0707.1250/ms.tex
1: % \documentclass[12pt,preprint]{aastex}
2: 
3: %
4:  \documentclass[10pt]{emulateapj}
5: %
6: 
7: \slugcomment{Accepted for publication in The Astrophysical 
8:              Journal (Date: July 9, 2007)}
9: 
10: \shorttitle{UNSTABLE DISK GALAXIES. I. MODAL PROPERTIES}
11: \shortauthors{M. A. Jalali}
12: 
13: \begin{document}
14: 
15: \title{UNSTABLE DISK GALAXIES. I. MODAL PROPERTIES}
16: 
17: 
18: \author{Mir Abbas Jalali
19: %\altaffilmark{1}
20: }
21: \affil{Sharif University of Technology, Azadi Avenue, Tehran, Iran;
22: mjalali@sharif.edu}
23: 
24: 
25: %\altaffiltext{1}{E-mail: mjalali@sharif.edu}
26: 
27: \begin{abstract}
28: 
29: I utilize the Petrov-Galerkin formulation and develop a new method
30: for solving the unsteady collisionless Boltzmann equation in both
31: the linear and nonlinear regimes. In the first order approximation, 
32: the method reduces to a linear eigenvalue problem which is solved
33: using standard numerical methods. I apply the method to the dynamics
34: of a model stellar disk which is embedded in the field of a
35: soft-centered logarithmic potential. The outcome is the full
36: spectrum of eigenfrequencies and their conjugate normal modes for
37: prescribed azimuthal wavenumbers. The results show that the
38: fundamental bar mode is isolated in the frequency space while
39: spiral modes belong to discrete families that bifurcate from the
40: continuous family of van Kampen modes. The population of spiral
41: modes in the bifurcating family increases by cooling the disk and
42: declines by increasing the fraction of dark to luminous matter. It
43: is shown that the variety of unstable modes is controlled by the
44: shape of the dark matter density profile. 
45: \end{abstract} 
46: 
47: \keywords{stellar dynamics,
48:           instabilities,
49:           methods: analytical,
50:           galaxies: kinematics and dynamics,
51:           galaxies: spiral,
52:           galaxies: structure}
53: 
54: 
55: 
56: \section{INTRODUCTION}
57: 
58: Dynamics of self-gravitating stellar systems and plasma fluids
59: are governed by the collisionless Boltzmann equation (CBE) \citep{BT87}.
60: Finding a general solution of the CBE has been a challenging problem
61: in various disciplines of physical sciences. Due to existing difficulties
62: of the general problem, finding a solution to the linearized CBE became
63: the center of attraction in the twentieth century when \citet{L46} and
64: van Kampen (1955) discovered the normal modes of collisionless ensembles.
65: Later in 1970's, \citet{K71,K77} developed a matrix theory that was
66: capable of computing normal modes of stellar systems through solving a
67: nonlinear eigenvalue problem. His theory remained as the only analytical
68: perturbation theory used by the community of galactic dynamicists over
69: the past three decades.
70: 
71: \citet{K77} assumed an exponential form
72: $\exp(-{\rm i}\omega t)$ for the time-dependent part
73: of physical quantities where ${\rm i}=\sqrt{-1}$, and
74: solved the linearized CBE for the perturbed distribution
75: function (DF) $f_1$ in terms of the perturbed
76: potential $V_1$. After expanding the potential and
77: density functions in terms of bi-orthogonal basis sets
78: in the configuration space, he used the weighted residual
79: form of the fundamental equation
80: %
81: \begin{equation}
82: f_1 d \textbf{\textit{v}} d \textbf{\textit{x}}=
83: \Sigma_1 d \textbf{\textit{x}},
84: \end{equation}
85: %
86: to obtain a nonlinear eigenvalue problem for the complex
87: eigenfrequency $\omega$. For self-consistent perturbations
88: the surface density $\Sigma_1=\int f_1 d \textbf{\textit{v}}$ is
89: related to $V_1$ through Poisson's integral, and the symbols
90: $ d \textbf{\textit{v}}$ and $ d \textbf{\textit{x}}$ denote the
91: elements of velocity and position vectors.
92: 
93: \citet{Z76} used Kalnajs's theory to compute the modes of the
94: isothermal disk \citep{M63}, which has the astrophysically important
95: property of a flat rotation curve. His analysis was then extended by
96: \citet{ER98a,ER98b} to general scale-free disks with arbitrary cusp
97: slopes. Application of Kalnajs's theory to soft-centered models of
98: stellar disks has been mainly focused on the isochrone and
99: Kuzmin-Toomre disks \citep{K78,H92,PC97}.
100: A disk with exponential light profile and an approximately
101: flat rotation curve was also investigated by \citet{VD96}.
102: More recently Jalali \& Hunter (2005a, hereafter JH) gave
103: new results for soft-centered models of stellar disks. They showed
104: the importance of a boundary integral in the modal properties of
105: unidirectional disks and computed a fundamental bar mode and a
106: secondary spiral mode for the isochrone, Kuzmin-Toomre and a newly
107: introduced family of cored exponential disks. JH also extended Kalnajs's
108: first order perturbation theory to the second order, and illustrated
109: energy and angular momentum content of different Fourier components.
110: Their bar charts showed that only a few number of expansion terms in
111: the radial angle govern the perturbed dynamics.
112: 
113: Implementation of Kalnajs's (1977) theory, however, has some technical
114: problems due to the nonlinear dependency of his matrix equations
115: on $\omega$. Most computational methods that deal with nonlinear
116: eigenvalue problems are iterative. They start with an initial guess
117: of the solution and continue with a search scheme in the frequency
118: space. Newton's method is perhaps the most efficient technique that
119: guarantees a quadratic convergence should the initial guess be close
120: enough to the solution. The key issue in success of any iterative
121: scheme is the attracting or repelling nature of an eigenvalue.
122: It is obvious that only attracting eigenvalues can be captured by
123: iterative methods while we have no priori knowledge of their
124: basins of attraction in order to make our initial guess.
125: The mentioned computational difficulties make it a formidable task 
126: to explore all normal modes of a stellar system, which include 
127: growing modes as well as stationary van Kampen modes. Moreover, 
128: it is not easy to develop a general nonlinear theory based on 
129: Kalnajs's method for studying the interaction of modes. 
130: 
131: \citet{P04,P05} introduced an alternative method for the normal mode
132: calculation of stellar disks whose outcome was a linear eigenvalue
133: problem for $\omega$. His method is capable of finding all
134: eigenmodes of a stellar disk should one use fine grids in the action
135: space. Polyachenko's method is somehow costly because it results in
136: a large linear system of equations to assure point-wise convergence
137: in the action space and mean convergence of Fourier expansions in
138: the space of the radial angle. Extension of his method to nonlinear
139: regime is another challenging problem yet to be investigated.
140: \citet{T05} has also followed an approach similar to \citet{P05} and
141: studied the instability of stellar disks surrounding massive
142: objects. His eigenvalue equations involve action variables, and
143: practically, need to be solved over a discretized grid in the space
144: of actions.
145: 
146: Recent developments in fluid mechanics \citep{DG95,MS99} inspired me
147: to formulate the dynamics of stellar systems in a new framework,
148: which is capable of solving the CBE not only in the linear regime,
149: but also in its full nonlinear form. The method systematically
150: searches for smooth solutions of the CBE by expanding the perturbed
151: DF using Fourier series of angle variables and an appropriate set of
152: trial functions in the space of actions. Coefficients of expansion
153: are unknown time-dependent amplitude functions whose evolution
154: equations are obtained by the Petrov-Galerkin projection \citep{F72} 
155: of the CBE. That is indeed taking the weighted residual form of the 
156: CBE by integration over the action-angle space and deriving a system 
157: of nonlinear ordinary differential equations (ODEs) for the amplitude
158: functions. The associated first order system of ODEs leads to a
159: linear eigenvalue problem, which is solved using standard numerical
160: methods.
161:  
162: In this paper I present my new method and apply it to explore the
163: modal properties of a model galaxy. In a second paper, I will
164: address the nonlinear evolution of modes and wave interactions. The
165: paper is organized as follows. In sections
166: \ref{sec::dynamical-theory} and \ref{sec::linear-theory}, I use the
167: Petrov-Galerkin method to project the CBE to a system of ODEs in the
168: time domain and derive a system of linear eigenvalue equations. Section
169: \ref{sec:modes-exp-disk} presents the eigenfrequency spectra and
170: their corresponding mode shapes of the cored exponential disk of JH.
171: The stars of this model move in the field of a soft-centered
172: logarithmic potential. I also investigate the effect of physical
173: parameters of the equilibrium model on the modal content. 
174: In section \ref{sec:final-discussions}, I discuss on the nature of 
175: a self-gravitating mode and compare the performance of my method 
176: with other theories. Some fundamental achievements of this work 
177: are summarized in section \ref{sec:conclusions}. 
178:   
179:  
180: \section{NONLINEAR THEORY}
181: \label{sec::dynamical-theory}
182: 
183: I use the usual polar coordinates $\textbf{\textit{x}}=(R,\phi)$ and assume
184: that the temporal evolution of the DF and gravitational potential
185: starts from an axisymmetric equilibrium state described by
186: $f_0(\textbf{\textit{x}},\textbf{\textit{p}})$ and $V_0(R)$ so that
187: %
188: \begin{eqnarray}
189: f(\textbf{\textit{x}},\textbf{\textit{p}},t) &=&
190: f_0(\textbf{\textit{x}},\textbf{\textit{p}})+
191: f_1(\textbf{\textit{x}},\textbf{\textit{p}},t), \\
192: %
193: V(\textbf{\textit{x}},\textbf{\textit{p}},t) &=&
194: V_0(R)+V_1(\textbf{\textit{x}},\textbf{\textit{p}},t).
195: \label{eq:perturbations}
196: \end{eqnarray}
197: %
198: Here $\textbf{\textit{p}}=\left (p_R,p_\phi \right )$ is the
199: momentum vector conjugate to $\textbf{\textit{x}}=\left (R,\phi \right )$.
200: Motion of stars in the equilibrium state is governed by the
201: zeroth order Hamiltonian
202: %
203: \begin{equation}
204: {\cal H}_0=\frac 12 \left ( p_R^2+\frac {p_\phi^2}{R^2}\right
205: )+V_0(R). \label{eq:zeroth-order-H}
206: \end{equation}
207: %
208: For bounded orbits, $R$ and $\phi$ become librating and rotating,
209: respectively. One can therefore describe the dynamics using the
210: action variables $\textbf{\textit{J}}=\left (J_R,J_\phi \right )$,
211: %
212: \begin{equation}
213: J_R=\oint p_R  d R,~~J_\phi=\oint p_\phi
214: d \phi=p_\phi,\label{eq:define-actions}
215: \end{equation}
216: %
217: and their conjugate angles $\Theta=(\theta_R,\theta_\phi)$. A
218: transformation $(\textbf{\textit{x}},\textbf{\textit{p}})\rightarrow
219: (\Theta,\textbf{\textit{J}})$ leaves the Hamiltonian ${\cal H}_0$ as
220: a function of actions only, ${\cal H}_0(\textbf{\textit{J}})$, and
221: therefore, the phase space flows of the equilibrium state lie on a
222: two dimensional torus $\textbf{\textit{J}}=\textbf{\textit{c}}$ with
223: $\textbf{\textit{c}}$ being a constant 2-vector. An action-angle
224: transformation can locally be found for any bounded regular orbit,
225: but it is a global transformation if only one orbit family occupies
226: the phase space. The axisymmetric potential $V_0(R)$ supports only
227: rosette orbits. Radial and circular orbits are the limiting cases of
228: rosette orbits with $J_\phi=0$ and $J_R=0$, respectively. By
229: representing $f$ in terms of the action-angle variables, the CBE
230: reads
231: %
232: \begin{equation}
233: \frac{\partial f}{\partial t}+[ f,{\cal H} ]=0,
234: \label{eq:CBE-action-angle}
235: \end{equation}
236: %
237: where $[,]$ denotes a Poisson bracket taken over the action-angle
238: space. According to Jeans theorem \citep{J1915,L-B62} $f_0$ depends
239: on the phase space coordinates through the integrals of motion,
240: which are the actions in the present formulation, and one obtains
241: $[f_0,{\cal H}_0]=0$. Subsequently, equation (\ref{eq:CBE-action-angle})
242: may be rewritten as
243: %
244: \begin{equation}
245: \frac{\partial f_1}{\partial t}=-\left [f_1,{\cal H}_0 \right ]-
246: \left [ f_0,{\cal H}_1 \right ]-\left [f_1,{\cal H}_1 \right ],
247: \label{eq:CBE-perturbed}
248: \end{equation}
249: %
250: where ${\cal H}_1$ is the perturbed Hamiltonian. A dark matter halo
251: contributes both to ${\cal H}_0$ and to ${\cal H}_1$ if it is live,
252: i.e., if it exchanges momentum/energy with the luminous stellar component.
253: In this paper I confine myself to a rigid halo that only contributes
254: to ${\cal H}_0$ through $V_0$ and assume that ${\cal H}_1=V_1$ is the
255: perturbed potential due to self-gravity.
256: 
257: Let me expand $f_1$ and $V_1$ in Fourier series of angle variables
258: and write
259: %
260: \begin{eqnarray}
261: f_1(\Theta,\textbf{\textit{J}},t) &=& \!\!
262: \sum_{m,l=-\infty}^{\infty}\sum_{j=0}^{\infty} d^{ml}_j(t)
263: \Phi^{ml}_j(\textbf{\textit{J}})e^{{\rm i}\left(m\theta_\phi+l\theta_R
264: \right)},
265: \label{eq:expansion-f1} \\
266: V_1(\Theta,\textbf{\textit{J}},t) &=& \!\!
267: \sum_{m,l=-\infty}^{\infty}\sum_{j=0}^{\infty} b^{ml}_j(t)
268: \Psi^{ml}_j(\textbf{\textit{J}})e^{{\rm i}\left(m\theta_\phi+l\theta_R
269: \right)}, \label{eq:expansion-V1}
270: \end{eqnarray}
271: %
272: where $\Phi^{ml}_j(\textbf{\textit{J}})$ and
273: $\Psi^{ml}_j(\textbf{\textit{J}})$ are
274: some trial functions in the space of action variables, and
275: $d^{ml}_j(t)$ and $b^{ml}_j(t)$ are time-dependent amplitude
276: functions. On the other hand, one can expand $V_1$ and its
277: corresponding surface density $\Sigma_1$ in the configuration space
278: as
279: %
280: \begin{eqnarray}
281: \Sigma_1(R,\phi,t) &=& \!\!
282: \sum_{m=-\infty}^{\infty}\sum_{j=0}^{\infty} a^m_j(t)
283: \sigma^{|m|}_j(R)e^{{\rm i}m\phi},
284: \label{eq:expansion-sigma1-config} \\
285: V_1(R,\phi,t) &=& \!\! \sum_{m=-\infty}^{\infty}\sum_{j=0}^{\infty}
286: a^m_j(t) \psi^{|m|}_j(R)e^{{\rm i}m\phi}. \label{eq:expansion-V1-config}
287: \end{eqnarray}
288: %
289: Here $\psi^{|m|}_j(R)$ and $\sigma^{|m|}_j(R)$ are bi-orthogonal
290: potential--surface density pairs that satisfy the relation
291: %
292: \begin{equation}
293: \label{eq:Djk-definition}
294: 2\pi \int\limits_{0}^{\infty}
295: \psi^{|m|}_j(R)\sigma^{|m'|}_{j'}(R)R d R = D_j(m)\delta
296: _{m,m'}\delta_{j,j'},
297: \end{equation}
298: %
299: where $\delta_{m,m'}$ is the Kronecker delta and $D_j(m)$ are some
300: constants. It is remarked that the real parts of $\Sigma_1$, $V_1$
301: and $f_1$ describe physical solutions. In this work I utilize the
302: \citet{CB72} functions
303: %
304: \begin{eqnarray}
305: \psi^{|m|}_j &=&
306:       -\frac {1}{b} \left (\! {1-\xi \over 2}\! \right )^{1/2} \!\!
307:              P^{|m|}_i(\xi),~~\xi={R^2-b^2\over R^2+b^2}, \\
308: \sigma^{|m|}_j &=& \left (\! {2|m|+2j+1\over 2\pi b^2}\! \right )
309:                \left ( {1-\xi \over 2}\right )^{3/2}\!\!
310:                P^{|m|}_i(\xi),
311: \end{eqnarray}
312: %
313: that yield \citep{AI78,H80}
314: %
315: \begin{equation}
316: D_j(m)=D_j(-m)= -{(2|m|+j)!\over 2b j!}.
317: \end{equation}
318: %
319: $P^{|m|}_i(\xi)$ are associated Legendre functions with $i=|m|+j$.
320: Clutton-Brock functions have a length scale $b$, which makes
321: them suitable for reproducing the potential and surface density of
322: soft-centered models. The choice of this parameter is an important
323: step in the calculation of normal modes. I will discuss on this issue
324: later in \S\ref{sec:modes-exp-disk}.
325: 
326: Equating (\ref{eq:expansion-V1}) and (\ref{eq:expansion-V1-config}),
327: multiplying both sides of the resulting equation by 
328: $\exp[-{\rm i}(l\theta_R+m\theta_{\phi})]$ and integrating over the
329: $\left (\theta_R,\theta_{\phi} \right )$-space, lead to 
330: (see also Kalnajs 1977 and Tremaine \& Weinberg 1984)  
331: %
332: \begin{eqnarray}
333: &{}& \!\!\! \sum_{j=0}^{\infty} b^{ml}_j(t) \Psi^{ml}_j(\textbf{\textit{J}})
334:   \!=\! \sum_{j=0}^{\infty} a^m_j(t) \tilde \Psi^{ml}_j(\textbf{\textit{J}}),
335:  \label{eq:relation-b-and-a} \\
336: %
337: &{}& \!\!\! \tilde \Psi^{ml}_j(\textbf{\textit{J}}) \!=\!
338:      {1\over \pi}\int\limits_{0}^{\pi}
339: \psi^{|m|}_j(R) \cos [l\theta_R+m(\theta_{\phi}-\phi)] d \theta _R,
340:  \label{eq:fourier-coeffs}
341: \end{eqnarray}
342: %
343: where $\tilde \Psi^{ml}_j$ are the Fourier coefficients of the basis
344: potential functions in the configuration space. The trial functions
345: used in the expansion of $V_1$ in the action-angle space are not
346: necessarily identical to $\tilde \Psi^{ml}_j$. However, subsequent
347: mathematical derivations are greatly simplified by setting
348: $\Psi^{ml}_j(\textbf{\textit{J}})= \tilde
349: \Psi^{ml}_j(\textbf{\textit{J}})$, which implies
350: $b^{ml}_j(t)=a^m_j(t)$. To build a relation between $d^{ml}_j(t)$
351: and $a^m_j(t)$, I use the fundamental equation
352: %
353: \begin{equation}
354: f_1(\Theta,\textbf{\textit{J}},t) d \textbf{\textit{J}} d \Theta
355: =\Sigma_1(R,\phi,t) R dR d \phi,
356: \label{eq:fundamental-equation}
357: \end{equation}
358: %
359: where $ d \textbf{\textit{J}} d \Theta$ is the volume of
360: an infinitesimal phase space element. On substituting
361: (\ref{eq:expansion-f1}) and (\ref{eq:expansion-sigma1-config})
362: in (\ref{eq:fundamental-equation}), multiplying both sides of
363: the resulting equation by $\psi^{|m|}_j(R)e^{-{\rm i}m\phi}$ and
364: integrating, one obtains
365: %
366: \begin{eqnarray}
367: a^m_j(t)&=& \frac
368: {4\pi^2}{D_j(m)}\sum_{l=-\infty}^{\infty}\sum_{p=0}^{\infty}
369: \Lambda^{ml}_{jp}d^{ml}_p(t),\label{eq:a-versus-d} \\
370: %
371: \Lambda^{ml}_{jp} &=& \int
372: \Psi^{ml}_j(\textbf{\textit{J}})
373: \Phi^{ml}_p(\textbf{\textit{J}}) d \textbf{\textit{J}},
374: \end{eqnarray}
375: %
376: which is inserted in (\ref{eq:expansion-V1}) to represent $V_1$
377: in terms of the amplitude functions $d^{ml}_j(t)$ as
378: %
379: \begin{equation}
380: V_1 \! =\!\!\!\!
381: \sum_{m,l,k=-\infty}^{\infty}\sum_{j,p=0}^{\infty}\frac
382: {4\pi^2}{D_j(m)}\Lambda^{mk}_{jp}\Psi^{ml}_j(\textbf{\textit{J}})
383: d^{mk}_p(t)
384: e^{{\rm i}\left(m\theta_\phi+l\theta_R \right)}.
385: \label{eq:expansion-V1-vs-dmlj}
386: \end{equation}
387: %
388: 
389: It would be computationally favorable to collect $d^{ml}_j(t)$
390: in a single vector $\textbf{\textit{z}}(t)=\{z_n(t)\}$ by defining
391: a map $(m,l,j)\rightarrow n$. In practice the infinite sums in
392: (\ref{eq:expansion-f1}) are truncated and approximated by finite
393: sums so that $-l_{\rm max}\le l \le l_{\rm max}$,
394: $-m_{\rm max}\le m\le m_{\rm max}$ and $0\le j\le
395: j_{\rm max}$. For $1\le n\le n_{\rm max}$,
396: a simple map between indices will be
397: %
398: \begin{eqnarray}
399: n &=& \left ( m+m_{\rm max}\right )
400: \left ( 2l_{\rm max}+1\right )(j_{\rm max}+1)
401: \nonumber \\
402: &{}& + \left (l+l_{\rm max}\right )(j_{\rm max}+1)+j+1,
403: \label{eq:map-mlj-to-i} \\
404: n_{\rm max} &=& \left ( 2m_{\rm max}+1\right )
405:       \left ( 2l_{\rm max}+1\right )
406:       (j_{\rm max}+1). \label{eq:maximum-n}
407: \end{eqnarray}
408: %
409: One can now use (\ref{eq:expansion-f1}) and
410: (\ref{eq:expansion-V1-vs-dmlj}) in (\ref{eq:CBE-perturbed}) and
411: apply the Petrov-Galerkin method to construct the weighted residual
412: form of the CBE. That is to multiply (\ref{eq:CBE-perturbed}) by
413: some weighting functions $W^{ml}_j(\Theta,\textbf{\textit{J}})$ and to
414: integrate the identity over the action-angle space. The outcome is
415: the following system of nonlinear ODEs 
416: %
417: \begin{equation}
418: {\rm i} { d z_p\over  d t}\!=\! \sum_{q=1}^{n_{\rm max}}
419: \! A_{pq} z_q \!+ \!\! \sum_{q,r=1}^{n_{\rm max}}
420: \! B_{pqr} z_q z_r,~~p=1,2,\cdots,n_{\rm max},
421: \label{eq:nonlinear-ODE}
422: \end{equation}
423: %
424: for the amplitude functions $z_n(t) \equiv d^{ml}_j(t)$. The
425: elements of $A_{pq}$ and $B_{pqr}$ have been determined in Appendix
426: A. Each equation in (\ref{eq:nonlinear-ODE}) is the projection of
427: the CBE on a subspace spanned by a weighting function. 
428: Therefore, the left
429: hand side of (\ref{eq:nonlinear-ODE}) is the projection of $\partial
430: f_1/\partial t$, the summation over first order terms is the
431: projection of $-\left [f_1,{\cal H}_0 \right ]- \left [ f_0,{\cal
432: H}_1 \right ]$, and the second order terms are the projections of
433: $-\left [f_1,{\cal H}_1 \right ]$. The second order terms of
434: amplitude functions, characterized by $B_{pqr}$, show the
435: interaction of modes in both the radial and azimuthal directions.  
436: 
437: Distribution of angular momentum between different Fourier components
438: provides useful information of the disk dynamics. I compute the rate
439: of change of the total angular momentum ${\cal L}$ using
440: (see Appendix B in JH)
441: %
442: \begin{equation}
443: { d {\cal L}\over  d t}=-\frac {1}{4}
444: \int\int \left ( f_1+\overline{f}_1 \right )
445: {\partial \over \partial \theta_{\phi} }
446: \left ( V_1+\overline{V}_1 \right ) d \textbf{\textit{J}} d \Theta,
447: \label{eq:rate-change-L-one}
448: \end{equation}
449: %
450: where a bar denotes complex conjugate. Substituting
451: (\ref{eq:expansion-f1}) and (\ref{eq:expansion-V1-vs-dmlj})
452: in (\ref{eq:rate-change-L-one}) and evaluating the integrals,
453: yield
454: %
455: \begin{eqnarray}
456: { d {\cal L}\over  d t} &=& {\rm i}\pi^2 \!\!\!
457: \sum_{m,l=-\infty}^{\infty}\sum_{j,p=0}^{\infty}
458: \biggl \{  m\left [ a^{(-m)}_{j}(t)
459: +\overline{a^{m}_{j}}(t) \right ] \Lambda^{ml}_{jp}
460: d^{ml}_{p}(t) \nonumber \\
461: &{}& - m\left [ a^{m}_{j}(t)
462: +\overline{a^{(-m)}_{j}}(t) \right ] \Lambda^{ml}_{jp}
463: \overline{d^{ml}_{p}}(t) \biggr \}. \label{eq:rate-change-L-two}
464: \end{eqnarray}
465: %
466: Define $a^m_j(t)=u^m_j(t)+{\rm i}v^m_j(t)$ with $u^m_j(t)$ and
467: $v^m_j(t)$ being real functions of time. According to identity
468: (\ref{eq:a-versus-d}), one may further simplify equation
469: (\ref{eq:rate-change-L-two}) to
470: %
471: \begin{eqnarray}
472: { d {\cal L}\over  d t} &=&
473: \sum_{m=-\infty}^{\infty} L_m(t),
474: \label{eq:rate-change-L-three} \\
475: L_m(t) &=& -\frac {m}{2}\sum_{j=0}^{\infty}
476: D_j(m) \nonumber \\
477: &{}& \times \left [ u^m_j(t)v^{-m}_j(t)+u^{-m}_j(t)v^m_j(t) \right ].
478: \label{eq:rate-change-L-four}
479: \end{eqnarray}
480: %
481: The share of the $m$th mode from $ d {\cal L}/ d t$ is thus
482: determined by $L_m(t)$. As one could anticipate for an isolated
483: stellar disk, $ d {\cal L}/ d t$ vanishes and the total
484: angular momentum remains constant because the terms $L_m(t)$ and
485: $L_{-m}(t)$ cancel each other in (\ref{eq:rate-change-L-three}),
486: and $L_0(t)$ is annulled by the factor $m$
487: in (\ref{eq:rate-change-L-four}).
488: 
489: 
490: \subsection{Trial and Weighting Functions}
491: \label{sec::trial-and-test-fncs}
492: 
493: Choosing the trial functions $\Phi^{ml}_j(\textbf{\textit{J}})$ is
494: the most delicate step in the reduction of the CBE to a system of ODEs.
495: One possible way is to set $\partial f_1/\partial t=0$ in
496: (\ref{eq:CBE-perturbed}) and solve the first order equation
497: %
498: \begin{equation}
499: \left [f_1,{\cal H}_0 \right ]+\left [f_0,{\cal H}_1 \right ]=0,
500: \label{eq:CBE-linear-equilibrium}
501: \end{equation}
502: %
503: for $f_1$. Substituting (\ref{eq:expansion-f1}) and
504: (\ref{eq:expansion-V1}) in (\ref{eq:CBE-linear-equilibrium}) gives
505: %
506: \begin{eqnarray}
507: d^{ml}_j\Phi^{ml}_j(\textbf{\textit{J}}) &=& b^{ml}_j
508: \varrho^{ml}_{0}(\textbf{\textit{J}}) \Psi^{ml}_j(\textbf{\textit{J}}),
509: \label{eq:trial-functions-one} \\
510: \varrho^{ml}_{0}(\textbf{\textit{J}}) &=& {l\frac {\partial f_0}{\partial
511: J_R}+ m\frac {\partial f_0}{\partial J_\phi} \over
512: l\Omega_R+m\Omega_\phi },
513: \end{eqnarray}
514: %
515: where
516: %
517: \begin{equation}
518: \Omega_R(\textbf{\textit{J}})=\frac {\partial {\cal H}_0}{\partial J_R},~~
519: \Omega_\phi(\textbf{\textit{J}})=\frac {\partial {\cal H}_0}{\partial J_\phi}.
520: \end{equation}
521: %
522: Equation (\ref{eq:trial-functions-one}) suggests to choose
523: %
524: \begin{equation}
525: \Phi^{ml}_j(\textbf{\textit{J}})=\varrho^{ml}_{0}(\textbf{\textit{J}})
526: \Psi^{ml}_j(\textbf{\textit{J}}),\label{eq:trial-functions-two}
527: \end{equation}
528: %
529: as the trial functions (in the space of actions) for an unsteady
530: $f_1(\Theta,\textbf{\textit{J}},t)$. These functions have integrable
531: singularities for resonant orbits with $l\Omega_R+m\Omega_{\phi}=0$.
532: For unidirectional disks with only prograde orbits, they also
533: include a term with the Dirac delta function $\delta(J_\phi)$
534: (see JH). One should therefore avoid the partial derivatives of
535: $\Phi^{ml}_j$ with respect to the actions by evaluating the weighted
536: residual form of $[f_1,{\cal H}_1]$ through integration by parts
537: (Appendix A).
538:  
539: For deriving the relation between $a^m_j$ and $d^{ml}_{j}$
540: in (\ref{eq:a-versus-d}), the fundamental equation
541: (\ref{eq:fundamental-equation}) was multiplied by the complex
542: conjugates of the basis functions used in the expansion of
543: $V_1(R,\phi,t)$. One may follow a similar approach for
544: obtaining the weighted residual form of the CBE and set
545: %
546: \begin{equation}
547: W^{ml}_j(\Theta,\textbf{\textit{J}})=\Psi^{ml}_j(\textbf{\textit{J}})
548: e^{-{\rm i}(m\theta_\phi+l\theta_R)},\label{eq:weight-functions}
549: \end{equation}
550: %
551: which are the complex conjugates of the basis functions used in the 
552: expansion of $V_1(\Theta,\textbf{\textit{J}},t)$ in the action-angle 
553: space. The trial and weighting functions introduced as above, are not 
554: orthogonal but they result in a simple form for the linear part of 
555: the reduced CBE as I explain in \S\ref{sec::linear-theory}. 
556:  
557: 
558: 
559: \section{LINEAR THEORY}
560: \label{sec::linear-theory}
561: 
562: In a first order perturbation analysis, the second order terms of
563: the amplitude functions are ignored. The evolution of modes is
564: then governed by the linear parts of (\ref{eq:ODE-for-z-one}) as
565: %
566: \begin{equation}
567: {\rm i} \textbf{\textit{M}}\cdot { d \over  d t}
568: \textbf{\textit{z}}(t) =
569: \textbf{\textit{C}}\cdot \textbf{\textit{z}}(t).
570: \label{eq:ODE-for-z-linear}
571: \end{equation}
572: %
573: A general solution of (\ref{eq:ODE-for-z-linear}) has the
574: form $\textbf{\textit{z}}(t)=e^{-{\rm i}\omega t}\textbf{\textit{z}}_0$,
575: which leads to the following linear eigenvalue problem
576: %
577: \begin{equation}
578: \textbf{\textit{C}}(m)\cdot \textbf{\textit{z}}_0=
579: \omega \textbf{\textit{M}}(m)\cdot \textbf{\textit{z}}_0,
580: \label{eq:eigenvalue-problem-one}
581: \end{equation}
582: %
583: for a prescribed azimuthal wavenumber $m$.
584: Operating $\textbf{\textit{M}}^{-1}$ on (\ref{eq:eigenvalue-problem-one})
585: yields
586: %
587: \begin{equation}
588: \textbf{\textit{A}}(m)\cdot \textbf{\textit{z}}_0=
589: \omega \textbf{\textit{z}}_0,
590: \label{eq:eigenvalue-problem-two}
591: \end{equation}
592: %
593: where $\textbf{\textit{A}}$ is a general non-symmetric matrix. A reduction
594: to Hessenberg form followed by the QR algorithm \citep{Press01}
595: gives all real and complex eigenvalues. Real eigenvalues correspond
596: to van Kampen modes and complex eigenvalues, which occur in conjugate
597: pairs, give growing/damping modes. I utilize the method of singular value
598: decomposition for finding the eigenvectors and perform the decomposition
599: $\textbf{\textit{A}}-\omega \textbf{\textit{I}}=\textbf{\textit{U}}\cdot
600: \textbf{\textit{S}}\cdot \textbf{\textit{V}}^T$
601: where $\textbf{\textit{I}}$ is the identity matrix and the diagonal matrix
602: $\textbf{\textit{S}}$ is composed of the singular values $S_j$
603: ($j=1,2,\cdots,n_{\rm max}$). The column of $\textbf{\textit{V}}$ that
604: corresponds to the smallest $S_j$ is the eigenvector associated
605: with $\omega$.
606: 
607: Calculation of $\textbf{\textit{C}}$ and $\textbf{\textit{M}}$
608: involves evaluation of some definite integrals in the action space.
609: There will be two types of such integrals (instead of three) if one
610: uses the trial functions defined in (\ref{eq:trial-functions-two}).
611: Let me introduce the auxiliary integral
612: %
613: \begin{equation}
614: {\cal I}^{ml}_{jk}=
615: \int  d \textbf{\textit{J}} \left ( l{\partial f_0\over \partial J_R}+
616: m{\partial f_0\over \partial J_\phi} \right )
617: \Psi^{ml}_{j}(\textbf{\textit{J}})\Psi^{ml}_{k}(\textbf{\textit{J}}),
618: \end{equation}
619: %
620: and apply the trial functions $\Phi^{ml}_{j}=\varrho^{ml}_0\Psi^{ml}_{j}$
621: in (\ref{eq:A-tensor}). The elements of $\textbf{\textit{M}}$ and
622: $\textbf{\textit{C}}$ are thus computed from
623: %
624: \begin{eqnarray}
625: M_{pq} &=& \delta_{l,l'}\Lambda^{ml}_{jj'}, \label{eq:bar-M-matrix} \\
626: C_{pq} &=& \delta_{l,l'}{\cal I}^{ml}_{jj'}-
627: \sum_{k=0}^{j_{\rm max}} \left [ {4\pi^2\over D_{k}(m)}\right ]
628: {\cal I}^{ml}_{jk}\Lambda^{ml'}_{kj'}. \label{eq:bar-A-matrix}
629: \end{eqnarray}
630: %
631: Both $\Lambda^{ml}_{jk}$ and ${\cal I}^{ml}_{jk}$ consist of
632: boundary integrals when the unperturbed stellar disk is
633: unidirectional with the DF
634: $f_0(\textbf{\textit{J}})=H(J_\phi)f^P_0(\textbf{\textit{J}})$.
635: Here $H$ is the Heaviside function. The boundary terms are
636: %
637: \begin{eqnarray}
638: \tilde \Lambda^{ml}_{jk} &=& \int_{0}^{\infty} d J_R\left [
639: {mf^P_0(\textbf{\textit{J}})
640: \Psi^{ml}_{j}(\textbf{\textit{J}})\Psi^{ml}_{k}(\textbf{\textit{J}}) \over
641: l\Omega_R(\textbf{\textit{J}})+m\Omega_{\phi}(\textbf{\textit{J}}) }
642: \right ]_{J_{\phi}=0}, \\
643: %
644: {\tilde {\cal I}}^{ml}_{jk} &=& \int_{0}^{\infty} d J_R\left [
645: mf^P_0(\textbf{\textit{J}})
646: \Psi^{ml}_{j}(\textbf{\textit{J}})\Psi^{ml}_{k}(\textbf{\textit{J}}) \right
647: ]_{J_{\phi}=0}.
648: \end{eqnarray}
649: %
650: 
651: Dynamics of modes with different azimuthal wavenumbers are decoupled
652: in the linear regime and the matrix $\textbf{\textit{A}}$ is an odd
653: function of the wavenumber $m$. i.e.,
654: $\textbf{\textit{A}}(-m)=-\textbf{\textit{A}}(m)$. An
655: immediate result of this property is $a^{-m}_j(t)=\overline {a^m_j}(t)$.
656: Consequently, $L_m(t)$ becomes equal to zero for all $|m|\ge 0$ and
657: each mode individually conserves the total angular momentum.
658: 
659: The present theory has three major advantages over Kalnajs's formulation.
660: Firstly, all eigenmodes relevant to a prescribed azimuthal wavenumber
661: are obtained at once with classical linear algebraic algorithms. This makes
662: it possible to explore and classify all families of growing modes beside
663: pure oscillatory van Kampen modes. Secondly, the constituting integrals
664: of the elements of $M_{pq}$, $C_{pq}$ and $K_{pqr}$
665: (Appendix \ref{app::Petrov-Galerkin}) are regular
666: at exact resonances when the condition $l\Omega_R+m\Omega_\phi-\omega=0$
667: holds. Finally, nonlinear interaction of modes, and the mass and angular
668: momentum exchange between them, can be readily monitored by integrating
669: the system of nonlinear ODEs given in (\ref{eq:nonlinear-ODE}). In the
670: proceeding section I will be concerned with the calculation and
671: classification of modes in the linear regime.
672: 
673: %
674: \begin{figure*}
675: \plottwo{f1a.eps}{f1b.eps}
676: \plottwo{f1c.eps}{f1d.eps}
677: \plottwo{f1e.eps}{f1f.eps}
678: \caption{Eigenfrequency spectra of a cored exponential disk
679: with $v_0=1$, $R_C=1$ and $(N,\lambda,\alpha)=(6,1,0.42)$.
680: Eigenfrequencies have been displayed for the azimuthal wavenumbers
681: $0\le m \le 5$. The results correspond to $l_{\rm max}=10$ and
682: $j_{\rm max}=15$ in the series expansion of the perturbed
683: distribution function. \label{pic:full-spectra}}
684: \end{figure*}
685: %
686: 
687: %
688: \begin{figure*}
689: \plottwo{f2a.eps}{f2b.eps}
690:  \caption{({\em a}) Zoomed
691: eigenfrequency spectra (filled squares) of a cored exponential disk 
692: with $(N,\lambda,\alpha)=(6,1,0.42)$ for the azimuthal wavenumber $m=2$.
693: The (isolated) bar mode has been labeled B1. The most prominent
694: growing modes belong to a discrete family that bifurcates from a van
695: Kampen mode. The members of this family, labeled as S1, S2,
696: S3,$\cdots$, have spiral patterns. Circles show the eigenfrequencies
697: of the fundamental and secondary modes calculated using Kalnajs's
698: (1977) method (see Table 4 in JH). ({\em b}) The eigenfrequency loci
699: of the same model of panel {\em a} as $\alpha$ is increased
700: continuously from $0.2$ to $0.42$. A few sample eigenfrequencies
701: have been displayed on each locus.
702:  \label{pic:spectrum-m2-unstable}} 
703: \end{figure*}
704: %
705: %
706: \begin{figure*}
707: \plotone{f3a.eps}
708: \plotone{f3b.eps}
709: \caption{Mode shapes of the cored exponential disk for $R_C=1$,
710: $v_0=1$ and $(N,\lambda,\alpha)=(6,1,0.42)$. Panels have been
711: labeled by the corresponding mode name. The contour plots show the
712: positive part of $\Sigma_1(R,\phi,0)$. The contour levels range
713: from 10$\%$ to 90$\%$ of the maximum of $\Sigma_1(R,\phi,0)$ with
714: increments of 10$\%$. The panel below each mode shape shows the
715: amplitude of wave patterns as defined in equation (\ref{eq:mode-shape}).
716: \label{pic:mode-shape-m2}}
717: \end{figure*}
718: %
719: 
720: \section{MODES OF THE CORED EXPONENTIAL DISK}
721: \label{sec:modes-exp-disk}
722: 
723: JH calculated barred and spiral modes of certain stellar
724: disks for the wavenumber $m=2$. Among the models
725: studied in JH, the cored exponential disk with the surface
726: density profile
727: %
728: \begin{equation}
729: \Sigma_D(R)=\Sigma_s \exp \left ( -\lambda
730: \sqrt{1+R^2/R_C^2}\right ),~~\lambda=\frac{R_C}{R_D},
731: \label{eq:density-exp-disk}
732: \end{equation}
733: %
734: and embedded in the field of the soft-centered logarithmic
735: potential
736: %
737: \begin{equation}
738: V_0(R)=v_0^2\ln \sqrt{1+R^2/R_C^2},
739: \label{eq:potential-cored-logarithmic}
740: \end{equation}
741: %
742: is a viable model that resembles most features of realistic spirals.
743: Here $R_C$ is the core radius, $R_D$ is the length scale of
744: the exponential decay, and $\Sigma_s$ is a density scaling factor.
745: The velocity of circular orbits in this model rises from zero at
746: the galactic center and approaches to the constant value $v_0$
747: in outer regions where the light profile falls off exponentially.
748: \citet{JHb} have derived the gravitational potential corresponding
749: to $\Sigma_D(R)$. I denote this potential by $V_D(R)$. The gradient
750: %
751: \begin{equation}
752: F_H=\frac{d}{dR}\left [V_0(R)-V_D(R)\right ],
753: \end{equation}
754: %
755: will give the gravitational force of a spherical dark matter component,
756: computed inside the galactic disk. The density profile of the dark
757: component, $\rho_H$, can then be determined using $F_H$. The positiveness
758: of $\rho_H$ imposes some restrictions on the physical values of $\lambda$
759: and $\alpha=G\Sigma_s R_D/v_0^2$ as Figure 5 in JH shows. For a given
760: $\lambda$, $\alpha$ cannot exceed a critical value $\alpha_{cr}$.
761: The parameter $\lambda$ determines the shape of the dark matter density 
762: profile. A model with $\lambda=1$ and $\alpha=\alpha_{cr}$ is maximal 
763: in the region where the rotation curve is rising. i.e., there is no
764: dark matter in that region. Models with $\lambda >1$ and
765: $\alpha=\alpha_{cr}$ are still maximal but only in the vicinity of
766: the center for $R<R_D$. In such models the rotational velocity of
767: stars due to dark matter ($v_H=\sqrt{R F_H}$) has a monotonically
768: rising profile. For $\lambda<1$, dark matter penetrates into the 
769: galactic center and its density profile becomes cuspy in the limit 
770: of $\lambda \rightarrow 0$. The role of the parameter $\alpha$ is 
771: to control the fraction of dark to luminous matter. Models with 
772: $\alpha \ll \alpha_{cr}$ are dominated by dark matter. 
773:  
774: JH introduced a family of equilibrium DFs that reproduces
775: $\Sigma_D(R)$ and depends on an integer constant $N$.
776: This parameter controls the population of near-circular
777: orbits and the disk temperature: the parameter $Q$ of \citet{T64}
778: decreases by increasing $N$. The DFs of JH have an isotropic part
779: that determines the fraction of radial orbits. That isotropic part,
780: which reconstructs the central density of the equilibrium state,
781: shrinks to central regions of the galaxy as $N$ increases.
782: 
783: I apply my new method to the cored exponential disks of JH and calculate
784: the spectrum of $\omega=\omega_R+{\rm i}\omega_I$. Subsequently, 
785: the eigenvector $\textbf{\textit{z}}_0$ is calculated 
786: from (\ref{eq:eigenvalue-problem-two}) and it is used 
787: in (\ref{eq:a-versus-d}) to compute
788: $a^m_j(t)=e^{-{\rm i}\omega t}a^m_j(0)$ and the perturbed density
789: %
790: \begin{equation}
791: \Sigma_1(R,\phi,t)= e^{\omega_I t} P_m(R) \cos \left [
792: m\phi \!-\! \omega_R t \!+\! \vartheta_m(R)\right ],
793: \label{eq:mode-shape}
794: \end{equation}
795: %
796: which is the real part of (\ref{eq:expansion-sigma1-config}).
797: $P_m(R)$ and $\vartheta_m(R)$ are the amplitude and phase functions
798: of an $m$-fold circumferential wave that travels with the angular
799: velocity $\omega_R/m$. The factor $e^{\omega_I t}$ shows the exponential
800: growth/decay of the wave amplitude. I normalize all length, velocity, 
801: and time variables to $R_C$, $v_0$ and $R_C/v_0$, respectively, and 
802: set $G=R_C=v_0=1$.  
803: 
804: I begin my case studies in \S\ref{sec:maximal-disk} with a near 
805: maximal disk of $(N,\lambda,\alpha)$=$(6,1,0.42)$ and compute its 
806: eigenfrequency spectra for the wavenumbers $0\le m\le 5$. I then 
807: classify unstable $m=2$ modes of this model and investigate their 
808: evolution as the parameter $\alpha$ is varied. 
809: In \S\ref{sec:variations-of-lambda} and \S\ref{sec:variation-of-N},
810: I study the behavior of unstable $m=2$ waves as the parameters 
811: $\lambda$ and $N$ are changed. The eigenfrequency spectrum of a 
812: model with an inner cutout is also computed and discussed 
813: in \S\ref{sec:cutout-models}. 
814: 
815:  
816: \subsection{A Near Maximal Disk}
817: \label{sec:maximal-disk}
818: 
819: I pick up the first model from Table 4 of JH with
820: $(N,\lambda,\alpha)=(6,1,0.42)$ and start solving the
821: eigensystem (\ref{eq:eigenvalue-problem-two})
822: with $(l_{\rm max},j_{\rm max})=(2,4)$ and increase these limits
823: until complex eigenfrequencies converge. For an error threshold
824: of $1\%$ the program terminates when
825: $(l_{\rm max},j_{\rm max})=(10,15)$, which gives a size of
826: $336\times 336$ for the matrix $\textbf{\textit{A}}$. In such a
827: circumstance, out of $336$ eigenfrequencies of $\textbf{\textit{A}}$
828: (for each wavenumber $m$), less than 15 pair have non-zero
829: growth rates ($\omega_I\not =0$). Further increasing of
830: $l_{\rm max}$ and $j_{\rm max}$ does not alter the number and
831: location of complex eigenfrequencies in the $\omega$-plane.
832: This shows that unstable modes do not constitute a continuous
833: family.
834: 
835: Figure \ref{pic:full-spectra} displays the eigenfrequency spectra
836: for the azimuthal wavenumbers $0\le m\le 5$. Eigenfrequencies on
837: the real axis are oscillatory van Kampen modes. Their calculation
838: requires evaluation of Cauchy's principal value \citep{V03}
839: if one uses Kalnajs's first order theory. In the present formalism,
840: van Kampen modes are found together with growing modes without any
841: special treatment. More van Kampen modes are obtainable
842: by increasing the truncation limit $l_{\rm max}$ of Fourier terms
843: in the $\theta_R$-direction. Toomre's $Q$ is marginally greater
844: than 1 for the model (Figure 7{\em b} in JH), and therefore,
845: one could expect that the disk is stable for $m=0$ excitations
846: (see top-left panel in Figure \ref{pic:full-spectra}).
847: The model is highly unstable for $m>0$ excitations although the
848: average growth rate of unstable modes decreases for larger
849: wavenumbers. It is evident that either unstable modes are isolated
850: or they are grouped in {\it discrete families}. 
851: Depending on the wavenumber, there may be one or more discrete 
852: families. The most prominent family bifurcates from van Kampen modes. 
853: Members of this family have spiral patterns with multiple peaks in 
854: their $P_m(R)$ functions. The (global) fastest growing mode belongs 
855: to the spectrum of $m=2$. That is the bar mode of a two-member 
856: unstable family. 
857: 
858: The length scale of Clutton-Brock functions has been set to $b=1.5$
859: for $m=2$ and $b=2$ for other wavenumbers. Changing this length
860: scale slightly displaces the eigenfrequencies although the spectrum
861: maintains its global pattern. Large values of $b$ lead to a better
862: computation accuracy of extensive modes (with smaller pattern
863: speeds), while compact bar modes show a rapid convergence for small
864: values of $b$. Moreover, the suitable value of $b$ differs from one
865: azimuthal wavenumber to another. Finding an optimum length scale
866: that gives the best results for all modes and wavenumbers is an open
867: problem yet to be investigated precisely. For the cored exponential 
868: disks with $0.5 \le R_C,R_D\le 2$, working in the range $1\le b\le 2.5$ 
869: gives reasonable results.
870: 
871: %
872: \begin{figure*}
873: \plottwo{f4a.eps}{f4b.eps} 
874: \caption{({\em a}) Eigenfrequency loci (solid lines) of cored 
875: exponential disks with $(N,\lambda)=(6,0.625)$ for $0.2\le \alpha \le 0.36$. 
876: Large squares show the eigenfrequency spectrum of a model 
877: with $\alpha=0.34$, and circles show the eigenfrequencies of the 
878: fundamental and secondary modes calculated using Kalnajs's method 
879: (see Table 4 in JH) for $\alpha=0.34$. ({\em b}) Same as panel {\em a} 
880: but for models with $(N,\lambda)=(6,2)$ and $0.25 \le \alpha \le 0.6$.
881: Some sample eigenfrequencies have been demonstrated on each locus.
882: \label{pic:spectrum-m2-vary-lambda}}
883: \end{figure*}
884: %       
885: 
886: For $m=2$, I have zoomed out and plotted in Figure
887: \ref{pic:spectrum-m2-unstable}{\em a} the portion of the spectrum
888: that contains growing modes. The first and second modes reported in
889: Table 4 of JH have been shown by circles in the same figure. The
890: most unstable mode (labeled as B1) is a compact, rapidly rotating
891: bar. The majority of unstable modes belong to a discrete spiral
892: family that bifurcates from a van Kampen mode with $\omega \approx
893: 0.43$. I have labeled these modes by S1,$\cdots$,S6. The number of
894: density peaks along the spiral arms is proportional to the integer
895: number in the mode name. Both B2 and G are double peaked spirals but
896: I have classified B2 as a bar mode, and collected it with B1 in a
897: two member family, for it takes a bar-like structure when it is 
898: stabilized by decreasing $\alpha$. I classify mode G as an isolated 
899: mode because it does not behave similar to either of S- or B-modes 
900: as the model parameters vary. There is another isolated mode in 
901: the spectrum, C1, which exhibits a spiral pattern. By decreasing
902: $\lambda$, mode C1 joins a new family of spiral modes, which are
903: accumulated near the galactic center (see
904: \S\ref{sec:variations-of-lambda}). 
905: 
906: Reducing $\alpha$ increases the abundance of dark matter and
907: according to \cite{T81} and JH the growth rate of modes should decrease.
908: My calculations show that by reducing $\alpha$, spiral modes are
909: affected sooner and more effective than the bar mode, and they join
910: to the stationary modes, one by one from the location of the
911: bifurcation point until the whole S-family disappears. This is
912: a generic scenario for all $\lambda \ge 1$ models regardless of the
913: disk temperature controlled by $N$. Solid lines in
914: Figure \ref{pic:spectrum-m2-unstable}{\em b} show the {\it eigenfrequency
915: loci} of a model with $(N,\lambda)=(6,1)$ as $\alpha$ increases from
916: $0.2$ to $0.42$. It is evident that mode B1 is destabilized through
917: a pitchfork bifurcation while the loci of modes B2 and C1, and S-modes
918: exhibit a tangent bifurcation. All modes except mode G are stable
919: for $\alpha <0.23$. Surprisingly, mode G resists against stabilization
920: even for very small values of $\alpha$. This indicates that 
921: mode G is not characterized by the fraction of dark to luminous 
922: matter. In \S\ref{sec:variations-of-lambda} and \S\ref{sec:variation-of-N}, 
923: I will show that this mode is highly sensitive to the variations of 
924: $\lambda$ and $N$. According to my computations 
925: (e.g., Figure \ref{pic:spectrum-m2-unstable}{\em b}), by increasing 
926: $\alpha$ all S-modes are born at the same bifurcation frequency 
927: $\omega_{S}\approx 0.43$, but mode B2 comes out from a van Kampen 
928: mode with $\omega_{B}\approx 0.83$. This result completely rules out
929: any skepticism that mode B2 is a member of S-family.  
930: 
931: % 
932: \begin{figure*}
933: \plotone{f5a.eps}
934: \plotone{f5b.eps}
935: \caption{Same as Figure \ref{pic:mode-shape-m2} but for a model
936: with $(\lambda,\alpha)=(0.625,0.34)$. The fastest growing spiral
937: mode, S1, has also been found by JH. Modes C1, C2 and C3 have
938: emerged due to dark matter presence at the galactic center.
939: They shrink to central regions and grow slower as their pattern
940: speed increases.
941: \label{pic:mode-shape-m2-SR34}}
942: \end{figure*}
943: %
944:  
945: Figure \ref{pic:mode-shape-m2} displays the wave patterns and
946: amplitude functions of modes B1, B2, G, S1, S2 and S3. It is seen
947: that the patterns of S-modes rotate slower and become more extensive 
948: as the mode number increases. Mode S6, which is at the bifurcation 
949: point of the spiral family, has the largest extent. Eight wave packets 
950: of this mode are distributed by a phase shift of 90 degrees along 
951: major spiral arms. Mode G has at most two density peaks on its major 
952: spiral arms but the magnitude of its second peak increases as the disk 
953: is cooled. Modes B1, B2, S1, S2 and S3 are, respectively, analogous to 
954: modes A, B, C, E and F of a Gaussian disk explored in \citet{T81}. 
955: There are three low-speed modes in Figure 11 of \citet{T81} that 
956: have not been labeled, but they are analogs of modes S4, S5 and S6. 
957: Mode G and Toomre's mode D also have some similarities but are of 
958: different origins (see \S\ref{sec:final-discussions}). None of them
959: can be stabilized only by increasing the fraction of dark matter. 
960: 
961: Figure \ref{pic:spectrum-m2-unstable}{\em a} shows that the 
962: fundamental mode obtained by JH coincides 
963: with mode B1. The wave pattern of mode B1 (displayed in Figure
964: \ref{pic:mode-shape-m2}) is identical to the mode shape computed
965: using Kalnajs's theory and demonstrated in Figure 8 of JH. JH found a
966: secondary mode which lies between modes B2 and G. That mode is also
967: a double-peaked spiral and it is not easy to identify its true
968: nature unless we investigate its evolution as the model parameters 
969: vary. By comparing Figure 10{\em a} of JH with 
970: Figure \ref{pic:spectrum-m2-unstable}{\em b}, one can see that both 
971: mode B2 and the secondary mode of JH are destabilized through a 
972: tangent bifurcation while mode G has a different nature. 
973: The bifurcation frequency of mode B2 that I find  
974: ($\omega_{B}\approx 0.83$) matches very well with the frequency 
975: of the stabilized secondary mode of JH (see Figure 10{\em a} in JH
976: but note that their vertical axis indicates $\Omega_p=\omega_R/2$). 
977: Therefore, I conclude that the secondary mode of JH is indeed 
978: mode B2 although it seems to be closer to mode G. The existing 
979: discrepancy is due to different length scale of Clutton-Brock 
980: functions that JH have used for finding the secondary mode. 
981: By adjusting $b$ one can improve the location of B2. However, 
982: this is an unnecessary attempt given the fact that mode B2 has 
983: already been identified, and the computation accuracy of other 
984: eigenfrequencies has an impressive level.  
985: 
986: \subsection{Variations of
987: $\lambda$} \label{sec:variations-of-lambda}
988: 
989: The parameter $\lambda$ controls the density profile of the dark
990: matter component, specifically near the galactic center. The
991: fraction of dark to luminous matter has its minimum value in
992: marginal models with $\alpha\approx \alpha_{cr}$. I choose a
993: marginal $\lambda <1$ model with
994: $(N,\lambda,\alpha)=(6,0.625,0.34)$, which has also been investigated 
995: by JH. Figure \ref{pic:spectrum-m2-vary-lambda}{\em a} shows the portion 
996: of the spectrum that contains complex eigenfrequencies of this model. 
997: The spectrum has been computed for $b=1.5$. 
998: Although $m=2$ bar and spiral modes survive 
999: in this model, their pattern speeds and growth rates drop considerably.
1000: Mode G has been wiped out of existence by dark matter penetration into 
1001: the center, and four unstable modes (C1, C2, C3 and C4) have emerged 
1002: that constitute a new family of spiral modes. They populate the central 
1003: regions of the disk in most of $\lambda <1$ models. Again, the location 
1004: of eigenfrequencies obtained by JH have been marked by circles. 
1005: The agreement between the results of JH and the present study is very 
1006: good and the variance is less than $2\%$.
1007: 
1008: The population of spiral modes is changed by varying $\alpha$, 
1009: and the eigenfrequencies of unstable modes are altered 
1010: significantly. Solid lines in Figure
1011: \ref{pic:spectrum-m2-vary-lambda}{\em a} show the loci of growing
1012: modes as $\alpha$ increases from $0.2$ to $\alpha_{cr}\approx 0.36$.
1013: Similar to the previous $\lambda=1$ model, S-modes and mode B1 are
1014: destabilized through tangent and pitchfork bifurcations,
1015: respectively. All C-modes are born by a pitchfork bifurcation 
1016: although some minor modes of the same nature come and go as $\alpha$
1017: varies. The loci of modes B1 and S1 (in Figure 
1018: \ref{pic:spectrum-m2-vary-lambda}{\em a}) are in harmony with the
1019: results of Kalnajs's method plotted in Figure 10{\em b} of JH. 
1020: It is noted that the locus of mode B1 steeply joins the real axis,
1021: well before stabilizing the S-modes. This is how slowly growing 
1022: spirals may dominate a stellar disk.   
1023: 
1024: There are no new families of growing modes in $\lambda>1$ models.
1025: Dark matter in these models induces a rising rotation curve on the
1026: disk stars (see Figure 6 in JH) and the population of S-modes declines. 
1027: The growth rate of mode B2 increases proportional to $\lambda$, but 
1028: that of mode G falls off although mode G is still robust against 
1029: the variations of $\alpha$. Mode G has its maximum growth rate in
1030: $\lambda=1$ models, which suggests that it must be a self-gravitating 
1031: response of the luminous matter that involves only the potential 
1032: of the disk, $V_D$. The function $P_2(R)$ of mode C1 loses 
1033: its minor peaks and becomes smoother as $\lambda$ increases.  
1034: Figure \ref{pic:spectrum-m2-vary-lambda}{\em b} shows the  
1035: eigenfrequency loci of models with $(N,\lambda)=(6,2)$ as 
1036: $\alpha$ increases from $0.25$ to $\alpha_{cr}\approx 0.6$. 
1037: The eigenfrequency loci of modes C1 and B2, and S-modes (as $\alpha$ 
1038: varies) are similar to $\lambda=1$ models, but the locus of mode B1 
1039: loses its steepness and stretches towards small pattern speeds in 
1040: an approximately linear form until it joins the real $\omega_R$-axis.
1041: The bifurcation frequency of mode B2 differs from S-modes and it is 
1042: larger. 
1043:  
1044: Figure \ref{pic:mode-shape-m2-SR34} shows the wave patterns of modes
1045: B1, B2, S1, C1, C2 and C3 for the model with 
1046: $(N,\lambda,\alpha)=(6,0.625,0.34)$. The (isolated) mode B1 is still 
1047: a single-peaked bar although its edge is more extensive as the flat
1048: part of its $P_2(R)$ plot indicates. Mode S1 is a triple-peaked
1049: spiral (as before) and mode B2 is being stabilized ($\omega_{\rm
1050: B2}=0.775+0.007{\rm i}$). It is seen that mode B2 has a bar-like
1051: structure, which justifies its classification as the secondary bar
1052: mode. The pattern of S1 and its $P_2(R)$ plot can be compared with
1053: Figure 9 in JH. The agreement is quite satisfactory. There is a
1054: remarkable difference between the patterns of C- and S-modes
1055: although both families have spiral structures. In contrast to
1056: S-modes that become more extensive as their growth rate decays,
1057: C-modes are shrunk to central regions because their pattern speed
1058: increases. C-modes are also a bifurcating family, but their
1059: bifurcation point lies on large pattern speeds associated with the
1060: azimuthal frequency ($\Omega_\phi$) of central stars.  
1061:  
1062: The parameters $\alpha$ and $\lambda$ are essentially controlling 
1063: the fraction and density profile of dark matter component, respectively. 
1064: However, the radial velocity dispersion $\sigma_R$ of the equilibrium
1065: state is also playing an important role in the perturbed dynamics. 
1066: $\sigma_R$ is an indicator of the initial temperature of the disk. 
1067: \citet{ER98b} had already pointed out that the pitch angle of spiral 
1068: patterns decreases when $\sigma_R\rightarrow 0$ (see their Figure 7). 
1069: Apart from this morphological implication, can the variation in the 
1070: disk temperature affect the modal content? In the following subsection, 
1071: I trace the evolution of growing waves by changing the disk temperature 
1072: and show that the population of S-modes is larger in rotationally 
1073: supported, cold disks.   
1074: 
1075: 
1076: %
1077: \begin{figure*}
1078: \plottwo{f6a.eps}{f6b.eps}
1079:  \plottwo{f6c.eps}{f6d.eps}
1080:  \caption{The evolution of the eigenfrequency spectra as the
1081: disk temperature drops from left ($N=4$) to right ($N=8$) panels.
1082: Top and bottom panels correspond to $(\lambda,\alpha)=(1,0.42)$
1083: and $(\lambda,\alpha)=(0.625,0.34)$, respectively.
1084: \label{pic:spectrum-vary-N}}
1085: \end{figure*}
1086: % 
1087: %
1088: \begin{figure}
1089: \plotone{f7.eps}
1090:  \caption{Eigenfrequency spectrum
1091: of a cutout model with $L_0=0.1$ and $(N,\lambda,\alpha)=(6,1,0.42)$. 
1092: Circles show the eigenfrequencies found by JH using Kalnajs's theory. 
1093: \label{pic:spectrum-cutout}}
1094: \end{figure} 
1095: % 
1096: 
1097: \subsection{Variations of the Disk Temperature}
1098:  \label{sec:variation-of-N}
1099: 
1100: The parameter $N$ of the DFs of JH controls the disk temperature by
1101: adjusting the size of the isotropic core and the population of near
1102: circular orbits. As $N$ increases, the streaming velocity $\langle
1103: v_{\phi}\rangle$ approaches the rotational velocity of circular
1104: orbits and the stellar disk is cooled. Figure \ref{pic:spectrum-vary-N}
1105: displays the eigenfrequencies of previous $(\lambda,\alpha)=(1,0.42)$
1106: and $(\lambda,\alpha)=(0.625,0.34)$ models for $N=4$ and $N=8$. The
1107: spectra for the intermediate value of $N=6$ have already been shown
1108: in Figures \ref{pic:spectrum-m2-unstable}{\em a} and
1109: \ref{pic:spectrum-m2-vary-lambda}{\em a}.
1110: 
1111: Increasing $N$ gives birth to more S-modes while the bifurcation
1112: point of the family is preserved. As a new member is born at the
1113: bifurcation point, other members including mode S1, are pushed away
1114: from the real axis on a curved path. This behavior is observed in
1115: both models but the branch of S-family in the model with
1116: $(\lambda,\alpha)=(0.625,0.34)$ stays closer to the real axis than
1117: the other model. The growth rates of C-modes increase remarkably as
1118: the disk is cooled. Despite mode B1 which rotates and grows faster
1119: in cold disks, mode B2 grows faster in warmer disks. Variation
1120: in the disk temperature changes the eigenfrequency of mode G more
1121: effective than what $\alpha$ could, but nothing is more influential
1122: than the role of $\lambda$.  
1123: 
1124: Another consequence of cooling the stellar disk is that $m=0$ waves
1125: are no longer stable. The parameter $Q$ of Toomre was marginally
1126: greater than unity for $N=6$ models. For $N=8$, I find $Q<1$
1127: over an annular region because the plot of $Q$ versus $R$ exhibits a
1128: minimum at some finite radius (e.g., Figure 7 of JH). For instance, 
1129: I find three growing
1130: $m=0$ modes for the model $(N,\lambda,\alpha)=(8,1,0.42)$. They
1131: correspond to pure complex eigenfrequencies $\omega_1=0.621{\rm i}$,
1132: $\omega_2=0.494{\rm i}$ and $\omega_3=0.238{\rm i}$. Mode shapes
1133: have (obviously) ringed structures but the number of rings, which is
1134: identical to the number of peaks of $P_0(R)$, depends on the growth
1135: rate. The modes associated with $\omega_1$, $\omega_2$ and
1136: $\omega_3$ have three, four and five rings, respectively. Ring modes
1137: are very sensitive to the variations of model parameters and they
1138: are suppressed by decreasing $\lambda$ and $\alpha$. 
1139: 
1140: 
1141: \subsection{The Effect of an Inner Cutout}
1142: \label{sec:cutout-models}
1143: 
1144: In order to simulate an immobile bulge, which does not respond to
1145: density perturbations, JH utilized an inner cutout function of the
1146: form
1147: %
1148: \begin{equation}
1149: H_{\rm cut}=1-e^{-\left(J_\phi/L_0\right)^2},
1150: \end{equation}
1151: %
1152: where $L_0$ is an angular momentum scale. Multiplying $H_{\rm cut}$
1153: by the self-consistent DF of the equilibrium state, prohibits the
1154: stars with $J_{\phi}< L_0$ from participating in the perturbed
1155: dynamics. Consequently, incoming waves are reflected at some finite
1156: radius and the innermost wave packets of multiple-peaked modes are
1157: diminished. My calculations show that all S-modes survive in cutout 
1158: models, mode G disappears, and the growth rate of mode B2 increases.
1159: The pattern speed of mode B1 is boosted so that the corotation
1160: resonance is destroyed, but its growth rate drops drastically.
1161: Figure \ref{pic:spectrum-cutout} shows the eigenfrequency spectrum
1162: of a model with $L_0=0.1$ and $(N,\lambda,\alpha)=(6,1,0.42)$. 
1163: Circles show the eigenfrequencies found by JH. Again, the agreement
1164: between the results of JH and the present work is very good. The
1165: reason that I have identified mode B2 as the second member of
1166: B-family, and not the most unstable S-mode, is that its $P_2(R)$
1167: function has an evolved double-peaked structure (see Figure 11 in JH)
1168: and its locus versus $\alpha$ does not emerge from the same bifurcation 
1169: frequency of S-modes. Disappearance of mode G in cutout models confirms 
1170: my earlier note that it is a self-gravitating mode. 
1171: 
1172: 
1173: \section{DISCUSSIONS}
1174: \label{sec:final-discussions}
1175: 
1176: There are similarities between mode G of this study and Toomre's (1981) 
1177: mode D. Both of these modes resist against stabilization by increasing 
1178: the fraction of dark to luminous matter and they have at most double 
1179: peaks on their spiral arms. Nonetheless, these modes are not the same 
1180: because mode G is amplified through a feedback from the galactic center 
1181: but Toomre's mode D has been identified as an edge mode. A question 
1182: remains to be answered: why Toomre (1981) did not detect mode G and I 
1183: do not find an edge mode? The most convincing explanation is that to 
1184: excite a self-gravitating wave inside the core of the stellar component, 
1185: the governing potential in that region should mainly come from the 
1186: self-gravity of stars. This requirement is fulfilled in my $\lambda\ge 1$ 
1187: models for $R<R_D$. However, the completely flat rotation curve imposed 
1188: by Toomre (1981) nowhere follows the rotational velocity induced by the 
1189: self-gravity of stars and it prohibits the Gaussian disk from developing 
1190: a G-like mode. On the other hand, I don't find an edge mode because the 
1191: density profile of the cored exponential disk does not decay as steep as 
1192: the Gaussian disk to create an outer boundary at some finite radius for 
1193: reflecting the outgoing waves.    
1194: 
1195: Similar to the first order analysis of \S\ref{sec::linear-theory},
1196: Polyachenko's (2005) approach results in the full spectrum of
1197: eigenfrequencies for a given azimuthal wavenumber. There are some
1198: differences between his method and the present formulation.
1199: Polyachenko directly uses Poisson's integral to establish a
1200: point-wise relation in the action space between the Fourier
1201: components of the perturbed DF and its self-consistent
1202: potential. Combination of equations (5) and (9)
1203: in his paper is analogous to equation (\ref{eq:expansion-V1-vs-dmlj})
1204: in this paper. The main departure of the two theories is in the way
1205: that the linearized CBE is treated. Polyachenko forces a point-wise
1206: fulfillment of the CBE in the action space while the present method
1207: works with a weighted residual form of the CBE.
1208: 
1209: A point-wise formulation poses a challenge for the numerical
1210: calculation of the eigenvalues and their conjugate eigenvectors.
1211: According to the bar charts of JH, at least ten Fourier
1212: components ($-3 \le l \le 6$) are needed in the $\theta_R$-direction
1213: to assure a credible convergence of $f_1$ in a typical soft-centered
1214: galaxy model. Therefore, if one chooses a grid of $n_a\times n_a$ in the
1215: action space, Polyachenko's eigenvector $\textbf{\textit{F}}$ will have a
1216: dimension of $10\times n_a\times n_a$. Therefore, for a {\it very coarse}
1217: grid with $n_a=21$ that Polyachenko uses, the unknown eigenvector
1218: will have a dimension of 4410. This number must be compared with
1219: the dimension of $\textbf{\textit{z}}_0$ in equation
1220: (\ref{eq:eigenvalue-problem-two}). That is indeed $n_{\rm max}=336$
1221: for the most accurate calculations carried out by setting
1222: $(l_{\rm max},j_{\rm max})=(10,15)$ which means that $21$
1223: Fourier components in the $\theta_R$-direction and $16$
1224: expansion terms in the $R$-direction have been taken into account.
1225: Noting that the definite integrals ${\cal I}^{ml}_{jk}$ and
1226: $\Lambda^{ml}_{jk}$ are independently evaluated over the action
1227: space with any desired accuracy, the present theory proves to
1228: be more efficient for eigenmode calculation (in the linear regime)
1229: than other existing alternatives.
1230: 
1231: The agreement between the results of this work and those of JH, who
1232: have used Kalnajs's method, is impressive. There is only a
1233: discrepancy in the results for a double-peaked spiral mode of
1234: $\lambda=1$ models. In fact, these models have two double-peaked
1235: modes, modes B2 and G, and JH find mode B2. The origin of
1236: discrepancies was attributed to the length scale of Clutton-Brock
1237: functions, $b$, which is a fixed number for the whole spectrum of 
1238: a given azimuthal wavenumber. Provided that JH optimized $b$ for each
1239: growing mode that they calculated (see also \S\ref{sec:maximal-disk}), 
1240: some minor deviations from the results of this paper are reasonable. 
1241: In most cases the algorithm used by JH converges to mode B1 and the 
1242: fastest rotating S-mode. They capture mode B2 only if
1243: its growth rate is large enough. Other modes remain unexplored
1244: because Newton's method needs an initial guess of $\omega$,
1245: which has a little chance to be in the basin of attraction of the
1246: other members of S-family. The separation of eigenfrequencies near
1247: the bifurcation point of S-modes is very small and one could
1248: anticipate complex boundaries for the basins of attraction of these
1249: eigenfrequencies. Thus, there is no guarantee that successive
1250: Newton's iterations keep an estimated eigenfrequency on the same
1251: basin that it was initially. Nevertheless, in Kalnajs's formulation,
1252: a systematic search for all growing modes is possible by introducing
1253: the mathematical eigenvalue \citep{Z76,ER98b} and investigating its
1254: loci as the pattern speed and growth rate vary.
1255: 
1256: 
1257:    
1258: \section{CONCLUSIONS}
1259: \label{sec:conclusions}
1260: 
1261: After three decades of Kalnajs's (1977) publication, it was not
1262: known exactly whether growing modes of stellar systems appear as
1263: distinct roots in the eigenfrequency space or they belong to
1264: continuous families as van Kampen modes do. In this paper, I
1265: attempted to answer this question using the Galerkin projection
1266: of the CBE and unveiled the full eigenfrequency spectrum of a
1267: stellar disk. I showed that similar to gaseous disks \citep{AJ06},
1268: majority of growing modes emerge as {\it discrete families} 
1269: through a bifurcation from stationary modes. There are some 
1270: exceptions for this rule, the most important of which are 
1271: the isolated bar and G modes.  
1272: 
1273: The model that I used to test my method allows for dark matter
1274: presence as a spherical component, whose potential inside the
1275: galactic disk contributes to the rotational velocity of stars.
1276: By varying the parameters of the model, and investigating the
1277: eigenfrequencies and their associated mode shapes, I showed that
1278: it is not the fraction of dark to luminous matter that controls
1279: the variety of growing modes. What determines that variety is
1280: indeed the shape of the dark matter density profile controlled
1281: by the parameter $\lambda=R_C/R_D$. My survey in the parameter
1282: space revealed that the concentration of dark matter in the 
1283: galactic center ($\lambda <1$) destroys mode G and weakens the 
1284: growth of B-modes substantially. Emergence of spiral C-modes 
1285: that accumulate near the galactic center is another remarkable 
1286: consequence of dark matter presence in central regions of a 
1287: cored stellar disk.  
1288: 
1289: Although the solution of the Galerkin system showed a credible
1290: convergence of the series expansions, the existence of strong
1291: solutions for the CBE, in its full nonlinear form, is still an
1292: open problem. It has been known for years that van Kampen modes
1293: make a complete set \citep{C59}, and therefore, they may be used
1294: for a series representation of stationary oscillations. But there is
1295: not a mathematical proof for the completeness of the discrete families
1296: of growing modes. In other words, whether an observed galaxy can be
1297: assembled using the modes of a linear eigensystem, requires further
1298: analysis.
1299: 
1300: In the second part of this study, I will investigate the mechanisms
1301: of wave interactions in the nonlinear regime and will probe the mass
1302: and angular momentum transfer between waves of different Fourier
1303: numbers.
1304: 
1305: 
1306: \acknowledgments I am indebted to Chris Hunter for his instructive
1307: and valuable comments since the beginning of this work. I also thank
1308: the referee for helpful suggestions that improved the presentation
1309: of the results. This work was partially supported by the Research 
1310: Vice-Presidency at Sharif University of Technology. 
1311:  
1312: 
1313: \appendix
1314: 
1315: \section{WEIGHTED RESIDUAL FORM OF THE COLLISIONLESS BOLTZMANN EQUATION} 
1316: \label{app::Petrov-Galerkin}
1317: 
1318: Let me define a nonlinear operator ${\cal A}$ and denote 
1319: $\textbf{\textit{u}}^{(\ell)}$ as the $\ell$th prolongation \citep{OL93}
1320: of the physical quantity $u$ in the domain of independent 
1321: variables. Assume a (nonlinear) partial differential equation
1322: % 
1323: \begin{equation}
1324: {\cal A} \left ( u^{(\ell)}, x,t \right )=0, 
1325: \label{eq:general-operator-prolonged}
1326: \end{equation}
1327: %
1328: and its associated initial and boundary conditions that govern 
1329: the evolution of $u(x,t)$ in the domain of the 
1330: spatial variable $x$ and the time $t$. A weighted
1331: residual method \citep{F72} attempts to find an approximate solution 
1332: of the form 
1333: % 
1334: \begin{equation}
1335: u(x,t)=
1336: \sum_{k=1}^{k_{\rm max}} a_{k}(t) 
1337: \varphi_{k}(x),
1338: \label{eq:expansion-for-u-general}
1339: \end{equation}
1340: %
1341: through determining the time-dependent functions $a_k(t)$ for a given 
1342: set of trial (basis) functions $\varphi_{k}(x)$.
1343: The trial functions should satisfy the boundary conditions and be 
1344: linearly independent. Using (\ref{eq:expansion-for-u-general}) and taking 
1345: the inner product of (\ref{eq:general-operator-prolonged}) by some weighting 
1346: functions $W_{k'}(x)$, yield the determining equations 
1347: of $a_k(t)$ as
1348: %
1349: \begin{equation}
1350: \left ( {\cal A},W_{k'} \right ) \equiv 
1351: \int {\cal A} W_{k'} dx=0,~~
1352: k'=1,2,\cdots,k_{\rm max}. 
1353: \label{eq:weighted-residual-form-general}
1354: \end{equation}
1355: %
1356: There are several procedures for choosing $W_{k'}(x)$
1357: and each procedure has its own name. The method with $W_{k'}=\varphi_{k'}$ 
1358: is called the Bubnov-Galerkin, or simply the Galerkin method. The 
1359: Petrov-Galerkin method is associated with $W_{k'}\not=\varphi_{k'}$.
1360: The well-known collocation method uses Dirac's delta functions for 
1361: the weighting purpose. There is an alternative interpretation for the 
1362: inner product $\left ({\cal A},W_{k'} \right )=0$. That is projecting 
1363: the equation ${\cal A}=0$ on a subspace spanned by the weighting function 
1364: $W_{k'}$. Therefore, equation (\ref{eq:weighted-residual-form-general}) 
1365: is often called the {\it Galerkin projection} of 
1366: (\ref{eq:general-operator-prolonged}). In what follows, I use the 
1367: Petrov-Galerkin method and construct the weighted residual form of 
1368: the CBE.   
1369:    
1370: Assume the functions $U(\Theta,\textbf{\textit{J}})$ and
1371: $V(\Theta,\textbf{\textit{J}})$, and define their inner product over the
1372: action-angle space as
1373: %
1374: \begin{equation}
1375: \left (U,V \right )=\int\int
1376: U(\Theta,\textbf{\textit{J}})V(\Theta,\textbf{\textit{J}})
1377:  d \textbf{\textit{J}} d \Theta.
1378: \end{equation}
1379: %
1380: Taking the inner product of the perturbed CBE by the weighting functions
1381: $W^{ml}_{j}(\Theta,\textbf{\textit{J}})=
1382: \Psi^{ml}_{j}(\textbf{\textit{J}})e^{-{\rm i}
1383: \left (l\theta_R+m\theta_\phi \right)}$ gives
1384: % 
1385: \begin{equation}
1386: \left (\frac{\partial f_1}{\partial t},W^{ml}_{j}\right )=
1387: -\left ( \left [f_1,{\cal H}_0 \right ],W^{ml}_{j}\right )
1388: -\left ( \left [f_0,{\cal H}_1 \right ],W^{ml}_{j} \right )-
1389: \left ( \left [f_1,{\cal H}_1 \right ],W^{ml}_{j}\right ).
1390: \label{eq:CBE-weighted-residual-1}
1391: \end{equation}
1392: %
1393: Note that the CBE is the governing equation of the perturbed DF whose
1394: trial functions are $\Phi^{ml}_j(\textbf{\textit{J}})$. With my choice 
1395: of the weighting function (as above) I am following the 
1396: Petrov-Galerkin method. On substituting (\ref{eq:expansion-f1}) and
1397: (\ref{eq:expansion-V1-vs-dmlj}) in
1398: (\ref{eq:CBE-weighted-residual-1}) and after some rearrangements of
1399: summations, one obtains  
1400: % 
1401: \begin{eqnarray}
1402: &{}& {\rm i}\sum_{m',l'}^{}\sum_{j'}^{} \delta_{m,m'}\delta_{l,l'}{{\rm
1403: d}\over  d t}d^{m'l'}_{j'}(t)\int  d \textbf{\textit{J}}
1404: \Psi^{ml}_{j}(\textbf{\textit{J}})\Phi^{m'l'}_{j'}(\textbf{\textit{J}})=
1405: \nonumber \\
1406: %
1407: &{}& \sum_{m',l'}^{}\sum_{j'}^{}
1408: \delta_{m,m'}\delta_{l,l'}d^{m'l'}_{j'}(t)
1409: \int  d \textbf{\textit{J}} \left ( l'\Omega_R+m'\Omega_\phi \right
1410: )\Psi^{ml}_{j}(\textbf{\textit{J}})\Phi^{m'l'}_{j'}(\textbf{\textit{J}})
1411: \nonumber \\
1412: %
1413: &-& \sum_{m',l'}^{}\sum_{j'}^{} \delta_{m,m'} d^{m'l'}_{j'}(t)
1414: \sum_{k}^{} \left [ {4\pi^2\over D_k(m')}\right ] \Lambda^{m'l'}_{kj'}
1415: \int  d \textbf{\textit{J}} \left ( l{\partial f_0\over \partial J_R}
1416: +m'{\partial f_0\over \partial J_{\phi}} \right
1417: )\Psi^{m'l}_{j}(\textbf{\textit{J}})\Psi^{m'l}_k(\textbf{\textit{J}})
1418: \nonumber \\
1419: %
1420: &+& \sum_{m',l'}^{}\sum_{j'}^{}
1421:     \sum_{m'',l''}^{}\sum_{j''}^{}
1422: \delta_{m'',(m-m')} d^{m'l'}_{j'}(t) d^{m''l''}_{j''}(t)
1423: \sum_{k}^{} \left [ {4\pi^2\over D_k(m'')} \right ]
1424: \Lambda ^{m''l''}_{kj''} \nonumber \\
1425: %
1426: &\times& \Biggl [ \int  d \textbf{\textit{J}}
1427: \Psi^{ml}_{j}(\textbf{\textit{J}})\Phi^{m'l'}_{j'}(\textbf{\textit{J}})
1428: \left ( l'{\partial \over \partial J_R}+
1429:         m'{\partial \over \partial J_\phi} \right )
1430: \Psi^{m''(l-l')}_{k}(\textbf{\textit{J}}) \nonumber \\
1431: %
1432: &{}& \qquad -\int  d \textbf{\textit{J}}
1433: \Psi^{ml}_{j}(\textbf{\textit{J}})
1434: \Psi^{m''(l-l')}_{k}(\textbf{\textit{J}})
1435: \left ( (l-l'){\partial \over \partial J_R} +
1436:         m''{\partial \over \partial J_\phi} \right )
1437: \Phi^{m'l'}_{j'}(\textbf{\textit{J}}) \Biggr ],
1438: \label{eq:weighted-residual-expan}
1439: %
1440: \end{eqnarray}
1441: %
1442: where $-m_{\rm max}\le m,m',m''<m_{\rm max}$,
1443: $-l_{\rm max}\le l,l',l''\le l_{\rm max}$ and $0 \le
1444: j,j',j'',k\le j_{\rm max}$. Using equation (\ref{eq:map-mlj-to-i})
1445: and carrying out the index mappings $(m,l,j)\rightarrow p$,
1446: $(m',l',j')\rightarrow q$ and $(m'',l'',j'')\rightarrow r$ one may
1447: introduce the arrays
1448: %
1449: \begin{eqnarray}
1450: M_{pq}&=& \delta_{m,m'}\delta_{l,l'}\int  d \textbf{\textit{J}}
1451: \Psi^{ml}_{j}(\textbf{\textit{J}})\Phi^{m'l'}_{j'}(\textbf{\textit{J}}),
1452: \label{eq:M-tensor} \\
1453: %
1454: C_{pq}&=& \delta_{m,m'} \Biggl [ \delta_{l,l'}
1455: \int  d \textbf{\textit{J}}
1456: \left ( l'\Omega_R+m'\Omega_\phi \right )\Psi^{ml}_{j}(\textbf{\textit{J}})
1457: \Phi^{m'l'}_{j'}(\textbf{\textit{J}}) \nonumber \\
1458: &{}& \qquad \qquad -\sum_{k=0}^{j_{\rm max}}
1459: \left [ {4\pi^2\over D_k(m')}\right ]
1460: \Lambda^{m'l'}_{kj'} \int  d \textbf{\textit{J}} \left ( l{\partial
1461: f_0\over \partial J_R} +m'{\partial f_0\over \partial J_{\phi}}
1462: \right )\Psi^{m'l}_{j}(\textbf{\textit{J}})
1463: \Psi^{m'l}_k(\textbf{\textit{J}})\Biggr ],
1464: \label{eq:A-tensor} \\
1465: %
1466: K_{pqr}&=& \delta_{m',(m-m'')} \delta_{l',(l-l'')}
1467: \sum_{k=0}^{j_{\rm max}} \left [ {4\pi^2\over
1468: D_k(m'')} \right ] \Lambda ^{m''l''}_{kj''} \nonumber \\
1469: &{}& \times \Biggl [ \int  d \textbf{\textit{J}}
1470: \Psi^{ml}_{j}(\textbf{\textit{J}})
1471: \Phi^{m'l'}_{j'}(\textbf{\textit{J}})
1472: \left ( l'{\partial \over \partial J_R}+
1473:         m'{\partial \over \partial J_\phi} \right )
1474: \Psi^{m''l''}_{k}(\textbf{\textit{J}}) \nonumber \\
1475: %
1476: &{}& \qquad -\int  d \textbf{\textit{J}}
1477: \Psi^{ml}_{j}(\textbf{\textit{J}})
1478: \Psi^{m''l''}_{k}(\textbf{\textit{J}})
1479: \left ( l''{\partial \over \partial J_R} +
1480:         m''{\partial \over \partial J_\phi} \right )
1481: \Phi^{m'l'}_{j'}(\textbf{\textit{J}}) \Biggr ].\label{eq:B-tensor}
1482: %
1483: \end{eqnarray}
1484: %
1485: Consequently, equation (\ref{eq:weighted-residual-expan}) takes the
1486: following matrix form
1487: %
1488: \begin{equation}
1489: {\rm i}\sum_{q=1}^{n_{\rm max}} M_{pq} { d \over  d t}z_q(t) =
1490: \sum_{q=1}^{n_{\rm max}} C_{pq}z_q(t)+\sum_{q,r=1}^{n_{\rm max}}
1491: K_{pqr}z_q(t)z_r(t),~~z_p(t)\equiv d^{ml}_j(t),~~
1492: p=1,2,\cdots,n_{\rm max}.
1493: \label{eq:ODE-for-z-one}
1494: \end{equation}
1495: %
1496: Let the matrix $\textbf{\textit{M}}^{-1}=[M^{-1}_{pq}]$ be the inverse of
1497: $\textbf{\textit{M}}=[M_{pq}]$ and left-multiply (\ref{eq:ODE-for-z-one})
1498: by $\textbf{\textit{M}}^{-1}$ to get
1499: %
1500: \begin{equation}
1501: {\rm i} { d \over  d t}z_p(t) =
1502: \sum_{q=1}^{n_{\rm max}} A_{pq}z_q(t)+\sum_{q,r=1}^{n_{\rm max}} B_{pqr}
1503: z_q(t)z_r(t),~~\textbf{\textit{A}}=\textbf{\textit{M}}^{-1}
1504: \cdot \textbf{\textit{C}},~~
1505: B_{pqr}=\sum_{s=1}^{n_{\rm max}}M^{-1}_{ps}K_{sqr}.
1506: \label{eq:ODE-for-z-two}
1507: \end{equation}
1508: %
1509: Evaluation of the integrands in (\ref{eq:B-tensor}) will be
1510: considerably simplified if one avoids the partial derivatives
1511: of $\Phi^{ml}_j(\textbf{\textit{J}})$ through integrating (\ref{eq:B-tensor})
1512: by parts. That gives
1513: %
1514: \begin{eqnarray}
1515: K_{pqr}= &{}& \delta_{m',(m-m'')} \delta_{l',(l-l'')}
1516: \sum_{k=0}^{j_{\rm max}} \left [ {4\pi^2\over
1517: D_k(m'')} \right ] \Lambda ^{m''l''}_{kj''} \nonumber \\
1518: \times \biggl \{ &{}& \int {\rm
1519: d}\textbf{\textit{J}} \Psi^{ml}_{j}(\textbf{\textit{J}})
1520: \Phi^{m'l'}_{j'}(\textbf{\textit{J}})
1521: \left ( l{\partial \over \partial J_R}+
1522:         m{\partial \over \partial J_\phi} \right )
1523: \Psi^{m''l''}_{k}(\textbf{\textit{J}}) \nonumber \\
1524: %
1525: &+& \int  d \textbf{\textit{J}}
1526: \Phi^{m'l'}_{j'}(\textbf{\textit{J}})
1527: \Psi^{m''l''}_{k}(\textbf{\textit{J}})
1528: \left ( l''{\partial \over \partial J_R} +
1529:         m''{\partial \over \partial J_\phi} \right )
1530: \Psi^{ml}_{j}(\textbf{\textit{J}})
1531: \biggr \}.\label{eq:B-tensor-by-parts}
1532: \end{eqnarray}
1533: %
1534: When all stars move on prograde orbits, the equilibrium DF takes the
1535: form $f_0(\textbf{\textit{J}})=H(J_\phi)f^P_0(\textbf{\textit{J}})$
1536: where $H$ is the
1537: Heaviside function. Upon using (\ref{eq:trial-functions-two}),
1538: this contributes a term including Dirac's delta function
1539: $\delta(J_\phi)$ to the trial functions. Thus, the following
1540: boundary terms
1541: %
1542: \begin{eqnarray}
1543: K^b_{pqr} \! &=& \! \delta_{m',(m-m'')} \delta_{l',(l-l'')}
1544: \sum_{k=0}^{j_{\rm max}}
1545: \left [ {4\pi^2\over D_k(m'')} \right ]
1546: \Lambda ^{m''l''}_{kj''} \nonumber \\
1547: %
1548: \times \Biggl \{ \int_{0}^{\infty} \!\! &  d J_R &
1549: \left [ {m' f^P_0(\textbf{\textit{J}})\Psi^{ml}_{j}(\textbf{\textit{J}})
1550:                           \Psi^{m'l'}_{j'}(\textbf{\textit{J}}) \over
1551:          l'\Omega_R(\textbf{\textit{J}})+m'\Omega_\phi(\textbf{\textit{J}})
1552:         }
1553: \left ( l{\partial \over \partial J_R}+
1554:         m{\partial \over \partial J_\phi} \right )
1555: \Psi^{m''l''}_{k}(\textbf{\textit{J}}) \right ]_{J_{\phi}=0} \nonumber \\
1556: %
1557: + \int_{0}^{\infty} \!\! &  d J_R & \left [
1558:         {m' f^P_0(\textbf{\textit{J}}) \Psi^{m''l''}_{k}(\textbf{\textit{J}})
1559:                              \Psi^{m'l'}_{j'}(\textbf{\textit{J}})
1560:    \over l'\Omega_R(\textbf{\textit{J}})+m'\Omega_\phi(\textbf{\textit{J}})
1561:         }
1562: \left ( l''{\partial \over
1563: \partial J_R} +
1564:         m''{\partial \over \partial J_\phi} \right )
1565: \Psi^{ml}_{j}(\textbf{\textit{J}}) \right ]_{J_\phi=0}
1566:  \Biggr \}, \label{eq:B-tensor-boundary}
1567: \end{eqnarray}
1568: %
1569: must be added to $K_{pqr}$ when the equilibrium
1570: disk is unidirectional. The partial derivatives of
1571: $\Psi^{ml}_j(\textbf{\textit{J}})$ needed for equations
1572: (\ref{eq:B-tensor-by-parts}) and (\ref{eq:B-tensor-boundary})
1573: are calculated by differentiating equation (\ref{eq:fourier-coeffs})
1574: partially with respect to an action:
1575: %
1576: \begin{eqnarray}
1577: {\partial \Psi^{ml}_j\over \partial J_{\nu}} &=&
1578: {1\over \pi} \int\limits_{0}^{\pi} \biggl \{
1579: {\partial \psi^{|m|}_j \over \partial R}{\partial R \over
1580:  \partial J_{\nu}}\cos [l\theta _R+m(\theta _\phi -\phi)] \nonumber \\
1581: &{}& \qquad -m\psi^{|m|}_j(R){\partial \over \partial J_\nu}
1582:      \left(\theta_{\phi}-\phi\right)
1583:      \sin [l\theta _R+m(\theta _\phi -\phi)] \biggr \}
1584:  d \theta _R,~~\nu \equiv R,\phi. \label{eq::deriv-fourier-coeff}
1585: \end{eqnarray}
1586: %
1587: Jalali \& Hunter (2005b) encountered these partial derivatives
1588: in their second order perturbation theory devised for computing
1589: the energy of eigenmodes. I adopt their technique for calculating
1590: the quantities $\partial R/\partial J_{\nu}$ and
1591: $\partial \left( \theta_{\phi}-\phi \right )/\partial J_{\nu}$.
1592: The variables $R$, $(\theta _\phi -\phi)$, and $p_R$ are regarded
1593: as functions of $(J_R,J_{\phi},\theta_R)$ because the action-angle
1594: transformation
1595: $(\textbf{\textit{x}},\textbf{\textit{p}})\rightarrow
1596: (\textbf{\textit{J}},\Theta)$
1597: is defined in the phase space of an axisymmetric state. From
1598: $v_R=dR/dt=(\partial R/\partial \theta_R)\Omega_R$ one may write
1599: %
1600: \begin{equation}
1601: \label{eq::dRdJ}
1602: \frac{ d }{ d t}\left[{\partial R \over \partial J_\nu}\right]=
1603: {\partial^2 R \over \partial \theta_R \partial J_\nu}\frac{d\theta_R}{dt}
1604: =\Omega_R\frac{\partial}{\partial J_\nu}
1605: \left({\partial R \over \partial \theta_R}\right)
1606: =\Omega_R\frac{\partial}{\partial J_\nu}\left(\frac{v_R}{\Omega_R}\right)
1607: ={\partial v_R \over \partial J_\nu}-\frac{v_R}{\Omega_R}
1608: {\partial \Omega_R \over \partial J_\nu}.
1609: \end{equation}
1610: %
1611: Similarly, one obtains
1612: %
1613: \begin{eqnarray}
1614: \frac{ d }{ d t}\left[{\partial \over \partial J_\nu}
1615:  (\theta _\phi -\phi)\right]
1616: &=& {\partial \Omega_{\phi} \over \partial J_\nu}
1617: -\frac{\delta_{\nu,\phi}}{R^2}
1618: +{2J_{\phi} \over R^3}{\partial R \over \partial J_\nu}
1619: -\frac{1}{\Omega_R}\left[\Omega_{\phi}-\frac{J_{\phi}}{R^2}\right]
1620: {\partial \Omega_R \over \partial J_\nu}, \label{eq::dthetadJ} \\
1621: \frac{ d }{ d t}\left[{\partial v_R \over \partial J_\nu}\right]
1622: &=& {2J_{\phi} \over R^3}\delta_{\nu,\phi}
1623: -\left[{3J_{\phi}^2 \over R^4}+V_0^{\prime\prime}(R)\right]
1624: {\partial R \over \partial J_\nu}
1625: -\frac{1}{\Omega_R}\left[{J_{\phi}^2 \over R^3}-V_0^{\prime}(R)\right]
1626: {\partial \Omega_R \over \partial J_\nu}. \label{eq::dvdJ}
1627: \end{eqnarray}
1628: %
1629: The set of three equations (\ref{eq::dRdJ}) through (\ref{eq::dvdJ})
1630: can be integrated along an orbit, and they
1631: provide the additional values needed to evaluate
1632: the partial derivatives (\ref{eq::deriv-fourier-coeff}).
1633: Initial values are $\partial v_R/\partial J_\nu=
1634: \partial(\theta_{\phi}-\phi)/\partial J_\nu=0$ at $\theta_R=t=0$
1635: where $R=R_{\rm min}$ because $v_R=\theta_{\phi}-\phi=0$ for
1636: all orbits. However the initial $R_{\rm min}$ values change with the
1637: actions, and initial values for the derivatives of $R$ with respect
1638: to the actions are
1639: %
1640: \begin{equation}
1641: \left[ {\partial R \over \partial J_R}\right]_{R=R_{\rm min}}=
1642: \frac{R^3_{\rm min}\Omega_R}
1643: {R^3_{\rm min}V_0^{\prime}(R_{\rm min})-J^2_{\phi}}, ~~
1644: \left[ {\partial R \over \partial J_{\phi}}\right]_{R=R_{\rm min}}=
1645: \frac{R_{\rm min}(R^2_{\rm min}\Omega_{\phi}-J_{\phi})}
1646: {R^3_{\rm min}V_0^{\prime}(R_{\rm min})-J^2_{\phi}}.
1647: \end{equation}
1648: %
1649: They are obtained by differentiating the zeroth order energy equation.
1650: 
1651: 
1652: 
1653: \begin{thebibliography}{}
1654: 
1655: \bibitem[Aoki \& Iye (1978)]{AI78} Aoki, S., \& Iye, M. 1978,
1656:     \pasj, 30, 519
1657: 
1658: \bibitem[Asghari \& Jalali (2006)]{AJ06}
1659:    Asghari, N. M., \& Jalali, M. A. 2006, \mnras, 373, 337
1660: 
1661: \bibitem[Binney \& Tremaine (1987)]{BT87} Binney, J., \& Tremaine, S.
1662:    1987, Galactic Dynamics (Princeton: Princeton Univ. Press)
1663: 
1664: \bibitem[Case (1959)]{C59} Case, K. M. 1959, Annals of Physics, 7, 349
1665: 
1666: \bibitem[Clutton-Brock (1972)]{CB72} Clutton-Brock, M.  1972,
1667:     Ap\&SS, 16, 101
1668: 
1669: \bibitem[Doering \& Gibbon (1995)]{DG95} Doering, C. R., \& Gibbon, J. D.
1670:     1995, Applied Analysis of the Navier-Stokes Equations
1671:     (Cambridge: Cambridge Univ. Press)
1672: 
1673: \bibitem[Evans \& Read (1998a)]{ER98a} Evans, N. W., \& Read, J. C. A.
1674:     1998a, \mnras, 300, 83
1675: 
1676: \bibitem[Evans \& Read (1998b)]{ER98b} Evans, N. W., \& Read, J. C. A.
1677:     1998b, \mnras, 300, 106
1678: 
1679: \bibitem[Finlayson (1972)]{F72} Finlayson, B. A. 1972, 
1680:     The Method of Weighted Residuals and Variational Principles
1681:     (New York: Academic Press)
1682:  
1683: \bibitem[Hunter (1980)]{H80} Hunter, C.  1980, \pasj, 32, 33
1684: 
1685: \bibitem[Hunter (1992)]{H92} Hunter, C.  1992, in
1686:      Astrophysical Disks, ed S. F. Dermott, J. H. Hunter Jr.,
1687:      \& R. E. Wilson (New York: Annals NY Acad. Sci. 675), 22
1688: 
1689: \bibitem[Jalali \& Hunter (2005a)]{JHa} Jalali M. A.,
1690:     \& Hunter, C. 2005a, \apj, 630, 804
1691: 
1692: \bibitem[Jalali \& Hunter (2005b)]{JHb} Jalali M. A.,
1693:     \& Hunter, C. 2005b, astro-ph/0503255
1694: 
1695: \bibitem[Jeans (1915)]{J1915} Jeans, J. H. 1915, \mnras, 76, 70
1696: 
1697: \bibitem[Kalnajs (1971)]{K71} Kalnajs, A. J.  1971, \apj, 166, 275
1698: 
1699: \bibitem[Kalnajs (1977)]{K77} Kalnajs, A. J.  1977, \apj, 212, 637
1700: 
1701: \bibitem[Kalnajs (1978)]{K78} Kalnajs, A. J.  1978, in IAU Symp. 77,
1702:      Structure and Properties of Nearby Galaxies, ed. E. M. Berhuijsen \&
1703:      R. Wielebinski (Dordrecht: Reidel) 113
1704: 
1705: \bibitem[Landau (1946)]{L46} Landau, L. D. 1946, J. Phys. USSR, 10, 25
1706: 
1707: \bibitem[Lynden-Bell (1962)]{L-B62} Lynden-Bell, D. 1962, \mnras, 124, 1
1708: 
1709: \bibitem[Mattingly \& Sinai (1999)]{MS99} Mattingly, J. C.,
1710:     \& Sinai, Ya. G. 1999, Commun. Contemp. Math., 1, 497
1711: 
1712: \bibitem[Mestel (1963)]{M63} Mestel, L. 1963, \mnras, 126, 553
1713: 
1714: \bibitem[Olver (1993)]{OL93} Olver, P. J. 1993, 
1715:     Applications of Lie Groups to Differential Equations
1716:     (New York: Springer-Verlag) 
1717: 
1718: \bibitem[Pichon \& Cannon (1997)]{PC97} Pichon, C.,
1719:     \& Cannon, R. C.  1997, \mnras, 291, 616
1720: 
1721: \bibitem[Polyachenko (2004)]{P04} Polyachenko, E. V. 2004,
1722:     \mnras, 348, 345
1723: 
1724: \bibitem[Polyachenko (2005)]{P05} Polyachenko, E. V. 2005,
1725:     \mnras, 357, 559
1726: 
1727: \bibitem[Press et al. (2001)]{Press01} Press, W. H., Teukolsky, S. A.,
1728:     Vetterling, W. T., \& Flannery, B. P. 2001, Numerical Recipes
1729:     in Fortran 77 (Cambridge: Cambridge Univ. Press)
1730: 
1731: \bibitem[Toomre (1964)]{T64} Toomre, A. 1964, \apj, 139, 1217
1732: 
1733: \bibitem[Toomre (1981)]{T81} Toomre, A.  1981, in Structure
1734:     and Evolution of Normal Galaxies, ed S. M. Fall \& D. Lynden-Bell
1735:     (Cambridge: Cambridge Univ. Press), 111
1736: 
1737: \bibitem[Tremaine (2005)]{T05} Tremaine, S., 2005, \apj, 625, 143
1738: 
1739: \bibitem[Tremaine \& Weinberg (1984)]{TW84} Tremaine, S., \&
1740:     Weinberg, M. D.  1984, \mnras, 209, 729
1741: 
1742: \bibitem[Vandervoort (2003)]{V03} Vandervoort, P., 2003,
1743:     \mnras, 339, 537
1744: 
1745: \bibitem[van Kampen (1955)]{vK55} van Kampen, N. G. 1955,
1746:     Physica, 31, 949
1747: 
1748: \bibitem[Vauterin \& Dejonghe (1996)]{VD96} Vauterin, P., \& Dejonghe, H.
1749:     1996, A\&A, 313, 465
1750: 
1751: \bibitem[Zang (1976)]{Z76} Zang, T. A.  1976, PhD Thesis,
1752:     Massachusetts Institute of Technology, Cambridge, MA
1753: 
1754: \end{thebibliography}
1755: 
1756: \end{document}
1757: 
1758: