1: % ----------------------------------------------------------------
2: % AMS-LaTeX Paper ************************************************
3: % **** -----------------------------------------------------------
4: \documentclass[aps,eqsecnum,twocolumn,amsmath]{revtex4}
5: \usepackage{graphics}
6: \usepackage[letterpaper,dvips,width=7.5in,includemp=false]{geometry}
7:
8:
9: \newcommand{\dslash}{D\!\!\!\!\slash}
10: %\setlength{\textheight}{8.85in}
11: \setlength{\topmargin}{.01in}
12: % ----------------------------------------------------------------
13: \vfuzz2pt % Don't report over-full v-boxes if over-edge is small
14: \hfuzz2pt % Don't report over-full h-boxes if over-edge is small
15: % MATH -----------------------------------------------------------
16: \newcommand{\norm}[1]{\left\Vert#1\right\Vert}
17: \newcommand{\abs}[1]{\left\vert#1\right\vert}
18: \newcommand{\set}[1]{\left\{#1\right\}}
19: \newcommand{\Real}{\mathbb R}
20: \newcommand{\eps}{\varepsilon}
21: \newcommand{\To}{\longrightarrow}
22: \newcommand{\BX}{\mathbf{B}(X)}
23: \newcommand{\A}{\mathcal{A}}
24: % ----------------------------------------------------------------
25: \begin{document}
26:
27: \title[title]{The quantization of exotic states in $SU(3)$ soliton models: A solvable quantum mechanical analog}%
28: \author{Aleksey Cherman, Thomas D. Cohen, Abhinav Nellore}%
29:
30: \affiliation{Department of Physics, University of Maryland,
31: College Park, MD 20742-4111}
32: %\email{placeholder}%
33:
34:
35: %\date{}%
36: %\dedicatory{}%
37: %\commby{}%
38: % ----------------------------------------------------------------
39: \begin{abstract}
40: The distinction between the rigid rotor and Callan-Klebanov
41: approaches to the quantization of $SU(3)$ solitons is considered in
42: the context of exotic baryons. A numerically tractable quantum
43: mechanical analog system is introduced to test the reliability of
44: the two quantization schemes. We find that in the equivalent of the
45: large $N_c$ limit of QCD, the Callan-Klebanov approach agrees with a
46: numerical solution of the quantum mechanical analog. Rigid rotor
47: quantization generally does not. The implications for exotic
48: baryons are briefly discussed.
49: \end{abstract}
50: \maketitle
51: % ----------------------------------------------------------------
52: \section{Introduction}
53:
54: Recent experimental reports\cite{exp} of the observation of an
55: exotic $\theta^{+}$ baryon have rekindled interest in the
56: quantization of exotic states in $SU(3)$ soliton
57: models\cite{Coh03}. The soliton models are unique among the
58: theoretical tools used to analyze these reported states because
59: they predate the experiment and predict the mass of the state
60: very closely to the claimed experimental values
61: \cite{Pres,DiaPetPoly}.
62:
63: A commonly employed technique for quantizing solitons treats the
64: system as a rigid rotor in which the soliton collectively rotates,
65: with its internal degrees of freedom fixed from the classical
66: solution\cite{Guad,SU3Quant}. This technique is clearly correct
67: for $SU(2)$ chiral solitons when treated in the large $N_c$ limit
68: \cite{ANW} (where $N_c$ is the number of colors in QCD). However,
69: the validity of applying the rigid rotor approach to the
70: quantization of $SU(3)$ solitons has been questioned in
71: situations where the Wess-Zumino term can play an active role in
72: the dynamics of the system --- as it does for exotic baryon states
73: \cite{Coh03,Coh04,IKOR,Pob}. This issue remains controversial
74: \cite{Coh03,IKOR, Pob,DP,Coh04}. In this paper we study a
75: numerically tractable quantum mechanical system that has many
76: critical features in common with $SU(3)$ solitons to obtain
77: insight about the underlying issues.
78:
79:
80: It should be noted that there are {\it two} major approaches in
81: the literature for quantizing $SU(3)$ solitons. One is the rigid
82: rotor approach mentioned above, in which the collective
83: rotational and vibrational modes of the soliton are assumed to be
84: decoupled, and only the rotational modes are
85: quantized\cite{SU3Quant}. This approach has a number of benefits:
86: it is relatively simple to implement and is a straightforward
87: generalization of the Adkins, Nappi and Witten procedure which is
88: known to be justified at large $N_c$ for the case of $SU(2)$
89: solitons\cite{ANW}. Moreover, it is known to be justified at
90: large $N_c$ for non-exotic collective states in $SU(3)$ models.
91: The vast majority of the published studies of exotic baryons in
92: the context of soliton models have used this approach. The other
93: approach to quantizing $SU(3)$ solitons is the Callan-Klebanov
94: approach \cite{CK,IKOR}. This scheme is most easily understood in
95: the case of broken $SU(3)$, in which case excitations carrying
96: strangeness are unambiguously vibrational states, and should be
97: quantized as harmonic vibrations. It has been argued that this
98: method remains valid as one approaches the $SU(3)$ limit
99: \cite{IKOR}. In fact, for the non-exotic states one of the
100: vibrational modes becomes softer and goes to zero frequency as the
101: $SU(3)$ limit is approached, and thus reproduces the results of
102: rigid rotor quantization. Therefore it seems that this method
103: should be valid for both broken and unbroken $SU(3)$. However,
104: for exotic states the Callan-Klebanov approach does {\it not}
105: reproduce the rigid rotor result; indeed when applied to the
106: original Skyrme model it gives {\it no} exotic resonant states at
107: all\cite{IKOR}. The Callan-Klebanov approach has the obvious
108: disadvantage of being more difficult to implement; this may
109: explain the fact that it has not been widely used in studies of
110: exotic baryon states. Indeed, to the best of our knowledge only
111: two such calculations have been reported\cite{IKOR,PR}.
112:
113: The fact that the Callan-Klebanov method gives different results
114: from rigid-rotor quantization for exotic states implies that at
115: least one of these methods is wrong. Assuming that one is correct,
116: it is critical to know which one. It has been argued elsewhere
117: on a number of grounds that the rigid-rotor approach is the
118: culprit\cite{Coh03,IKOR,Coh04,Pob}, and the Callan-Klebanov method
119: is correct, at least at large $N_c$. While there have been
120: attempts in the literature to rebut at least some of these
121: arguments \cite{DP}, it has been argued that these rebuttals are
122: fundamentally flawed\cite{Coh04}. We will not attempt to recap
123: these arguments but instead refer the reader to the original
124: literature.
125:
126: Our purpose here is simply to consider a tractable quantum
127: mechanical system that has states analogous to the exotic states of
128: the $SU(3)$ soliton, which is analytically intractable. This system
129: can be solved numerically (to essentially any desired degree of
130: accuracy) and via both approximate methods --- the Callan-Klebanov
131: approach and the rigid-rotor approach in the analog of the large
132: $N_c$ limit. One can then explicitly see which approach works. As
133: we will show below, the Callan-Klebanov method reproduces the
134: numerical solutions for this model up to expected errors of order
135: $1/N_c$, while the rigid-rotor approach generally fails.
136:
137: The model considered here was introduced in
138: refs.~\cite{Coh04,Pob}. In those works it was observed that the
139: Callan-Klebanov approach and rigid-rotor approach gave different
140: answers for the excitation spectrum of the model, and it was
141: argued on semiclassical grounds that the assumptions underlying
142: the rigid-rotor quantization were not self consistent. However,
143: these works did not show explicitly that the Callan-Klebanov
144: approach actually produces the correct spectrum, and that the
145: rigid rotor approach does not.
146:
147: This paper is organized as follows. In the next section, the model
148: will be introduced. The two subsequent sections will implement the
149: rigid rotor quantization and the Callan-Klebanov quantization for
150: this model. (Some details of the Callan-Klebanov treatment are
151: relegated to an appendix). The next section contains a brief
152: discussion of the numerical solution of the model and a comparison
153: of the numerical solution with the two approximation methods.
154: Finally, we discuss the implications of our results for soliton
155: treatments of exotic baryons.
156:
157:
158:
159:
160: \section{A Tractable model}
161:
162:
163: We wish to study a numerically soluble model that incorporates the
164: relevant features of the excited states of $SU(3)$ solitons. Given
165: the nature of the critique of the rigid rotor treatment
166: \cite{Coh03,IKOR,Coh04,Pob}, in order to mimic the soliton
167: problem, we need a system in which there are both collective
168: rotational and vibrational degrees with the same quantum numbers.
169: There should also be a force on the system that is topological in
170: nature and velocity dependent to mimic the effect of the
171: Wess-Zumino term (which is topological and first order in time
172: and thus acts on velocities).
173:
174: To do this, let us consider the problem of a composite charged
175: particle moving nonrelativistically on the surface of a sphere of
176: radius $R$, which has a magnetic monopole of strength $g$ at its
177: center. It was observed long ago by Witten that the motion of a
178: charged particle on the surface of a sphere in the field of a
179: magnetic monopole is topological in essentially the same way as
180: the motion of chiral fields in the presence of the Wess-Zumino
181: term\cite{Wit1}. Indeed, the original rigid rotor quantization of
182: $SU(3)$ solitons by Guadagnini was explicitly done in analogy to
183: the monopole problem\cite{Guad}. It is important that we consider
184: a composite system, {\it i.e.}, one with internal degrees of
185: freedom. The dynamics of the composite system are analogous to
186: the internal dynamics of the soliton and the key issues are
187: associated with the possible interplay of internal and collective
188: degrees of freedom.
189:
190:
191: To be concrete, we consider our composite particle as being made of
192: two point-like constituent particles\cite{Coh04,Pob}. One
193: constituent is a charged particle (with charge $q$). The other
194: constituent particle is electrically neutral. We take both
195: constituent particles to have the same mass ($M$). The particles
196: interact via a nonsingular potential that binds the particles
197: together. To ensure that the system is rotationally invariant, the
198: potential can depend only on the separation between the particles.
199: Since the particles are strongly bound, and thus spend most of their
200: time near the minimum of the potential, we can take the interaction
201: to be due to an approximately harmonic potential of spring constant
202: $k$. As noted above, the magnetic field due to the monopole serves
203: to provide the desired velocity-dependent topological force. The
204: analog of the classical static soliton is simply the classical
205: configuration which minimizes the energy---namely, the two particles
206: on top of each other at the minimum of the potential. This
207: configuration will be referred to as the ``soliton''.
208:
209: The semiclassical treatment of the $SU(3)$ soliton is only
210: justified in the large $N_c$ limit of QCD. Thus, it is important
211: that the various parameters in the toy model are chosen to scale
212: with $N_c$ in a manner that emulates the soliton case:
213: \begin{equation}\label{toyscale}
214: q \sim N_c^0 \; \; \; R \sim N_c^0 \; \; \; g \sim N_c^1 \; \;
215: \;
216: M \sim N_c^1 \; \; \; k \sim N_c^1 \; \; .
217: \end{equation}
218: These scaling rules ensure that energy of the classical ``soliton''
219: scales as $N_{c}^{1}$, the characteristic frequencies associated
220: with internal excitations of the ``soliton'' ($\sqrt{2k/M}$) scale
221: as $N_c^0$, and that the excitations associated with exotic motion
222: also scale as $N_c^0$. This behavior is analogous to the $SU(3)$
223: soliton system \cite{Coh04}.
224:
225:
226:
227:
228: \section{Rigid-Rotor Quantization\label{RRQ}}
229:
230: To develop some intuition about the rigid-rotor approach, first
231: consider a simpler problem: the charged composite particle moving
232: on the surface of a sphere {\it without} a magnetic monopole. Both
233: constituents of the particle have mass $M$ and interact via a
234: (nearly) harmonic potential with spring constant $k$. The
235: classical ground state of the ``soliton'' has the two particles
236: on top of one another located at some point (say the north
237: pole). Of course, since the ``soliton'' breaks rotational
238: symmetry, this classical state is highly degenerate---the
239: ``soliton'' can be localized at any point on the sphere. Thus,
240: there exists a collective manifold of configurations, all with
241: the same classical ground state energy.
242:
243: One can consider the system slowly moving through this
244: manifold---{\it i.e.}, the two particles move together in lock-step,
245: with the center of mass moving collectively around the sphere.
246: Without solving the quantum equations of motion one can see that
247: this approximation is justified quantum mechanically in the large
248: $N_c$ limit due to the large spring constant and the large mass in
249: Eq.~(\ref{toyscale}). To see this, first one assumes that the
250: intrinsic vibrations are decoupled from the collective motion, and
251: subsequently checks for self consistency. The characteristic energy
252: scale for low-lying collective excitations is just the inverse of
253: the moment of inertia $E_{\rm rot} \sim \frac{1}{2 M R^{2}} \sim
254: 1/N_c $. This is very small compared to the characteristic energy
255: associated with the intrinsic vibrations $E_{\rm vib} \sim \sqrt{2
256: k/M} \sim N_c^0$. The fundamentally different scales at large $N_c$
257: allow the collective rotational motion to be essentially decoupled
258: from the intrinsic vibrational motion. Thus the composite
259: system---our ``soliton''---behaves as if it were a single charged
260: particle of mass $2 M$ at large $N_{c}$. Indeed this is hardly
261: surprising: the strong spring constant ensures that in the large
262: $N_c$ limit the two particles are tightly bound, and thus move
263: together collectively. This is rigid rotor quantization since the
264: internal structure of the ``soliton'' is approximated as being rigid
265: and corrections to this are higher order in $1/N_c$.
266:
267: Now consider what happens when the monopole field is added to the
268: system. The spring constant remains strong and the wave function
269: for the composite system remains highly localized. Thus it is
270: plausible that the two particles continue to move collectively
271: together in the same way that they did in the absence of the
272: monopole. This reasoning suggests that the collective excitations
273: will be those of a single particle of mass of $2 M$ and charge $q$
274: moving in the field of the monopole.
275:
276: To obtain the collective energy spectrum, we need only consider the
277: dynamics of a single charged particle (of mass $2M$ and charge $q$)
278: on a sphere of radius $R$ with a magnetic monopole of strength $g$
279: at its center. This system is well described in refs.
280: \cite{Wit1,Guad}. Any point on a sphere can be labeled by an element
281: of $SU(2)$. (Technically, a point on a sphere $S^2$ corresponds to
282: a fiber in $SU(2)$; elements in a fiber are related by unitary
283: phases.) Thus the states of a single particle on a sphere correspond
284: to the irreducible representations of $SU(2)$, and these can be
285: written in terms of Wigner D-functions $D^{J}_{m,m'}$. Without a
286: magnetic monopole, a particle sitting on a sphere has no intrinsic
287: angular momentum and the $m'$ of the Wigner D-function is always
288: zero. The allowed states are then simply the spherical harmonics
289: $Y^{J}_{m}$. However, the presence of a magnetic monopole gives the
290: particle an intrinsic angular momentum of $q g$. The reason for
291: this is simple. A system which has both an electric and magnetic
292: field has momentum carried in the fields. A calculation of
293: $\vec{J}_{\rm field} = \vec{r} \times \vec{p}_{\rm field}$ for a
294: static classical field configuration yields $\vec{J} = q g
295: \hat{r}$---the angular momentum along the intrinsic $z$ axis is $q
296: g$. This restricts the allowed representations to those where $J
297: \geq q g$. Thus the states of the particle can be written as Wigner
298: D-functions $D^{J}_{m, q g}$, with $J \geq q g$.
299:
300: It is easy to find the energy of the system. At the classical
301: level, the Hamiltonian is given by
302: \begin{equation}
303: H = \frac{J^2 - (q g)^2}{2 I} \; .\label{Hclass}
304: \end{equation}
305: where $I$ is the moment of inertia. The energy is purely kinetic
306: and is zero when the system is at rest, namely, for $J = g q$.
307: For the present system $I = 2 M R^2$ where the factor of two
308: reflects the fact that the mass of the composite is $2 M$. To
309: quantize the system one simply promotes $J$ to a quantum
310: operator. The energy of the system is then given by
311: \begin{equation}
312: E(J) = \frac{J (J+1) - (q g)^{2}}{4 M R^2} \; \; {\rm with}\; \; J
313: \ge q g \; . \label {ERR}
314: \end{equation}
315: The excitation energy of the lowest-lying collective state is then
316: given by
317: \begin{equation}
318: \Delta E_{R} \equiv {E(q g+1) - E(q g)} = \frac{q g}{2 M R^2} \; .
319: \label{DeltaER}
320: \end{equation}
321:
322:
323: \section{Callan-Klebanov quantization}
324: \begin{figure}[t]
325: \begin{minipage}{2.5in}
326: \includegraphics{analytic.eps} \label{anFig}
327: \caption{Semiclassical solutions of the system based on the
328: Callan-Klebanov approach as given in \cite{Coh04}. The energy is
329: measured in units of $\omega_r$ and the solutions are plotted as a
330: function of the ratio $\omega_v / \omega_r$, with $\omega_r =
331: \frac{q g}{2 M R^2}$ and $\omega_v = \sqrt{\frac{2 k}{M}}$. The plot
332: shows $\Delta E = E - E_{ground}$. Each solution is labeled by the
333: $n_1 n_2 n_3$ of Eq.~(\ref{n1n2n3}), giving its decomposition in
334: terms of the three non-zero positive normal modes of the system.}
335: \end{minipage}
336: \end{figure}
337:
338: There is another way to look at the problem. First, consider the
339: case of a single charged particle moving on a sphere in the field
340: of a strong magnetic monopole from a classical perspective. The
341: system is a charged particle moving in a strong magnetic field.
342: The paths of particles moving in magnetic fields bend, and in
343: strong fields they bend into classical orbits of small radius.
344: Thus, the particle does not make great circle orbits around the
345: sphere but instead makes tight local orbits. Moreover, it is well
346: known that the quantization of these localized orbits leads to
347: Landau levels---namely, the excitation spectrum of a harmonic
348: oscillator \cite{Landau}. Now consider the classical dynamics of
349: the complete toy model, with two interacting particles on a
350: sphere in the presence of a magnetic monopole. There are two
351: types of harmonic dynamics: the orbits associated with motion of
352: the center of mass in the magnetic field, and the motion
353: associated with excitations of the two particles relative to one
354: another. These two types of motion each have characteristic
355: frequencies associated with them. The orbits of the center of
356: mass due to the magnetic field have a characteristic frequency,
357: \begin{equation}
358: \omega_r \equiv \frac{q g}{2 M R^2} \label{omegar} \; ,
359: \end{equation}
360: which is the cyclotron frequency of a particle of mass $2M$ and
361: charge $q$ moving in the magnetic field of a monopole of strength
362: $g$ a distance $R$ away. It is worth observing that $\omega_r$ is
363: precisely equal to the excitation energy in the rigid rotor
364: approximation of Eq.~(\ref{DeltaER}). The characteristic frequency
365: of the intrinsic vibrations is
366: \begin{equation}
367: \omega_{v} \equiv \sqrt{ \frac{2 k}{M} } \; . \label{omegav}
368: \end{equation}
369: The key point is that the classical equations of motion can induce
370: mixing between these two types of motion.
371:
372: The classical equations can be truncated at harmonic order and
373: then solved for the normal mode frequencies. Due to the presence
374: of velocity-dependent forces, it is useful to formulate the
375: problem in terms of coupled first-order differential equations for
376: positions and velocities. This classical problem was analyzed in
377: refs.~\cite{Coh04,Pob} and this analysis is briefly recapitulated
378: in the appendix. The problem has four degrees of freedom and
379: hence there are four normal mode frequencies (which come paired
380: as positive and negative frequency solutions to the equations of
381: motion). Three of the normal mode frequencies are nonzero and
382: the fourth is a zero mode. The Callan-Klebanov approach amounts
383: to the quantization of these harmonic modes.
384:
385: The zero mode is non-dynamical in nature: it corresponds to
386: relocating the position of the ``soliton'' to a new point on the
387: sphere but has no velocity associated with it. Quantum mechanically
388: this mode corresponds to the nonexotic states of the system. For
389: the present case it represents the ``excitation'' of one of the $2
390: qg +1$ degenerate ground states. In the context of the $SU(3)$
391: soliton case it corresponds to the excitation of the usual hyperons
392: from the nucleon (which, of course, are exactly degenerate in the
393: SU(3) limit).
394:
395: The non-zero modes correspond to physical harmonic excitations. The
396: excitation spectrum to leading order in the $1/N_c$ expansion is
397: thus given by
398: \begin{equation}\label{n1n2n3}
399: \Delta E= E - E_{ground} = n_1 \omega_1 + n_2 \omega_2 + n_3
400: \omega_3
401: \end{equation}
402: where the $\omega_i$ are the three non-zero positive
403: eigenfrequencies. In Fig. 1, we plot a few representative
404: low-lying states. In this figure $\Delta E$ is given in units of
405: $\omega_r$ and it is plotted as a function of differing spring
406: constants which are reflected in the dimensionless ratio
407: $\frac{\omega_v}{\omega_r} = \frac{R^2 \sqrt{8 M k}}{q g}$. The
408: advantages of working with these two dimensionless ratios should
409: be clear; the results only depend on particular combinations of
410: the parameters greatly simplifying the analysis. Thus, it is
411: sufficient to work with fixed values of $M$, $R$, $q$ and $g$
412: while varying $k$ to explore all of the physics at large $N_c$.
413: It is worth noting that $\frac{\omega_v}{\omega_r}$ is of order
414: $N_c^0$ and thus the large $N_c$ limit can be taken for any value
415: of this ratio.
416:
417: For the sake of comparison, in Fig.~1 we also plot the excitation
418: energy of the collective state predicted by the rigid rotor
419: quantization as a dashed line (located at $\frac{\Delta
420: E}{\omega_r} = 1$) . One key observation needs to be made at this
421: point. Clearly these two approaches predict different excitation
422: spectra for generic values of $\frac{\omega_v}{\omega_r}$. The two
423: approaches do agree in the limit $\frac{\omega_v}{\omega_r}
424: \rightarrow \infty$ where $\omega_1 \rightarrow 1$, but as noted
425: above nothing in the large $N_c$ scaling rules tells us that this
426: ratio should be large.
427:
428: It is apparent from this analysis that there are two conflicting
429: pictures for quantization in this problem. At this stage it is
430: worth noting that the rigid-rotor approach describes far fewer
431: states in the spectrum than the Callan-Klebanov approach does. It
432: is important to determine which of these two quantization schemes
433: actually describes this system as the large $N_c$ limit is
434: approached.
435:
436:
437: \section{Numerical solution of the model}
438:
439: To numerically solve the quantum system one must specify the
440: Hamiltonian in a particular basis. As shown in ref.~\cite{Guad},
441: Wigner D-functions with $J \geq eg$ form a basis of states for the
442: charged particle. The states in this basis are labeled by $J$
443: ($J \ge q g$), $m_J$ ($J \ge m_J \ge -J$) and ${m'}_J= q g$. Of
444: course, the standard spherical harmonics form a basis of states
445: for the neutral particle and are labeled by $L$ and $m_L$ ($L \ge
446: m_L \ge -L$). We can obtain a basis of states for the composite
447: system by taking tensor products of the basis states for the
448: charged and neutral particles.
449:
450: \begin{figure*}[t]
451: \label{numFig}
452: \begin{tabular}{ccc}
453: \begin{minipage}{2in}
454: \includegraphics{p100.eps}
455: \end{minipage} &
456: \begin{minipage}{2in}
457: \includegraphics{p200.eps}
458: \end{minipage} &
459: \begin{minipage}{2in}
460: \includegraphics{p300.eps}
461: \end{minipage} \\
462: \begin{minipage}{2in}
463: \includegraphics{p010.eps}
464: \end{minipage} &
465: \begin{minipage}{2in}
466: \includegraphics{p110.eps}
467: \end{minipage} &
468: \begin{minipage}{2in}
469: \includegraphics{p001.eps}
470: \end{minipage} \\
471: \end{tabular}
472: \caption{We plot the numerical simulation data (dots) against the
473: semiclassical approximation solutions based on the
474: Callan-Klebanov approach (solid lines) for each of the energy
475: levels shown in Fig.~1. The plots are of ${\Delta}E = E -
476: E_{ground}$. The numerical computation was done with $ e g = 40$
477: and $J_{max} = L_{max} = 48$.}
478: \end{figure*}
479:
480: In the semiclassical analysis it is sufficient to consider a
481: harmonic potential in the large $N_c$ limit; anharmonic effects
482: only come in as $1/N_c$ corrections. Thus in the numerical model
483: studied, the interaction potential between the particles should be
484: approximately quadratic at short distances to mimic the soliton
485: problem. However, since the problem is posed on a sphere and we are
486: using angular variables, it is necessary that the potential be
487: periodic in the angular separation. The simplest form for the
488: potential with these properties is $k R^2 (1 - \cos{\gamma})$, where
489: $\gamma$ is the angular separation between the charged and neutral
490: particles.
491:
492: The Hamiltonian for this system is the sum of the kinetic
493: energies of the charged particle, the kinetic energy of the
494: neutral particles and the interaction potential between them:
495: \begin{equation}\label{H}
496: \hat{H}=\hat{H}_q+ \hat{H}_n+\hat{H}_{\rm int} \; .
497: \end{equation}
498: The charged particle kinetic energy is
499: \begin{equation}\label{Hq}
500: \hat{H}_{q} = \frac{\hat{J}^2 \otimes \hat{\mathrm{1}}_n - (eg)^{2}}{2 M
501: R^{2}} \; ,
502: \end{equation}
503: for the reasons discussed earlier. The neutral particle term is
504: \begin{equation}\label{Hn}
505: \hat{H}_{n} = \frac{ \hat{\mathrm{1}}_q \otimes \hat{L}^2}{2 M R^{2}}
506: \end{equation}
507: and the interaction term is
508: \begin{equation}\label{Hint}
509: \hat{H}_{\rm int} = R^2 k (1 - \cos{\gamma}) \; .
510: \end{equation}
511:
512: A straightforward computation using standard identities produces
513: the matrix elements of the total system Hamiltonian:
514: \begin{widetext}
515: \begin{eqnarray} \label{matElemEq}
516: &\langle J', {m'}_J, e g; L', {m'}_L|\hat{H}|J, m_J, e g; L, m_L
517: \rangle = \frac{J(J+1) +
518: L(L+1) - (eg)^{2}}{2 M R^{2}}\delta(J,J')\delta(m_J,-{m'}_J)\delta(L,L')\delta(m_L,-{m'}_L) + \nonumber \\
519: & k R^2 \, \delta(J,J')\delta(m_J,-{m'}_J)\delta(L,L')\delta(m_L,-{m'}_L) - (-1)^{e g - {m'}_J +
520: {m'}_L}\sqrt{(2J+1)(2J'+1)(2L+1)(2L'+1)}\times \nonumber \\
521: &\sum_{c=-1}^{1}{
522: \begin{pmatrix}
523: J' & 1 & J \\
524: - e g & 0 & e g \\
525: \end{pmatrix}
526: \begin{pmatrix}
527: J' & 1 & J \\
528: -{m'}_J & -c & m_J \\
529: \end{pmatrix}
530: \begin{pmatrix}
531: L' & 1 & L \\
532: 0 & 0 & 0 \\
533: \end{pmatrix}
534: \begin{pmatrix}
535: L' & 1 & L \\
536: -{m'}_L & c & m_L \\
537: \end{pmatrix}
538: }
539: \end{eqnarray}
540: \end{widetext}
541: Here $\delta(i,j)$ is the Kronecker delta function, and the terms in
542: the summation are Wigner $3j$ symbols. Armed with
543: Eq.~(\ref{matElemEq}), we can calculate the matrix of $\hat{H}$ in a
544: truncated basis (working up to some cutoff values of $J$ and $L$),
545: and find the lowest few eigenvalues. Of course the system is
546: rotationally invariant, and hence the matrix block diagonalizes into
547: blocks of good total angular momentum and good z component of the
548: total angular momentum. In principle, one can exploit this symmetry
549: to greatly reduce the size of the matrices considered. It is quite
550: straightforward to exploit the third component of the total angular
551: momentum: to find the spectrum it is sufficient to study states of
552: total $m=0$ since all multiplets have an $m=0$ member. Since the
553: $z$ component of the angular momentum is additive it is sufficient
554: to study basis states which have $m_J=-m_L$. Imposing good total
555: angular momentum is in principle straightforward, but in practice is
556: rather cumbersome due to the large number of terms in the
557: Clebsch-Gordan series (as a result of the large cutoffs needed for
558: convergence for the cases of numerical interest). Thus it was
559: simpler to work with the full matrices for total $m=0$. Taking $q
560: g=40$ to ensure large $N_c$, our matrices were of dimension
561: approximately $21,000$. Fortunately they are quite sparse and easily
562: amenable to sparse matrix techniques. With these methods the lowest
563: several eigenvalues of the system were easily calculated.
564:
565: In Fig.~2, numerical solutions for several low-lying energy levels
566: are presented for the case $q g = 40$. They are compared to the
567: semiclassical predictions based on the Callan-Klebanov approach; a
568: wide variety of spring constants are considered. The plots are
569: expressed in terms of the same dimensionless ratios as used in
570: Fig.~1. To keep the graphs uncluttered, we have presented each
571: energy level as a function of the strength of the spring constant
572: on a separate graph. We used the level ordering given in the
573: semiclassical expression, Eq.~(\ref{n1n2n3}), to associate
574: particular numerically computed energies with semiclassical
575: states; {\it i.e.}, the $n^{th}$ lowest numerical computed state
576: is associated with the $n^{th}$ lowest state in
577: Eq.~(\ref{n1n2n3}). There is some ambiguity with this method in
578: the immediate vicinity of level crossings in which case we used
579: numerical smoothness to determine the association of levels. It
580: is quite apparent that the actual energy levels of the system
581: closely follow the semiclassical treatment based on the
582: Callan-Klebanov approach. Moreover, generically they are rather
583: far from the prediction of rigid rotor quantization; namely, that
584: a collective state should exist at $\frac{\Delta E}{\omega_r} = 1$
585: (which is indicated in both Fig.~1 and Fig.~2 as a dashed line).
586:
587:
588: Of course, the semiclassical predictions based on the
589: Callan-Klebanov method are not expected to precisely reproduce the
590: spectra. One expects $1/N_c$ corrections with a characteristic
591: scale of $(q g)^{-1}$ and that such corrections will increase with
592: increasing excitation energy due to anharmonicities. Thus it seems
593: highly plausible that the small but discernible deviations of the
594: semiclassically predicted $2\, 0\, 0$ and $3\, 0\, 0$ modes from the
595: numerical simulation are due to such effects. Indeed it is easy to
596: see that the scale of these deviations is typical of what one
597: expects. While it is generically nontrivial to compute the $1/N_c$
598: corrections, it is quite straightforward to do so in the limit
599: $\frac{\omega_v}{\omega_r} \rightarrow \infty$, which is the
600: infinitely strong coupling limit. In this case, the dynamics
601: undoubtedly do reduce to that of a single particle of mass $2 M$,
602: and Eq.~(\ref{ERR}) holds for any $N_c$. This implies that in the
603: limit $\frac{\omega_v}{\omega_r} \rightarrow \infty$ the actual
604: value for $\frac{\Delta E}{\omega_r}$ in a state $n 0 0$ will exceed
605: the Callan-Klebanov prediction by an amount given by $\frac{n^2 +
606: 1}{2 q g}$. This is clearly a $1/N_c$ correction. It gives a
607: correction of $0.025$, $0.0625$ and $0.125$ for the $1\, 0\, 0$,
608: $2\, 0\, 0$ and $3\, 0\, 0$ states, respectively, in the strong
609: coupling limit. For the largest values of
610: $\frac{\omega_v}{\omega_r}$ we computed ($\frac{\omega_v}{\omega_r}
611: = 3.84$), the actual amount by which the Callan-Klebanov formula
612: underpredicted the numerical result was $0.0250$, $0.0739$, and
613: $0.147$, respectively, and it is highly plausible that these values
614: will asymptote to the known $1/N_c$ corrections in the strong
615: coupling limit. Thus, the size of the deviations from the
616: Callan-Klebanov predictions is of the scale expected from $1/N_c$
617: corrections.
618:
619: Upon the completion of this work we learned of a calculation of the
620: spectrum of this model by Diakonov and Petrov. Their results agree
621: with ours---the Callan-Klebanov results accurately describe the
622: spectrum while the rigid rotor does not\cite{private}.
623:
624:
625: \section{Conclusion}
626:
627:
628: The analysis of the toy system considered above shows explicitly
629: that for a system with topological velocity-dependent interactions
630: analogous to a dynamically active Wess-Zumino term, the
631: Callan-Klebanov method is the correct way to implement
632: semiclassical quantization. On the other hand, rigid-rotor
633: quantization does not generally work in this system.
634:
635: It is worth understanding why the plausible sounding argument given
636: in Sect.~\ref{RRQ} for rigid-rotor quantization fails. The key point
637: is that the intuition gained from the case without the
638: monopole---that the strong coupling present at large $N_c$ implies
639: that the center of mass of the system moves collectively---does not
640: automatically translate to the case where the monopole {\it is}
641: present. While the strength of the coupling remains large, the
642: effect of the monopole on the internal dynamics is also
643: parametrically large for large $N_c$. The reason that the monopole
644: plays such a role should be clear. If one imagines the center of
645: mass of the composite system slowly moving in an apparently
646: collective way, one finds that the monopole exerts a force only on
647: the charged constituent, but {\it not} on the neutral one. The
648: neutral constituent can only follow the charged one due to the force
649: exerted by the spring. This implies possible mixing of the internal
650: dynamics associated with the spring and the collective dynamics.
651: Such mixing will be strong if the characteristic scale of the
652: internal vibrations $\omega_v$ is comparable to the scale associated
653: with collective motion $\omega_r$. As both of these scales are
654: parametrically of order $N_c^0$ there is no reason associated with
655: $1/N_c$ physics for them not to mix strongly and thereby ruin the
656: rigid rotor dynamics. This is the semiclassical argument outlined in
657: ref.~\cite{Coh04}. Underlying this is the simple time scale argument
658: of ref.~\cite{Coh03}.
659:
660: An alternative way to see this is simply to look at the classical
661: dynamics of the system discussed in the appendix. The essential
662: point there is to note that although there are two center of mass
663: degrees of freedom (say, moving in the $x$ and $y$ directions from
664: the north pole) there is only {\it one} zero mode and it is
665: purely static. Rigid-rotor quantization is only legitimate for
666: true collective motion associated with zero modes. Since there
667: is no dynamical zero mode in the problem, one does not expect
668: rigid-rotor quantization to apply. It is worth noting here that
669: the analogous thing occurs for the chiral soliton case: it is
670: precisely the lack of dynamical zero modes which accounts for the
671: difference between the Callan-Klebanov and rigid-rotor
672: approaches.
673:
674:
675: Of course, the model considered here is just a toy. The real
676: question of interest is what form of quantization is correct for the
677: soliton models. However, the analogy with the actual models is, in
678: fact, very close: the key issues of topology, velocity-dependent
679: forces and the existence of collective and vibrational modes with
680: the same quantum numbers are present in both the toy problem and in
681: the problem of physical interest. Thus, this calculation should be
682: viewed as strong evidence that in the physically interesting
683: problems arising in soliton models the Callan-Klebanov approach is
684: correct at large $N_c$ and the rigid-rotor approach is generally
685: incorrect.
686:
687: There are models for which the rigid rotor approximation does
688: work, such as the model considered in ref.~\cite{DP}. However, in
689: that case it is easy to see that the Callan-Klebanov approach
690: gives the same result as the rigid rotor method at large
691: $N_c$.\cite{Pob} Thus, the present situation is one in which the
692: Callan-Klebanov approach has been shown to be valid at large
693: $N_c$ for all cases studied, while the rigid rotor method is only
694: valid when it agrees with the Callan-Klebanov approach. This
695: strongly suggests that where the two methods disagree the
696: Callan-Klebanov quantization is the correct one. Since the two
697: approaches are known to disagree for chiral soliton
698: models\cite{IKOR}, it seems highly unlikely that the rigid rotor
699: quantization is valid; at large $N_c$ Callan-Klebanov
700: quantization is the correct approach. This strongly suggests that
701: the many calculations done using rigid rotor quantization for
702: exotic states of $SU(3)$ solitons are unjustified.
703:
704: \acknowledgments The authors acknowledge constructive comments from
705: I.~ Klebanov, D.~Diakonov, D.~Dakin and P.~V.~Pobylitsa. This work
706: was supported by the U.S.~Department of Energy through grant
707: DE-FG02-93ER-40762. One of the authors (AN) acknowledges the support
708: of the University of Maryland through the Senior Summer Scholars
709: program.
710:
711:
712: \section{Appendix}\label{append}
713: \begin{figure}
714: \label{threeOmegas}
715: \begin{minipage}{2.5in}
716: \includegraphics{threeomega.eps}
717: \caption{We plot the non-zero semiclassical eigenmodes of the
718: system, $\omega_1$, $\omega_2$, $\omega_3$.}
719: \end{minipage}
720: \end{figure}
721:
722: Here we review the semiclassical treatment of the toy problem,
723: following ref.~\cite{Coh04}. The first step is the treatment of
724: classical motion at small amplitude centered on the north pole.
725: For small amplitude motion, the system looks like a particle on a
726: plane with a constant perpendicular magnetic field. In this
727: regime, we can write the linearized equations of motion for the
728: particles and their velocities to first order as
729: \begin{equation}\label{eom}
730: \frac{d \mathbf{v}}{d t} = i \overline{M} \mathbf{v}
731: \end{equation} with
732: \begin{widetext}
733: \begin{equation}
734: \overline{M}= -i \left( \begin{array}{c c c c c c c c}
735: 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \\
736: 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \\
737: 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \\
738: 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \\
739: \frac{-\omega_v^2}{2} & 0 & \frac{\omega_v^2}{2}& 0 & 0 & 2 \omega_r & 0 & 0 \\
740: 0 & \frac{-\omega_v^2}{2} & 0 & \frac{\omega_v^2}{2} & -2\omega_r & 0 & 0 & 0 \\
741: \frac{\omega_v^2}{2} & 0 & \frac{-\omega_v^2}{2}& 0 & 0 & 0 & 0 & 0 \\
742: 0 & \frac{\omega_v^2}{2} & 0 & \frac{-\omega_v^2}{2} & 0 & 0 & 0 & 0 \end{array} \right)\; \mathrm{and} \;
743: \mathbf{v}= \left( \begin{array}{c} x_q\\ y_q\\ x_n \\
744: y_n\\\dot{x}_q\\\dot{y}_q\\\dot{x}_n\\\dot{y}_n \end{array} \right)
745: \; \; \; {\rm with} \; \; \; \omega_v=\sqrt{\frac{2 k}{m}} \; \;, \;
746: \; \omega_r=\frac{g q}{2 m R^2} \; . \label{m}\end{equation}
747: \end{widetext}
748: The subscripts $q$ and $n$ refer to the charged and neutral
749: particles respectively, and $x$ and $y$ refer to the Cartesian
750: coordinates. Note that since there are velocity-dependent terms
751: it is natural to work in terms of coupled first-order differential
752: equations. The canceling factors of $i$ in Eqs.~(\ref{eom}) and
753: (\ref{m}) are put in for later convenience.
754:
755: We are interested in finding the eigenmodes of the system; {\it
756: i.e.}, harmonic solutions of the form $\mathbf{v}(t) = \mathbf{v}_j
757: \exp(-i \omega_j t)$. Inserting this ansatz into Eq.~(\ref{eom})
758: yields a simple eigenvalue equation:
759: \begin{equation}\label{modeEq}
760: \omega_j \mathbf{v}_j = \overline{M} \mathbf{v}_j \; .
761: \end{equation}
762: On physical grounds, we know the $\omega_j$ are real. Moreover the
763: matrix $\overline{M}$ is purely imaginary which implies that if
764: ${\mathbf v}_j$ is an eigenvector with eigenvalue $\omega_j$ then
765: ${\mathbf v}_j^*$ is an eigenvector with eigenvalue $-\omega_j$.
766: Thus, the eigenvectors form pairs associated with positive and
767: negative frequency solutions. We refer to one of these pairs
768: together an an eigenmode.
769:
770: One of these modes is a zero frequency mode:
771: \begin{equation}
772: \mathbf{v}_0= \left( \begin{array}{c} 1\\ i \\ 1 \\ i \\
773: 0\\ 0 \\ 0 \\ 0 \end{array} \right ) \;\;.
774: \end{equation}
775: A striking feature of this zero mode is the absence of any time
776: dependence in it: the four lower components associated with the
777: time derivatives are zero. Thus, this mode is associated with
778: pure static rotations. By extracting the real and imaginary
779: parts ($\mathbf{v}_0$ and $\mathbf{v}_0^*$ are degenerate since
780: $\omega=0$ and hence one can form linear combinations of the two)
781: we see that this mode corresponds to a collective
782: time-independent rotation of the two particles in either the $x$
783: or $y$ directions. As this mode is completely non-dynamical,
784: when quantized it is associated with ``excitations'' which move
785: from one of the highly degenerate ground states of the theory to
786: another. In this context it is useful to recall that the
787: degeneracy of the ground state is $2 qg +1$ which diverges at
788: large $N_c$.
789:
790:
791: It is extremely important to note that the one zero mode found
792: above is the {\it only} zero mode for the system. We can solve Eq.
793: (\ref{modeEq}) to find the three non-zero frequency eigenmodes of
794: the system. The three eigenvalues can be computed analytically,
795: but the form of the solutions is cumbersome and not particularly
796: illuminating. We denote the three eigenfrequencies as $\omega_1$,
797: $\omega_2$, and $\omega_3$ defined such that $\omega_3 > \omega_2
798: > \omega_1$. These are plotted in terms of convenient
799: combinations of variables in Fig.~3.
800:
801: As these modes are harmonic in the large $N_c$ limit, they can be
802: quantized trivially yielding Eq.~(\ref{n1n2n3}).
803:
804:
805:
806:
807: % ----------------------------------------------------------------
808: \bibliographystyle{amsplain}
809: \begin{thebibliography}{99}
810: \bibitem{exp}The first report was in T.~ Nakano {\it et al.},
811: Phys.~Rev.~Lett.~{\bf 91}, 012002 (2003). A number of other
812: groups have also reported evidence for the $\theta^+$. A summary
813: of the situation as of the end of 2003 can be found in S.
814: Eidelman {\it et al.}, Phys. Lett. {\bf B 592}, 1 (2004).
815:
816: \bibitem{Coh03} T.~D.~Cohen, Phys.Lett. B {\bf 581}, 175-181 (2004).
817:
818: \bibitem{Pres} M.~Praszalowicz, talk at
819: ``Workshop on Skyrmions and Anomalies'', M. Je˙zabek and M.
820: Praszalowicz, editors (World Scientific, 1987), page 112. This work
821: was not published in a journal at that time. An updated analysis
822: based on this work may be found in hep-ph/0308114.
823:
824: \bibitem{DiaPetPoly} D.~Diakonov, V.~Petrov and M.~Polyakov,
825: Z.~Phys. ~A{\bf \ 359}, 305 (1997).
826:
827: \bibitem{Guad} E.~Guadagnini, Nucl.~Phys.~B{\bf 236}, 35 (1984).
828:
829: \bibitem{SU3Quant}
830: P.~O.~Mazur, M.~A.~Nowak and M.~Praszalowicz, Phys.~Lett.~B{\bf
831: 147},137 (1984); A.~V.~Manohar, Nucl.~Phys.~{\bf B248}, 19 (1984);
832: M.~Chemtob, Nucl.~Phys.~B{\bf 256}, 600 (1985); S.~Jain and
833: S.~R.~Wadia, Nucl.~Phys.~B{\bf 258}, 713 (1985).
834:
835: \bibitem{ANW} G.~Adkins, C.~Nappi and E.~Witten, Nucl.~Phys.~B{\bf
836: 228}, 522 (1983).
837:
838: \bibitem{Coh04} T.~D.~Cohen, hep-ph/0312191.
839:
840: \bibitem{IKOR} N.~Itzhaki, I.~R.~Klebanov, P.~Ouyang,
841: L.~Rastelli, hep-ph/0309305.
842:
843: \bibitem{Pob} P.~V.~Pobylitsa, hep-ph/0310221.
844:
845: \bibitem{DP} D.~Diakonov and V.~Petrov, Phys.Rev. D {\bf 69}, 056002 (2004).
846:
847: \bibitem{CK}C.~G.~Callan, Jr.~and I.~R.~Klebanov,
848: Nucl.~Phys.~{\bf B262}, 365 (1985); C.~G.~Callan,
849: Jr.,K.~Hornbostel and I.~R.~Klebanov, Phys.~Lett.~B {\bf B202},
850: 269 (1988).
851:
852: \bibitem{PR} B.-Y.~Park, M.~Rho and D.-P.~Min hep-ph/0405246.
853:
854:
855: \bibitem{Wit1} E.~Witten, Nucl.~Phys.~{\bf B223}, 422 (1983).
856:
857: \bibitem{Landau} This is a standard topic in quantum mechanics;
858: see, for example, L.~D.~Landau and E.~M.~Lifshitz, {\it Quantum
859: Mechanics}, (Pergamon Press, Oxford, 1977).
860:
861: \bibitem{private} Dimitri Diakonov, private communication.
862:
863: \end{thebibliography}
864: \end{document}
865: % ----------------------------------------------------------------
866: