). Angular momentum space is
784: the Poisson manifold that results from Poisson reduction of the phase
785: space ${\mathbb R}^4$ under the $U(1)$ action (\ref{Iflow}) generated
786: by $I$. It is not by itself an ordinary phase space (symplectic
787: manifold), which would have an even dimensionality, but it is foliated
788: into symplectic submanifolds (the symplectic leaves). In this case
789: the symplectic leaves are the 2-spheres in angular momentum space,
790: that is, the level sets ${\bf J}^2 = j^2$, the images under $\pi$ of
791: the 3-spheres $I=j$ in ${\mathbb R}^4$ or ${\mathbb C}^2$. Canonical
792: coordinates on a given 2-sphere ${\bf J}^2 = j^2$ are $(\phi,J_z)$, a
793: $(q,p)$ pair, where $J_z=j\cos\theta$ and where $(\theta,\phi)$ are
794: the usual spherical angles in angular momentum space. Thus we have
795: \begin{equation}
796: dq\wedge dp = d\phi \wedge d(j\cos\theta) = j\,\sin\theta\,d\theta
797: \wedge d\phi = j\,d\Omega,
798: \label{S2symplform}
799: \end{equation}
800: and the symplectic form on a given sphere is $j\,d\Omega$, where
801: $d\Omega$ is the element of solid angle. This is not the geometrical
802: solid angle in a Euclidean geometry on angular momentum space, which
803: would be $j^2 \, d\Omega$. Another interpretation of angular momentum
804: space is that it is the dual of the Lie algebra of $SU(2)$, while
805: $\pi$, given by (\ref{classJidef}), is the momentum map of the
806: $SU(2)$ action (\ref{ndotJflow}).
807:
808: We now have three spaces, the ``large phase space'' ${\mathbb R}^4$ or
809: ${\mathbb C}^2$, its image under $\pi$, ``angular momentum space''
810: ${\mathbb R}^3$, and its symplectic leaves, the ``small phase
811: spaces,'' the 2-spheres ${\bf J}^2 = j^2$. Angular momentum space is
812: useful for visualizing classical angular momentum vectors, but by
813: considering inverse projections under $\pi$ the corresponding
814: geometrical objects in the large phase space can be constructed.
815: Angular momentum space has been used since the time of the old quantum
816: theory for visualizing the classical limit of quantum angular momentum
817: operators; for example, one spoke of an angular momentum vector
818: ``precessing'' around the $z$-direction. In reality, the
819: ``precession'' defines a manifold of classical states in the small
820: phase space that is a level set of a complete set of commuting
821: observables, that is, it is an invariant torus of an integrable system
822: (just a circle in the $1j$-model, where the commuting observables are
823: $I$ and $J_z$).
824:
825: \subsection{The $Nj$-model}
826:
827: We now consider the classical mechanics of the $Nj$-model, which is
828: mostly a simple generalization of the $1j$-model. We have $2N$
829: classical oscillators $H_{r\mu} = (1/2)(x_{r\mu}^2 + p_{r\mu}^2)$; the
830: configuration space is $({\mathbb R}^2)^N ={\mathbb R}^{2N}$ and the
831: ``large'' phase space is $({\mathbb R}^4)^N={\mathbb R}^{4N}$ or
832: $({\mathbb C}^2)^N={\mathbb C}^{2N}$. We define $z_{r\mu}=(x_{r\mu} +
833: i p_{r\mu})/\sqrt{2}$, ${\bar z}_{r\mu} = (x_{r\mu} -i
834: p_{r\mu})/\sqrt{2}$, so a point in phase space can be thought of as a
835: collection of $N$ 2-spinors, $(z_{r1},z_{r2})$, $r=1,\ldots,N$. We
836: make the obvious definitions (classical versions of equations
837: (\ref{Irdef})--(\ref{Jtotalsquareddef})),
838: \begin{eqnarray}
839: I_r &=\frac{1}{2}\sum_\mu |z_{r\mu}|^2,
840: \label{classIrdef} \\
841: J_{ri} &= \frac{1}{2}\sum_{\mu\nu} {\bar z}_{r\mu}
842: \, \sigma^i_{\mu\nu} \, z_{r\nu},
843: \label{classJridef}
844: \end{eqnarray}
845: as well as ${\bf J}_r^2 = \sum_i J_{ri}^2$, $J_i = \sum_r J_{ri}$ or
846: ${\bf J} = \sum_r {\bf J}_r$, and ${\bf J}^2 = \sum_i J_i^2$.
847:
848: We denote a value of the functions $I_r$ by $j_r \ge 0$; for positive
849: values $j_r>0$, $r=1,\ldots,N$, the level set $I_r = j_r$ (or ${\bf
850: J}_r^2 = j_r^2$) in the phase space $({\mathbb R}^4)^N$ is $S^3 \times
851: \ldots \times S^3= (S^3)^N$. The flow generated by $I_r$ for a
852: specific value of $r$ is just multiplication of the $r$-th spinor
853: $(z_{r1},z_{r2})$ by a phase factor $\exp(-i\psi_r/2)$, as in
854: (\ref{Iflow}); the other spinors are not affected. Thus the $N$
855: commuting flows generated by all the $I_r$'s constitute a $U(1)^N =
856: T^N$ action on the large phase space ($T^N$ is the $N$-torus).
857:
858: Equation~(\ref{classJridef}) defines the projection map $\pi:({\mathbb
859: C}^2)^N \to ({\mathbb R}^3)^N$, the latter space being ``angular
860: momentum space'' for the $Nj$-model, with one copy of ${\mathbb R}^3$
861: for each classical angular momentum vector ${\bf J}_r$. In view of
862: its importance, we write out the components of this map explicitly:
863: \begin{eqnarray}
864: J_{rx} &= \frac{1}{2}({\bar z}_{r1} z_{r2} + {\bar z}_{r2} z_{r1})
865: = {\mathop {\rm Re}}({\bar z}_{r1} z_{r2}),
866: \label{Jxeqn} \\
867: J_{ry} &= -\frac{i}{2}({\bar z}_{r1} z_{r2} - {\bar z}_{r2} z_{r1})
868: ={\mathop {\rm Im}}({\bar z}_{r1} z_{r2}),
869: \label{Jyeqn} \\
870: J_{rz} &= \frac{1}{2}(|z_{r1}|^2 - |z_{r2}|^2).
871: \label{Jzeqn}
872: \end{eqnarray}
873: Points of angular momentum space can be visualized as $N$ classical angular
874: momentum vectors, each living in its own angular momentum space, or $N$
875: such vectors all in the same 3-dimensional angular momentum space. The
876: inverse image under $\pi$ of a set of $N$ nonvanishing classical
877: angular momentum vectors is an $N$-torus in the large phase space,
878: generated by taking any point in the inverse image (a collection of
879: $N$ 2-spinors), and multiplying them by $N$ independent, overall phase
880: factors. We denote the angles on this torus by $\psi_r$,
881: $r=1,\ldots,N$, which are the evolution parameters corresponding to
882: the $I_r$, as in (\ref{Iflow}); thus their periods are $4\pi$. As
883: in the $1j$-model, angular momentum space $({\mathbb
884: R}^3)^N$ is a Poisson manifold, now with Poisson bracket
885: \begin{equation}
886: \{f,g\} = \sum_r {\bf J}_r \cdot \left(
887: \frac{\partial f}{\partial {\bf J}_r} \times
888: \frac{\partial g}{\partial {\bf J}_r} \right).
889: \label{NjLiePoisson}
890: \end{equation}
891: The symplectic leaves (the ``small phase spaces'') are the spaces $S^2
892: \times \ldots \times S^2 = (S^2)^N$ obtained by fixing the values of
893: $j_1, \ldots, j_N$, with canonical coordinates $(\phi_r,J_{rz})$ on
894: each sphere.
895:
896: In the classical $Nj$-model any partial or total sum of the angular
897: momenta ${\bf J}_r$ generates an $SU(2)$ action on the large phase
898: space, generalizing equations~(\ref{ndotJeqns}) and (\ref{ndotJflow}) in
899: the $1j$-model, in that the $SU(2)$ matrix $u$ is applied to all
900: spinors $(z_{r1},z_{r2})$ whose $r$ values lie in the sum. For
901: example, the total ${\bf J}$ rotates all spinors.
902:
903: These $SU(2)$ actions on the large phase space project to $SO(3)$ actions
904: on angular momentum space. Consider, for example, the $SU(2)$ action
905: generated by the total ${\bf J}$. Along an orbit in the large phase
906: space generated by ${\bf n}\cdot {\bf J}$, parameterized by $\theta$,
907: we can follow the value of ${\bf J}_r$, giving us ${\bf J}_r(\theta)$,
908: an orbit in the small phase space (the projection under $\pi$ of the
909: first orbit). The latter orbit is
910: \begin{equation}
911: J_{ri}(\theta) = \sum_j R({\bf n},\theta)_{ij} \, J_{rj}(0),
912: \label{Jrrotation}
913: \end{equation}
914: where $R({\bf n},\theta)$ is the $3 \times 3$ rotation associated with
915: $u({\bf n},\theta)$ according to (\ref{Ruproj}). This is the
916: classical analog of (\ref{rotateJr}). It follows
917: from (\ref{classJridef}) and the spinor adjoint equation,
918: $u^\dagger \sigma^i u = \sum_j R_{ij} \, \sigma^j$,
919: itself equivalent to (\ref{Ruproj}). Thus, under the $SU(2)$
920: action on the large phase space generated by ${\bf J}$, the individual
921: vectors ${\bf J}_r$ rotate in their individual angular momentum spaces
922: by the corresponding $3 \times 3$ rotation. For example, $J_z$
923: rotates all vectors ${\bf J}_r$ about the $z$-axis. Because of the
924: two-to-one relation between $SU(2)$ and $SO(3)$, when the orbit in the
925: large phase space goes around once ($\theta$ goes from 0 to $4\pi$),
926: the angular momentum vectors go around twice in their individual
927: angular momentum spaces.
928:
929: We may visualize this action as in Fig.~\ref{SU2action}, where $A$
930: represents a point of angular momentum space (a set of $N$ classical
931: angular momentum vectors ${\bf J}_r$ in the $Nj$-model). To obtain
932: the generic case we assume these vectors are linearly independent (in
933: particular, none of them vanishes). In the figure, $T$ is the inverse
934: image of $A$ under $\pi$, an $N$-torus. Point $a$ is any specific
935: point in the large phase space on this $N$-torus, to which the $SU(2)$
936: rotation $u({\bf n},\theta)$ is applied for $0\le\theta < 4\pi$. That
937: is, we treat $a$ as initial conditions for the Hamiltonian flow
938: generated by ${\bf n}\cdot {\bf J}$, with $\theta$ as the parameter.
939: This generates the circle $C$ in the large phase space, which amounts
940: to rotating all $N$ spinors by the same $u({\bf n},\theta)$. The
941: projection of the circle $C$ is a circle $D$ in angular momentum space
942: $({\mathbb R}^3)^N$, that is, all classical vectors ${\bf J}_r$ rotate
943: about ${\bf n}$ by angle $\theta$. However, when the circle $C$ is
944: covered once, circle $D$ is covered twice. This is because when
945: $\theta=2\pi$, the spinor rotation $u({\bf n},\theta)=-1$, so all
946: spinors in the large phase space are just multiplied by $-1$. This is
947: illustrated as point $a'$ in the figure, where all spinors are $-1$
948: times their values at $a$. Since $-1$ is just a phase factor, both
949: $a$ and $a'$ project onto the same point $A$ in angular momentum
950: space. These are the only two points on $C$ that project onto $A$;
951: for $\theta$ not a multiple of $2\pi$, the spinor rotation $u({\bf
952: n},\theta)$ is not a multiplication by a phase factor.
953:
954: Alternatively, we may apply the entire group $SU(2)$ to the original
955: point $a$ (not just rotations along a fixed axis). Then the manifold
956: $C$ is the orbit of the $SU(2)$ action which is diffeomorphic to
957: $SU(2)$. The point $a'$ is the image of $a$ under $u=-1$, a specific
958: element in $SU(2)$, and once again it projects onto the original point
959: $A$ in the small phase space. The manifold $D$ is the orbit of point
960: $A$ under the group $SO(3)$. In the $1j$-model, it is just a sphere
961: in angular momentum space (all vectors that can be reached from the
962: original one by applying all rotations), while in the $Nj$-model for
963: $N>1$ $D$ is generically diffeomorphic to $SO(3)$ (it is the set of
964: all classical configurations of $N$ angular momentum vectors that can
965: be obtained from the original one by applying rigid rotations).
966:
967: \section{The invariant $jm$-tori}
968: \label{jmtori}
969:
970: In this section we continue with the classical point of view,
971: examining the classical manifolds corresponding to the left side of
972: the matrix element (\ref{3jdef}). For this exercise and the rest of
973: the paper we adopt a $3j$-model ($N=3$). The manifolds in question
974: are the level sets of the commuting functions $I_r$, $J_{rz}$,
975: $r=1,2,3$, or, equally well, of the functions $I_{r\mu} = (1/2)
976: |z_{r\mu}|^2$ for $r=1,2,3$, $\mu=1,2$, since $I_r = I_{r1}+I_{r2}$
977: and $J_{rz} = I_{r1} - I_{r2}$. We denote the level sets by $I_r =
978: j_r$, $J_{rz} = m_r$ for contour values $j_r$, $m_r$, $r=1,2,3$, or,
979: equivalently, by
980: \begin{equation}
981: I_{r1} = \frac{1}{2}(j_r+m_r), \qquad
982: I_{r2} = \frac{1}{2}(j_r-m_r).
983: \label{ia1ia2}
984: \end{equation}
985: In spite of the notation, $j_r$ and $m_r$ take on continuous values
986: and are not necessarily the eigenvalues of any quantum operators.
987: Since the $I_{r\mu}$ are all nonnegative, we have
988: \begin{equation}
989: j_r \ge 0, \qquad -j_r \le m_r \le j_r,
990: \label{jmranges}
991: \end{equation}
992: the classical analogs of the usual inequalities in quantum mechanics.
993:
994: Since each of the six $I_{r\mu}$ is a harmonic oscillator (times
995: $1/2$), the level set of the $I_{r\mu}$'s is an invariant torus of a
996: collection of harmonic oscillators. Generically (for nonzero
997: amplitude in each oscillator, that is, when none of the quantities
998: $j_r \pm m_r$ vanishes) this is a 6-torus, upon which the coordinates
999: may be taken to be the six angles $\theta_{r\mu}$, the variables of
1000: evolution of the $I_{r\mu}$. The Hamiltonian flow generated by
1001: $I_{r\mu}$ for a specific value of $r$ and $\mu$ just multiplies
1002: $z_{r\mu}$ for the same values of $r$ and $\mu$ by
1003: $\exp(-i\theta_{r\mu}/2)$, while leaving all other $z$'s unaffected.
1004: This is not an overall spinor rotation since the other half of the
1005: spinor containing the given $z_{r\mu}$ is not affected. If viewed in
1006: the Cartesian $x_{r\mu}$-$p_{r\mu}$ phase plane, this flow is a
1007: clockwise rotation by angle $\theta_{r\mu}/2$, as illustrated in
1008: Fig.~\ref{xpplane}. The period of the angles $\theta_{r\mu}$ is
1009: $4\pi$. We agree to measure the angles $\theta_{r\mu}$ from the
1010: positive $x_{r\mu}$ axis, as in the figure, where $z_{r\mu}$ is real
1011: and positive (or zero); this is a specific convention for a set of
1012: canonical coordinates $(\theta_{r\mu}, I_{r\mu})$, $r=1,2,3$,
1013: $\mu=1,2$ on the large phase space. The volume of the 6-torus with
1014: respect to the measure $d\theta_{11} \wedge \ldots
1015: \wedge d\theta_{32}$ is $(4\pi)^6$.
1016:
1017: These tori are also the orbits of the flows generated by the
1018: observables $I_r$, $J_{rz}$. We denote the evolution variables of the
1019: $I_r$ and the $J_{rz}$ by $\psi_r$ (as above) and $\phi_r$,
1020: respectively. Each $J_{rz}$ generates an $SU(2)$ rotation about the
1021: $z$-axis on the spinor with the given value of $r$; thus each of the
1022: six angles $(\psi_r,\phi_r)$ has period $4\pi$. However, when we
1023: allow all six angles $(\psi_r,\phi_r)$ to range from 0 to $4\pi$, the
1024: torus is actually covered eight times. This can be seen from
1025: Fig.~\ref{SU2action}: a rotation by $2\pi$ in one of the $\phi$'s and
1026: one of the $\psi$'s returns us to the initial point (the path is $a$
1027: to $a'$ along $C$ in Fig.~\ref{SU2action}, then $a'$ to $a$ along
1028: $T$.) Alternatively, we may consider the canonical transformation
1029: $(\theta_{r\mu}; I_{r\mu}) \to (\psi_r, \phi_r; I_r, J_{rz})$,
1030: generated by
1031: \begin{equation}
1032: F_2(\theta_{r1}, \theta_{r2}, I_r, J_{rz}) =
1033: \frac{1}{2}[\theta_{r1}(I_r + J_{rz}) +
1034: \theta_{r2}(I_r - J_{rz})],
1035: \label{F2gf}
1036: \end{equation}
1037: which generates $I_r=I_{r1}+I_{r2}$, $J_{rz}=I_{r1}-I_{r2}$ and
1038: \begin{equation}
1039: \psi_r = \frac{1}{2}(\theta_{r1} + \theta_{r2}), \qquad
1040: \phi_r = \frac{1}{2}(\theta_{r1} - \theta_{r2}),
1041: \label{psiphiCT}
1042: \end{equation}
1043: so that the Jacobian in the angles is $(1/2)^3=1/8$. To cover the
1044: torus precisely once we may let the $\psi_r$'s range from $0$ to
1045: $4\pi$ and the $\phi_r$'s from $0$ to $2\pi$, or vice versa; thus the
1046: volume of the torus with respect to $d\psi_1 \wedge d\psi_2 \wedge
1047: d\psi_3 \wedge d\phi_1 \wedge d\phi_2 \wedge d\phi_3$ is
1048: \begin{equation}
1049: V_{jm} = (2\pi)^3 (4\pi)^3.
1050: \label{Vjm}
1051: \end{equation}
1052:
1053: The angles $\phi_r$ defined in this way on the large phase space can
1054: be projected onto the small phase space, whereupon they coincide with
1055: the usual azimuthal spherical angles in the individual angular
1056: momentum spaces. It is clear this must be so to within an additive
1057: constant, since the $J_{rz}$-flow is just an $SO(3)$ rotation about the
1058: $z$-axis in the $r$-th angular momentum space, but by our conventions
1059: even the additive constant comes out right. To see this we note first
1060: of all that $\phi_r$ for a given $r$ is constant along the $I_s$-flows
1061: for all $s$, since the variables in question are members of a
1062: canonical coordinate system on the large phase space and satisfy
1063: $\{\phi_r, I_s\}=0$. Thus, $\phi_r$, defined in the large phase
1064: space, projects onto a meaningful function in angular momentum space.
1065: Next, to compute the value of $\phi_r$ for a specific angular momentum
1066: vector ${\bf J}_r$, it suffices to take any point in the 3-torus that
1067: is the inverse image, that is, any value of the angles $\psi_s$ may be
1068: chosen. For simplicity we take $\psi_r=0$, which implies
1069: $\theta_{r1}=\phi_r$ and $\theta_{r2}=-\phi_r$. This in turn implies
1070: $z_{r1} = |z_{r1}| \exp(-i\phi_r/2)$, $z_{r2} = |z_{r2}|
1071: \exp(i\phi_r/2)$. But by equations~(\ref{Jxeqn}) and (\ref{Jyeqn}), these
1072: imply $J_{rx} = J_{r\perp} \cos\phi_r$, $J_{ry} = J_{r\perp}
1073: \sin\phi_r$, where $J_{r\perp} = |z_{r1}||z_{r2}|$.
1074:
1075: We shall henceforth call the level set $I_r=j_r$, $J_{rz}=m_r$ the
1076: ``$jm$-torus.'' This torus can be projected onto angular momentum
1077: space; we consider the generic case when $j_r \pm m_r \ne 0$ for all
1078: $r$, in which case the $jm$-torus is a 6-torus. In this case, its
1079: image in angular momentum space is a 3-torus, which, since it is a
1080: surface on which $I_r=j_r$, is also a submanifold of the small phase
1081: space. This is because the three $\psi_r$ angles just change the
1082: overall phases of the three spinors, without changing their image
1083: under $\pi$, so the three coordinates on the projected 3-torus are the
1084: angles $\phi_r$. The 3-torus in angular momentum space can be
1085: visualized as three classical vectors ${\bf J}_r$ in a single angular
1086: momentum space, with specified values of $m_r = J_{rz}$,
1087: ``precessing'' about the $z$-axis. See Fig.~\ref{3Jcones}. This is
1088: an example of how we shall visualize manifolds in the large phase
1089: space: The $jm$-torus, a six-dimensional manifold in the large phase
1090: space (itself with twelve dimensions), is visualized as three angular
1091: momentum vectors in three dimensional space, as in Fig.~\ref{3Jcones},
1092: defining a 3-torus by varying their azimuthal angles independently,
1093: and each point of this 3-torus is associated with another 3-torus, the
1094: inverse projection under $\pi$ of the given point, which consists of
1095: independently changing the overall phases of the three spinors. The
1096: 6-dimensional $jm$-torus is thus conceived of as the Cartesian product
1097: $T^3 \times T^3$ (it is actually a trivial bundle).
1098:
1099: \section{The Wigner manifold}
1100: \label{Wignermanifold}
1101:
1102: Now we turn to the right hand side of the matrix element
1103: (\ref{3jdef}), containing the state $\vert j_1j_2j_3 {\bf 0}\rangle$.
1104: This state suggests that we examine the classical manifold upon which
1105: the $I_r$ have definite values, say, $I_r=j_r$, and upon which ${\bf
1106: J}^2=0$. Again, we do not necessarily identify the $j_r$ with any
1107: quantum numbers, but it is convenient in the following to assume that
1108: none of the $j_r$'s vanishes.
1109:
1110: \subsection{Properties of the Wigner manifold}
1111:
1112: Usually the dimensionality of a manifold can be guessed by counting
1113: the constraints that define it, for example, we expect the manifold in
1114: the large (twelve-dimensional) phase space upon which $I_r=j_r$, ${\bf
1115: J}_{12}^2 = j_{12}^2$, $J_z = m$ and ${\bf J}^2 = j^2$, for given
1116: contour values, to be six-dimensional (six constraints on twelve
1117: variables). Indeed, for most values of $j$ this is correct, and the
1118: manifold in question is a 6-torus (by the Liouville-Arnold theorem,
1119: for certain ranges of the contour values). These are the invariant
1120: tori that would be involved in the semiclassical treatment of the
1121: addition of three angular momenta, producing a nonzero result (the
1122: case $j\ne0$). But this naive dimension count only works when the
1123: differentials of the functions in question are linearly independent
1124: (in particular, nonvanishing) on the manifold. This condition breaks
1125: down when ${\bf J}^2 = j^2 = 0$, since ${\bf J}^2=0$ and ${\bf J}=0$
1126: imply one another, and $d({\bf J}^2)= 2{\bf J}\cdot d{\bf J}=0$. In
1127: fact, just the four conditions $I_r=j_r>0$, ${\bf J}^2=j^2=0$ define a
1128: six-dimensional manifold in the twelve-dimensional large phase space
1129: (for certain ranges of the $j_r$). To see this we notice first that
1130: since ${\bf J}^2=0$ implies $J_i=0$, $i=1,2,3$, and $J_z=m=0$ in
1131: particular, the $J_z$ constraint is not independent; neither is the
1132: ${\bf J}_{12}^2=j_{12}^2$ constraint, since when ${\bf J}=0$, $j_{12}
1133: = j_3$. This is just as in the quantum case. In fact, the manifold
1134: $I_r=j_r$, ${\bf J}^2=0$ is characterized equivalently but better by
1135: $I_r=j_r$, ${\bf J}=0$, since the six differentials $dI_r$ and $dJ_i$
1136: are linearly independent on it. (Although the $J_i$ vanish on the
1137: manifold in question, their differentials do not.) Thus, the naive
1138: count of dimensions works with the set $I_r$, $J_i$.
1139:
1140: We shall call the manifold $I_r=j_r$, ${\bf J}=0$ in the
1141: large phase space the ``Wigner manifold,'' because
1142: it corresponds to the rotationally invariant state $\vert
1143: j_1j_2j_3 {\bf 0} \rangle$ introduced by Wigner in his definition of
1144: the $3j$-symbols. The dimensionality of this manifold (six, in the
1145: appropriate ranges of the $j_r$'s) is the same as that of the
1146: invariant tori of any integrable system of six degrees of freedom, and
1147: indeed the same as that of the nearby invariant tori in phase space
1148: corresponding to the level sets of the functions $(I_1,I_2,I_3,{\bf
1149: J}_{12}^2,J_z,{\bf J}^2)$ when $j>0$. The Wigner manifold, however,
1150: is not a torus. This is not a contradiction of the Liouville-Arnold
1151: theorem, which requires that the classical observables making up the
1152: level set should Poisson commute. In the case of the Wigner manifold,
1153: we do have $\{I_r,I_s\}=0$ and $\{I_r,J_i\}=0$, but $\{J_i, J_j\} =
1154: \sum_k \epsilon_{ijk}\, J_k$. The Hamiltonian vector fields
1155: corresponding to the functions $I_r$, $J_i$ are linearly independent
1156: on the Wigner manifold, but the three $J_i$-flows do not commute (two
1157: Hamiltonian flows commute if and only if their Poisson bracket is a
1158: constant). The Wigner manifold is, however, an orbit of the
1159: collective action of these Hamiltonian flows (any point can be reached
1160: from any other point by following the flows in some order). These
1161: facts, information about the topology of the Wigner manifold, and the
1162: required ranges on the contour values $j_r$ will be clarified
1163: momentarily.
1164:
1165: The Wigner manifold is also a Lagrangian manifold, like the invariant
1166: tori of an integrable system. This means that the integral of $p\,dx$
1167: along the manifold is locally independent of path, so an action
1168: function $S(x,A)$ can be defined. This function in turn is the
1169: solution of the simultaneous Hamilton-Jacobi equations for the
1170: observables $(I_1,I_2,I_3,J_x,J_y,J_z)$, call them $A_i$,
1171: $i=1,\ldots,6$ for short, of which the Wigner manifold is the level
1172: set.
1173:
1174: To prove that the Wigner manifold is Lagrangian, we note that the
1175: differentials $dA_i$ are linearly independent, so the vector fields
1176: $X_i$ are, too, and span the tangent space to the Wigner manifold at
1177: each point. Evaluating the symplectic form on these vector fields, we
1178: have $\omega(X_i,X_j) = -\{A_i,A_j\}$. These Poisson brackets all
1179: vanish except for the $\{J_i,J_j\}$; the latter are nonzero at most
1180: points in phase space, but on the Wigner manifold where ${\bf J}=0$,
1181: these also vanish. Thus the symplectic form restricted to the Wigner
1182: manifold vanishes, the condition that the Wigner manifold be
1183: Lagrangian.
1184:
1185: To visualize the Wigner manifold we work our way up from angular
1186: momentum space to the large phase space. First we attempt to
1187: construct three angular momentum vectors of given positive lengths
1188: $j_1,j_2,j_3$ that add up to the zero vector. This can be done if and
1189: only if the $j_r$ satisfy the triangle inequalities, whereupon the
1190: values of the $j_r$'s (the lengths of the sides) specify a triangle
1191: that is unique to within orientation. If we choose a standard or
1192: reference orientation for the triangle, then the three desired vectors
1193: are the vectors running along its sides. Let us therefore assume the
1194: triangle inequalities are satisfied, and let us choose a standard
1195: orientation for the triangle by placing the ${\bf J}_3$ along the
1196: $z$-axis, ${\bf J}_1$ in the $x$-$z$ plane with $J_{1x}>0$, and ${\bf
1197: J}_2$ in the $x$-$z$ plane with $J_{2x}<0$, as illustrated in
1198: Fig.~\ref{Jtriangle}. Given any two triangles with the same
1199: (positive) sides, there exists a unique rotation that maps one into
1200: the other; this fact and others regarding triangles are discussed in
1201: the context of the 3-body problem by Littlejohn and Reinsch (1995).
1202: In the present context this means that all classical configurations of
1203: three classical angular momentum vectors of fixed lengths that add up
1204: to the zero vector are related to any one such configuration, such as
1205: the one shown in Fig.~\ref{Jtriangle}, by a unique rotation. Thus the
1206: manifold of such classical configurations in angular momentum space
1207: ${\mathbb R}^3 \times {\mathbb R}^3 \times {\mathbb R}^3$ or in the
1208: small phase space $S^2 \times S^2 \times S^2$ is diffeomorphic to
1209: $SO(3)$.
1210:
1211: The Wigner manifold in the large phase space is now the inverse
1212: projection under $\pi$ of this $SO(3)$ manifold in angular momentum
1213: space. Since the inverse image of any given point of angular momentum
1214: space is a 3-torus in the large phase space (obtained by varying the
1215: overall phases of the three spinors), the Wigner manifold is a 3-torus
1216: bundle over $SO(3)$, and is six-dimensional. The bundle is
1217: nontrivial.
1218:
1219: The Wigner manifold may also be visualized with the help of
1220: Fig.~\ref{Wman}, an elaboration of Fig.~\ref{SU2action}. It is
1221: assumed that the three $j$'s are positive and satisfy the triangle
1222: inequality. The lower part of this figure refers to angular momentum
1223: space, while the upper part refers to the large phase space.
1224: Projection $\pi$ maps between the two spaces. Point $A$ in angular
1225: momentum space is a state of three classical angular momenta of the
1226: given lengths $j_r$ whose vector sum is zero (that is, the angular
1227: momenta define a triangle), in a definite orientation. To be
1228: specific, let us say that $A$ is the configuration shown in
1229: Fig.~\ref{Jtriangle}. By applying all $SO(3)$ rotations to $A$ we
1230: generate all orientations of the triangle, of which $B$ in the figure
1231: is one. The lower circle in the figure represents the manifold of
1232: such configurations, diffeomorphic to $SO(3)$.
1233:
1234: The inverse image of any point on this manifold under $\pi$ is a
1235: 3-torus in the large phase space. The 3-tori above points $A$ and $B$
1236: are indicated schematically as lines $T_A$ and $T_B$ in the figure.
1237: Let $a$ be some point on $T_A$. To be specific, if $A$ is the
1238: configuration of angular momentum vectors shown in
1239: Fig.~\ref{Jtriangle} and the stated conditions on the $j_r$ hold, then
1240: none of the three vectors lies on the negative $z$-axis. This means
1241: that for any point on $T_A$, $|z_{r1}|^2$ is never zero, since by
1242: equations~(\ref{classIrdef}) and (\ref{Jzeqn}), we have
1243: \begin{equation}
1244: |z_{r1}|^2 = j_r + J_{rz}, \qquad
1245: |z_{r2}|^2 = j_r - J_{rz},
1246: \label{zrsquare}
1247: \end{equation}
1248: for $r=1,2,3$. Thus by adjusting the overall phases of the three
1249: spinors $(z_{r1},z_{r2})$, we can make $z_{r1}$ real and positive for
1250: all $r=1,2,3$. Let this be the point $a$ in Fig.~\ref{Wman}.
1251:
1252: It is notationally tempting to write $m_r$ for the value of $J_{rz}$,
1253: but we shall not do this in the context of the Wigner manifold,
1254: instead reserving the symbol $m_r$ for the contour value of $J_{rz}$
1255: on the $jm$-torus.
1256:
1257: Now we apply spinor rotations to point $a$, that is, simultaneous
1258: multiplication of all three spinors $(z_{r1}, z_{r2})$ by the same
1259: element of $SU(2)$. The orbit thereby generated is a manifold
1260: diffeomorphic to $SU(2)$, as indicated in Fig.~\ref{Wman}. The
1261: projection of this manifold onto angular momentum space is the surface
1262: $SO(3)$ shown in the figure, that is, all orientations of the triangle
1263: are generated. For example, in the 3-torus $T_B$ over the angular
1264: momentum triangle with orientation $B$, there is a point $b$ that can
1265: be reached from the given point $a$ by some spinor rotation. The
1266: spinor rotation in question is one of the two that projects onto the
1267: $SO(3)$ rotation that maps $A$ into $B$, according to (\ref{Ruproj}).
1268: The orbit of reference point $a$ under the $SU(2)$ action therefore
1269: passes through the 3-tori over every possible orientation of the
1270: triangle. In fact, it passes through each 3-torus twice, since the
1271: $SU(2)$ rotation $u=-1$ is just a phase factor. This is the meaning
1272: of points $a'$ and $b'$ in the figure, which are related to points $a$
1273: and $b$ by multiplying all three spinors by $-1$.
1274:
1275: Thus any point on the Wigner manifold can be reached from the
1276: reference point $a$ by applying some $SU(2)$ rotation, and then
1277: adjusting the overall phases of the three spinors $(z_{r1},z_{r2})$.
1278: The first step is equivalent to following along the Hamiltonian flows
1279: in the large phase space generated by the three $J_i$ (this creates the
1280: $SU(2)$ rotation), while the second is equivalent to following the
1281: Hamiltonian flows generated by the three $I_r$ (this changes the
1282: overall phases of the three spinors). By letting the rotations range
1283: over all of $SU(2)$ and the three angles $\psi_r$ range from $0$ to
1284: $4\pi$, the Wigner manifold is covered twice. Thus we obtain
1285: coordinates on the Wigner manifold $(\alpha,\beta,\gamma,
1286: \psi_1,\psi_2,\psi_3)$ (the first three of which are Euler angles on
1287: $SU(2)$).
1288:
1289: Solving the simulataneous amplitude transport equations for the six
1290: observables $I_r$, $J_i$, $r,i=1,2,3$ requires us to find an invariant
1291: measure on the Wigner manifold, that is, one invariant under all of
1292: the corresponding Hamiltonian flows. Details are presented in
1293: Sec.~\ref{ampdet}; for now we just guess that this measure is the
1294: Haar measure on the group $SU(2)$ times the obvious measure on the
1295: 3-tori generated by the $I_r$, namely,
1296: \begin{equation}
1297: \sin\beta \, d\alpha \wedge d\beta \wedge d\gamma \wedge d\psi_1
1298: \wedge d\psi_2 \wedge d\psi_3,
1299: \label{Wignermeasure}
1300: \end{equation}
1301: where $(\alpha,\beta,\gamma)$ are Euler angles on $SU(2)$. The
1302: integral of this measure over the Wigner manifold is
1303: \begin{equation}
1304: V_W = \frac{1}{2} (16\pi^2) (4\pi)^3 = 2^9 \pi^5,
1305: \label{Wignervolume}
1306: \end{equation}
1307: where the $1/2$ compensates for the fact that the Wigner manifold is
1308: covered twice when the Euler angles run over $SU(2)$ and each $\psi_r$
1309: runs from 0 to $4\pi$.
1310:
1311: \subsection{Angles related to the shape of the triangle}
1312:
1313: Figure~\ref{Jtriangle} defines the angles $\eta_r$ as the angles
1314: opposite vectors ${\bf J}_r$. Under our assumptions, these angles lie
1315: in the range $0 \le \eta_r \le \pi$. By projecting all three vectors
1316: onto the directions parallel and orthogonal to each of the vectors in
1317: turn, we obtain a series of identities,
1318: \begin{eqnarray}
1319: j_1 \cos\eta_2 + j_2 \cos \eta_1 + j_3 &=0,
1320: \label{jcosident} \\
1321: j_1 \sin\eta_2 -j_2 \sin\eta_1 &=0,
1322: \label{jsinident}
1323: \end{eqnarray}
1324: and four more obtained by cycling indices $1,2,3$. These allow us to
1325: solve for the cosines of the angles $\eta_r$,
1326: \begin{equation}
1327: \cos\eta_1 = \frac{j_1^2 -j_2^2 -j_3^2}{2j_2j_3},
1328: \label{cosetadef}
1329: \end{equation}
1330: and cyclic permutations, which in view of the stated ranges on the
1331: angles allows all three angles $\eta_r$ to be uniquely determined as
1332: functions of $(j_1,j_2,j_3)$. We shall regard angles $\eta_r$ as
1333: convenient substitutions for these definite functions of the lengths
1334: of the angular momentum vectors.
1335:
1336: For the stated ranges on the $\eta_r$, the sines of the angles are
1337: nonnegative and are related to the area $\Delta$ of the triangle, as
1338: follows:
1339: \begin{eqnarray}
1340: \fl\Delta&= \frac{1}{2} |{\bf J}_1 \times {\bf J}_2| =
1341: \frac{1}{2} j_1j_2 \sin\eta_3 \nonumber \\
1342: \fl &= \frac{1}{4} \sqrt{(j_1+j_2+j_3) (-j_1+j_2+j_3) (j_1-j_2+j_3)
1343: (j_1+j_2-j_3)},
1344: \label{Deltadef}
1345: \end{eqnarray}
1346: and cyclic permutations. Some authors define $\Delta$ as the final
1347: square root (without the $1/4$).
1348:
1349: \section{Intersections of manifolds}
1350: \label{intersections}
1351:
1352: The stationary phase points of the matrix element (\ref{3jdef}) are
1353: the intersections of the $jm$-manifold and the Wigner manifold in the
1354: large phase space. Thus we must use a version of
1355: (\ref{abmatrixelement1}) for the matrix element, rather than
1356: (\ref{abmatrixelement}). In this section we study the intersections
1357: of the manifolds, continuing with a classical picture.
1358:
1359: If the $jm$-torus and the Wigner manifold in the large phase space
1360: have a common point of intersection, then the projections of these two
1361: manifolds onto angular momentum space must have a common point of
1362: intersection. The converse is also true: if the projections have a
1363: point in common, then the inverse image of this point under $\pi$, a
1364: 3-torus which is the orbit of the three $I_r$-flows, must contain two
1365: points, one of which belongs to the $jm$-torus, and the other to the
1366: Wigner manifold. But the 3-torus is the orbit of the $I_r$-flows, and
1367: these flows confine one to both the $jm$-torus and the Wigner
1368: manifold. Therefore the entire 3-torus is common to both the
1369: $jm$-torus and the Wigner manifold. Therefore to find intersections
1370: of the $jm$-torus and the Wigner manifold, we may first find the
1371: intersections of their projections under $\pi$.
1372:
1373: The $jm$-torus projects onto a set of configurations of
1374: three angular momenta with given lengths and fixed values of $m_r =
1375: J_{rz}$, with arbitrary azimuthal angles, while the Wigner manifold
1376: projects onto configurations with the same lengths in which the vector
1377: sum of the angular momenta vanishes, forming a triangle, with
1378: arbitrary orientation. Therefore to find a common point between these
1379: two sets of classical angular momenta configurations, we can either
1380: adjust the azimuthal angles of the angular momenta with given $m$
1381: values until a triangle is formed (total ${\bf J}=0$), or we can
1382: rotate a triangle from a given, reference orientation until the $m$
1383: values are the desired ones. We choose the latter procedure.
1384:
1385: Our reference orientation of the triangle is shown in
1386: Fig.~\ref{Jtriangle}, which is indicated schematically as the point
1387: $A$ in Fig.~\ref{Wman}. We must rotate this reference orientation
1388: to obtain some prescribed values of $m_r$. These values satisfy the
1389: relations (\ref{jmranges}), so in particular $|m_3|\le j_3$. Thus by
1390: rotating the triangle in the reference orientation about the $y$-axis
1391: by a unique angle $\beta$, $0\le \beta \le \pi$, defined by
1392: \begin{equation}
1393: m_3=j_3\cos\beta,
1394: \label{betadef}
1395: \end{equation}
1396: we guarantee that ${\bf J}_3$ has the right projection. The result of
1397: this rotation is shown in Fig.~\ref{beta}, for a certain negative
1398: value of $m_3$.
1399:
1400: Next we rotate the triangle about the axis ${\bf J}_3$ by an angle
1401: $\gamma$, which does not change ${\bf J}_3$ or its projection, but
1402: which rotates ${\bf J_1}$ and ${\bf J}_2$ in a cone, as illustrated in
1403: Fig.~\ref{gamma}. We wish to choose the angle $\gamma$ so that ${\bf
1404: J}_2$ has the desired projection $m_2$ onto the $z$-axis. In
1405: Fig.~\ref{gamma}, ${\bf J}_1$ is not shown, but ${\bf J}_2$ rotates
1406: about the ${\bf J}_3$ direction, its tip sweeping out circle $C_3$.
1407: The circle $C_z$ in the figure is swept out by a vector of length
1408: $j_2$ and projection $m_2$ onto the $z$-axis (this vector is not
1409: shown). Circles $C_3$ and $C_z$ intersect in two points $Q$ and $Q'$
1410: in the figure, which represent two orientations of the triangle that
1411: have the correct values of both $m_3$ and $m_2$. Now the orientation
1412: of the triangle is fixed, so there is no more freedom to rotate ${\bf
1413: J}_1$. In this final orientation the value of $J_{1z}$ is
1414: $-J_{2z}-J_{3z} = -m_2-m_3$, since ${\bf J}=0$ for the triangle.
1415: Either this value of $J_{1z}$ equals the value of $m_1$ associated
1416: with the $jm$-torus, or it does not. If it does not, then there are
1417: no intersections between the $jm$-torus and the Wigner manifold.
1418: This is just the classical expression of the condition that the matrix
1419: element (\ref{3jdef}) vanishes unless $\sum_r m_r=0$. Henceforth we
1420: assume that the $m_r$ values for the $jm$-torus do satisfy this
1421: condition.
1422:
1423: In this case we may solve for the values of $\gamma$ associated with
1424: points $Q$ and $Q'$ in Fig.~\ref{gamma}. Writing $R({\bf n},\theta)$
1425: for a 3-dimensional rotation by angle $\theta$ about axis ${\bf n}$,
1426: we have applied the rotation
1427: \begin{equation}
1428: R({\bf j}_3, \gamma) R({\bf y},\beta) = R({\bf y},\beta)
1429: R({\bf z},\gamma)
1430: \label{Rprod}
1431: \end{equation}
1432: to the reference orientation in Fig.~\ref{Jtriangle}, where ${\bf
1433: j}_3$ is the unit vector in the direction ${\bf J}_3$ shown in
1434: Fig.~\ref{beta} (after the first rotation). In the reference
1435: orientation the vectors are
1436: \begin{equation}
1437: \fl{\bf J}_1 = j_1 \left(
1438: \begin{array}{c}
1439: \sin\eta_2 \\
1440: 0 \\
1441: \cos\eta_2
1442: \end{array}\right), \qquad
1443: {\bf J}_2 = j_2 \left(
1444: \begin{array}{c}
1445: -\sin\eta_1 \\
1446: 0 \\
1447: \cos\eta_1
1448: \end{array} \right), \qquad
1449: {\bf J_3}=j_3 \left(
1450: \begin{array}{c}
1451: 0 \\ 0 \\ 1
1452: \end{array}\right).
1453: \label{Jreference}
1454: \end{equation}
1455: After applying rotation (\ref{Rprod}) these become
1456: \begin{eqnarray}
1457: {\bf J}_1 &= j_1 \left(
1458: \begin{array}{c}
1459: \cos\beta\cos\gamma\sin\eta_2 + \sin\beta\cos\eta_2 \\
1460: \sin\gamma \sin\eta_2 \\
1461: -\sin\beta\cos\gamma\sin\eta_2 + \cos\beta\cos\eta_2
1462: \end{array}\right),
1463: \nonumber \\
1464: {\bf J}_2 &= j_2 \left(
1465: \begin{array}{c}
1466: -\cos\beta\cos\gamma\sin\eta_1 + \sin\beta\cos\eta_1 \\
1467: -\sin\gamma\sin\eta_1 \\
1468: \sin\beta\cos\gamma\sin\eta_1 + \cos\beta\cos\eta_1
1469: \end{array}\right),
1470: \nonumber \\
1471: {\bf J}_3 &= j_3 \left(
1472: \begin{array}{c}
1473: \sin\beta \\
1474: 0 \\
1475: \cos\beta
1476: \end{array}\right).
1477: \label{Jrotated}
1478: \end{eqnarray}
1479: We have already solved for $\beta$ in (\ref{betadef}); we may now
1480: solve for $\gamma$ by demanding either $J_{1z}=m_1$ or $J_{2z}=m_2$.
1481: These lead to
1482: \begin{equation}
1483: \cos\gamma =
1484: \frac{j_1 \cos\beta\cos\eta_2 -m_1}{j_1\sin\beta\sin\eta_2} =
1485: \frac{m_2-j_2 \cos\beta\cos\eta_1}{j_2\sin\beta\sin\eta_1}.
1486: \label{gammadef}
1487: \end{equation}
1488: These two conditions are equivalent (under the assumption $\sum_r
1489: m_r=0$), as follows from the identities
1490: (\ref{jcosident})--(\ref{cosetadef}). If the common value of the two
1491: expressions on the right hand side of (\ref{gammadef}) lies in the
1492: range $(-1,+1)$, then there are two real angles $\gamma$ satisfying
1493: (\ref{gammadef}), corresponding to the two points $Q$ and $Q'$ in
1494: Fig.~\ref{gamma}. In this case the two manifolds have real
1495: intersections, and we are in the classically allowed region for the
1496: $3j$-symbol. We let $\gamma$ represent the root (the ``principal
1497: branch'') in the range $[0,\pi]$, and $-\gamma$ the root (the
1498: ``secondary branch'') in the range $[-\pi,0]$. Note that
1499: $\sin\gamma\ge0$ ($\le0$) on the principal (secondary) branch. If the
1500: right hand side of (\ref{gammadef}) lies outside the range
1501: $[-1,1]$, then there are two complex roots for $\gamma$. In this case
1502: the two manifolds have no real intersections, but they do have complex
1503: ones. Only one of the two complex roots is picked up by the contour
1504: of integration used in obtaining the matrix element
1505: (\ref{abmatrixelement}) or (\ref{abmatrixelement1}), resulting in an
1506: exponentially decaying expression for the matrix element. In this
1507: case we are in the classically forbidden region of the $3j$-symbol.
1508: In the following for simplicity we assume we are in the classically
1509: allowed region.
1510:
1511: The points $Q$ and $Q'$ in Fig.~\ref{gamma} represent values of ${\bf
1512: J}_2$ in a single angular momentum space ${\mathbb R}^3$. Taken with
1513: the values of ${\bf J}_1$ and ${\bf J}_3$, they specify points, call
1514: them $P$ and $P'$, in the combined angular momentum space ${\mathbb
1515: R}^3 \times {\mathbb R}^3 \times {\mathbb R}^3$ that lie on the
1516: intersection of the projections of the $jm$-torus and the Wigner
1517: manifold onto that space. Then by applying rotations about the
1518: $z$-axis to $P$ and $P'$, we generate a pair of circles in angular
1519: momentum space. Such rotations change neither the $z$-projections
1520: of the three vectors nor their vector sum (zero). This is obviously a
1521: reflection of the fact that the operator $J_z$ defining the state on
1522: the right side of (\ref{3jdef}) is a function of the operators
1523: $(J_{1z}, J_{2z}, J_{3z})$ defining the state on the left. Thus the
1524: projections of the $jm$-manifold and Wigner manifold under $\pi$
1525: intersect generically in a pair of circles.
1526:
1527: Thus the intersection of the $jm$-torus and the Wigner manifold in
1528: the large phase space is the inverse image of this pair of circles
1529: under $\pi$, generically a pair of 3-torus bundles over a circle.
1530: Since the $I_r$-flows and the $J_z$-flow commute, these bundles are
1531: trivial, in fact each is a 4-torus, on which coordinates are
1532: $(\psi_1,\psi_2,\psi_3,\phi)$, where $\phi$ is the angle of evolution
1533: along the $J_z$-flow. The volume of either one of the 4-tori with
1534: respect to the measure $d\psi_1 \wedge d\psi_2 \wedge d\psi_3 \wedge
1535: d\phi$ is
1536: \begin{equation}\
1537: V_I = \frac{1}{2} (4\pi)^4,
1538: \label{VIdef}
1539: \end{equation}
1540: the factor of $1/2$ being explained by Fig.~\ref{SU2action}.
1541:
1542: The $jm$-torus and the Wigner manifold, both six-dimensional,
1543: intersect in a 4-torus because the the lists of functions defining the
1544: two manifolds, $(I_1,I_2,I_3,J_{1z},J_{2z},J_{3z})$ and
1545: $(I_1,I_2,I_3,J_x,J_y,J_z)$, have three functions in common while
1546: $J_z$ in the second list is a function of $(J_{1z},J_{2z},J_{3z})$ in
1547: the first list. Below we will transform the functions to make both
1548: lists have explicitly four variables in common (see Eq.~(\ref{JzCT})).
1549:
1550: \section{Action integrals}
1551: \label{actionintegrals}
1552:
1553: Action integrals on the $jm$-torus and the Wigner manifold are needed
1554: for the phases in expresssions such as (\ref{abmatrixelement1}). We
1555: only need the action function at some point on the intersection
1556: between the two manifolds, which gives us a lot of choice since that
1557: intersection is a 4-torus. We continue with a classical picture in
1558: this section.
1559:
1560: \subsection{Choosing reference points}
1561:
1562: Action integrals are defined relative to some initial or reference
1563: point on each manifold. For the $jm$-torus, a convenient point is the
1564: one where $\theta_{r\mu}=0$, $r=1,2,3$, $\mu=1,2$, that is, the point
1565: where each $z_{r\mu}$ is real and nonnegative, as explained in
1566: Sec.~\ref{jmtori}. According to (\ref{ia1ia2}), the spinors at
1567: this reference point are given explicitly by
1568: \begin{equation}
1569: \left(\begin{array}{c}
1570: z_{r1} \\ z_{r2}
1571: \end{array}\right) =
1572: \left(\begin{array}{c}
1573: \sqrt{j_r + m_r} \\
1574: \sqrt{j_r - m_r}
1575: \end{array}\right).
1576: \label{jmref}
1577: \end{equation}
1578: The projection of this point onto angular momentum space is a set of
1579: vectors ${\bf J}_r$, $r=1,2,3$ of given lengths $j_r$ that lie in the
1580: $x$-$z$ plane, with $J_{rz}=m_r$ and $J_x \ge 0$, as shown by
1581: (\ref{Jxeqn})--(\ref{Jzeqn}). Such vectors are illustrated in
1582: Fig.~\ref{3Jcones}.
1583:
1584: As for the Wigner manifold, it is convenient to take the reference point
1585: to be point $a$ in Fig.~\ref{Wman}, which is discussed in
1586: Sec.~\ref{Wignermanifold}. This point projects onto the standard
1587: orientation of the triangle, point $A$ in Fig.~\ref{Wman}, where the
1588: angular momentum vectors have the values shown in
1589: (\ref{Jreference}). At the point $a$, $z_{r1}$ is real and
1590: positive for all $r$, as explained in Sec.~\ref{Wignermanifold}. For
1591: example, for $r=1$ this assumption combined with (\ref{zrsquare})
1592: implies $z_{11}=\sqrt{j_1+J_{1z}}$, which by (\ref{Jreference})
1593: becomes $z_{11}=\sqrt{j_1(1+\cos\eta_2)}=\sqrt{2j_1}\cos\eta_2/2$.
1594: Then (\ref{Jyeqn}) and $J_{1y}=0$ imply that $z_{12}$ is purely
1595: real, and (\ref{Jxeqn}) allows us to solve for $z_{12}$ in terms
1596: of $J_{1x}$, given by (\ref{Jreference}), producing finally
1597: $z_{12}=\sqrt{2j_1}\sin\eta_2$. Proceeding similarly with the other
1598: two spinors $r=2,3$, we obtain the three spinors at the reference
1599: point $a$ on the Wigner manifold,
1600: \begin{eqnarray}
1601: \fl\left(\begin{array}{c}
1602: z_{11} \\ z_{12}
1603: \end{array}\right) &=
1604: \sqrt{2j_1} \left(\begin{array}{c}
1605: \cos\eta_2/2 \\
1606: \sin\eta_2/2
1607: \end{array}\right), \qquad
1608: \left(\begin{array}{c}
1609: z_{21} \\ z_{22}
1610: \end{array}\right) =
1611: \sqrt{2j_2} \left(\begin{array}{c}
1612: \cos\eta_1/2 \\
1613: -\sin\eta_1/2
1614: \end{array}\right),
1615: \nonumber \\
1616: [3pt]
1617: \fl\left(\begin{array}{c}
1618: z_{31} \\ z_{32}
1619: \end{array}\right) &=
1620: \sqrt{2j_3} \left(\begin{array}{c}
1621: 1 \\ 0
1622: \end{array}\right).
1623: \label{Wref}
1624: \end{eqnarray}
1625:
1626: Now to obtain a point common to both the $jm$-torus and the Wigner
1627: manifold, we apply the spinor rotation
1628: \begin{equation}
1629: u({\bf y},\beta) u({\bf z},\gamma)= \left(
1630: \begin{array}{cc}
1631: e^{-i\gamma/2}\cos\beta/2 & -e^{i\gamma/2}\sin\beta/2 \\
1632: e^{-i\gamma/2}\sin\beta/2 & e^{i\gamma/2}\cos\beta/2
1633: \end{array} \right)
1634: \label{uprod}
1635: \end{equation}
1636: to the reference spinors $(z_{r1},z_{r2})$, $r=1,2,3$,
1637: in (\ref{Wref}), where Euler angles $\beta$
1638: and $\gamma$ are defined by (\ref{betadef}) and
1639: (\ref{gammadef}). We obtain either the principal branch or the
1640: secondary one by taking $\gamma\ge0$ or $\gamma\le0$, respectively.
1641: The spinor rotation (\ref{uprod}) induces the $3\times 3$
1642: rotation on the angular momentum vectors shown in (\ref{Rprod}).
1643: Thus we obtain the spinors at the common point of intersection between
1644: the $jm$-torus and the Wigner manifold,
1645: \begin{eqnarray}
1646: \fl\left(\begin{array}{c}
1647: z_{11} \\ z_{12} \end{array} \right) &=
1648: \sqrt{2j_1} \left( \begin{array}{c}
1649: e^{-i\gamma/2} \cos\beta/2 \cos\eta_2/2 -
1650: e^{i\gamma/2} \sin\beta/2 \sin\eta_2/2
1651: \label{ip1} \\
1652: [3pt]
1653: e^{-i\gamma/2} \sin\beta/2 \cos\eta_2/2 +
1654: e^{i\gamma/2 } \cos\beta/2 \sin\eta_2/2
1655: \end{array}\right), \\
1656: \fl\left(\begin{array}{c}
1657: z_{21} \\ z_{22} \end{array}\right) &=
1658: \sqrt{2j_2} \left( \begin{array}{c}
1659: e^{-i\gamma/2} \cos\beta/2 \cos\eta_1/2 +
1660: e^{i\gamma/2} \sin\beta/2 \sin\eta_1/2
1661: \label{ip2} \\
1662: [3pt]
1663: e^{-i\gamma/2} \sin\beta/2 \cos\eta_1/2 -
1664: e^{i\gamma/2} \cos\beta/2 \sin\eta_1/2
1665: \end{array}\right), \\
1666: \fl\left(\begin{array}{c}
1667: z_{31} \\ z_{32} \end{array}\right) &=
1668: e^{-i\gamma/2} \sqrt{2j_3} \left(
1669: \begin{array}{c}
1670: \cos\beta/2 \\ \sin\beta/2
1671: \end{array}\right).
1672: \label{ip3}
1673: \end{eqnarray}
1674: One can easily check using (\ref{Jxeqn})--(\ref{Jzeqn}) that
1675: these spinors project onto the angular momentum vectors in
1676: (\ref{Jrotated}).
1677:
1678: \subsection{Computing the actions}
1679:
1680: In computing action integrals we use the identity,
1681: \begin{equation}
1682: \sum_{r\mu} p_{r\mu} \, dx_{r\mu} =
1683: \frac{i}{2} \sum_{r\mu}
1684: ({\bar z}_{r\mu} \, dz_{r\mu} -
1685: z_{r\mu} \, d{\bar z}_{r\mu}) +
1686: \frac{1}{2} d \sum_{r\mu}
1687: x_{r\mu} p_{r\mu}.
1688: \label{diffformident}
1689: \end{equation}
1690: The integral of the left hand side is the usual action one would need
1691: for wave functions $\psi(x_{11}, \ldots, x_{32})$, but it can be
1692: replaced by the integral of the first differential form on the right,
1693: for the following reason. First, the integral of the exact
1694: differential on the right contributes the difference in the function
1695: $(1/2)\sum_{r\mu} x_{r\mu} p_{r\mu}$ between the initial and final
1696: points. But the final point is the common point of intersection
1697: between the $jm$-torus and the Wigner manifold, so this contribution
1698: cancels when we subtract actions as in (\ref{abmatrixelement1}). As
1699: for the initial points on the two manifolds, these have been chosen
1700: (see Eqs.~(\ref{jmref}) and (\ref{Wref})) so that all $z_{r\mu}$ are
1701: purely real, or $p_{r\mu}=0$. Thus the function in question vanishes
1702: at the initial points. As for the integral of the first term on the
1703: right of (\ref{diffformident}), it can be written
1704: \begin{equation}
1705: S=
1706: \mathop{\rm Im} \int \sum_{r\mu}
1707: z_{r\mu} \, d{\bar z}_{r\mu}.
1708: \label{zaction}
1709: \end{equation}
1710:
1711: For the action on the $jm$-torus between initial point (\ref{jmref})
1712: and final point (\ref{ip1})--(\ref{ip3}), we follow a path consisting
1713: of flows of the functions $I_{r\mu} = (1/2)|z_{r\mu}|^2$ taken one at
1714: a time by angles $\theta_{r\mu}$. Along the $I_{r\mu}$-flow we have
1715: $d{\bar z}_{r\mu}/ d\theta_{r\mu} = (i/2) {\bar z}_{r\mu}$, so the
1716: contribution to $S$ is
1717: \begin{equation}
1718: \mathop{\rm Im} \int_0^{\theta_{r\mu}}
1719: \frac{i}{2} |z_{r\mu}|^2 \, d\theta_{r\mu} =
1720: I_{r\mu} \theta_{r\mu},
1721: \label{Irmucontrib}
1722: \end{equation}
1723: since $I_{r\mu}$ is constant along its own flow and since
1724: $\theta_{r\mu}=0$ at the reference point. Thus the total action
1725: between initial and final points on the $jm$-torus is
1726: \begin{equation}
1727: S_{jm}=\sum_{r\mu} I_{r\mu} \theta_{r\mu}.
1728: \label{jmaction}
1729: \end{equation}
1730: Under the canonical transformation (\ref{F2gf}) this becomes
1731: \begin{equation}
1732: S_{jm} = \sum_r (I_r \psi_r + J_{rz} \phi_r) =
1733: \sum_r (j_r \psi_r + m_r \phi_r),
1734: \label{jmaction1}
1735: \end{equation}
1736: where in the final form we replace $I_r$ and $J_{rz}$ by their values
1737: on a given $jm$-torus.
1738:
1739: The angles $\theta_{r\mu}$ or $(\psi_r,\phi_r)$ are the coordinates
1740: of the final point specified by Eqs.~(\ref{ip1})--(\ref{ip3}). The
1741: solutions of Hamilton's equations for the $I_r$-flow can be written
1742: $z_{r\mu}(\theta_{r\mu}) = z_{r\mu}(0) \exp(-i\theta_{r\mu}/2)$ (see
1743: Fig.~\ref{xpplane}) where the initial conditions are real and
1744: nonnegative, so we have $\theta_{r\mu} = 2 \mathop{\rm arg} {\bar
1745: z}_{r\mu}$. Combining this and (\ref{psiphiCT}), we can write the
1746: action on the $jm$-torus as
1747: \begin{equation}
1748: \fl S_{jm} = 2 \sum_{r\mu} I_{r\mu}
1749: \mathop{\rm arg} {\bar z}_{r\mu} =
1750: \sum_r j_r \mathop{\rm arg}
1751: ({\bar z}_{r1} {\bar z}_{r2}) +
1752: \sum_r m_r \mathop{\rm arg}
1753: ({\bar z}_{r1} z_{r2}).
1754: \label{jmaction2}
1755: \end{equation}
1756: Using Eqs.~(\ref{ip1})--(\ref{ip3}), this can be written,
1757: \begin{eqnarray}
1758: S_{jm} &= j_3\gamma
1759: +j_1 \mathop{\rm arg}(
1760: \cos\beta \sin\eta_2 +
1761: \sin\beta \cos\gamma \cos\eta_2 +
1762: i\sin\beta \sin\gamma) \nonumber \\
1763: &+j_2 \mathop{\rm arg}(
1764: -\cos\beta \sin\eta_1 +
1765: \sin\beta \cos\gamma \cos\eta_1 +
1766: i\sin\beta \sin\gamma) \nonumber \\
1767: &+m_1 \mathop{\rm arg}(
1768: \sin\beta \cos\eta_2 +
1769: \cos\beta \cos\gamma \sin\eta_2 +
1770: i\sin\gamma \sin\eta_2) \nonumber \\
1771: &+m_2 \mathop{\rm arg}(
1772: \sin\beta \cos\eta_1 -
1773: \cos\beta \cos\gamma \sin\eta_1 -
1774: i\sin\gamma\sin\eta_1).
1775: \label{jmaction3}
1776: \end{eqnarray}
1777:
1778: Here we have used the rule $\mathop{\rm arg}(ab) = \mathop{\rm arg}a +
1779: \mathop{\rm arg} b$, which is only valid for certain choices of branch
1780: of the arg function. A more careful analysis shows that
1781: (\ref{jmaction3}) is the correct action along a certain path from the
1782: initial to final point (the principal branch) on the $jm$-torus if the
1783: range of the arg function is taken to be $[-\pi,\pi)$. (The path is
1784: defined by $\gamma \le \theta_{11}, \theta_{22} \le 2\pi$, $-\gamma
1785: \le \theta_{12}, \theta_{21} \le \gamma$, that is, one integrates from
1786: 0 to these final $\theta$ values.) In particular, this means that
1787: $\psi_1, \psi_2,\psi_3,\phi_1$ all lie in $[0,\pi]$, while $\phi_2$
1788: lies in $[-\pi,0]$. These ranges on angles $\phi_1$, $\phi_2$ are
1789: also evident from Fig.~\ref{gamma}. Similarly, for the secondary
1790: branch ($-\pi \le \gamma \le 0$, $\sin\gamma \le 0$) there exists a
1791: path such that with the same range on the arg function
1792: (\ref{jmaction3}) is still correct. With these understandings, the
1793: values of $S_{jm}$ on the two branches differ by a sign. We shall
1794: henceforth write $S_{jm}$ ($-S_{jm}$) for the principal (secondary)
1795: branch.
1796:
1797: Equation~(\ref{jmaction}) can also be written in terms of $\cos^{-1}$
1798: functions. We note that $\mathop{\rm arg}({\bar z}_{11} {\bar
1799: z}_{12}) = \cos^{-1}[ \mathop{\rm Re} ({\bar z}_{11} {\bar z}_{12})
1800: /|z_{11} z_{12}|]$ and that $|z_{11} z_{12}| = (j_1^2 -m_1^2)^{1/2}$,
1801: etc. We can also use (\ref{gammadef}) to eliminate $\cos\gamma$. For
1802: the principal branch ($\gamma\ge0$) this gives
1803: \begin{eqnarray}
1804: \fl S_{jm} &= j_1 \cos^{-1} \left(
1805: \frac{j_1 \cos\beta - m_1 \cos\eta_2}
1806: {\sin\eta_2 J_{1\perp}}\right) +
1807: j_2 \cos^{-1} \left(
1808: \frac{m_2\cos\eta_1 - j_2 \cos\beta}
1809: {\sin\eta_1 J_{2\perp}}\right) \nonumber \\
1810: [3pt]
1811: \fl &+j_3 \cos^{-1} \left(
1812: \frac{j_1\cos\beta\cos\eta_2 -m_1}
1813: {j_1\sin\beta\sin\eta_2}\right) +
1814: m_1\cos^{-1}\left(
1815: \frac{j_1\cos\eta_2 -m_1\cos\beta}
1816: {\sin\beta J_{1\perp}}\right) \nonumber \\
1817: [3pt]
1818: \fl &-m_2 \cos^{-1} \left(
1819: \frac{j_2\cos\eta_1 -m_2 \cos\beta}
1820: {\sin\beta J_{2\perp}}\right),
1821: \label{jmaction4}
1822: \end{eqnarray}
1823: where
1824: \begin{equation}
1825: J_{r\perp} = \sqrt{j_r^2 - m_r^2},
1826: \label{Jrperpdef}
1827: \end{equation}
1828: and where the range of the $\cos^{-1}$ function is $[0,\pi]$.
1829: Finally, by using Eqs.~(\ref{cosetadef}) and (\ref{Deltadef}) these can
1830: be written explicitly in terms of the parameters $j_r$, $m_r$. The
1831: result has the form of (\ref{jmaction1}), where
1832: \begin{equation}
1833: \psi_1 = \cos^{-1} \left(
1834: \frac{j_1^2(m_3 - m_2) + m_1(j_3^2 - j_2^2)}
1835: {4\Delta J_{1\perp}}\right),
1836: \label{psi1def}
1837: \end{equation}
1838: and cyclic permutations of indices, and where
1839: \begin{equation}
1840: \fl \phi_1 = \cos^{-1}\left(
1841: \frac{j_2^2 - j_3^2 -j_1^2 - 2m_1m_3}
1842: {2J_{1\perp} J_{3\perp}}\right),
1843: \; \phi_2 = -\cos^{-1}\left(
1844: \frac{j_1^2 -j_3^2 -j_2^2 -2m_2m_3}
1845: {2 J_{2\perp} J_{3\perp}}\right),
1846: \label{phi12def}
1847: \end{equation}
1848: and where $\phi_3=0$.
1849:
1850: Now we consider the action on the Wigner manifold between the initial
1851: point (\ref{Wref}) and the final point (\ref{ip1})--(\ref{ip3}). The
1852: path between these points is made up of the product of rotations
1853: (\ref{uprod}), so we consider the action integral (\ref{zaction})
1854: along a rotation by angle $\theta$ generated by ${\bf n}\cdot {\bf
1855: J}$. Hamilton's equations (see Eq.~(\ref{ndotJeqns})) are $d{\bar
1856: z}_{r\mu}/d\theta = (i/2) \sum_\nu {\bar z}_{r\nu} ({\bf n} \cdot
1857: \bsigma)_{\nu\mu}$, so by (\ref{zaction}) we have
1858: \begin{equation}
1859: \fl S=\mathop{\rm Im} \int_0^\theta
1860: \frac{i}{2} \sum_{r\mu\nu} {\bar z}_{r\nu}
1861: ({\bf n} \cdot \bsigma)_{\nu\mu} z_{r\mu} \,
1862: d\theta =
1863: \int_0^\theta ({\bf n} \cdot {\bf J})\,
1864: d\theta =
1865: ({\bf n} \cdot {\bf J})\theta=0,
1866: \label{rotaction}
1867: \end{equation}
1868: where we use (\ref{classJridef}), the fact that ${\bf n} \cdot
1869: {\bf J}$ is constant along its own flow, and the fact that ${\bf J}=0$
1870: on the Wigner manifold. The rotational action vanishes.
1871:
1872: Thus the phase of the matrix element (\ref{3jdef}) is determined
1873: entirely by the action integral along the $jm$-torus, that is, to
1874: within a sign it is given by Eqs.~(\ref{jmaction})--(\ref{phi12def}).
1875: This is the phase function determined previously by Ponzano and Regge,
1876: Miller, and others, and we see that it is essentially a simple
1877: combination of the phases of the Schwinger oscillators. We have, however,
1878: determined this phase function entirely within a classical model, that
1879: is, without imposing any quantization conditions on the manifolds.
1880:
1881: \section{Bohr-Sommerfeld quantization}
1882: \label{BSquant}
1883:
1884: We do not need Bohr-Sommerfeld approximations to the eigenvalues of
1885: the operators involved in the $3j$-symbols because those eigenvalues
1886: are known exactly. We must, however, quantize the $jm$-torus and the
1887: Wigner manifold, to obtain the wave functions whose scalar product is
1888: the $3j$-symbol. We also need the Bohr-Sommerfeld rules to make the
1889: connection between the contour values for various classical functions
1890: and the standard quantum numbers of the associated operators.
1891:
1892: To quantize a Lagrangian manifold we must first find the generators of
1893: the fundamental or first homotopy group of the manifold, that is, a
1894: set of closed contours in terms of which all closed contours can be
1895: generated by concatening curves. In the following we shall call these
1896: generators ``basis contours,'' although technically the fundamental
1897: group, even when Abelian, is a group and not a vector space. For
1898: example, in the familiar case of the invariant $n$-tori of integrable
1899: systems of $n$ degrees of freedom, the fundamental group is ${\mathbb
1900: Z}^n$, that is, an arbitrary closed contour is expressed as a ``linear
1901: combination'' of the $n$ basis contours with integer coefficients.
1902: The $n$ basis contours themselves go around the torus once in the $n$
1903: different directions.
1904:
1905: After finding the basis contours, we compute the total phase
1906: associated with each of them, the sum of an action, the integral of
1907: $p\,dq$ around the contour, and a Maslov phase, which is $-\pi/2$
1908: times the Maslov index of the loop. Both these phases are topological
1909: invariants and are additive when loops are concatenated. Then we
1910: demand that the total phase be a multiple of $2\pi$; this is the
1911: consistency condition on the semiclassical wave function that selects
1912: out certain manifolds as being ``quantized.''
1913:
1914: The Lagrangian manifolds we are interested in are level sets of a set
1915: of classical functions that are the principal symbols of a set of
1916: operators, which in our application need not commute. Quantized
1917: Lagrangian manifolds support wave functions that are approximate
1918: eigenfunctions of the set of operators. The corresponding eigenvalues
1919: are the contour values of the principal symbols, to within errors of
1920: order $\hbar^2$.
1921:
1922: \subsection{Quantizing the $jm$-tori}
1923:
1924: In the case of the $jm$-tori, whose fundamental group is ${\mathbb
1925: Z}^6$, we are dealing with the eigenfunctions of a set of independent
1926: harmonic oscillators, so the problem could not be more elementary from
1927: a semiclassical standpoint. There are, however, interesting issues
1928: that arise. Let the complete set of commuting quantum observables be
1929: $({\hat I}_r, {\hat J}_{rz})$, $r=1,2,3$, defined in (\ref{Irdef})
1930: and (\ref{Jridef}). Let us denote the Weyl symbol of an operator
1931: ${\hat A}$ by $\mathop{\rm sym}({\hat A})$. Then we have
1932: \begin{equation}
1933: \mathop{\rm sym}({\hat I}_r) =
1934: I_r -\frac{1}{2}, \qquad
1935: \mathop{\rm sym}({\hat J}_{rz}) =
1936: J_{rz},
1937: \label{IJsymbols}
1938: \end{equation}
1939: where $I_r$ and $J_{rz}$ (the classical functions) are defined by
1940: Eqs.~(\ref{classIrdef}) and (\ref{classJridef}). The operator ${\hat
1941: I}_r$ violates our assumption in Sec.~\ref{scwfis} that the commuting
1942: operators defining our integrable system should have Weyl symbols that
1943: are even power series in $\hbar$ (since the $-1/2$ is of order
1944: $\hbar$). Our assumption is valid for the harmonic oscillators ${\hat
1945: H}_r = (1/2)\sum_\mu ({\hat x}_{r\mu}^2 + {\hat p}_{r\mu}^2)$, but in
1946: defining ${\hat I}_r$ in (\ref{Irdef}) we have subtracted the zero
1947: point energy, a constant of order $\hbar$, ${\hat I}_r = (1/2)({\hat
1948: H}_r - 1)$, so that the eigenvalues of ${\hat I}_r$ would be the
1949: conventional quantum numbers $j_r$ for an angular momentum, and so
1950: that the identity (\ref{JIidentity}) would have a familiar form. In
1951: the following we shall take the principal symbol of ${\hat I}_r$ to be
1952: the whole symbol, including the $-1/2$. This achieves the same
1953: results we would have had if we had worked with ${\hat H}_r$ instead
1954: of ${\hat I}_r$ and defined the principal symbol as the leading term
1955: in $\hbar$, as in Sec.~\ref{scwfis}, since $\mathop{\rm sym}({\hat
1956: H}_r) = 2I_r$. We must be careful, however, since the principal
1957: symbol of ${\hat I}_r$ is not $I_r$. For most of the other operators
1958: we shall use, the principal symbol is obtained simply by removing the
1959: hat (for example, $J_{rz}$ above).
1960:
1961: The basis contours on the $jm$-torus are most easily expressed as the
1962: contours on which each of the angles $\theta_{r\mu}$ is allowed to go
1963: from 0 to $4\pi$ while all other $\theta_{r\mu}$'s are held fixed. We
1964: may also use any linear combination of these contours with integer
1965: coefficients and unit determinant. In terms of the angles $\psi_r$,
1966: $\phi_r$, given by (\ref{psiphiCT}), a convenient choice is to
1967: take one basis contour as the path on which one $\psi_r$ goes from $0$
1968: to $4\pi$ while all others $\psi_r$'s and all $\phi_r$'s are held
1969: fixed; this is following the $I_r$-flow for elapsed angle $\psi_r=4\pi$.
1970: A second basis contour may be taken to be the path on which $\psi_r$
1971: goes from 0 to $2\pi$, and then $\phi_r$ goes from 0 to $2\pi$; the
1972: two legs involve following the $I_r$-flow and then the $J_{rz}$-flow,
1973: each for elapsed angle $2\pi$. Doing this for $r=1,2,3$ gives us
1974: six basis contours on the $jm$-torus.
1975:
1976: The action along the first basis contour is computed as in
1977: (\ref{Irmucontrib}). Hamilton's equations for $I_r$ are $d{\bar
1978: z}_{r\mu}/d\psi_r = (i/2){\bar z}_{r\mu}$, so we obtain $S_{r1}=4\pi
1979: I_r$, where $S_{r1}$ refers to the action along the first basis
1980: contour. For the second basis contour, the first leg contributes an
1981: action $2\pi I_r$, while the second leg, which follows the flow
1982: generated by $J_{rz}$, is a rotation whose action may be computed as
1983: in (\ref{rotaction}), but with ${\bf J}$ replaced by ${\bf J}_r$ since
1984: we do not sum over $r$. Thus the final answer does not vanish (${\bf
1985: J}_r$ is nonzero on the $jm$-torus), and the contribution from the
1986: second leg is $2\pi J_{rz}$, or $S_{r2} = 2\pi(I_r + J_{rz})$.
1987: Altogether, we have
1988: \begin{equation}
1989: S_{r1} = 4\pi j_r, \qquad
1990: S_{r2} = 2\pi(j_r + m_r),
1991: \label{jmactions}
1992: \end{equation}
1993: where the 1 and 2 refer to the first and second basis contours
1994: associated with a particular value of $r$, and where we have replaced
1995: $I_r$ and $J_{rz}$ by their contour values $j_r$ and $m_r$ on the
1996: $jm$-torus.
1997:
1998: Next we need the Maslov indices along the two basis contours. Here we
1999: follow the computational method described in Littlejohn and Robbins
2000: (1987), which uses the determinant of complex matrices and which is
2001: based ultimately on Arnold (1967). Similar techniques are discussed
2002: by Mishchenko \etal\ (1990). The method works for finding Maslov
2003: indices along closed curves on orientable Lagrangian manifolds in
2004: ${\mathbb R}^{2n}$. To describe the method we adopt a general
2005: notation, in which global coordinates on phase space are $(q_1,
2006: \ldots, q_n,p_1, \ldots, p_n)$. We suppose that there exists a set of
2007: $n$ vector fields on the Lagrangian manifold, linearly independent at
2008: each point, so that they span the Lagrangian tangent
2009: plane at each point. In our applications, these are the Hamiltonian
2010: vector fields associated with a set of functions $(A_1, \ldots, A_n)$.
2011: We consider the rate of change of the quantities $q_i -i p_i$ along
2012: the $j$-th vector field, which is the Poisson bracket
2013: $\{q_i-ip_i,A_j\}$. The set of these Poisson brackets forms an $n
2014: \times n$ complex matrix $M_{ij}$ that is never singular, so $\det M$
2015: traces out a closed loop in the complex plane without passing through
2016: the origin when we go around a closed loop on the Lagrangian manifold.
2017: Then the Maslov index $\mu$ associated with this loop is given by
2018: \begin{equation}
2019: \mu = 2 \mathop{\rm wn} \det M,
2020: \label{mudef}
2021: \end{equation}
2022: where wn refers to the winding number of the loop in the complex
2023: plane, reckoned as positive in the counterclockwise direction. The
2024: winding number is invariant when $M_{ij}$ is multiplied by any nonzero
2025: complex constant (or constant matrix), so such constants can be
2026: dropped in the calculation.
2027:
2028: For the $jm$-torus, we identify the $q$'s and $p$'s with the
2029: coordinates $x_{r\mu}$ and $p_{r\mu}$, and the $A$'s with the
2030: functions $(I_1,I_2,I_3, J_{1z},J_{2z},J_{3z})$. Then we can replace
2031: $q_i-ip_i$ by ${\bar z}_{r\mu}$, dropping the $1/\sqrt{2}$. The
2032: needed matrix elements are
2033: \begin{equation}
2034: \eqalign{
2035: \{{\bar z}_{r\mu}, I_s\} &=
2036: i\frac{\partial I_s}{\partial z_{r\mu}} =
2037: \frac{i}{2} \delta_{rs} \, {\bar z}_{r\mu}, \\
2038: \{{\bar z}_{r1}, J_{sz}\} &=
2039: i\frac{\partial J_{sz}}{\partial z_{r1}} =
2040: \frac{i}{2} \delta_{rs} \, {\bar z}_{r1}, \\
2041: \{{\bar z}_{r2}, J_{sz}\} &=
2042: i\frac{\partial J_{sz}}{\partial z_{r2}} =
2043: -\frac{i}{2} \delta_{rs} \, {\bar z}_{r2}.}
2044: \label{Mijjm}
2045: \end{equation}
2046: We drop the constant $i/2$ on the right hand side, and choose the
2047: ordering $(I_1,J_{1z}, I_2,J_{2z}, I_3,J_{3z})$ for the functions.
2048: Then the $6\times 6$ matrix block diagonalizes into three $2 \times 2$
2049: blocks, and we find
2050: \begin{equation}
2051: \det M = {\bar z}_{11} {\bar z}_{12} {\bar z}_{21}
2052: {\bar z}_{22} {\bar z}_{31} {\bar z}_{32},
2053: \label{detMjm}
2054: \end{equation}
2055: to within a constant.
2056:
2057: Along the flow of one of the $I_r$'s we have ${\bar z}_{r\mu}(\psi_r)
2058: = \exp(i\psi_r/2) {\bar z}_{r\mu}(0)$, or $\det M(\psi_r) =
2059: \exp(i\psi_r) \det M(0)$. Therefore when the given $\psi_r$ goes
2060: from 0 to $4\pi$, the other $\psi_r$'s being held fixed (this is the
2061: first basis contour), $\det M$ circles the origin twice and we have
2062: $\mu=4$. Along the flow of one of the $J_{rz}$'s, however, we have
2063: ${\bar z}_{r1}(\phi_r) = \exp(i\phi_r/2) {\bar z}_{r1}(0)$ and ${\bar
2064: z}_{r2}(\phi_r) = \exp(-i\phi_r/2) {\bar z}_{r2}(0)$, or $\det
2065: M(\phi_r) = \det M(0)$. Therefore along the second basis contour the
2066: $I_r$-flow takes $\det M$ once around the origin (elapsed parameter
2067: $\psi_r=2\pi$), while the $J_{rz}$-flow does nothing. Therefore the
2068: Maslov index of the second basis contour is $\mu=2$. There are easier
2069: ways to find the Maslov indices of harmonic oscillators, but this
2070: calculation is useful practice for the case of the Wigner manifold
2071: that we take up momentarily.
2072:
2073: Now we apply the quantization conditions. For the first basis
2074: contour the total phase is $4\pi j_r -4(\pi/2)$, which we set to
2075: $2n_r\pi$ where $n_r$ is an integer. Thus the quantized tori must
2076: satisfy $j_r = (n_r+1)/2$. The allowed values of $n_r$ are determined
2077: by the fact $n_r < -1$ is impossible in view of the fact that $I_r$ is
2078: nonnegative definite, and $n_r=-1$ corresponds to a torus of less than
2079: full dimensionality (six), so the wave function (\ref{scpsi}) is not
2080: meaningful. Thus we must have $n_r=0,1,\ldots$. The $j_r$ in these
2081: formulas, and throughout all of the classical analysis from
2082: Sec.~\ref{cmSchwinger} up to this point, has referred to a contour
2083: value for the function $I_r$; the only difference now is that we are
2084: restricting the value of $j_r$ in order that the torus be quantized.
2085: This $j_r$, however, is not value of the principal symbol of the
2086: operator ${\hat I}_r$ (see Eq.~(\ref{IJsymbols})), so the
2087: Bohr-Sommerfeld or EBK quantization rule gives the semiclassical
2088: eigenvalue of ${\hat I}_r$, call it $j_r^{\rm qu}$, as
2089: \begin{equation}
2090: j_r^{\rm qu} = j_r-\frac{1}{2} = \frac{n_r}{2}.
2091: \label{jrquant}
2092: \end{equation}
2093: The semiclassical eigenvalues of ${\hat I}_r$ are nonnegative integers
2094: or half-integers, the exact answer (not surprising in view of the fact
2095: that semiclassical quantization of quadratic Hamiltonians is exact).
2096: If we use the operator identity (\ref{JIidentity}) to find the
2097: eigenvalues of operators ${\bf J}_r^2$, these are also exact.
2098:
2099: Equation~(\ref{jrquant}) shows that the classical level set
2100: corresponding to quantum number $j_r^{\rm qu}$ is $j_r = j_r^{\rm qu}
2101: + 1/2$. The extra $1/2$ in this formula has caused some discussion in
2102: the past and merits a little more now. Ponzano and Regge (1968) used
2103: intuition and numerical evidence to argue for the presence of the
2104: $1/2$. Miller, without knowing about Ponzano and Regge, also included
2105: the $1/2$, referring to the ``usual'' semiclassical replacement for
2106: angular momenta. Presumably he was referring to the similar
2107: replacement that occurs in the treatment of radial wave equations (the
2108: Langer modification, see Berry and Mount (1972), Morehead (1995)). It
2109: is not obvious to us what the Langer modification has to do with the
2110: $1/2$ that occurs in the present context, nor are we aware of any
2111: general rules about when in the asymptotics of angular momentum theory
2112: it is correct to replace a classical $j$ by $j+1/2$ (instead of
2113: $[j(j+1)]^{1/2}$ or something else). Schulten and Gordon (1975b) and
2114: Reinsch and Morehead (1999) obtain the $1/2$ as a part of their proper
2115: semiclassical analyses. Biedenharn and Louck (1981b) also speculate
2116: on the significance of the $1/2$. Roberts' (1999) derivation of the
2117: asymptotics of the $6j$-symbols does not produce the $1/2$. He argues
2118: that in the asymptotic limit there is no confusion about whether a
2119: given point lies in the classically allowed or forbidden region,
2120: whether the $1/2$ is included or not. The omission of the $1/2$ does,
2121: however, cause an error in the phase function that is of order unity,
2122: so the oscillations are not even approximately represented (this point
2123: was also made by Biedenharn and Louck, 1981b). We suspect that a
2124: modification of Roberts' method would produce the $1/2$'s. According
2125: to Girelli and Livine (2005), different choices for the semiclassical
2126: replacement for the quantum number $j$ have been made by various
2127: researchers in the field of quantum gravity. Here we have shown that
2128: the extra $1/2$ is a necessary consequence of standard semiclassical
2129: theory. We remark in addition that with the inclusion of the $1/2$,
2130: the quantized spheres in angular momentum space are those with an area
2131: of $(2j^{\rm qu}+1)2\pi$, that is, they contain a number of Planck
2132: cells exactly equal to the dimension of the irrep, obviously a form of
2133: geometric quantization. In particular, the $s$-wave $j^{\rm qu}=0$ is
2134: represented by a sphere of nonzero radius, a case for which the
2135: replacement $j_r = j_r^{\rm qu} + 1/2$ is declared by Biedenharn and
2136: Louck (1981b) to be ``clearly invalid.''
2137:
2138: For the second basis contour on the $jm$-torus, the quantization
2139: condition is $2\pi(j_r + m_r) - 2(\pi/2) = 2\pi n'_r$, where $n'_r$ is
2140: an integer. With (\ref{jrquant}), this implies $m_r = -j_r^{\rm
2141: qu} + n'_r$. Combined with the classical restriction
2142: (\ref{jmranges}), this gives the usual range on magnetic quantum
2143: numbers. Again the semiclassical quantization is exact. In the case
2144: of $J_{rz}$, the eigenvalue of the operator is equal to the classical
2145: contour value $m_r$ on the quantized torus (without any correction such as
2146: we see in (\ref{jrquant})).
2147:
2148: \subsection{Quantizing the Wigner manifold}
2149:
2150: We begin the quantization of the Wigner manifold by guessing the basis
2151: contours of the fundamental group by inspection of Fig.\ref{Wman}.
2152: Taking the base (initial) point of the loops to be point $a$ in the
2153: figure, we get three independent basis contours (call them $C_1$,
2154: $C_2$, $C_3$) by going around the 3-torus $T_A$ in the three different
2155: directions. A fourth contour (call it $C_4$) is created by following
2156: an $SU(2)$ rotation about some axis by angle $2\pi$, taking us along
2157: the path $aba'$, which puts us half way around the torus $T_A$ from
2158: the starting point, and then by applying half rotations along each of
2159: the three directions on the torus, taking us down along $T_A$ in the
2160: diagram back to the starting point $a$. These four contours are not
2161: independent, since
2162: \begin{equation}
2163: 2C_4 = C_1 + C_2 + C_3,
2164: \label{homotopyrelation}
2165: \end{equation}
2166: where addition of contours means concatenation, but they are
2167: convenient for studying the quanitzation conditions since a minimal
2168: set of three contours (not $(C_1,C_2,C_3)$, but for example
2169: $(C_1,C_2,C_4)$) is less symmetrical. The fundamental group is
2170: ${\mathbb Z}^3$.
2171:
2172: We may show the correctness of this guess by a topological argument.
2173: First, the Wigner manifold is the orbit of a $SU(2) \times T^3$ group
2174: action on the large phase space. Let $(\psi_1,\psi_2,\psi_3)$ be
2175: coordinates on $T^3$, where $0\le \psi_r \le 4\pi$. The action of
2176: $(u,(\psi_1, \psi_2, \psi_3)) \in SU(2) \times T^3$ on the large phase
2177: space is to multiply each spinor by $u$ and then by
2178: $\exp(-i\psi_r/2)$. The isotropy subgroup of this action consists of the
2179: identity $(1,(0,0,0))$ and the element $(-1,(2\pi,2\pi,2\pi))$, a
2180: normal subgroup isomorphic to ${\mathbb Z}_2$. Thus the Wigner
2181: manifold is diffeomorphic to $(SU(2) \times T^3)/{\mathbb Z}_2$,
2182: itself a group manifold, of which the original group $SU(2) \times
2183: T^3$ is a double cover. This cover is topologically simple since
2184: $SU(2)$ is simply connected (we can go to the universal cover if we
2185: wish by replacing $T^3$ by ${\mathbb R}^3$). Therefore the homotopy
2186: classes on the Wigner manifold are in one-to-one correspondence with
2187: classes of topologically inequivalent curves that go from the identity
2188: in $SU(2) \times T^3$ to one of the elements of the isotropy
2189: subgroup. Since $SU(2)$ is simply connected, such paths are
2190: characterized by choice of the end point (the element of the isotropy
2191: subgroup), and the winding numbers around the torus $T^3$. They are
2192: thus all ``linear combinations'' of the four contours $C_k, k=1,
2193: \ldots, 4$ defined above. Because of the relation
2194: (\ref{homotopyrelation}), however, the fundamental group is not ${\mathbb
2195: Z}^4$, but only ${\mathbb Z}^3$.
2196:
2197: It is easy to compute the action along these contours. Contours
2198: $C_r$, $r=1,2,3$ follow the flows of the $I_r$'s, and the action is
2199: the same as on the $jm$-torus, namely, $S_r = 4\pi j_r$.
2200: Along contour $C_4$ the spinor rotation makes no contribution to the
2201: action while the half rotation around the torus in all three angles
2202: gives the action
2203: \begin{equation}
2204: S_4 = 2\pi \sum_r j_r.
2205: \label{2Zaction}
2206: \end{equation}
2207:
2208: As for the Maslov indices, we compute the complex matrix of Poisson
2209: brackets whose rows are indexed by the functions
2210: $(I_1,I_2,I_3,J_x,J_y,J_z)$, and whose columns are indexed by $({\bar
2211: z}_{11}, {\bar z}_{12}, {\bar z}_{21}, {\bar z}_{22}, {\bar z}_{31},
2212: {\bar z}_{32})$. To within a multiplicative constant that we drop, the
2213: determinant is
2214: \begin{eqnarray}
2215: &\left| \begin{array}{cccccc}
2216: {\bar z}_{11} & {\bar z}_{12} & 0 & 0 & 0 & 0 \\
2217: 0 & 0 & {\bar z}_{21} & {\bar z}_{22} & 0 & 0 \\
2218: 0 & 0 & 0 & 0 & {\bar z}_{31} & {\bar z}_{32} \\
2219: {\bar z}_{12} & {\bar z}_{11} & {\bar z}_{22} &
2220: {\bar z}_{21} & {\bar z}_{32} & {\bar z}_{31} \\
2221: -{\bar z}_{12} & {\bar z}_{11} & -{\bar z}_{22} &
2222: {\bar z}_{21} & -{\bar z}_{32} & {\bar z}_{31} \\
2223: {\bar z}_{11} & -{\bar z}_{12} & {\bar z}_{21} &
2224: -{\bar z}_{22} & {\bar z}_{31} & -{\bar z}_{32}
2225: \end{array}\right| \nonumber \\
2226: &={\rm const.} \times
2227: \left| \begin{array}{cc}
2228: {\bar z}_{11} & {\bar z}_{12} \\
2229: {\bar z}_{21} & {\bar z}_{22}
2230: \end{array}\right|
2231: \left| \begin{array}{cc}
2232: {\bar z}_{21} & {\bar z}_{22} \\
2233: {\bar z}_{31} & {\bar z}_{32}
2234: \end{array}\right|
2235: \left| \begin{array}{cc}
2236: {\bar z}_{31} & {\bar z}_{32} \\
2237: {\bar z}_{11} & {\bar z}_{12}
2238: \end{array}\right|.
2239: \label{2ZMi}
2240: \end{eqnarray}
2241: The final product of determinants is interesting, since these are the
2242: $SU(2)$ invariants that Schwinger (Biedenharn and van Dam, 1965) used
2243: to construct the rotationally invariant state $\vert j_1 j_2 j_3 {\bf
2244: 0} \rangle$ (with ${\bar z}_{r\mu}$ replaced by $a^\dagger_{r\mu}$).
2245: Bargmann (1962) and Roberts (1999) make use of the same invariants.
2246: For our purposes we need the winding number of the loop traced out in
2247: the complex plane by the product of the three determinants as we
2248: follow the four basis contours on the Wigner manifold.
2249:
2250: Proceeding as we did on the $jm$-torus, we find
2251: that the Maslov indices along the contours $C_r, r=1,2,3$ are 4.
2252: For example, along the $I_1$-flow ${\bar z}_{11}$ and ${\bar z}_{12}$
2253: get multiplied by $\exp(i\psi_1/2)$, which causes two of the three
2254: determinants to be multiplied by the same factor, so the product gets
2255: multiplied by $\exp(i\psi_1)$, which has winding number 2 and hence
2256: Maslov index 4 when $\psi_1$ goes from 0 to $4\pi$. This is the same
2257: answer we found along the $I_r$-flows on the $jm$-torus; this was not
2258: exactly a foregone conclusion, even though the contours are the same,
2259: because the tangent planes are different. On the other hand, in both
2260: cases we find the result (\ref{jrquant}) for the eigenvalues $j_r^{\rm
2261: qu}$ of the operators ${\hat I}_r$, which of course must not depend on
2262: how we compute them. As for contour $C_4$ the first leg, a rotation
2263: by angle $2\pi$ about some axis, leaves all three $2 \times 2$
2264: determinants in (\ref{2ZMi}) invariant, so the big determinant in
2265: the complex plane does not move. As for the second leg, since each
2266: $\psi_r$ only goes from 0 to $2\pi$ we get a winding number of 1 along
2267: each $I_r$-flow, but since there are three of them the total winding
2268: number is 3 and Maslov index is 6.
2269:
2270: Combining this result with (\ref{2Zaction}), we obtain the
2271: Bohr-Sommerfeld quantization condition for contour $C_4$
2272: in the form
2273: \begin{equation}
2274: 2\pi \sum_r j_r - 6\frac{\pi}{2} = 2\pi \times {\rm integer},
2275: \label{2Zquant}
2276: \end{equation}
2277: or, with (\ref{jrquant}),
2278: \begin{equation}
2279: \sum_r j_r^{\rm qu} = {\rm integer}.
2280: \label{2zquant1}
2281: \end{equation}
2282: This is precisely the condition that the three quantum angular momenta
2283: must satisfy, in addition to the triangle inequalities, that they may
2284: add up to zero. It emerges in a semiclassical analysis because the
2285: Wigner manifold is not quantized otherwise.
2286:
2287: In conclusion, the Bohr-Sommerfeld quantization conditions applied to
2288: the $jm$-torus and the Wigner manifold give us a complete (and exact)
2289: accounting of all the quantum numbers and the restrictions on them that
2290: appear in the coupling of three angular momenta with a resultant of
2291: zero. It also allows us to identify the classical manifold (that is,
2292: its contour values) with a given set of quantum numbers.
2293:
2294: \section{The amplitude determinant}
2295: \label{ampdet}
2296:
2297: The generic semiclassical eigenfunction of a complete set of commuting
2298: observables is given by (\ref{scpsi}), with the amplitude
2299: determinant expressed in terms of Poisson brackets by
2300: (\ref{ampldet}). These formulas apply in particular to the state
2301: $\vert j_1 j_2 j_3 m_1 m_2 m_3\rangle$ on the left of the matrix
2302: element (\ref{3jdef}), which is supported by the $jm$-torus in the
2303: large phase space. The state on the right, $\vert j_1 j_2 j_3 {\bf
2304: 0}\rangle$, however, which is supported by the Wigner manifold, is an
2305: eigenfunction of observables that do not commute. Therefore we must
2306: rethink the derivation of Eqs.~(\ref{scpsi}) and (\ref{ampldet}) to
2307: see what changes in this case. In particular, we must see what
2308: happens to the Poisson bracket expression for the amplitude
2309: determinant, which is the solution of the simultaneous amplitude
2310: transport equations for the collection of observables. As it turns
2311: out, nothing changes, the wave function is still given by
2312: Eqs.~(\ref{scpsi}) and (\ref{ampldet}), with the (now noncommuting)
2313: observables used in the amplitude determinant. In addition, there is
2314: a certain understanding about how the volume $V$ in (\ref{scpsi})
2315: is computed, since the angles $\alpha$ conjugate to the $A$'s are no
2316: longer meaningful.
2317:
2318: Once this is done, we must evaluate the scalar product of the two wave
2319: functions by stationary phase. If both states were eigenstates of
2320: complete sets of commuting observables, then the answer would be
2321: (\ref{abmatrixelement1}) with amplitude determinant (\ref{PBdet}), but
2322: again we must rethink the derivation of this result since the
2323: observables for one of the wave functions do not commute. Again, the
2324: answer turns out to be given by formulas (\ref{abmatrixelement1}) and
2325: (\ref{PBdet}) of Sec.~\ref{scwfis}, with a proper understanding of the
2326: meanings of the volume factors.
2327:
2328: Having established these facts, we can then proceed to the (easy)
2329: calculation of the amplitude determinant for the $3j$-symbol in terms
2330: of Poisson brackets, and finally put the remaining pieces together to
2331: get the leading asymptotic form of the $3j$-symbols.
2332:
2333: \subsection{Amplitude determinant for noncommuting observables}
2334:
2335: We begin showing that Eqs.~(\ref{scpsi}) and (\ref{ampldet}) are valid
2336: for the state $\vert j_1 j_2 j_3 {\bf 0} \rangle$, with a proper
2337: definition of the volume factors. The classical functions defining
2338: the Wigner manifold are $(I_1, I_2, I_3, J_x, J_y, J_z)$. Let us
2339: refer to these collectively as $A_k$, $k=1, \ldots, 6$, let us write
2340: $x^i$, $i=1, \ldots, 6$ for the configuration space coordinates
2341: instead of the notation $(x_{11}, \ldots, x_{32})$ used above, and let
2342: us adopt the summation convention. The functions $A_k$ form a Lie
2343: algebra, that is, $\{A_k, A_l \} = c^m_{kl} \, A_m$, where $c^m_{kl}$
2344: are the structure constants. The Wigner manifold is a compact group
2345: manifold with this Lie algebra, on which the Haar measure is both
2346: left- and right-invariant. This density is also invariant under the
2347: flows generated by the right-invariant vector fields, which in our
2348: case are the Hamiltonian flows of the functions $A_i$. The projection
2349: of this density onto configuration space is the density that provides
2350: the solution of the simultaneous amplitude transport equations for the
2351: functions $A_i$. These are the basic geometrical facts, which we now
2352: present more explicitly in coordinate language.
2353:
2354: The amplitude transport equations for the functions $A_k$, $k=1,
2355: \ldots, 6$ are
2356: \begin{equation}
2357: \frac{\partial}{\partial x^i}
2358: \left[ \Omega(x)
2359: \frac{\partial A_k}{\partial p_i}\right]=0,
2360: \label{ATeqn}
2361: \end{equation}
2362: where $p_i$ are the momenta conjugate to $x^i$. These are six
2363: simultaneous equations that must be solved for the density $\Omega(x)$
2364: on configuration space. Notice that $\partial A_k/\partial p_i =
2365: \{x^i,A_k\} = {\dot x}^i_{(k)}$, the latter being notation we shall
2366: use for the velocity in configuration space along the Hamiltonian flow
2367: generated by $A_k$. The amplitude transport equation is a continuity
2368: equation, which is form-invariant under general coordinate
2369: transformations.
2370:
2371: Let us pick one of the branches of the inverse projection from
2372: configuration space onto the Lagrangian (Wigner) manifold. We shall
2373: suppress the branch index in the following. Let $u^i$, $i=1,\ldots,6$
2374: be an arbitrary set of local coordinates on the Wigner manifold, which
2375: we extend in a smooth but arbitrary manner into some small
2376: neighborhood of the Wigner manifold, so that partial derivatives of
2377: the $u^i$ with respect to all phase space coordinates are defined.
2378: Assuming we are not at a caustic, the transformation from $x^i$ to
2379: $u^i$ is locally one-to-one, and the Jacobian $\partial u^i/\partial
2380: x^j$ is nonsingular. Under the inverse projection or coordinate
2381: transformation $x\to u$, the flow velocity transforms according to
2382: \begin{equation}
2383: {\dot u}^i_{(k)} =
2384: \frac{\partial u^i}{\partial x^j} \,
2385: {\dot x}^j_{(k)} = \{ u^i, A_k\} = X^i_{(k)},
2386: \label{velxfm}
2387: \end{equation}
2388: which defines the quantities $X^i_{(k)}$. As a matrix, $X^i_{(k)}$ is
2389: nonsingular because the flow vectors are linearly independent on the
2390: Wigner manifold. As for the density, it transforms according to
2391: \begin{equation}
2392: \sigma(u) = \Omega(x) \left| \det
2393: \frac{\partial x^l}{\partial u^m}\right|,
2394: \label{densityxfm}
2395: \end{equation}
2396: so that the amplitude transport equations, lifted to the the Wigner
2397: manifold, become
2398: \begin{equation}
2399: \frac{\partial}{\partial u^i}
2400: [ \sigma(x) X^i_{(k)}]=0.
2401: \label{ATWm}
2402: \end{equation}
2403: Now define $\Lambda^{(k)}_j$ as the matrix inverse to $X^i_{(k)}$,
2404: \begin{equation}
2405: \Lambda^{(k)}_i X^i_{(l)} = \delta^k_l.
2406: \label{Lambdadef}
2407: \end{equation}
2408: As we will prove momentarily, the solution of Eqs.~(\ref{ATWm}) is
2409: \begin{equation}
2410: \sigma(u)= |\det \Lambda^{(k)}_j|,
2411: \label{sigmasoln}
2412: \end{equation}
2413: which, by (\ref{densityxfm}), gives us the solution of
2414: (\ref{ATeqn}),
2415: \begin{eqnarray}
2416: \Omega(x) &= \left| {\mathop{\rm det}\nolimits}_{kl}
2417: \left(
2418: \Lambda^{(k)}_i
2419: \frac{\partial u^i}{\partial x^l}\right)\right| =
2420: \left| {\mathop{\rm det}\nolimits}_{kl}
2421: \left(
2422: X^i_{(k)}
2423: \frac{\partial x^l}{\partial u^i}\right)\right|^{-1}
2424: \nonumber \\
2425: &= \left| {\mathop{\rm det}\nolimits}_{kl}\,
2426: \left({\dot u}^i_{(k)}
2427: \frac{\partial x^l}{\partial u^i}\right)\right|^{-1}
2428: =\left| {\mathop{\rm det}\nolimits}_{kl}\,
2429: {\dot x}^l_{(k)} \right|=
2430: \left| {\mathop{\rm det}\nolimits}_{kl}\,
2431: \{x^l,A_k \} \right|^{-1}.
2432: \label{Omegasoln}
2433: \end{eqnarray}
2434: In carrying out these manipulations it is important to note that
2435: $\partial u^i/\partial x^j$ is taken at constant $A_k$, not $p_k$.
2436: Thus the amplitude determinant for the wave function $\psi(x)$
2437: associated with the Wigner manifold has the same Poisson bracket form
2438: shown in (\ref{ampldet}), that is, in spite of the fact that the
2439: $A_i$ do not commute.
2440:
2441: The essential differential geometry of these manipulations is that
2442: $X_{(k)} = X^i_{(k)} \partial/\partial u^i$ are the Hamiltonian vector
2443: fields on the Wigner manifold associated with functions $A_k$,
2444: $\lambda^{(k)} = \Lambda^{(k)}_i \, du^i$ are the dual forms,
2445: $\lambda^{(k)} X_{(l)} = \delta^k_l$, and $\sigma=\lambda^1 \wedge
2446: \ldots \wedge \lambda^6 = \sigma(u) \,du^1 \wedge \ldots \wedge du^6$
2447: is the Haar measure. The condition (\ref{ATWm}) is equivalent to
2448: $L_{X_{(k)}} \sigma = 0$.
2449:
2450: To prove (\ref{sigmasoln}) in coordinates we substitute it into
2451: (\ref{ATWm}) and expand out the derivative, obtaining an
2452: expression proportional to
2453: \begin{equation}
2454: X^i_{(k),i} - X^i_{(k)} X^l_{(m),i} \Lambda^{(m)}_l
2455: \label{intqty}
2456: \end{equation}
2457: using commas for derivatives. Then we use the Lie bracket of the
2458: vector fields $X_{(k)}$,
2459: \begin{equation}
2460: [X_{(k)}, X_{(m)}]^l = X^i_{(k)} X^l_{(m),i} -
2461: X^i_{(m)} X^l_{(k),i} = -c^n_{km} X^l_{(n)},
2462: \label{Liebracket}
2463: \end{equation}
2464: where $c^n_{km}$ are the structure constants. Here we use the
2465: identity expressing the Lie bracket of Hamiltonian vector fields for
2466: two functions in terms of the Hamiltonian vector field of their
2467: Poisson bracket, $[X_H, X_K] = -X_{\{H,K\}}$ (Arnold, 1989). Thus
2468: (\ref{intqty}) becomes simply $c^m_{km}$, which vanishes since for the
2469: group in question the structure constants are completely
2470: antisymmetric.
2471:
2472: Finally to normalize the semiclassical eigenfunction supported by the
2473: Wigner manifold we use the stationary phase approximation to compute
2474: the integral
2475: \begin{equation}
2476: \int dx \, \left| \sum_{\rm br} \Omega(x)^{1/2}
2477: \exp[iS(x)-i\mu\pi/2]\right|^2,
2478: \label{Wignorm}
2479: \end{equation}
2480: where the sum is over branches and the branch index is suppressed.
2481: Cross terms do not contribute, and when the integral is lifted to the
2482: Wigner manifold it just gives the volume of that manifold with respect
2483: to the Haar measure,
2484: \begin{equation}
2485: \sum_{\rm br} \int \frac{dx}
2486: {|{\mathop{\rm det}\nolimits} \{x^i, A_j \}|}
2487: = \int du\,\sigma(u) = V_W,
2488: \label{Wignorm1}
2489: \end{equation}
2490: where $V_W$ is given by (\ref{Wignervolume}).
2491:
2492: \subsection{Matrix elements for noncommuting observables}
2493:
2494: Now we write the $3j$ matrix element (\ref{3jdef}) as $\langle b \vert
2495: a \rangle$, where $A=(I_1, I_2, I_3, J_x, J_y, J_z)$ and $B=(I_1, I_2, I_3,
2496: J_{1z}, J_{2z}, J_{3z})$. Actually this is not the most convenient
2497: form, since $J_z$ in the $A$-list is a function of the $J_{rz}$ in the
2498: $B$-list. We fix this by performing a canonical transformation
2499: $(\phi_1, \phi_2, \phi_3, J_{1z}, J_{2z}, J_{3z}) \to ({\tilde
2500: \phi}_1, {\tilde \phi}_2, {\tilde \phi}_3, {\tilde J}_{1z}, {\tilde
2501: J}_{2z}, J_z)$ on the functions in the $B$-list, generated by
2502: \begin{equation}
2503: \fl F_2(\phi_1,\phi_2,\phi_3, {\tilde J}_{1z}, {\tilde J}_{2z},
2504: J_z) = \phi_1 {\tilde J}_{1z} + \phi_2 {\tilde J}_{2z}
2505: + \phi_3(J_z - {\tilde J}_{1z} - {\tilde J}_{2z}).
2506: \label{JzCT}
2507: \end{equation}
2508: This gives ${\tilde J}_{1z} = J_{1z}$, ${\tilde J}_{2z} = J_{2z}$,
2509: $J_z = J_{1z} + J_{2z} + J_{3z}$, and ${\tilde \phi}_1 =
2510: \phi_1 - \phi_3$, ${\tilde \phi}_2 = \phi_2-\phi_3$, ${\tilde \phi}_3
2511: = \phi_3$. The linear transformation in the angles has unit
2512: determinant, so the volume of the $jm$-torus is still given by
2513: (\ref{Vjm}). Dropping the tildes, the $B$-list is now
2514: $(I_1,I_2,I_3, J_{1z}, J_{2z}, J_z)$, which has four functions in
2515: common with the $A$-list.
2516:
2517: Now the integral we must evaluate is
2518: \begin{eqnarray}
2519: \langle b \vert a \rangle &=
2520: \frac{1}{\sqrt{V_A V_B}} \int dx
2521: \frac{1}{|{\mathop {\rm det}} \{ x^i, A_j \}|^{1/2}}
2522: \frac{1}{|{\mathop {\rm det}} \{ x^i, B_j \}|^{1/2}}
2523: \nonumber \\
2524: &\qquad \times
2525: \sum_{\rm br}
2526: \exp\{i[S_A(x)-S_B(x) - \mu\pi/2]\},
2527: \label{abintegral}
2528: \end{eqnarray}
2529: where the sum is over all branches of the projections of the two
2530: manifolds, and where $\mu$ just stands for whatever Maslov index
2531: appears in a given term (different $\mu$'s are not necessarily
2532: equal). An integral like this was evaluated by Littlejohn (1990),
2533: using the angles conjugate to the $A$'s and $B$'s, but those do not
2534: all exist in the present circumstances and we must evaluate the
2535: integral in a different way.
2536:
2537: Let us write $A=(C,D)$ and $B=(C,E)$, where $C=(I_1, I_2, I_3, J_z)$
2538: are the four observables in common in the $A$- and $B$-lists, and
2539: where $D=(J_x, J_y)$ and $E=(J_{1z}, J_{2z})$ are the two pairs of
2540: observables that are distinct. The stationary phase set of the
2541: integral (\ref{abintegral}) consists of points $x$ where
2542: $\partial(S_A-S_B)/\partial x^i=0$, that is, it is the projection onto
2543: configuration space of the intersection of the $A$-manifold and the
2544: $B$-manifold. That intersection, which we denote by $I$, was
2545: studied in Sec.~\ref{intersections} (it is a 4-torus). It is the
2546: simultaneous level set of all of the $A$'s and $B$'s, and at the same
2547: time the orbit of the commuting Hamiltonian flows generated by the
2548: $C$'s. Its projection onto configuration space is a 4-dimensional
2549: region.
2550:
2551: We introduce a local coordinate transformation in configuration space
2552: $x \to (y,z)$ where the four $y$'s are coordinates along the stationary
2553: phase set and the two $z$'s are transverse to it. We let the stationary
2554: phase set itself be specified by $z=0$. We let $(u,v)$ be the momenta
2555: conjugate to $(y,z)$. Then the two amplitude determinants in
2556: (\ref{abintegral}) may be combined with the Jacobian of
2557: the coordinate transformation to result in the square root of the
2558: product of two determinants, one of which is
2559: \begin{equation}
2560: \det \frac{\partial(y,z)}{\partial x}
2561: \det \{ x, A \} =
2562: \det \left(\begin{array}{cc}
2563: \{ y, C \} & \{ y, D \} \\
2564: \{ z, C \} & \{ z, D \}
2565: \end{array}
2566: \right),
2567: \label{yzCDPB}
2568: \end{equation}
2569: and the other of which is the same but with the substitutions $A\to
2570: B$, $D \to E$. But since the $C$'s generate flows along $I$, we have
2571: $\{z^i,C_j\}=0$, and the lower left block of the two matrices
2572: vanishes. Thus, the product of the two determinants becomes
2573: \begin{equation}
2574: [\det \{ y, C \}]^2 \det \{ z, D\} \det \{ z, E\},
2575: \label{detprod}
2576: \end{equation}
2577: the square root of which appears in the denominator of the integrand.
2578: Evaluating the final two Poisson brackets in the $(y,z;u,v)$ canonical
2579: coordinates, we have
2580: \begin{equation}
2581: \{z^i, D_j \} = \frac{\partial D_j}{\partial v_i},
2582: \qquad
2583: \{z^i, E_j \} = \frac{\partial E_j}{\partial v_i}.
2584: \label{zDzEPBs}
2585: \end{equation}
2586:
2587: We perform the $z$-integration by stationary phase, expanding $S_A$
2588: and $S_B$, regarded as functions of $(y,z)$, to second order in $z$
2589: for a fixed value of $y$, and simply evaluating the amplitude at $z=0$
2590: (that is, on $I$). To within a phase, the $z$-integration gives
2591: \begin{equation}
2592: 2\pi
2593: \left| \det \left(
2594: \frac{\partial^2 S_A}{\partial z\partial z}
2595: -\frac{\partial^2 S_B}{\partial z\partial z}\right)
2596: \right|^{-1/2}.
2597: \label{zintegral}
2598: \end{equation}
2599: The determinant in this result must be multiplied by the determinants
2600: of the matrices (\ref{zDzEPBs}) to get the overall determinant in the
2601: denominator after the $z$-integration. The product of these three
2602: determinants is the determinant of the matrix
2603: \begin{equation}
2604: \frac{\partial D}{\partial v}\left[
2605: \left(\frac{\partial v}{\partial z}\right)_{yCD} -
2606: \left(\frac{\partial v}{\partial z}\right)_{yCE}\right]
2607: \left(\frac{\partial E}{\partial v}\right)^T,
2608: \label{3matrixprod}
2609: \end{equation}
2610: where it is understood that a partial derivative stands for a matrix
2611: whose row index is given by the numerator and column index by the
2612: denominator, unless the matrix transpose or inverse is indicated, in
2613: which case the rule is reversed. Also, if a partial derivative is
2614: shown without subscripts, then it is assumed that it is computed in
2615: the canonical coordinates $(y,z;u,v)$, and otherwise the variables to
2616: be held fixed are explicitly indicated. In the two middle matrices in
2617: (\ref{3matrixprod}), the variables held fixed amount to
2618: differentiating $v$ with respect to $z$ along the $A$- and
2619: $B$-manifolds, respectively, since $\partial S_A/\partial z = v(x,A)$
2620: and $\partial S_B/\partial z=v(x,B)$. Notice that these two matrices
2621: are symmetric.
2622:
2623: Now we express the two matrices in the middle of
2624: (\ref{3matrixprod}) purely in terms of partial derivatives
2625: computed in the canonical $(y,z;u,v)$ coordinates. We do this by
2626: writing out the Jacobian matrix $\partial(y,z;C,D)/ \partial(y,z;u,v)$
2627: and the inverse Jacobian $\partial(y,z;u,v)/\partial(y,z;C,D)$,
2628: multiplying the two together to obtain a series of identities
2629: connecting the forward and inverse Jacobian blocks, and then solving
2630: for the inverse Jacobian blocks in terms of the forward ones. We note
2631: that the block $\partial C/\partial v$ of the forward Jacobian
2632: vanishes, since it is $\{z,C\}$. Thus we find
2633: \begin{equation}
2634: \fl\eqalign{
2635: \left(\frac{\partial v}{\partial z}\right)_{yCD} &=
2636: \left(\frac{\partial D}{\partial v}\right)^{-1} \left[
2637: \frac{\partial D}{\partial u}
2638: \left(\frac{\partial C}{\partial u}\right)^{-1}
2639: \frac{\partial C}{\partial z}-
2640: \frac{\partial D}{\partial z}\right], \cr
2641: \left(\frac{\partial v}{\partial z}\right)_{yCE} &=
2642: \left[\left(\frac{\partial C}{\partial z}\right)^T
2643: \left(\frac{\partial C}{\partial u}\right)^{-1T}
2644: \left(\frac{\partial E}{\partial u}\right)^T -
2645: \left(\frac{\partial E}{\partial z}\right)^T\right]
2646: \left(\frac{\partial E}{\partial v}\right)^{-1T}. \cr}
2647: \label{vzidents}
2648: \end{equation}
2649: Upon substituting these into (\ref{3matrixprod}), that matrix
2650: becomes
2651: \begin{eqnarray}
2652: &\frac{\partial D}{\partial v}
2653: \left(\frac{\partial E}{\partial z}\right)^T -
2654: \frac{\partial D}{\partial z}
2655: \left(\frac{\partial E}{\partial v}\right)^T
2656: +\frac{\partial D}{\partial u}
2657: \left(\frac{\partial C}{\partial u}\right)^{-1}
2658: \frac{\partial C}{\partial z}
2659: \left(\frac{\partial E}{\partial v}\right)^{T}
2660: \nonumber \\
2661: &-\frac{\partial D}{\partial v}
2662: \left(\frac{\partial C}{\partial z}\right)^T
2663: \left(\frac{\partial C}{\partial u}\right)^{-1T}
2664: \left(\frac{\partial E}{\partial u}\right)^T,
2665: \label{firsthalfDEPB}
2666: \end{eqnarray}
2667: where the first two terms are the beginning of the Poisson bracket
2668: $\{E,D\}$. As for the last two terms, we write out the vanishing
2669: Poisson brackets $\{C,E\}$ and $\{C,D\}$ in the $(y,z;u,v)$
2670: coordinates, making use of $\partial C/\partial v=0$, to obtain
2671: \begin{equation}
2672: \eqalign{
2673: \frac{\partial C}{\partial z}
2674: \left(\frac{\partial E}{\partial v}\right)^T
2675: &= \frac{\partial C}{\partial u}
2676: \left(\frac{\partial E}{\partial y}\right)^T -
2677: \frac{\partial C}{\partial y}
2678: \left(\frac{\partial E}{\partial u}\right)^T, \cr
2679: \frac{\partial D}{\partial v}
2680: \left(\frac{\partial C}{\partial z}\right)^T
2681: &= \frac{\partial D}{\partial y}
2682: \left(\frac{\partial C}{\partial u}\right)^T -
2683: \frac{\partial D}{\partial u}
2684: \left(\frac{\partial C}{\partial y}\right)^T. \cr}
2685: \label{CEPBident}
2686: \end{equation}
2687: Actually the matrix of Poisson brackets $\{C,D\}$ does not vanish
2688: everywhere in phase space, just on the $A$- (or Wigner) manifold, and
2689: in particular on the intersection $I$ which is where we are evaluating
2690: them. Now substituting Eqs.~(\ref{CEPBident}) into the last two terms
2691: of (\ref{firsthalfDEPB}), those terms become
2692: \begin{eqnarray}
2693: &\frac{\partial D}{\partial u}
2694: \left(\frac{\partial E}{\partial y}\right)^T -
2695: \frac{\partial D}{\partial y}
2696: \left(\frac{\partial E}{\partial u}\right)^T \\
2697: &+\frac{\partial D}{\partial u} \left[
2698: \left(\frac{\partial C}{\partial y}\right)^T
2699: \left(\frac{\partial C}{\partial u}\right)^{-1T} -
2700: \left(\frac{\partial C}{\partial u}\right)^{-1}
2701: \frac{\partial C}{\partial y}\right]
2702: \left(\frac{\partial E}{\partial u}\right)^T,
2703: \label{secondhalfDEPB}
2704: \end{eqnarray}
2705: in which the first two terms give us the remainder of the Poisson
2706: bracket $\{E,D\}$. As for the last major term, the factor in the
2707: square brackets vanishes, as we see by writing out the vanishing
2708: Poisson bracket $\{C,C\}$ in coordinates $(y,z;u,v)$ and using
2709: $\partial C/\partial v=0$.
2710:
2711: As a result the integral (\ref{abintegral}) becomes
2712: \begin{equation}
2713: \langle b \vert a \rangle =
2714: \frac{2\pi}{\sqrt{V_A V_B}}
2715: \sum_{\rm br}
2716: \int \frac{dy}{|\det\{y,C\}|}
2717: \frac{e^{i(S_I - \mu\pi/2)}}
2718: {|\det\{E,D\}|^{1/2}},
2719: \label{abyintegral}
2720: \end{equation}
2721: where the branch sum runs over all branches of the projection of $I$
2722: onto configuration space as well as the two disconnected components of $I$
2723: (the two 4-tori discussed in Sec.~\ref{intersections}), and where
2724: $S_I$ is the phase on a given connected component of $I$ (this is the
2725: phase $\pm S_{jm}$ computed in Sec.~\ref{actionintegrals}). We have also
2726: dropped an overall phase, and we are not attempting to compute the
2727: Maslov indices in detail. The amplitude determinant has been reduced
2728: to a $2\times 2$ matrix of Poisson brackets of the observables in the
2729: $A$- and $B$-lists that differ, exactly as in (\ref{PBdet}).
2730: Calculating this matrix explicitly, we find
2731: \begin{eqnarray}
2732: |\det\{E,D\}| &= \left|\begin{array}{cc}
2733: \{ J_{1z}, J_x \} & \{ J_{1z}, J_y \} \\
2734: \{ J_{2z}, J_x \} & \{ J_{2z}, J_y \}
2735: \end{array}\right| =
2736: \left|\begin{array}{cc}
2737: J_{y1} & -J_{x1} \\
2738: J_{y2} & -J_{x2}
2739: \end{array}\right|
2740: \nonumber \\
2741: [3pt]
2742: &= |J_{x1} J_{y2} - J_{x2} J_{y1}|
2743: = | {\bf z} \cdot ({\bf J}_1 \times {\bf J}_2)|
2744: = 2 \Delta_z,
2745: \label{3jampldet}
2746: \end{eqnarray}
2747: where $\Delta_z$ is the projection of the area of the triangle
2748: $\Delta$ onto the $x$-$y$ plane (see Eq.~(\ref{Deltadef}) and Fig.~2
2749: of Ponzano and Regge (1968)). This quantity is invariant under
2750: rotations about the $z$-axis, that is, it Poisson commutes with $J_z$.
2751: It also Poisson commutes with the other three variables in the
2752: $C$-list, $(I_1,I_2,I_3)$, and so is constant on the intersection
2753: $I$ and can be taken out of the $y$-integral. The same applies to the
2754: phase factor, since $S_I$ is also constant on the $I$-manifold. Then
2755: the $y$-integral can be done, since $|\det\{y,C\}|$ is just the
2756: Jacobian connecting $y$ with the angle variables conjugate to
2757: $C=(I_1,I_2,I_3,J_z)$, denoted above by $(\psi_1,\psi_2,\psi_3,\phi)$.
2758: Thus the $y$-integral just gives the volume $V_I$ of the intersection $I$
2759: with respect to these angles, see (\ref{VIdef}). In fact, had the
2760: variables $C$ not been commuting, but if they had formed a Lie
2761: algebra, then $V_I$ would be the volume of $I$ with respect to the
2762: Haar measure of the corresponding group. This circumstance arises,
2763: for example, in a similar treatment of the $6j$-symbol.
2764:
2765: As a result of these rather lengthy manipulations of amplitude
2766: determinants, we obtain the final, simple result,
2767: \begin{equation}
2768: \langle b \vert a \rangle =
2769: \frac{2\pi}{\sqrt{V_A V_B}} \sum_{\rm br} V_I
2770: \frac{e^{i(S_I-\mu\pi/2)}}{|\det\{E,D\}|^{1/2}},
2771: \label{integraldone}
2772: \end{equation}
2773: where the branches now run over just the two disconnected pieces of
2774: the intersection $I$. This is a version of
2775: (\ref{abmatrixelement1}), with the right understanding of the
2776: volume measures, generalized to the case at hand in which the
2777: observables do not commute. The actual calculation of the final amplitude
2778: determinant takes just one line, Eq.~(\ref{3jampldet}).
2779:
2780: In fact, for our application the volume $V_I$ and the remaining
2781: amplitude determinant are the same for both branches and can be taken
2782: out of the sum. The relative Maslov index between the two branches is
2783: 1; we will not belabor this point since the answer is already known.
2784: We simply note that by splitting the Maslov phase $i\pi/2$ between
2785: the two branches and subsitituting $V_A=V_W$, $V_B=V_{jm}$, we obtain
2786: to within an overall phase the result of Ponzano and Regge,
2787: \begin{equation}
2788: \left(
2789: \begin{array}{ccc}
2790: j_1 & j_2 & j_3 \\
2791: m_1 & m_2 & m_3
2792: \end{array}
2793: \right) = ({\rm phase}) \times
2794: \frac{\cos(S_{jm}+\pi/4)}{\sqrt{2\pi\Delta_z}}.
2795: \label{PRresult}
2796: \end{equation}
2797:
2798: \section{Conclusions}
2799: \label{conclusions}
2800:
2801: In many ways the $3j$-symbol is not as interesting as the $6j$-symbol,
2802: of which it is a limiting case. We intended our work on the
2803: $3j$-symbol as a warm-up exercise, expecting a routine application of
2804: semiclassical methods for integrable systems. The nongeneric
2805: Lagrangian (Wigner) manifold was a surprise. Similar nongeneric
2806: Lagrangian manifolds occur also in the semiclassical analysis of the
2807: $6j$- and $9j$-symbols.
2808:
2809: If all one wants is a derivation of an asymptotic formula, then there
2810: are many ways to proceed. For example, one can simply take the
2811: expression for the symbol due to Wigner ($3j$) or Racah ($6j$) as a
2812: sum over a single index, and apply standard asymptotic methods
2813: (Stirling's approximation, Poisson sum rule, etc). But if one wants a
2814: derivation that reveals the geometrical meaning of the classical
2815: objects that emerge (the triangle, the tetrahedron, etc), then an
2816: approach such as ours may be preferable.
2817:
2818: Our approach is more geometrical than earlier ones, and in that
2819: respect is closer in spirit to the work of Roberts (1999), Freidel and
2820: Louapre (2003) and later authors. It is likely that at some deeper
2821: level all these methods are the same, although superficially we see
2822: only a little similarity between our work and these others.
2823:
2824: One may also desire a method that makes the symmetries of the symbol
2825: manifest. Our analysis does not do this for the $3j$-symbol, but
2826: those symmetries are not manifest in Wigner's definition of the
2827: $3j$-symbol that we employ as our starting point, either. To bring
2828: the symmetries out it seems necessary to employ some construction
2829: related to Schwinger's generating functions, which involve lifting the
2830: definitions into higher dimensional spaces.
2831:
2832: Our method of calculating amplitude determinants in terms of Poisson
2833: brackets may have computational advantages in other applications, as
2834: well. The method can be remarkably easy to use. For example, the
2835: $6j$-symbol can be defined as a matrix element,
2836: \begin{equation}
2837: \left\{ \begin{array}{ccc}
2838: j_1 & j_2 & j_{12} \cr
2839: j_4 & j_3 & j_{23}
2840: \end{array}\right\}
2841: = {\rm const.} \times \langle j_1j_2j_3j_4 j_{23} {\bf 0} \vert
2842: j_1 j_2 j_3 j_4 j_{12} {\bf 0} \rangle,
2843: \label{6jdef}
2844: \end{equation}
2845: which is the unitary matrix in $(j_{12},j_{23})$ defining a change of
2846: basis in the subspace in which four angular momenta of given lengths
2847: add up to zero (${\bf 0}$ means ${\bf J}={\bf 0}$). In this case
2848: there are eight observables on each side of the matrix element, of
2849: which seven are common and one is different. Thus the amplitude of
2850: the $6j$-symbol is the inverse square root of the single Poisson
2851: bracket,
2852: \begin{equation}
2853: \{{\bf J}_{23}^2, {\bf J}_{12}^2\} = 4
2854: {\bf J}_1 \cdot ({\bf J}_2 \times {\bf J}_3),
2855: \label{6jPB}
2856: \end{equation}
2857: as follows immediately from (\ref{NjLiePoisson}). One sees
2858: immediately that it is proportional to the volume of the tetrahedron.
2859: A similarly easy calculation is possible for the $9j$-symbol. It is
2860: harder, however, to express these amplitudes in terms of the quantum
2861: numbers (the magnitudes $j_r$), that is, to translate these magnitudes
2862: into vectors ${\bf J}_r$ that lie on the stationary phase set. We
2863: shall report on these and other extensions of our work in future
2864: publications.
2865:
2866:
2867:
2868: \section*{References}
2869: \begin{harvard}
2870:
2871: \item[] Anderson R W and Aquilanti V 2006 {\it J. Chem. Phys.} {\bf
2872: 124} 214104
2873:
2874: \item[] Aquilanti V, Cavalli S and De~Fazio D 1995 {\it J. Phys. Chem.}
2875: {\bf 99} 15694
2876:
2877: \item[] Aquilanti V, Cavalli S and Coletti C 2001 {\it Chem. Phys. Lett.}
2878: {\bf 344} 587
2879:
2880: \item[] Aquilanti V and Coletti C 2001 {\it Chem. Phys. Lett.} {\bf
2881: 344} 601
2882:
2883: \item[] Arnold V I 1967 {\it Functional Anal.\ Appl.} {\bf 1} 1
2884:
2885: \item[] \dash 1989 {\it Mathematical Methods of Classical
2886: Mechanics} (New York: Springer-Verlag)
2887:
2888: \item[] Barrett J W and Steele C M 2002 {\it Class. Quant. Grav.} {\bf
2889: 20} 1341
2890:
2891: \item[] Baez J C, Christensen J D and Egan G 2002 {\it
2892: Class. Quant. Grav.} {\bf 19} 6489
2893:
2894: \item[] Balazs N L and Jennings B K 1984 {\it Phys. Reports} {\bf 104}
2895: 347
2896:
2897: \item[] Bargmann V 1962 {\it Rev. Mod. Phys.} {\bf 34} 829
2898:
2899: \item[] Berry M V 1977 {\it Phil. Trans. Roy. Soc.} {\bf 287} 237
2900:
2901: \item[] Berry M V and Mount K E 1972 {\it Rep.\ Prog.\ Phys.} {\bf 35}
2902: 315
2903:
2904: \item[] Berry M V and Tabor M 1976 {\it Proc.\ Roy.\ Soc.\ Lond.\ A}
2905: {\bf 349} 101
2906:
2907: \item[] Biedenharn L C and Louck J D 1981a {\it Angular Momentum in
2908: Quantum Physics} (Reading, Massachusetts: Addison-Wesley)
2909:
2910: \item[] \dash 1981b {\it The Racah-Wigner Algebra in Quantum Theory}
2911: (Reading, Massachusetts: Addison-Wesley)
2912:
2913: \item[] Biedenharn L C and van Dam H 1965 {\it Quantum Theory of
2914: Angular Momentum} (New York: Academic Press)
2915:
2916: \item[] Brack Matthias and Bhaduri Rajat K 1997 {\it Semiclassical
2917: Physics} (Reading, Massachusetts: Addison-Wesley)
2918:
2919: \item[] Brillouin M L 1926 {\it J.\ Phys.} {\bf 7} 353
2920:
2921: \item[] Cargo Matthew, Gracia-Saz Alfonso, Littlejohn R G , Reinsch
2922: M W and de M. Rios P 2005, {\it J. Phys.\ A} {\bf 38} 1977
2923:
2924: \item[] Cargo Matthew, Gracia-Saz Alfonso and Littlejohn R G 2005,
2925: preprint math-ph/0507032.
2926:
2927: \item[] Cushman R H and Bates L 1997 {\it Global Aspects of Classical
2928: Integrable Systems} (Basel: Birkh\"auser Verlag)
2929:
2930: \item[] De~Fazio D, Cavalli S and Aquilanti V 2003 {\it
2931: Int. J. Quant. Chem.} {\bf 93} 91
2932:
2933: \item[] de Gosson, M 1997 {\it Maslov Classes, Metaplectic
2934: Representation and Lagrangian Quantization (Mathematical Research
2935: Vol. 5)} (Berlin: Akademischer Verlag)
2936:
2937: \item[] Einstein A 1917 {\it Verh.\ dt.\ Phys.\ Ges.} {\bf 19} 82
2938:
2939: \item[] Estrada Ricardo, Gracia-Bond\'\i a J M and V\'arilly J C 1989
2940: {\it J. Math. Phys.} {\bf 30} 2789
2941:
2942: \item[] Frankel Theodore 1997 {\it The Geometry of Physics} (Cambridge)
2943:
2944: \item[] Freidel Laurent and Louapre David 2003 {\it Class. Quantum Grav.}
2945: {\bf 20} 1267
2946:
2947: \item[] Geronimo J S, Bruno O and Van Assche W 2004 {\it Oper. Theory
2948: Adv. Appl.} {\bf 154} 101
2949:
2950: \item[] Girelli F and Livine E R 2005 {\it Class. Quantum Grav.}
2951: {\bf 22} 3295
2952:
2953: \item[] Gracia-Bond\'\i a J M and V\'arilly J C 1995 {\it
2954: J. Math. Phys.} {\bf 36} 2691
2955:
2956: \item[] Groenewold H J 1946 {\it Physica} {\bf 12} 405
2957:
2958: \item[] Gutzwiller Martin C 1990 {\it Chaos in Classical and Quantum
2959: Mechanics} (New York: Springer-Verlag)
2960:
2961: \item[] Hillery M, O'Connell R F, Scully M O and Wigner E P 1984 {\it
2962: Phys. Reports} {\bf 106} 123
2963:
2964: \item[] Keller J B 1958 {\it Ann.\ Phys.} {\bf 4} 180
2965:
2966: \item[] Littlejohn R G 1986 {\it Phys. Reports} {\bf 138} 193
2967:
2968: \item[] \dash 1990 {\it J. Math.\ Phys.} {\bf 31} 2952
2969:
2970: \item[] Littlejohn R G and Reinsch M 1995 {\it Phys.\ Rev.\ A} {\bf52}
2971: 2035
2972:
2973: \item[] Littlejohn R G and Robbins J M 1987 {\it Phys.\ Rev.\ A}
2974: {\bf 36} 2953
2975:
2976: \item[] Marsden J E and Ratiu T 1999 {\it Introduction to Mechanics and
2977: Symmetry} (New York: Springer-Verlag)
2978:
2979: \item[] Marzuoli A and Rasetti M 2005 {\it Ann. Phys.} {\bf 318} 345
2980:
2981: \item[] Maslov V P and Fedoriuk M V 1981 {\it Semi-Classical
2982: Approximations in Quantum Mechanics} (Dordrecht: D. Reidel)
2983:
2984: \item[] McDonald S 1988 {\it Phys. Reports} {\bf 158} 377
2985:
2986: \item[] Miller W H 1974 {\it Adv. Chem. Phys.} {\bf 25} 69
2987:
2988: \item[] Mishchenko A S, Shatalov V E and Sternin B Yu 1990 {\it Lagrangian
2989: manifolds and the Maslov operator} (Berlin: Springer Verlag)
2990:
2991: \item[] Morehead J J 1995 {\it J.\ Math.\ Phys.} {\bf 36} 5431
2992:
2993: \item[] Moyal J E 1949 {\it Proc. Camb. Phil. Soc.} {\bf 45} 99
2994:
2995: \item[] Nakahara M 2003 {\it Geometry, Topology and Physics} (Bristol:
2996: IOP Publishing)
2997:
2998: \item[] Neville Donald 1971 {\it J. Math. Phys.} {\bf 12} 2438
2999:
3000: \item[] Ozorio de Almeida Alfredo M 1998 {\it Phys.\ Reports} {\bf
3001: 295} 265
3002:
3003: \item[] Percival I C 1973 {\it J.\ Phys.\ B} {\bf 6} L229
3004:
3005: \item[] Ponzano G and Regge T 1968 in {\it Spectroscopy and Group
3006: Theoretical Methods in Physics} ed F Bloch \etal\ (Amsterdam:
3007: North-Holland) p~1
3008:
3009: \item[] Reinsch M W and Morehead J J 1999 {\it J.\ Math.\ Phys.} {\bf
3010: 40} 4782
3011:
3012: \item[] Roberts J 1999 {\it Geometry and Topology} {\bf 3} 21
3013:
3014: \item[] Sakurai J J 1994 {\it Modern Quantum Mechanics} (New York:
3015: Addison-Wesley)
3016:
3017: \item[] Schulman L S 1981 {\it Techniques and Applications of Path
3018: Integration} (New York: John Wiley \& Sons)
3019:
3020: \item[] Schulten K and Gordon R G 1975a {\it J. Math. Phys.} {\bf 16} 1961
3021:
3022: \item[] \dash 1975b {\it J. Math. Phys.} {\bf 16} 1971
3023:
3024: \item[] Smorodinskii Ya A and Shelepin L A 1972 {\it Sov. Phys. Usp.}
3025: {\bf 15} 1
3026:
3027: \item[] Taylor Y U and Woodward C T 2005 {\it Selecta Math. (N.S.)}
3028: {\bf 11} 539
3029:
3030: \item[] Voros A 1977 {\it Ann. Inst. Henri Poincar\'e} {\bf 4} 343
3031:
3032: \item[] Weyl H 1927 {\it Z. Phys.} {\bf 46} 1
3033:
3034: \item[] Wigner E P 1932 {\it Phys. Rev.} {\bf 40} 749
3035:
3036: \item[] \dash 1959 {\it Group Theory} (Academic Press, New York)
3037:
3038: \end{harvard}
3039:
3040: \Figures
3041:
3042: \begin{figure}
3043: \caption{\label{SU2action} The action of an $SU(2)$ rotation about a
3044: fixed axis on a point of the large phase space, and its projection
3045: onto angular momentum space.}
3046: \end{figure}
3047:
3048: \begin{figure}
3049: \caption{\label{xpplane} Harmonic oscillator motion in the
3050: $x_{r\mu}$-$p_{r\mu}$ plane, generated by $I_{r\mu}$. Definition of
3051: angle $\theta_{r\mu}$ is shown.}
3052: \end{figure}
3053:
3054: \begin{figure}
3055: \caption{\label{3Jcones} The quantum state $\vert j_1j_2j_3 m_1m_2m_3
3056: \rangle$ corresponds classically to a 3-torus in the small phase
3057: space $S^2 \times S^2 \times S^2$, which can be visualized as three
3058: ${\bf J}$ vectors lying on three cones with fixed values of $J_z$. The
3059: three azimuthal angles are independent.}
3060: \end{figure}
3061:
3062: \begin{figure}
3063: \caption{\label{Jtriangle} If $j_1$, $j_2$ and $j_3$ satisfy the
3064: triangle inequalities, then they define a triangle that is unique
3065: apart from its orientation. A standard orientation places the
3066: triangle in the $x$-$z$ plane with sides ${\bf J}_1$, ${\bf J}_2$,
3067: ${\bf J}_3$ oriented as shown. The angle opposite ${\bf J}_r$ is
3068: $\eta_r$.}
3069: \end{figure}
3070:
3071: \begin{figure}
3072: \caption{\label{Wman} An schematic illustration showing how the Wigner
3073: manifold in the large phase space is the inverse image under $\pi$
3074: of a set of triangles formed from three angular momentum vectors
3075: with vanishing sum, all related by rigid rotations.}
3076: \end{figure}
3077:
3078: \begin{figure}
3079: \caption{\label{beta} By rotating the reference orientation of the
3080: triangle about the $y$-axis, we can give ${\bf J}_3$ the desired
3081: projection $m_3$ onto the $z$-axis.}
3082: \end{figure}
3083:
3084: \begin{figure}
3085: \caption{\label{gamma} Once vector ${\bf J_3}$ has the desired
3086: projection $m_3$, we rotate the triangle by angle $\gamma$ about the
3087: axis ${\bf J}_3$ to make ${\bf J}_2$ have its desired projection
3088: $m_2$. This cannot always be done for real angles $\gamma$, but
3089: when it can be done there are generically two angles that work,
3090: illustated by points $Q$ and $Q'$ in the figure.}
3091: \end{figure}
3092:
3093:
3094: \end{document}
3095: