astro-ph0211580/ms.tex
1: %\documentstyle[12pt,aasms4]{article}
2: \documentstyle[aaspp4]{article}
3: %\documentclass[12pt,preprint]{aastex}
4: 
5: \received{}
6: \accepted{}
7: \journalid{VOL}{JOURNAL DATE}
8: \articleid{START PAGE}{END PAGE}
9: \paperid{MANUSCRIPT ID}
10: 
11: %\lefthead{Lee \& Yoshida}
12: %\righthead{The r-mode instability of neutron stars with the superfluid core}
13: 
14: \begin{document}
15: 
16: %\magnification=1200
17: %\def\pmb#1{\setbox0=\hbox{$#1$}
18: % \kern-.025em\copy0\kern-\wd0
19: % \kern-.05em\copy0\kern-\wd0
20: % \kern-.025em\raise.0433em\box0 }
21: \def\gtsim {>\kern-1.2em\lower1.1ex\hbox{$\sim$}}
22: \def\ltsim {<\kern-1.2em\lower1.1ex\hbox{$\sim$}}
23: \def\ref{\hangindent=1pc \hangafter=1 \noindent}
24: %
25: \def\pmb#1{\mbox{\boldmath$#1$}}
26: %
27: %\baselineskip=20pt
28: 
29: \title{\bf R-modes of neutron stars with the superfluid core}
30: 
31: \author{Umin Lee$^1$ \& Shijun Yoshida$^{1,2}$}
32: 
33: \affil{$^1$Astronomical Institute, Tohoku University, Sendai, Miyagi 980-8578, Japan
34: \\ lee@astr.tohoku.ac.jp, yoshida@astr.tohoku.ac.jp}
35: 
36: \affil{$^2$Centro Multidisciplinar de Astrof\'{\i}sica -- CENTRA,
37: Departamento de F\'{\i}sica, Instituto Superior T\'ecnico, Av. Rovisco 
38: Pais 1,
39: 1049-001 Lisboa, Portugal}
40: 
41: \begin{abstract}
42: 
43: We investigate the modal properties of the $r$-modes 
44: of rotating neutron stars with the core filled with neutron and proton superfluids,
45: taking account of entrainment effects between the superfluids.
46: The stability of the $r$-modes against gravitational radiation reaction
47: is also examined considering
48: viscous dissipation due to shear and a damping mechanism called mutual 
49: friction between the superfluids in the core.
50: We find the $r$-modes in the superfluid core
51: are split into ordinary $r$-modes and
52: superfluid $r$-modes, which we call, respectively, $r^o$- and $r^s$-modes.
53: The two superfluids in the core flow together for the $r^o$-modes, while
54: they counter-move for the $r^s$-modes.
55: For the $r^o$-modes, the coefficient $\kappa_0\equiv\lim_{\Omega\rightarrow 0}\omega/\Omega$ 
56: is equal to $2m/[l^\prime(l^\prime+1)]$, almost independent of the parameter $\eta$ that
57: parameterizes the entrainment effects between the superfluids,
58: where $\Omega$ is the angular frequency of rotation, 
59: $\omega$ the oscillation frequency observed in the corotating frame of the star, and
60: $l^\prime$ and $m$ are the indices of
61: the spherical harmonic function representing the angular dependence of the $r$-modes.
62: For the $r^s$-modes, on the other hand, $\kappa_0$ is equal to $2m/[l^\prime(l^\prime+1)]$
63: at $\eta=0$ (no entrainment), and it almost linearly increases as $\eta$ is increased from $\eta=0$.
64: The $r^o$-modes, for which $\pmb{w}^\prime\equiv\pmb{v}^\prime_p-\pmb{v}^\prime_n\propto\Omega^3$,
65: correspond to the $r$-modes discussed by Lindblom \& Mendell (2000), where
66: $\pmb{v}^\prime_n$ and $\pmb{v}^\prime_p$ are the Eulerian velocity perturbations
67: of the neutron and proton superfluids, respectively.
68: The mutual friction in the superfluid core is found ineffective to stabilize the $r$-mode
69: instability caused by the $r^o$-mode except in a few narrow regions of $\eta$.
70: The $r$-mode instability caused by the $r^s$-modes, on the other hand, is extremely
71: weak and easily damped by dissipative processes in the star.
72: 
73: 
74: 
75: \end{abstract}
76: 
77: \keywords{instabilities --- stars: neutron --- stars: oscillations --- stars : rotation}
78: 
79: 
80: \section{Introduction}
81: 
82: 
83: One of the roles expected for the $r$-mode instability to play
84: (Andersson 1998, Friedman \& Morsink 1998) 
85: is deceleration of the spin of newly born hot neutron stars by emitting
86: gravitational waves that carry away the angular momentum of the star
87: (e.g., Lindblom et al 1998).
88: We know, however, that
89: among older and colder neutron stars as found in LMXB systems
90: there are many rapidly rotating neutron stars like a millisecond pulsar
91: (see, e.g, Phinney \& Kulkarni 1994).
92: This fact suggests the possibility that
93: the $r$-mode instability does not always work well
94: to spin down the rapid rotation of the stars.
95: For the $r$-modes in cold neutron stars with a solid crust, for example,
96: Bildsten \& Ushomirsky (2000) suggested a damping mechanism operating
97: in the viscous boundary layer at the interface between the solid crust and the
98: fluid core,
99: to explain the clustering of spin frequencies around the value of 300Hz 
100: for accreting neutron stars in LMXB systems (van der Klis 2000; see also
101: Andersson, Kokkotas, \& Stergioulas 1999).
102: For the modal properties of the $r$-modes in neutron stars
103: with a solid crust, see Yoshida \& Lee (2001), who
104: showed that the $r$-modes in the core are largely affected by resonance with
105: the toroidal sound modes propagating in the solid crust.
106: 
107: 
108: As neutron stars cool down below $T\sim 10^9{\rm K}$,  
109: neutrons and protons in the core are believed to
110: be in superfluid states (e.g., Shapiro \& Tuekolwsky 1983).
111: In a rotating system of superfluids, it is well known that scattering between 
112: vortices in the superfluids and normal fluid particles produces
113: dissipation called mutual friction (e.g., Khalatnikov 1965, Tilley \& Tilley 1990).
114: Therefore, for people who are interested in the $r$-modes instability, 
115: it was a serious concern whether the $r$-mode instability could survive the dissipation
116: due to mutual
117: friction in the superfluid core of cold neutron stars, e.g., in LMXBs.
118: It was Lindblom \& Mendell (2000) who first examined the damping effects of 
119: mutual friction in the core
120: on the $r$-mode instability, and concluded that the mutual friction could not be
121: strong enough to damp out the instability in most of the parameter domains 
122: which we are interested in.
123: In their analysis of the $r$-modes in neutron stars,
124: Lindblom \& Mendell (2000)
125: employed a perturbative method in which the spin angular frequency $\Omega$ is regarded
126: as the infinitesimal parameter to expand the eigenfrequencies and eigenfunctions
127: of the modes, and they looked for the $r$-modes with the scalings given by
128: $\beta^\prime\equiv\mu_p^\prime-\mu_n^\prime+m_e\mu_e^\prime/m_p\propto \Omega^4$ 
129: and $\pmb{w}^\prime\equiv\pmb{v}_p^\prime-\pmb{v}_n^\prime\propto\Omega^3$, where
130: $\mu_p$, $\mu_n$, and $\mu_e$ are
131: the chemical potentials of the proton, neutron,
132: and electron in the core, and $\pmb{v}_p$ and $\pmb{v}_n$ are the velocities of the
133: proton and neutron superfluids, and the prime $(^\prime)$ indicates the
134: Euler perturbation of the quantity.
135: 
136: Recently, Andersson \& Comer (2001) discussed the dynamics of superfluid neutron star cores, 
137: and confirmed an earlier result by Lee (1995)
138: that there are no $g$-modes propagating in the superfluid core.
139: They also applied their argument to the $r$-modes in the core 
140: filled with neutron and proton superfluids, and
141: suggested the existence of two distinct families of the $r$-modes in the core, i.e.,
142: $r$-modes for which
143: the neutrons and the protons flow together, and those
144: for which the neutrons and the protons are counter moving.
145: Lindblom \& Mendell (2000) considered the former family of the $r$-modes,
146: which are less strongly affected by the mutual friction than the latter.
147: 
148: 
149: In this paper, we employ a different method of calculation
150: to investigate the $r$-modes in rotating neutron stars, although
151: the basic equations describing the dynamics of superfluids in the core
152: are essentially the same as those given in Lindblom \& Mendell (1994).
153: Our method of solution is a variant of that used in Lee \& Saio (1986).
154: Because the separation of variables is not possible for perturbations in
155: rotating stars,
156: we expand the perturbations in terms of
157: spherical harmonic functions $Y_l^m(\theta,\phi)$ with different $l$'s for a given $m$.
158: We substitute the expansions into linearized basic equations to obtain a set of
159: simultaneous linear ordinary differential equations of the expansion coefficients, which is
160: to be solved as an eigenvalue problem of the oscillation frequency.
161: In this method, we do not have to assume apriori a form of solutions for the $r$-modes
162: in the lowest order of $\Omega$.
163: In \S 2 we present the basic equations employed in this paper
164: for the dynamics of superfluids in the core, and in \S 3
165: dissipation processes considered in this paper are described.
166: \S 4 gives numerical results, and \S 5 and \S 6 are for discussions and conclusions.
167: 
168: 
169: 
170: \section{Oscillation Equations in the Superfluid Core of Rotating Neutron Stars}
171: 
172: Microscopically, the superfluids in a rotating system
173: move irrotationally everywhere except within the core
174: of vortex lines.
175: Averaging over many vortices in the fluids, we may define the average superfluid velocities
176: $<\pmb{v}>$,
177: which can satisfy the usual relation for uniform rotation 
178: $\nabla\times <\pmb{v}>=2\pmb{\Omega}$ in the equilibrium state
179: (see, e.g., Feynman 1972).
180: In the following we simply use $\pmb{v}$, instead of $<\pmb{v}>$, to signify the 
181: superfluid velocities.
182: Hydrodynamic equations for a rotating superfluid based on the two-fluid model
183: with the normal fluid and superfluid components
184: are derived, for example, in Khalatnikov (1965).
185: 
186: 
187: We derive basic equations governing superfluid motions
188: in the neutron star core in the Newtonian dynamics, assuming uniform rotation of the star.
189: The core is assumed to be filled with neutron and proton superfluids and 
190: a normal fluid of electron.
191: We also assume perfect charge neutrality between the protons and electrons because
192: the plasma frequency is much higher than the oscillation frequencies considered in
193: this paper (see, e.g., Mendell 1991a).
194: Since the transition temperatures $T_c\sim 10^9$K to neutron and proton superfluids 
195: are much higher than the interior temperatures of old neutron stars (see, e.g., 
196: Epstein 1988), 
197: we assume that all the neutrons and protons in the core are in superfluid states and
198: the normal fluid components of the fluids can be ignored.
199: 
200: The basic hydrodynamic equations employed in this paper for the
201: neutron and proton superfluids in rotating neutron stars 
202: are essentially the same as those given in Mendell (1991a) and
203: Lindblom \& Mendell (1994).
204: In a fluid system in which two superfluids coexist, the entrainment between 
205: the two superfluid motions occurs because a Cooper pair of one
206: fluid particles is affected by the force field produced by the other fluid particles
207: (Andreev \& Bashkin 1975 for a system of $^3$He and $^4$He superfluids).
208: In the system of the neutron and proton superfluids in the core
209: the entrainment effects between them are mediated
210: by quantum mechanical nuclear force
211: between the neutrons and the protons (e.g., Alpar et al 1984).
212: Here, we introduce the entrainment effects in
213: mass conservation equations, which are given by
214: %
215: \begin{equation}
216: {\partial\rho_n\over\partial t}+\nabla\cdot\pmb{j}_n=0, 
217: \end{equation}
218: %
219: and
220: %
221: \begin{equation}
222: {\partial\rho_p\over\partial t}+\nabla\cdot\pmb{j}_p=0, 
223: \end{equation}
224: %
225: where $\rho_n$ and $\rho_p$ are the mass densities of the neutron and proton superfluids,
226: and the mass current vectors $\pmb{j}_n$ and $\pmb{j}_p$ are defined as
227: %
228: \begin{equation}
229: \pmb{j}_n=\rho_{nn}\pmb{v}_n+\rho_{np}\pmb{v}_p, 
230: \end{equation}
231: %
232: and
233: %
234: \begin{equation}
235: \pmb{j}_p=\rho_{pp}\pmb{v}_p+\rho_{pn}\pmb{v}_n, 
236: \end{equation}
237: %
238: where $\pmb{v}_n$ and $\pmb{v}_p$ denote the velocities of the neutron and the proton superfluids,
239: respectively, and the coefficients $\rho_{nn}$, $\rho_{np}$, $\rho_{pp}$, and $\rho_{pn}$
240: are defined to satisfy $\rho_n=\rho_{nn}+\rho_{np}$ and
241: $\rho_p=\rho_{pp}+\rho_{pn}$ and $\rho_{np}=\rho_{pn}$ under Galilean transformations.
242: The mass conservation equation for the electron fluid is given by
243: %
244: \begin{equation}
245: {\partial\rho_e\over\partial t}+\nabla\cdot\left(\rho_e\pmb{v}_e\right)=0,
246: \end{equation}
247: %
248: where $\rho_e$ and $\pmb{v}_e$ denote, respectively,
249: the mass density and the velocity of the electron fluid.
250: The velocity equation of the neutron superfluid is in an inertial frame given by
251: %
252: \begin{equation}
253: {\partial \pmb{v}_n\over\partial t}+\pmb{v}_n\cdot\nabla\pmb{v}_n=-\nabla(\mu_n+\Psi)
254: +{\rho_{np}\over\rho_n}\left(\pmb{v}_p-\pmb{v}_n\right)\times\left(\nabla\times\pmb{v}_n\right),
255: \end{equation}
256: %
257: where $\mu_n$ is the chemical potential of neutron per unit mass, and $\Psi$ is the 
258: gravitational potential.
259: The term proportional to $\rho_{np}$ on the right hand side of equation (6) represents
260: a drag force between neutrons and protons.
261: If we assume perfect charge neutrality of 
262: the proton and electron plasma, we may have
263: %
264: \begin{equation}
265: \pmb{j}_p/\rho_p=\pmb{v}_e, \quad {\rm and} \quad \rho_p/m_p=\rho_e/m_e,
266: \end{equation}
267: %
268: where $m_p$ and $m_e$ are the proton and the electron masses.
269: The velocity equation for the proton-electron fluid is then given by
270: %
271: \begin{equation}
272: {\partial\over\partial t}\left(\pmb{v}_p+{m_e\over m_p}\pmb{v}_e\right)
273: +\pmb{v}_p\cdot\nabla\pmb{v}_p
274: +{m_e\over m_p}{\pmb{v}_e}\cdot\nabla{\pmb{v}_e}
275: =-\nabla\left(\mu_p+{m_e\over m_p}\mu_e+\zeta\Psi\right)
276: -{\rho_{np}\over\rho_p}\left(\pmb{v}_p-\pmb{v}_n\right)\times\left(\nabla\times\pmb{v}_p\right), 
277: \end{equation}
278: %
279: where $\zeta=1+m_e/m_p$, and $\mu_p$ and $\mu_e$ are the chemical potentials
280: per unit mass for the proton and electron, respectively.
281: Note that we have neglected the entropy carried by the normal fluid of electron
282: for simplicity.
283: The Poisson equation is given by
284: %
285: \begin{equation}
286: \nabla^2\Psi=4\pi G\rho, 
287: \end{equation}
288: %
289: where $\rho=\rho_n+\rho_p+\rho_e$ and $G$ is the gravitational constant.
290: 
291: 
292: To linearise the hydrodynamic equations for the superfluids in
293: rotating neutron stars, we assume that
294: the neutron and proton superfluids and the electron normal fluid in an equilibrium state 
295: are in the same rotational motion
296: with the angular velocity $\Omega$ around the axis of rotation, which is along the z-axis.
297: In a perturbed state, however, the neutron and the proton superfluids move differently
298: from each other in the core, obeying their own governing equations.
299: The mass current vectors are linearized to be
300: %
301: \begin{equation}
302: \pmb{j}_n^\prime=\rho_n^\prime\pmb{v}_0+\tilde{\pmb{j}}_n^\prime,
303: \quad {\rm and} \quad
304: \pmb{j}_p^\prime=\rho_p^\prime\pmb{v}_0+\tilde{\pmb{j}}_p^\prime,
305: \end{equation}
306: %
307: where the prime $(^\prime)$ indicates the Eulerian perturbation of the quantity,
308: and $\pmb{v}_{n0}=\pmb{v}_{p0}=\pmb{v}_0=r\sin\theta\Omega\pmb{e}_\phi$
309: is the fluid velocity in the equilibrium state, and
310: %
311: \begin{equation}
312: \tilde{\pmb{j}}_n^\prime=\rho_{nn}\pmb{v}_n^\prime+\rho_{np}\pmb{v}_p^\prime,
313: \quad {\rm and} \quad
314: \tilde{\pmb{j}}_p^\prime=\rho_{pn}\pmb{v}_n^\prime+\rho_{pp}\pmb{v}_p^\prime
315: \end{equation}
316: %
317: are the perturbed mass current vectors in a corotating frame.
318: The perturbed superfluid velocities are then given in terms of $\tilde{\pmb{j}}_n^\prime$
319: and $\tilde{\pmb{j}}_p^\prime$ as
320: %
321: \begin{equation}
322: \pmb{v}_n^\prime={\rho_{11}\over\rho_n^2}\tilde{\pmb{j}}_n^\prime
323: +{\rho_{12}\over\rho_n\rho_p}\tilde{\pmb{j}}_p^\prime,
324: \quad {\rm and} \quad
325: \pmb{v}_p^\prime={\rho_{21}\over\rho_n\rho_p}\tilde{\pmb{j}}_n^\prime
326: +{\rho_{22}\over\rho_p^2}\tilde{\pmb{j}}_p^\prime,
327: \end{equation}
328: %
329: where
330: %
331: \begin{equation}
332: \rho_{11}={\rho_{pp}\rho_n^2\over\tilde{\rho}^2}, \quad
333: \rho_{22}={\rho_{nn}\rho_p^2\over\tilde\rho^2}, \quad 
334: \rho_{12}=\rho_{21}=-{\rho_{np}\rho_n\rho_p\over\tilde\rho^2},
335: \end{equation}
336: %
337: and
338: %
339: \begin{equation}
340: \tilde\rho^2=\rho_{nn}\rho_{pp}-\rho_{np}\rho_{pn}.
341: \end{equation}
342: %
343: Note that $\rho_{11}+\rho_{12}=\rho_n$ and $\rho_{22}+\rho_{21}=\rho_p$.
344: 
345: If the equilibrium structure is axisymmetric about the rotation axis, 
346: the time dependence and $\phi-$dependence of the perturbations can be given by
347: $\exp(i\sigma t+i m\phi)$ with $\sigma$ being the oscillation frequency observed 
348: in an inertial frame, and $m$ is an integer representing the azimuthal wavenumber.
349: Introducing the vectors $\pmb{\xi}^n$, $\pmb{\xi}^p$, and $\pmb{\xi}^e$
350: defined as
351: %
352: \begin{equation}
353: \pmb{\xi}^n\equiv{\tilde{\pmb{j}}_n^\prime\over i\omega\rho_n},
354: \quad
355: \pmb{\xi}^p\equiv{\tilde{\pmb{j}}_p^\prime\over i\omega\rho_p},
356: \quad {\rm and} \quad
357: \pmb{\xi}^e\equiv {\pmb{v}_e^\prime\over i\omega},
358: \end{equation}
359: %
360: where $\omega\equiv\sigma+m\Omega$ 
361: is the oscillation frequency observed in a
362: corotating frame of the star,
363: the mass conservation equations are linearised to be
364: %
365: \begin{equation}
366: \rho_n^\prime+\nabla\cdot\left(\rho_n\pmb{\xi}^n\right)=0,
367: \end{equation}
368: %
369: and
370: %
371: \begin{equation}
372: \rho_p^\prime+\nabla\cdot\left(\rho_p\pmb{\xi}^p\right)=0.
373: \end{equation}
374: %
375: Note that the mass conservation equation for the electron fluid becomes the same as that for
376: the proton fluid because of the assumption of perfect charge neutrality, that is, 
377: $\pmb{\xi}^p=\pmb{\xi}^e$, and $\rho_p^\prime/m_p=\rho_e^\prime/m_e$.
378: The velocity equations (6) and (8) are linearized as
379: %
380: \begin{equation}
381: {\rho_{11}\over\rho_n}\pmb{F}\left(\pmb{\xi}^n\right)+
382: {\rho_{12}\over\rho_n}\pmb{F}\left(\pmb{\xi}^p\right)=
383: -\nabla\left(\mu_n^\prime+\Psi^\prime\right)
384: +i\omega{\rho_{np}\rho_p\over\tilde\rho^2}\left(\pmb{\xi}^p-\pmb{\xi}^n\right)
385: \times\left(\nabla\times\pmb{v}_0\right),
386: \end{equation}
387: %
388: and
389: %
390: \begin{equation}
391: {\rho_{21}\over\rho_p}\pmb{F}\left(\pmb{\xi}^n\right)+
392: \left({\rho_{22}\over\rho_p}+{m_e\over m_p}\right)\pmb{F}\left(\pmb{\xi}^p\right)=
393: -\nabla\left(\tilde\mu^\prime+\zeta\Psi^\prime\right)
394: -i\omega{\rho_{np}\rho_n\over\tilde\rho^2}\left(\pmb{\xi}^p-\pmb{\xi}^n\right)
395: \times\left(\nabla\times\pmb{v}_0\right),
396: \end{equation}
397: %
398: where $\tilde\mu\equiv\mu_p+\mu_e{m_e}/m_p$, and 
399: %
400: \begin{equation}
401: \pmb{F}\left(\pmb{\xi}\right)\equiv-\omega\sigma\pmb{\xi}+i\omega\pmb{v}_0\cdot\nabla\pmb{\xi}
402: +i\omega\pmb{\xi}\cdot\nabla\pmb{v}_0.
403: \end{equation}
404: %
405: The Poisson equation is reduced to
406: %
407: \begin{equation}
408: \nabla^2\Psi^\prime=4\pi G\left(\rho_n^\prime+\zeta\rho_p^\prime\right).
409: \end{equation}
410: %
411: Using a variant of Gibbs-Duhem relation,
412: the pressure perturbation is given by
413: %
414: \begin{equation}
415: p^\prime=\rho_n\mu_n^\prime+\rho_p\tilde\mu^\prime, 
416: \end{equation}
417: %
418: where we have again ignored the entropy carried by the electron normal fluid.
419: Note also that the superfluids carry no entropy.
420: 
421: To obtain a relation between the densities $\rho_n$ and $\rho_p$ and the chemical potentials
422: $\mu_n$ and $\tilde\mu$, we begin with writing the energy density $e$ as
423: %
424: \begin{equation}
425: e=e\left(\rho_n,\rho_p\right), 
426: \end{equation}
427: %
428: with which the chemical potentials are defined as
429: %
430: \begin{equation}
431: \mu_n\left(\rho_n,\rho_p\right)=\left(\partial e/\partial \rho_n\right)_{\rho_p}, \quad
432: \mu_p\left(\rho_n,\rho_p\right)=\left(\partial e/\partial \rho_p\right)_{\rho_n}.
433: \end{equation}
434: %
435: If the chemical potential of the electron is given by
436: %
437: \begin{equation}
438: \mu_e(\rho_e)=c^2\sqrt{1+(3\pi^2\hbar^3\rho_e/m_e^4c^3)^{2/3}}, 
439: \end{equation}
440: %
441: we have, assuming $\rho_e=\rho_pm_e/m_p$,
442: %
443: \begin{equation}
444: \left(\matrix{\mu_n^\prime \cr \tilde\mu^\prime\cr}\right)
445: =\left(\matrix{{\cal P}_{11} & {\cal P}_{12}\cr {\cal P}_{21} & {\cal P}_{22}\cr}\right)
446: \left(\matrix{\rho_n^\prime \cr \rho_p^\prime \cr}\right),
447: \end{equation}
448: %
449: where
450: %
451: \begin{equation}
452: {\cal P}_{11}=\left({\partial\mu_n\over \partial\rho_n}\right)_{\rho_p}, \quad
453: {\cal P}_{12}=\left({\partial\mu_n\over\partial\rho_p}\right)_{\rho_n}
454: =\left({\partial\tilde\mu\over\partial\rho_n}\right)_{\rho_p}={\cal P}_{21}, \quad
455: {\cal P}_{22}=\left({\partial\tilde\mu\over\partial\rho_p}\right)_{\rho_n}.
456: \end{equation}
457: %
458: We write the inverse of equation (26) as
459: %
460: \begin{equation}
461: \left(\matrix{\rho_n^\prime \cr \rho_p^\prime \cr}\right)=
462: \left(\matrix{{\cal Q}_{11} & {\cal Q}_{12}\cr {\cal Q}_{21} & {\cal Q}_{22}\cr}\right)
463: \left(\matrix{\mu_n^\prime \cr \tilde\mu^\prime\cr}\right),
464: \end{equation}
465: %
466: where
467: %
468: \begin{equation}
469: \left(\matrix{{\cal Q}_{11} & {\cal Q}_{12}\cr {\cal Q}_{21} & {\cal Q}_{22}\cr}\right)
470: =\left(\matrix{{\cal P}_{11} & {\cal P}_{12}\cr {\cal P}_{21} & {\cal P}_{22}\cr}\right)^{-1}.
471: \end{equation}
472: %
473: 
474: To represent the rotationally deformed equipotential surfaces of a rotating star,
475: we employ a coordinate system $(a,\theta,\phi)$, the relation of which
476: to spherical polar coordinates $(r,\theta,\phi)$ is given by
477: $r=a[1+\epsilon(a,\theta)]$,
478: where $\epsilon$ is proportional to $\Omega^2$ and represents a small deviation of the
479: equipotential surface from the corresponding spherical equipotential surface of 
480: the non-rotating star.
481: We apply Chandrasekhar-Milne expansion (Chandrasekhar 1933a,b, 
482: Tassoul 1978) to the hydrostatic equations to determine the function $\epsilon$
483: in the form $\epsilon(a,\theta)=\alpha(a)+\beta(a)P_2(\cos\theta)$ 
484: with $P_2$ being the Legendre function.
485: See Lee (1993) for the definition of $\alpha(a)$ and $\beta(a)$.
486: Since for uniformly rotating stars
487: all the physical quantities in hydrostatic equilibrium 
488: are constant on the deformed equipotential surface,
489: labeled by the coordinate $a$, we write
490: the linearised basic equations using the coordinates $(a,\theta,\phi)$
491: (Saio 1981, Lee 1993).
492: In our formulation, the terms up to of order of $\Omega^3$ in the perturbed velocity 
493: equations are retained so that the eigenfrequencies of the $r$-modes are correctly
494: determined to the order of $\Omega^3$ (see Yoshida \& Lee 2000a).
495: 
496: 
497: The perturbations in a uniformly rotating star are expanded in terms of spherical harmonic
498: functions with different $l$'s for a given $m$ (e.g., Lee \& Saio 1986).
499: For example, the vector $\pmb{\xi}^n$ is given by
500: %
501: \begin{equation}
502: {\xi_a^n\over a}=\sum_{l\geq |m|} S_l^n(a)Y_l^m(\theta,\phi)e^{i\sigma t},
503: \end{equation}
504: %
505: \begin{equation}
506: {\xi_\theta^n\over a}=\sum_{l\geq|m|}\left[H_l^n(a){\partial\over\partial\theta}
507: Y_l^m(\theta,\phi)+T_{l^\prime}^n(a){1\over\sin\theta}{\partial\over\partial\phi}
508: Y_{l^\prime}^m(\theta,\phi)\right]e^{i\sigma t},
509: \end{equation}
510: %
511: \begin{equation}
512: {\xi_\phi^n\over a}=\sum_{l\geq|m|}\left[H_l^n(a){1\over\sin\theta}{\partial\over
513: \partial\phi}Y_l^m(\theta,\phi)-T_{l^\prime}^n(a){\partial\over\partial\theta}
514: Y_{l^\prime}^m(\theta,\phi)\right]e^{i\sigma t},
515: \end{equation}
516: %
517: and the Euler perturbation of the gravitational potential, $\Psi^\prime$ is given as
518: \begin{equation}
519: \Psi^\prime=\sum_{l\geq|m|}\Psi^\prime_l(a)Y_l^m(\theta,\phi)
520: e^{i\sigma t},
521: \end{equation}
522: %
523: where $l=|m|+2(k-1)$ and $l'=l+1$ for even modes, and $l=|m|+2k-1$ and $l'=l-1$
524: for odd modes, and $k=1,~2,~3,~\cdots$.
525: Substituting these expansions and the like
526: into the linearised basic equations (16) $\sim$ (19) and (21), 
527: we obtain oscillation equations given as a set of
528: simultaneous linear ordinary differential equations of the expansion coefficients
529: (see Appendix), which is to be integrated in the superfluid core.
530: The oscillation equations solved in the normal fluid envelope are the same as those
531: given in Yoshida \& Lee (2000a).
532: 
533: To obtain a complete solution of an oscillation mode, solutions in
534: the superfluid core and in the normal fluid envelope are matched at the interface between 
535: the two domains by imposing jump conditions given by
536: %
537: \begin{equation}
538: \xi_a=\xi_a^n, \quad \xi_a=\xi_a^p, \quad
539: [p^\prime]^+_-=0, \quad [\Psi^\prime]^+_-=0,
540: \quad {\rm and} \quad [d\Psi^\prime/da]^+_-=0, 
541: \end{equation}
542: %
543: where 
544: $[f(x)]^+_-\equiv\lim_{s\rightarrow 0}\{f(x+s)-f(x-s)\}$.
545: The boundary conditions at the stellar center is 
546: the regularity condition of the perturbations
547: $\xi_a^n$, $\xi_a^p$, $\mu_n^\prime/g$, $\tilde\mu^\prime /g$, and $\Psi^\prime/g$, 
548: where $g=GM_a/a^2$.
549: The boundary conditions at the stellar surface 
550: are $\delta p=0$ and $\Psi_l^\prime\propto a^{-(l+1)}$,
551: where $\delta p$ is the Lagrangian perturbation of the pressure.
552: 
553: 
554: For numerical computation, oscillation equations of a finite dimension are obtained
555: by disregarding the terms
556: with $l$ larger than $l_{max}=|m|+2k_{max}-1$ in the expansions of perturbations
557: such as given by (30) to (33).
558: For the $r$-modes with $l^\prime=|m|$ calculated in this paper,
559: we usually use $k_{max}=6$ so that we can get reasonable convergence of
560: the eigenfrequency and the eigenfunction.
561: We solve the oscillation equations of a finite dimension as an eigenvalue problem of the 
562: oscillation frequency $\omega$ using a Henyey type relaxation method
563: (see, e.g., Unno et al 1989).
564: 
565: 
566: 
567: \section{Dissipations}
568: 
569: The stability of an oscillation mode of a star is determined by summing up all contributions 
570: from various damping and excitation mechanisms.
571: If we consider the contributions from gravitational radiation reaction, 
572: viscous processes, and mutual friction in the superfluid core,
573: the energy loss (or gain) rate $dE/dt$ of a normal mode in a rotating neutron
574: star may be given by 
575: %
576: \begin{eqnarray}
577: {dE\over dt}=&&-\sigma\omega\sum_{l=2}^\infty N_l\sigma^{2l}
578: \left(\left|D_{lm}^\prime\right|^2+\left|J_{lm}^\prime\right|^2\right)
579: -\int d^3x \left(\sum_{ij}{\sigma^{\prime ij}
580: \sigma_{ij}^{\prime*}\over 2\zeta_S}+\zeta_B\left|\nabla\cdot\pmb{v}^\prime\right|^2\right) 
581: \nonumber \\
582: &&-2\Omega\int d^3x\rho_nB_n\left({\tilde\rho^2\over\rho_n\rho_p}\right)^2
583: \left(\pmb{w}^\prime\cdot\pmb{w}^{\prime *}-w_z^\prime w_z^{\prime *}\right) \nonumber\\
584: =&& \left({dE\over dt}\right)_{GD}+\left({dE\over dt}\right)_{GJ}+\left({dE\over dt}\right)_{S}
585: +\left({dE\over dt}\right)_{B}+\left({dE\over dt}\right)_{MF},
586: \end{eqnarray}
587: %
588: where the asterisk $(^*)$ indicates the complex conjugate of the quantity, and
589: the canonical energy $E$ of oscillation observed in the corotating frame of the star is
590: defined in the normal fluid envelope as (Friedman \& Schutz 1978)
591: %
592: \begin{equation}
593: E={1\over 2}\int d^3x \left(\rho\pmb{v}^\prime\cdot\pmb{v}^{\prime *}+
594: {p^\prime\over\rho}\rho^{\prime *}-{\nabla\Psi^\prime\cdot\nabla\Psi^{\prime *}\over 4\pi G}
595: \right),
596: \end{equation}
597: %
598: and in the superfluid core as (Mendell 1991b)
599: %
600: \begin{equation}
601: E={1\over 2}\int d^3x\left(\rho\pmb{v}^\prime\cdot\pmb{v}^{\prime *}+
602: {\tilde\rho^2\over\rho}\pmb{w}^\prime\cdot\pmb{w}^{\prime *}+
603: \sum_{ij}{\cal P}_{ij}\rho^\prime_i\rho^{\prime *}_j
604: -{\nabla\Psi^\prime\cdot\nabla\Psi^{\prime *}\over 4\pi G}\right),
605: \end{equation}
606: %
607: where 
608: $\pmb{v}^\prime=(\rho_n\pmb{v}_n^\prime+\rho_p\pmb{v}_p^\prime)/\rho$, 
609: $\pmb{w}^\prime=\pmb{v}_p^\prime-\pmb{v}_n^\prime$,   
610: $\rho^\prime_1=\rho^\prime_n$, and $\rho^\prime_2=\rho^\prime_p$ in the core.
611: 
612: The terms $(dE/dt)_{GD}$ and $(dE/dt)_{GJ}$ on the right hand side of equation (35)
613: denote the energy loss (or gain) rates due to gravitational
614: radiation reaction associated with the mass multipole moment $D_{lm}^\prime$ and
615: the mass current multipole moment $J_{lm}^\prime$, where
616: %
617: \begin{equation}
618: D_{lm}^\prime=\int d^3x\rho^\prime r^l Y_l^{m*}, 
619: \end{equation}
620: %
621: % 
622: \begin{equation}
623: J_{lm}^\prime={2\over c(l+1)}
624: \int d^3x r^l\left(\rho\pmb{v}^\prime+\rho^\prime\pmb{v}_0\right)\cdot
625: \left(\pmb{r}\times\nabla Y_l^{m*}\right),
626: \end{equation}
627: %
628: and
629: %
630: \begin{equation}
631: N_l={4\pi G\over c^{2l+1}}{(l+1)(l+2)\over l(l-1)[(2l+1)!!]^2}, 
632: \end{equation}
633: %
634: and $c$ is the velocity of light (Thorne 1980, Lindblom et al 1998).
635: 
636: The terms $(dE/dt)_{S}$ and $(dE/dt)_B$ are the energy dissipation rates 
637: due to shear and bulk viscosities, 
638: and $\zeta_S$ and $\zeta_B$ are the shear and the bulk viscosity coefficients, 
639: and $\sigma_{ij}^\prime$ is 
640: the traceless stress tensor for the perturbed velocity field
641: (e.g., Landau \& Lifshitz 1987).
642: In this paper, we ignore the contribution from the bulk viscosity, which is
643: important only for newly born hot neutron stars without superfluids in the core.
644: The shear viscosity coefficient we use in the superfluid core is
645: %
646: \begin{equation}
647: \zeta_S=6\times 10^{18}\left({\rho\over 10^{15}{\rm g/cm^3}}\right)^2
648: \left({10^9 K\over T}\right)^2 {\rm g/cm ~s}
649: \end{equation}
650: %
651: (Cutler \& Lindblom 1987, Sawyer 1989), and that in the normal fluid envelope is given by
652: %
653: \begin{equation}
654: \zeta_S=2\times 10^{18}\left({\rho\over 10^{15}{\rm g/cm^3}}\right)^{9/4}
655: \left({10^9 K\over T}\right)^2 {\rm g/cm ~s}
656: \end{equation}
657: %
658: (Cutler \& Lindblom 1987, Flowers \& Itoh 1979).
659: The stress tensor $\sigma_{ij}^\prime$ is evaluated by using $\pmb{v}_e^\prime$ 
660: in the superfluid core (Lindblom \& Mendell 2000).
661: 
662: The term $(dE/dt)_{MF}$ is the energy loss rate due to 
663: mutual friction in the superfluid core, 
664: and the dimensionless coefficient $B_n$ is given by (Mendell 1991b)
665: %
666: \begin{equation}
667: B_n=0.011\times {\rho_p\over\rho_n}
668: \left({\rho_{pp}\over\rho_{p}}\right)^{1/2}
669: \left({\rho_{pn}\over\rho_{pp}}\right)^2
670: \left({\rho_p\over 10^{14}{\rm gcm^{-3}}}\right)^{1/6}.
671: \end{equation}
672: %
673: Mutual friction is a dissipation mechanism inherent to a rotating system 
674: of superfluids, and it is caused by scattering of normal fluid particles off the vortices
675: in the superfluids.
676: Since we have assumed perfect charge neutrality between the electrons and protons, we consider
677: scattering between the normal electrons and vortices of the neutron superfluid
678: (Mendell 1991b, see also Alpar etal 1984).
679: 
680: As is indicated by the first two terms on the right hand side of equation (35), 
681: if a normal mode has an oscillation frequency that satisfies
682: $-\sigma\omega>0$, 
683: the oscillation energy $E$ in the corotating frame, in the absence of other damping mechanisms, 
684: increases as a result of
685: gravitational wave radiation, indicating instability of the mode (Friedman \& Schutz 1978).
686: It was Andersson (1998) and Friedman \& Morsink (1998) who realized that the $r$-modes have
687: oscillation frequencies that satisfy the instability condition.
688: 
689: 
690: 
691: 
692: The damping (or growth) time-scale $\tau$ of a normal mode may be given by
693: \begin{equation}
694: {1\over\tau}={1\over 2E}\left({dE\over dt}\right)={1\over\tau_{GD}}+{1\over\tau_{GJ}}+
695: {1\over\tau_S}+{1\over\tau_{MF}},
696: \end{equation}
697: %
698: where $\tau_i=2E/(dE/dt)_i$. For the $r$-modes of $l^\prime =|m|$, it is convenient to
699: derive an extrapolation formula of the time-scale $\tau$ given as a function of $\Omega$
700: and the interior temperature $T$ (e.g., Lindblom et al 1998, Lindblom \& Mendell 2000):
701: %
702: \begin{equation}
703: {1\over\tau (\Omega,T)}
704: ={1\over\tau^0_{GD}}\left({\Omega^2\over \pi G\bar\rho}\right)^{l+2}
705: +{1\over\tau^0_{GJ}}\left({\Omega^2\over \pi G\bar\rho}\right)^{l}
706: +{1\over\tau^0_S}\left({10^9 K\over T}\right)^2
707: +{1\over\tau^0_{MF}}\left({\Omega^2\over \pi G\bar\rho}\right)^\gamma,
708: \end{equation}
709: %
710: where $\bar\rho=M/(4\pi R^3/3)$, and 
711: only the dominant term in each of the dissipation processes with $l=|m|+1=3$
712: has been retained for the $r$-modes. 
713: The quantities $\tau^0_{GD}$, $\tau^0_{GJ}$, $\tau^0_S$, and $\tau^0_{MF}$
714: are assumed to be only weakly dependent on $\Omega$ and $T$. 
715: 
716: 
717: 
718: 
719: \section{Numerical Results}
720: 
721: 
722: Following Lindblom \& Mendell (2000), 
723: we employ a polytropic model of index $N=1$ with the mass $M=1.4M_\odot$ and the radius $R=12.57$km
724: as a background model for modal analysis.
725: The model is divided into a superfluid core and a normal fluid
726: envelope, the interface of which is set at $\rho=\rho_s=2.8\times 10^{14}$g/cm$^3$.
727: In the normal fluid envelope, we use the polytropic equation of state given by $p=K\rho^2$
728: both for the equilibrium structure and for the oscillation equations.
729: The core is assumed to be filled with neutron and proton superfluids and a normal fluid of electron,
730: for which an equation of state, labeled
731: A18+$\delta v$+UIX (Akmal, Pandharipande, and Ravenhall 1998), is used to
732: give the energy density (23) and the relation (26) used for the oscillation equations.
733: For the mass density coefficients $\rho_{nn}$, $\rho_{pp}$, and $\rho_{np}$ in the core,
734: we employ an empirical relation given by (see Lindblom \& Mendell 2000)
735: %
736: \begin{equation}
737: {\rho_p/\rho}\approx 0.031+8.8\times 10^{-17}\rho, 
738: \end{equation}
739: %
740: and a formula given by
741: %
742: \begin{equation}
743: \rho_{np}=-\eta\rho_n, 
744: \end{equation}
745: %
746: where $\eta$ is a parameter of order of $\sim 0.04$ that parametrizes the entrainment effects
747: between the two superfluids
748: (Borumand, Joynt, \& Klu\'zniak 1996).
749: 
750: We find that the $r$-modes of $l^\prime=m$ in the superfluid core are split into
751: ordinary $r$-modes and superfluid $r$-modes, which we call $r^o$-modes and $r^s$-modes.
752: The toroidal components $iT_{l^\prime}$ of the $r^o$- and $r^s$-modes of $l^\prime=m=2$
753: are plotted versus $a/R$ for the case of $\eta=0.04$ and $\bar\Omega\equiv\Omega/\sqrt{GM/R^3}=0.01$ 
754: in Figure 1, where
755: the solid, dotted, dashed, and dash-dotted curves are used to indicate respectively
756: the toroidal components $iT_m^n$, $iT_{m+2}^n$, $iT_m^p$, and $iT_{m+2}^p$ 
757: in the superfluid core, and
758: the amplitude normalization is given by $\max(|iT_m|)=1$.
759: The figure shows that the two superfluids in the core flow together for the $r^o$-modes, while
760: they counter-move for the $r^s$-modes.
761: Note that the amplitudes of $iT_{m+2}^p$ for the $r^s$-mode
762: are not necessarily negligibly small compared with
763: those of $iT_m^p$ in the core.
764: This splitting of the $r$-modes in the superfluid core into two distinct families has been
765: suggested by Andersson \& Comer (2001).
766: For the $r^o$-mode, the radial dependence of the difference $iT_m^p-iT_m^n$ in the core
767: is given in Figure 2, which shows that the difference is quite small 
768: compared with the normalization $\max(|iT_m|)=1$.
769: It is important to note that
770: the $l^\prime=m$ $r$-modes having basically nodeless and dominant $iT_m$ are the 
771: only $r$-modes we can find, as in the case of
772: the $r$-modes in isentropic models (see Yoshida \& Lee 2000a,b).
773: 
774: 
775: 
776: In Table 1, the expansion coefficients $\kappa_0$ and $\kappa_2$, and the scaled
777: damping (or growth) timescales $\tau^0_i$ for the $l'=m=2$ $r$-modes are tabulated
778: for the cases of $\eta=0$, $0.02$, $0.04$, and $0.06$, 
779: where the coefficients $\kappa_0$ and $\kappa_2$ are defined in the expansion:
780: %
781: \begin{equation}
782: \omega/\Omega=\kappa_0(\eta)+\kappa_2(\eta)\bar\Omega^2+O(\bar\Omega^4).
783: \end{equation}
784: %
785: Note that the exponent $\gamma$ employed to define $\tau^0_{MF}$ in equation (45) is 
786: $\gamma=2.5$ for the $r^o$-modes 
787: and $\gamma=0.5$ for the $r^s$-modes for the non-zero $\eta$'s in the table.
788: This is because the velocity difference 
789: $\pmb{w}^\prime=\pmb{v}^\prime_p-\pmb{v}^\prime_n$ approximately
790: scales as $\pmb{w}^\prime\propto\Omega^3$ for
791: the $r^o$-modes and as $\pmb{w}^\prime\propto\Omega$ for the $r^s$-modes 
792: (see Lindblom \& Mendell 2000).
793: The coefficient $\kappa_0$ for the $r^o$-modes is numerically equal to 
794: $2m/[l^\prime(l^\prime+1)]$, 
795: and $\kappa_2$ is almost independent of the entrainment parameter $\eta$.
796: The value of $\kappa_0$ is the same as the value found by Lindblom \& Mendell (2000)
797: and the value of $\kappa_2$ differs only by 2.5\%, suggesting that the $r^o$-modes are
798: the same modes found by Lindblom \& Mendell (2000).
799: (Because of different normalization conventions the value of $\kappa_2$ reported by
800: Lindblom \& Mendell must be multiplied by a factor of $4/3$ before comparing
801: with our results.)
802: The coefficients $\kappa_0$ and $\kappa_2$ for the $r^s$-modes, on the other hand, 
803: appreciably depends on $\eta$, and $\kappa_0$ deviates from $2m/[l^\prime(l^\prime+1)]$ 
804: as $\eta$ is increased from $\eta=0$.
805: This kind of deviation of $\kappa_0$ from $2m/[l^\prime(l^\prime+1)]$ for the $l^\prime=m$ $r$-modes
806: has been found for relativistic neutron stars where
807: the relativistic factor $GM/c^2R$ is regarded as a parameter (Yoshida 2001; Yoshida \& Lee 2002a).
808: Computing the $r^s$-mode of $l^\prime=m=2$
809: as a function of $\eta$, we find that a linear formula given by
810: %
811: \begin{equation}
812: \kappa_0^s(\eta)\approx0.667+9.35\eta
813: \end{equation}
814: %
815: gives a good fit to the $r^s$-mode frequency except at avoided crossings with inertial modes
816: (see below).
817: At $\eta=0$, the coefficients $\kappa_0$ for the $r^o$- and $r^s$-modes are both equal to 
818: $2m/[l^\prime(l^\prime+1)]$, which
819: suggests that in the lowest order of $\Omega$ the two $r$-modes are 
820: degenerate at $\eta=0$.
821: In Figure 3, the toroidal components $iT_m$ and $iT_{m+2}$ 
822: of the $r^o$- and $r^s$-modes of $l^\prime=m=2$ at
823: $\bar\Omega=0.01$ are given versus $a/R$ for the case of $\eta=0$.
824: The amplitudes of $iT_{m+2}$ are much smaller than those of $iT_m$ 
825: both for the $r^o$- and $r^s$-modes.
826: This figure also shows that $|iT_m^p-iT_m^n|\not\ll 1$ in the core
827: for the $r^o$-mode at $\eta=0$, for which we find that $\pmb{w}^\prime\propto\Omega$.
828: 
829: 
830: 
831: Inertial modes in the superfluid core are also split into 
832: ordinary and superfluid inertial modes, which we call $i^o$- and $i^s$-modes.
833: In Table 2, we have given $\kappa_0$ for $i^o$- and $i^s$-modes for $m=2$ and $\eta=0$, and
834: see, e.g., Lockitch \& Friedman (1999) and Yoshida \& Lee (2000a) for 
835: the classification scheme employed here for inertial modes.
836: For given $m$ and $l_0-|m|$,
837: we find at $\eta=0$ pairs of $i^o$- and $i^s$-modes that have close values of $\kappa_0$,
838: and the number of the pairs is equal to $l_0-|m|$.
839: As an example, 
840: for the case of $m=2$ and $\eta=0$, the eigenfunctions $iT_{l^\prime}$
841: are shown for the $i^o$- and $i^s$-modes 
842: of $\kappa_0=0.5180$ and $\kappa_0=0.5077$ in Figure 4, 
843: for those of $\kappa_0=0.4215$ and $\kappa_0=0.4060$ in Figure 5, and
844: for those of $\kappa_0=1.1046$ and $\kappa_0=1.1134$ in Figure 6.
845: The inertial modes in Figure 4 belong to $l_0-|m|=3$ and those in Figures 5 and 6 to
846: $l_0-|m|=5$.
847: Note that
848: the $r^o$- and $r^s$-modes belong to $l_0-|m|=1$ (e.g., Yoshida \& Lee 2000a).
849: It is generally observed for inertial modes with long radial wavelengths
850: that the two superfluids co-move in the core
851: for the $i^o$-modes, and they counter-move for the $i^s$-modes.
852: The coefficient $\kappa_0$ for the $i^o$-modes only weakly
853: depends on the entrainment parameter $\eta$ (see Figure 7).
854: The coefficient $\kappa_0$ for the $i^s$-modes, on the other hand, 
855: increases approximately linearly as $\eta$ is increased from $\eta=0$ (see Figure 8).
856: See Yoshida \& Lee (2002b) for extended discussions on inertial modes in the superfluid core.
857: 
858: 
859: The $r^s$-modes ($r^o$-modes) experience mode crossings with $i^o$-modes ($i^s$-modes)
860: as the parameter $\eta$ is increased from $\eta=0$.
861: For the case of $m=2$ and $\bar\Omega=0.01$, Figure 7 illustrates an avoided crossing
862: between the $r^s$-mode of $l^\prime=m$ and the $i^o$-mode that tends to 
863: $\kappa_0=1.1046$ as $\eta\rightarrow0$.
864: In this figure, the dashed line is given by equation (49).
865: For mode crossings between the $r^o$-mode and $i^s$-modes, on the other hand,
866: it is quite difficult to numerically discern whether the mode crossings result in
867: avoided crossing or degeneracy of the mode frequencies at the crossing point.
868: Most prominent among such mode crossings of the $r^o$-mode of $l^\prime=m=2$
869: are those with the $i^s$-modes that tend to 
870: $\kappa_0=0.5077$ and $\kappa_0=0.4060$ as $\eta\rightarrow 0$.
871: For the case of $m=2$ and $\bar\Omega=0.01$,
872: the two mode crossings which occur at 
873: $\eta\approx0.0230$ and 0.0484 are shown in Figure 8, where the solid lines and the dashed line 
874: are for the $i^s$-modes and the $r^o$-mode, respectively.
875: We note that the ratio $\omega/\Omega\approx \kappa_0$ for the $r^o$-mode is almost constant, 
876: equal to $2m/[l^\prime(l^\prime+1)]=0.6667$ for $l^\prime=m=2$, while the ratios
877: $\omega/\Omega$ for the $i^s$-modes increase approximately linearly with increasing $\eta$.
878: Figure 9 shows the differences $iT_m^p-iT_m^n$ versus $a/R$ for the $l^\prime=m=2$ $r^o$-modes
879: at $\eta=0.023$ (panel a) and 0.0484 (panel b), and Figure 10 the eigenfunctions $iT_{l^\prime}$
880: for the $i^s$-modes of $m=2$ at $\eta=0.023$ (panel a) and at $\eta=0.0484$ (panel b), 
881: where we have assumed $\bar\Omega=0.01$ and the normalization $\max(|iT_m|)=1$.
882: The amplitudes of the differences at the mode crossing points are much larger than 
883: amplitudes of the difference off the crossing points (see, e.g., Figure 2).
884: The resemblance between the $i^s$-modes of $l_0-|m|=3$ at $\eta=0$ (Figure 4b)
885: and at $\eta=0.023$ (Figure 10a) and between
886: the $i^s$-modes of $l_0-|m|=5$ at $\eta=0$ (Figure 5b) and $\eta=0.0484$ (Figure 10b)
887: is obvious.
888: 
889: 
890: 
891: 
892: >From Table 1, we find that
893: the growth timescales $\tau^0_{GJ}$ and $\tau^0_{GD}$ for
894: the $r^o$-modes of $l^\prime=m=2$ are almost the same as those for the $l^\prime=m=2$ $r$-modes 
895: of the $N=1$ polytropic model without superfluidity in the core 
896: (see, e.g., Yoshida \& Lee 2000a).
897: On the other hand, the growth timescales $\tau^0_{GJ}$ and $\tau^0_{GD}$ for the
898: $r^s$-modes are much longer than those of the $r^o$-modes, and
899: the instability caused by the $r^s$-modes
900: is much weaker than the instability by the $r^o$-modes.
901: This is because the amount of gravitational wave radiation 
902: emitted from the $r^s$-modes is much smaller than that from the $r^o$-modes
903: since $(\rho_piT_m^p+\rho_niT_m^n)/\rho\sim 0$ and 
904: $\rho^\prime\sim 0$ for the $r^s$-modes.
905: It is interesting to note that $\tau^0_{GJ}$ of the $r^s$-modes at $\eta\not=0$ is 
906: by several orders of magnitude longer than that at $\eta=0$.
907: Figure 11 shows, as a function of $a/R$, $\sqrt{|\omega\sigma^{2l+1}|N_l/2E}J_{mm}^\prime(a)$ for
908: the $r^s$-modes at $\eta=0$ (dashed line) and at $\eta=0.04$ (solid line), where
909: $J_{mm}^\prime(a)=\int_0^adadJ_{mm}^\prime/da$.
910: As found for the case of $\eta=0.04$, the negative contributions to
911: $J_{mm}^\prime(a=R)$ in the envelope almost completely cancel out the positive contributions 
912: in the core, which leads to extremely long growth timescales $\tau^0_{GJ}$
913: of the $r^s$-modes at $\eta\not=0$.
914: 
915: 
916: Figure 12 illustrates the dependence of $-\tau^0_{MF}$ on the entrainment parameter $\eta$ for
917: the $l^\prime=m=2$ $r^o$-mode, and we find that
918: $-\tau^0_{MF}$ has prominent and deep minimums at $\eta\approx0.230$ and $0.484$.,
919: which is consistent with the result by Lindblom \& Mendell (2000).
920: Note that we have assumed $\rho_s=2.8\times 10^{14}$g/cm$^3$ for the interface between the
921: core and the envelope.
922: These prominent minimums result from the mode crossings with the long radial wavelength
923: $i^s$-modes that tend to $\kappa_0=0.5077$ and $\kappa_0=0.4060$ as $\eta\rightarrow 0$ 
924: (Figure 8; see also Andersson \& Comer 20001).
925: Mode crossings of the $r^o$-mode with $i^s$-modes that have much shorter radial wavelengths result
926: in narrow and shallow minimums of $-\tau^0_{MF}$, as found for $\eta\gtsim 0.05$.
927: %
928: These minor dips were not found by Lindblom \& Mendell (2000).
929: We guess that the $r$-modes calculated by Lindblom \& Mendell (2000) 
930: are less strongly affected by inertial modes associated with large $l_0-|m|$ having
931: short radial wavelengths since they ignored terms higher than $\Omega^4$ in their 
932: perturbative method.
933: %
934: Comparing our results with those by Lindblom \& Mendell (2000), we find that
935: the damping timescale $-\tau^0_{MF}$ at a given $\eta$ is about by an order of magnitude
936: shorter than that they obtained, and that our values of $\eta\approx0.0230$ and 0.0484
937: for the local minimums of $-\tau^0_{MF}$ are slightly larger than their $critical$ values 
938: 0.02294 and 0.04817.
939: From Table 1 we also find that the value of $\kappa_2$ found here for the $r^o$-modes
940: is about by 2.5\% larger than the value they found.
941: We guess that these differences between the
942: two calculations partly come from the differences in the way of evaluating the
943: derivatives of the thermodynamical quantities that appear in the oscillation equations.
944: For example, Lindblom \& Mendell (2000) used
945: $\rho^\prime=(\partial \rho/\partial p)_\beta p^\prime+(\partial\rho/\partial\beta)_p\beta^\prime$,
946: for which they assumed $(\partial \rho/\partial p)_\beta=(\rho/2p)$ from
947: the polytropic relation $p=K\rho^2$ and employed
948: a fitting formula to $(\partial\rho/\partial\beta)_p$ obtained from
949: Akmal et al's equation of state (1998), where $\beta=\tilde{\mu}-\mu_n$.
950: In this paper, on the other hand, we used
951: $\rho^\prime=({\cal Q}_{11}+\zeta{\cal Q}_{21})\mu_n^\prime
952: +({\cal Q}_{12}+\zeta{\cal Q}_{22})\mu_p^\prime$,
953: and the coefficients ${\cal Q}_{ij}$ were all calculated by using equations (23) to (29)
954: with Akmal et al's equation of state (1998).
955: 
956: 
957: 
958: The eigenfunction $\delta\beta(r,\theta,\phi)=\delta\beta_0(r,\theta,\phi)
959: +(4/3)\delta\beta_2(r,\theta,\phi)\bar\Omega^2+O(\bar\Omega^4)$ in Lindblom \& Mendell (2000) may be
960: given in terms of the eigenfunctions $y^{p}_{2,k}$ and $y^{n}_{2,k}$ as
961: \begin{equation}
962: \delta \beta(r,\theta,\phi)=(M_a/M)(R/a)\bar\Omega^{-2}
963: \sum_{k\ge 1}\Delta y_{2,k}(a)Y_{l_km}(\theta,\phi),
964: \end{equation}
965: where $r=a[1+\epsilon(a,\theta)]$, and $\Delta y_{2,k}(a)=y_{2,k}^p(a)-y_{2,k}^n(a)$ 
966: and see Appendix for definition of the functions $y_{2,k}^{p(n)}(a)$.
967: Assuming $\delta\beta_0=0$, we obtain
968: \begin{equation}
969: \delta \beta_2(r,\theta,\phi)\approx 0.75\times(M_r/M)(R/r)\bar\Omega^{-4}
970: \sum_{k\ge 1}\Delta y_{2,k}(r)f_{l_km}P_{l_k}^m(\cos\theta)e^{im\phi}
971: \equiv\sum_{k\ge1}\delta\beta_{2,k}P_{l_k}^m(\cos\theta)e^{im\phi},
972: \end{equation}
973: where the mean radial distance $a$ have been replaced by the radial distance $r$, and
974: the factor $f_{lm}$ is defined by the relation 
975: $Y_l^m(\theta,\phi)=f_{lm}P^m_l(\cos\theta)e^{im\phi}$, and $l_k=|m|+2k-1$.
976: In Figure 13, we plot the functions $\delta\beta_{2,k}(r)$ 
977: for the $r^o$-mode of $l^\prime=m=2$ at $\eta=0.04$, applying amplitude normalization
978: given by $y_{2,k=1}(R)=f_{m+1,m}\bar\Omega^2$ at the surface, 
979: where the solid, dotted, and dashed lines
980: are for $\delta\beta_{2,1}$, $100\times\delta\beta_{2,2}$, and $100\times\delta\beta_{2,3}$, 
981: respectively.
982: Since $|\delta\beta_{2,1}(r)|>>|\delta\beta_{2,k}(r)|$ for $k\ge2$, the $\theta$ depenence 
983: of the function 
984: $\delta\beta_2(r,\theta,\phi)$ is well represented by a single associated Legendre function 
985: $P^m_{m+1}(\cos\theta)$, and 
986: is not necessarily the same $\theta$ dependence of the function $\delta\beta_2$ 
987: found by Lindblom \& Mendell (2000).
988: The amplitude of the function $\delta\beta_{2,1}(r)$ is about by a factor of $3$ larger than
989: the amplitude of $\delta\beta_2$ calculated by Lindblom \& Mendell (2000), which is
990: consistent with the result that $\tau_{MF}^0$ in this paper 
991: is about by an order of magnitude
992: shorter than $\tau_{MF}^0$ they obtained.
993: 
994: \section{Discussion}
995: 
996: 
997: If we employ a set of the dependent variables defined as (see Lindblom \& Mendell 2000)
998: %
999: \begin{equation}
1000: \pmb{\xi}={\rho_n\pmb{v}_n^\prime+\rho_p\pmb{v}_p^\prime\over i\omega\hat\rho}, \quad
1001: \pmb{\xi}^w={\pmb{w}^\prime\over i\omega}, \quad 
1002: U={p^\prime\over\hat\rho}+\Psi^\prime, \quad
1003: \beta^\prime=\tilde{\mu}^\prime-\mu^\prime_n,
1004: \end{equation}
1005: %
1006: %={\rho_n\pmb{\xi}^n+\rho_p\pmb{\xi}^p\over\hat\rho}
1007: %={\rho_n\rho_p\over\tilde\rho^2}(\pmb{\xi}^p-\pmb{\xi}^n)
1008: %
1009: where $\hat\rho=\rho_n+\rho_p$,
1010: the continuity equations (16) and (17) and the velocity equations (18) and (19)
1011: are rewritten as
1012: %
1013: \begin{equation}
1014: \rho_n^\prime+\rho_p^\prime+\nabla\cdot\left(\hat\rho\pmb{\xi}\right)=0,
1015: \end{equation}
1016: %
1017: \begin{equation}
1018: {\rho_p^\prime\over\rho_p}-{\rho_n^\prime\over\rho_n}
1019: +\pmb{\xi}\cdot\nabla\ln{\rho_p\over\rho_n}
1020: +{\tilde\rho^2\over\rho_n\rho_p}\left(
1021: \pmb{\xi}^w\cdot\nabla\ln{\tilde\rho^2\over\hat\rho}+\nabla\cdot\pmb{\xi}^w\right)=0,
1022: \end{equation}
1023: %
1024: \begin{equation}
1025: \pmb{F}\left(\pmb{\xi}\right)
1026: =-\nabla U
1027: +{\rho_n\rho_p\over\hat\rho^2}\left(\nabla\ln{\rho_p\over\rho_n}\right)\beta^\prime,
1028: \end{equation}
1029: %
1030: \begin{equation}
1031: \pmb{F}\left(\pmb{\xi}^w\right)+i\omega{\rho_{np}\hat\rho\over\rho_n\rho_p}
1032: \pmb{D}\left(\pmb{\xi}^w\right)=-\nabla\beta^\prime,
1033: \end{equation}
1034: %
1035: and the linearized Poisson equation remains the same:
1036: %
1037: \begin{equation}
1038: \nabla^2\Psi^\prime=4\pi G\left(\rho_n^\prime+\rho_p^\prime\right),
1039: \end{equation}
1040: %
1041: where $\pmb{D}(\pmb{\xi})=\pmb{\xi}\times(\nabla\times\pmb{v}_0)$, and
1042: the terms of order of $m_e/m_p$ have been ignored for simplicity.
1043: The term proportional to $\pmb{D}\left(\pmb{\xi}^w\right)$ in equation (56) 
1044: represents the drag force between the two superfluids.
1045: In a perturbative treatment of the $r$-modes,
1046: we may expand the eigenfunctions and eigenfrequencies in terms of $\Omega$ as
1047: %
1048: \begin{equation}
1049: \pmb{\xi}=\pmb{\xi}_0+\pmb{\xi}_2\Omega^2+O(\Omega^4), \quad
1050: \pmb{\xi}^w=\pmb{\xi}_0^w+\pmb{\xi}_2^w\Omega^2+O(\Omega^4), 
1051: \end{equation}
1052: %
1053: \begin{equation}
1054: U=U_2\Omega^2+U_4\Omega^4+O(\Omega^6), \quad
1055: \beta^\prime=\beta_2^\prime\Omega^2+\beta_4^\prime\Omega^4+O(\Omega^6), 
1056: \end{equation}
1057: %
1058: and
1059: %
1060: \begin{equation}
1061: \omega=\kappa_0\Omega+\kappa_2\Omega^3+O(\Omega^5), \quad
1062: \sigma=s_0\Omega+s_2\Omega^3+O(\Omega^5),
1063: \end{equation}
1064: %
1065: where $|\pmb{\xi}_0|\sim |iT_m|\sim O(1)$.
1066: Note that the quantities like $\nabla\cdot\pmb{\xi}$,
1067: $\pmb{\xi}\cdot\nabla\rho$, $\rho_n^\prime$, 
1068: $\rho_p^\prime$, and $\Psi^\prime$ are of order of $\Omega^2$ in the
1069: lowest order for the $r$-modes.
1070: For the expansions given above, we have
1071: %
1072: \begin{equation}
1073: \pmb{F}(\pmb{\xi})=\pmb{F}_2(\pmb{\xi}_0,\kappa_0,s_0)\Omega^2
1074: +\pmb{F}_4(\pmb{\xi}_0,\pmb{\xi}_2,\kappa_0,\kappa_2,s_0,s_2)\Omega^4
1075: +O(\Omega^6), 
1076: \end{equation}
1077: %
1078: and we may not need to give the full expressions of $\pmb{F}_2$ and $\pmb{F}_4$ here.
1079: Equations (55) and (56) suggest the existence of two families of $r$-modes, that is,
1080: $r$-modes, which are governed by equation (55) and have
1081: $\kappa_0=2m/[l^\prime(l^\prime+1)]$, and $r$-modes, which are governed by equation (56)
1082: and have $\kappa_0\not=2m/[l^\prime(l^\prime+1)]$.
1083: If we assume $\pmb{F}_2(\pmb{\xi}_0)=-\nabla U_2$ in equation (55), we obtain 
1084: the lowest order $r$-mode solution given by $iT_m\propto a^{|m|-1}$ and
1085: $\kappa_0=2m/[l^\prime(l^\prime+1)]$, which leads to $\beta_2^\prime=0$ from equation (55),
1086: and to $\pmb{\xi}^w_0=0$ from equation (56).
1087: This suggests the existence of the $r$-mode solutions, for which 
1088: $\kappa_0=2m/[l^\prime(l^\prime+1)]$, and $\pmb{\xi}_0\not=0$ and $U_2\not=0$, but
1089: $\pmb{\xi}^w_0=0$ and $\beta^\prime_2=0$.
1090: The solutions of this kind correspond to 
1091: the $r$-modes Lindblom \& Mendell (2000) obtained in their perturbative treatment, and
1092: to the $r^o$-modes found in this paper.
1093: On the other hand, because of the drag force, i.e.,
1094: the term proportional to $\pmb{D}(\pmb{\xi}^w)$,
1095: $\kappa_0$ of the $r$-modes governed by equation (56)
1096: is not necessarily equal to $2m/[l^\prime(l^\prime+1)]$.
1097: For these solutions we may expect 
1098: $\pmb{\xi}^w_0\not=0$ and $\beta_2^\prime\not=0$ from equation (56), and 
1099: $\pmb{\xi}_0\not=0$ and $U_2\not=0$ from equation (55).
1100: The solutions of this kind
1101: correspond to the $r^s$-modes found in this paper (see also Andersson \& Comer 2001).
1102: 
1103: 
1104: To show the importance of the drag force for the existence of the $r$-mode solutions
1105: with the scaling $\pmb{w}^\prime\propto\Omega^3$, we calculate
1106: the $r^o$- and $r^s$-modes by ignoring the drag force terms proportional to $\pmb{D}(\pmb{\xi})$ 
1107: on the right hand side of the velocity equations (18) and (19).
1108: The results are summarized in Table 3, in which
1109: the coefficients $\kappa_0$, $\kappa_2$, and $\tau^0_i$'s are tabulated
1110: for the $r^o$- and $r^s$-modes of $l^\prime=m=2$ for the case of $\eta=0.04$.
1111: When we ignore the drag force terms, we can not find the $r^o$-modes with the
1112: scaling of $\pmb{w}^\prime\propto\Omega^3$ and
1113: the $r$-modes with the scaling $\pmb{w}^\prime\propto\Omega$ are the only $r$-mode
1114: solutions we can find, and hence
1115: the exponent $\gamma$ to define $\tau^0_{MF}$ in the table
1116: is equal to $0.5$ for both the $r^o$- and $r^s$-modes.
1117: In Figure 14, the toroidal components $iT_m$ and $iT_{m+2}$ of the two $r$-modes
1118: are given for $\eta=0.04$, where we have assumed $\bar\Omega=0.01$ and 
1119: the normalization $\max(|iT_m|)=1$.
1120: The amplitudes of $iT_{m+2}$ are completely negligible compared to those of $iT_m$.
1121: In the absence of the drag force,
1122: $\kappa_0$'s for the two $r$-modes are both equal to
1123: $2m/[l^\prime(l^\prime+1)]$, independent of $\eta$, which means that the frequencies of 
1124: the two $r$-modes are degenerate in the lowest order of $\Omega$.
1125: These results suggest that
1126: the $r$-mode solutions with $\pmb{w}^\prime\propto\Omega^3$ can be found
1127: only when the frequencies of the $r$-modes are not degenerate in the lowest order in $\Omega$.
1128: In this sense, it is reasonable to find $\pmb{w}^\prime\propto\Omega$ for the $r^o$-modes
1129: at $\eta=0$ since $\kappa_0$'s of the $r^o$- and $r^s$-modes are equal to each other 
1130: at $\eta=0$, as shown by Table 1.
1131: 
1132: 
1133: 
1134: \section{Conclusion}
1135: 
1136: 
1137: 
1138: 
1139: In this paper we have discussed the modal properties of the $r$-modes of neutron stars with
1140: the core filled with neutron and proton superfluids.
1141: We numerically find that
1142: the $r$-modes of rotating neutron stars with the superfluid core are split into ordinary $r^o$-modes 
1143: and superfluid $r^s$-modes, and that
1144: the two superfluids in the core flow together for the $r^o$-modes and they counter move
1145: for the $r^s$-modes.
1146: These findings are consistent with earlier suggestions made analytically by Andersoon \& Comer (2001).
1147: We also find that,
1148: although $\kappa_0$ of the $r^o$-modes is numerically equal to $2m/[l^\prime(l^\prime+1)]$,
1149: almost independent of the entrainment parameter $\eta$,
1150: $\kappa_0$ of the $r^s$-modes approximately linearly increases from
1151: $2m/[l^\prime(l^\prime+1)]$ as $\eta$ is increased from $\eta=0$.
1152: The $r^o$-modes have the scaling 
1153: of $\pmb{w}^\prime\propto\Omega^3$ and
1154: are the same $r$-modes discussed by Lindblom \& Mendell (2000), while 
1155: the $r^s$-modes have the scaling of $\pmb{w}^\prime\propto\Omega$.
1156: The instability caused by the $r^s$-modes is found much weaker than the instability 
1157: by the $r^o$-mode and will be
1158: easily stabilized by various dissipation processes in the star.
1159: 
1160: 
1161: We have confirmed that,
1162: except in a few narrow regions of $\eta$ around the prominent local minimums of $\tau^0_{MF}$,  
1163: the dissipation due to mutual friction in the superfluid core is ineffective 
1164: to stabilize the instability by the $r^o$-modes, as first shown by
1165: Lindblom \& Mendell (2000).
1166: We have shown that these prominent local minimums of $-\tau^0_{MF}$ are caused by
1167: mode crossings between the $r^o$-mode and 
1168: the superfluid inertial $i^s$-modes
1169: with long radial wavelengths comparable to those of the $r^o$-mode 
1170: (see also Andersson \& Comer 2001).
1171: Since mutual friction is almost always very strong for the $r^s$-modes and the instability
1172: by the $r^s$-modes itself is quite weak,
1173: only the instability by the $r^o$-modes will be of direct observational importance, e.g., as 
1174: mechanisms that generate
1175: gravitational waves from normal modes of rotating neutron stars and/or
1176: cause the clustering of spin frequencies around 300Hz for neutron stars in
1177: LMXB's, unless there exist other strong damping mechanisms for the $r$-modes 
1178: (see Jones 2001a,b;
1179: Lindblom \& Owen 2002; Haensel, Levenfish \& Yakovlev 2002).
1180: 
1181: 
1182: In this paper, we have made a brief report on a numerical result of inertial modes in
1183: the superfluid core.
1184: We find that inertial modes in superfluid neutron stars are also split into ordinary 
1185: inertial $i^o$-modes and superfluid inertial $i^s$-modes.
1186: It is generally observed for inertial modes with long radial wavelengths
1187: that the two superfluids co-move in the core
1188: for the $i^o$-modes, and they counter-move for the $i^s$-modes
1189: (see Yoshida \& Lee 2002b for
1190: more complete discussions on inertial modes in the superfluid core).
1191: Inertial modes are expected to work as a mechanism limiting 
1192: the amplitude growth of the $r$-modes (Morsink 2002).
1193: 
1194: There are many interesting and challenging problems we have to deal with
1195: before we can conclude definitely about the neutron star oscillations.
1196: Confronting their cooling calculations of neutron stars
1197: with observational data for the surface temperature of several neutron
1198: stars, Kaminker, Haensel, \& Yakovlev (2001), and 
1199: Kaminker, Yakovlev, \& Gnedin (2002) suggested that neutron superfluidity
1200: in the core of middle-aged neutron stars should be weak in the sense that
1201: the critical temperature $T_c$ is less than $10^8$K. 
1202: If this is the case, there exist no neutron superfluids in the core of
1203: temperatures $T>10^8$K.
1204: If neutrons in the core are in normal state, modal properties of low frequency oscillations 
1205: propagating in the core will be different
1206: from those for the core filled with neutron and proton superfluids,
1207: since buoyant force in the core, produced by thermal and/or chemical stratification 
1208: (Reisenegger \& Goldreich 1992), comes into play.
1209: The normal neutron fluid core with stratification will support $g$-mode propagation, 
1210: and the buoyant force in it
1211: will affect inertial modes in slowly rotating neutron stars.
1212: The nodeless $r$-modes of $l^\prime=m$, however, will remain almost the same even 
1213: in the presence of buoyant force (see Yoshida \& Lee 2000b), and 
1214: damping due to mutual friction in the core will remain weak for the $r^o$-modes,
1215: for which neutron and proton fluids co-move.
1216: Let us point out a possibility of using neutron star binaries as a probe 
1217: to investigate existence or non-existence of neutron superfludity in the core,
1218: since tidally excited low frequency modes will be $g$-modes in the normal fluid core but
1219: inertial modes in the superfluid core, and the different tidal responses 
1220: will result in different binary evolutions.
1221: The presence of a solid crust and a magnetic field in neutron star interior 
1222: is another factor that makes complicated the problems of global oscillations of the stars.
1223: The solid crust is not a rigid body and supports its own normal modes 
1224: (see, e.g., McDermott et al 1988, Lee \& Strohmayer 1996).
1225: For example, there exist torsional modes propagating in the crust, which
1226: can be resonantly coupled with the $r$-modes in the normal fluid core
1227: (Yoshida \& Lee 2001; Levin \& Ushomirsky 2001).
1228: Quite recently, using a local analysis of perturbations, Kinney \& Mendell (2002) 
1229: have suggested that
1230: no $r$-mode solution to the magnetohydrodynamic equations (e.g., Mendell 1998)
1231: exists in the superfluid core when
1232: both the neutron and proton vortices are pinned to the solid crust.
1233: This may suggest that a careful formulation of boundary conditions at the core-crust interfaces
1234: will be necessary to obtain reliable solutions to global oscillations of neutron stars,
1235: particularly when a magnetic field is essential for the oscillations.
1236: 
1237: \appendix
1238: 
1239: \section{Oscillation Equations in the Superfluid Core}
1240: 
1241: 
1242: For the oscillation equations in the superfluid core,
1243: we employ vectors $\pmb{y}_1^n$, $\pmb{y}_2^n$, $\pmb{y}_1^p$, $\pmb{y}_2^p$,
1244: $\pmb{y}_3$, and $\pmb{y}_4$, whose components are given by
1245: %
1246: \begin{equation}
1247: y_{1,k}^n=S_l^n, \quad y_{2,k}^n={\mu_{n,l}^\prime+\Psi^\prime_l\over ga}, \quad
1248: y_{1,k}^p=S_l^p, \quad y_{2,k}^p={\tilde\mu_{l}^\prime+\zeta\Psi^\prime_l\over ga}, \quad
1249: y_{3,k}={\Psi^\prime_l\over ga}, \quad y_{4,k}={1\over g}{d\Psi^\prime_l\over da},
1250: \end{equation}
1251: %
1252: where $l=|m|+2(k-1)$ for even modes and $l=|m|+2k-1$ for odd modes, and
1253: $k=1,~2,~3, \cdots$.
1254: We also introduce vectors $\pmb{h}^n$, $\pmb{h}^p$, $i\pmb{t}^n$, and $i\pmb{t}^p$,
1255: the components of which are
1256: %
1257: \begin{equation}
1258: h_k^n=H_l^n, \quad it_k^n=iT_{l^\prime}^n, \quad 
1259: h_k^p=H_l^p, \quad it_k^p=iT_{l^\prime}^p, 
1260: \end{equation}
1261: %
1262: where $l^\prime=l+1$ for even modes and $l^\prime=l-1$ for odd modes.
1263: Using these vectors,
1264: the perturbed continuity equations (16) and (17) are written as
1265: %
1266: \begin{eqnarray}
1267: a{d\pmb{y}_1^n\over da}&=&\left(-{d\ln\rho_n\over d\ln a}-3-a{d\chi_3(\alpha)\over da}
1268: -a{d\chi_3(\beta)\over da}A_0\right)\pmb{y}_1^n-{ga\over\rho_n}{\cal Q}_{11}\pmb{y}_2^n
1269: -{ga\over\rho_n}{\cal Q}_{12}\pmb{y}_2^p
1270: +{ga\over\rho_n}\left({\cal Q}_{11}+\zeta {\cal Q}_{12}\right)\pmb{y}_3 \nonumber  \\
1271: &+& \left(\Lambda_0+3\chi_3(\beta)B_0\right)\pmb{h}^n
1272: +3m\chi_3(\beta)Q_0i\pmb{t}^n, 
1273: \end{eqnarray}
1274: %
1275: \begin{eqnarray}
1276: a{d\pmb{y}_1^p\over da}&=&\left(-{d\ln\rho_p\over d\ln a}-3-a{d\chi_3(\alpha)\over da}
1277: -a{d\chi_3(\beta)\over da}A_0\right)\pmb{y}_1^p-{ga\over\rho_p}{\cal Q}_{21}\pmb{y}_2^n
1278: -{ga\over\rho_p}{\cal Q}_{22}\pmb{y}_2^p
1279: +{ga\over\rho_p}\left({\cal Q}_{21}+\zeta {\cal Q}_{22}\right)\pmb{y}_3 \nonumber  \\
1280: &+& \left(\Lambda_0+3\chi_3(\beta)B_0\right)\pmb{h}^p
1281: +3m\chi_3(\beta)Q_0i\pmb{t}^p.
1282: \end{eqnarray}
1283: %
1284: The radial components of equations (18) and (19) are reduced to
1285: %
1286: \begin{eqnarray}
1287: a{d\pmb{y}_2^n\over da}&=&c_1\bar\omega^2{\rho_{11}\over\rho_n}E_0\pmb{y}_1^n+(1-U)\pmb{y}_2^n
1288: +c_1\bar\omega^2{\rho_{12}\over\rho_n}E_0\pmb{y}_1^p \nonumber \\
1289: &-&c_1\bar\omega^2\left({\rho_{11}\over\rho_n}3\beta B_0+m\nu E_1\right)\pmb{h}^n
1290: -c_1\bar\omega^2{\rho_{12}\over\rho_n}3\beta B_0\pmb{h}^p  \nonumber \\
1291: &-&c_1\bar\omega^2\left({\rho_{11}\over\rho_n}3m\beta Q_0+\nu E_1C_0\right)i\pmb{t}^n
1292: -c_1\bar\omega^2{\rho_{12}\over\rho_n}3m\beta Q_0i\pmb{t}^p, 
1293: \end{eqnarray}
1294: %
1295: \begin{eqnarray}
1296: a{d\pmb{y}_2^p\over da}&=&c_1\bar\omega^2\left({\rho_{22}\over\rho_p}+{m_e\over m_p}\right)
1297: E_0\pmb{y}_1^p+(1-U)\pmb{y}_2^p
1298: +c_1\bar\omega^2{\rho_{21}\over\rho_p}E_0\pmb{y}_1^n \nonumber \\
1299: &-&c_1\bar\omega^2\left[\left({\rho_{22}\over\rho_p}+{m_e\over m_p}\right)3\beta B_0
1300: +m\nu\zeta E_1\right]\pmb{h}^p
1301: -c_1\bar\omega^2{\rho_{21}\over\rho_p}3\beta B_0\pmb{h}^n \nonumber \\
1302: &-&c_1\bar\omega^2\left[\left({\rho_{22}\over\rho_p}+{m_e\over m_p}\right)3m\beta Q_0
1303: +\nu\zeta E_1C_0\right]i\pmb{t}^p
1304: -c_1\bar\omega^2{\rho_{21}\over\rho_p}3m\beta Q_0i\pmb{t}^n.
1305: \end{eqnarray}
1306: %
1307: The perturbed Poisson equation (21) is reduced to
1308: \begin{equation}
1309: a{d\pmb{y}_3\over da}=\left(1-U\right)\pmb{y}_3+\pmb{y}_4,
1310: \end{equation}
1311: %
1312: \begin{eqnarray}
1313: a{d\pmb{y}_4\over da}&= & 4\pi G a^2\left({\cal Q}_{11}+\zeta{\cal Q}_{21}\right)I^{-1}\pmb{y}_2^n
1314: +4\pi G a^2\left({\cal Q}_{12}+\zeta{\cal Q}_{22}\right)I^{-1}\pmb{y}_2^p \nonumber \\
1315: &+&I^{-1}\left[\Lambda_0-4\pi G a^2\left({\cal Q}_{11}
1316: +\zeta{\cal Q}_{21}+\zeta({\cal Q}_{12}+\zeta{\cal Q}_{22})\right)\pmb{1}
1317: -2\left(\alpha\pmb{1}+\beta A_0\right)\Lambda_0\right]\pmb{y}_3  \nonumber \\
1318: &+& I^{-1}\left[(-U+\chi_0(\alpha))I
1319: +\chi_0(\beta)A_0-6\beta (A_0+B_0)\right]\pmb{y}_4.
1320: \end{eqnarray}
1321: %
1322: Algebraic equations that determine the relations between the variables $\pmb{h}$, $i\pmb{t}$,
1323: $\pmb{y}_1$, and $\pmb{y}_2$ are derived by making use of the horizontal components of
1324: the velocity equations (18) and (19), and they are
1325: %
1326: \begin{equation}
1327: \tilde L_0\pmb{h}^n-\tilde M_1i\pmb{t}^n
1328: -f_nmF_0(\pmb{h}^p-\pmb{h}^n)-f_n\tilde M_1^+(i\pmb{t}^p-i\pmb{t}^n)
1329: = \tilde J\pmb{y}_1^n +f_n \tilde J^+(\pmb{y}_1^p-\pmb{y}_1^n)
1330: +b_{11}{\pmb{y}_2^n\over c_1\bar\omega^2}+b_{12}{\pmb{y}_2^p\over c_1\bar\omega^2},
1331: \end{equation}
1332: %
1333: \begin{equation}
1334: \tilde L_0\pmb{h}^p-\tilde M_1i\pmb{t}^p
1335: + f_pmF_0(\pmb{h}^p-\pmb{h}^n)+f_p\tilde M_1^+(i\pmb{t}^p-i\pmb{t}^n)
1336: = \tilde J\pmb{y}_1^p-f_p \tilde J^+(\pmb{y}_1^p-\pmb{y}_1^n)
1337: +b_{21}{\pmb{y}_2^n\over c_1\bar\omega^2}+b_{22}{\pmb{y}_2^p\over c_1\bar\omega^2},
1338: \end{equation}
1339: %
1340: \begin{equation}
1341: -\tilde M_0\pmb{h}^n+\tilde L_1i\pmb{t}^n
1342: -f_n\tilde M_0^+(\pmb{h}^p-\pmb{h}^n)
1343: -f_nmF_1(i\pmb{t}^p-i\pmb{t}^n)
1344: =-\tilde K\pmb{y}_1^n-f_n\tilde K^+(\pmb{y}_1^p-\pmb{y}_1^n),
1345: \end{equation}
1346: %
1347: \begin{equation}
1348: -\tilde M_0\pmb{h}^p+\tilde L_1i\pmb{t}^p
1349: +f_p\tilde M_0^+ (\pmb{h}^p-\pmb{h}^n)
1350: +f_pmF_1(i\pmb{t}^p-i\pmb{t}^n)
1351: =-\tilde K\pmb{y}_1^p+f_p\tilde K^+(\pmb{y}_1^p-\pmb{y}_1^n),
1352: \end{equation}
1353: %
1354: where
1355: \begin{equation}
1356: f_n={\rho_{np}\over\rho_n}{\zeta\over\tilde\zeta}, \quad
1357: f_p={\rho_{np}\over\rho_p}{1\over\tilde\zeta},
1358: \end{equation}
1359: %
1360: \begin{equation}
1361: \zeta=1+{m_e\over m_p}, \quad
1362: \tilde\zeta=1+{m_e\over m_p}{\rho_{pp}\over\rho_p},
1363: \end{equation}
1364: and
1365: \begin{equation}
1366: \left(\matrix{b_{11}&b_{12}\cr b_{21}&b_{22}\cr}\right)=
1367: \left(\matrix{\rho_{11}/\rho_n&\rho_{12}/\rho_n\cr
1368: \rho_{21}/\rho_p&(\rho_{22}/\rho_p)+(m_e/m_p)\cr}
1369: \right)^{-1}.
1370: \end{equation}
1371: The functions $\chi_1(\alpha)$, $\chi_2(\alpha)$, $\chi_3(\alpha)$, and $\chi_0(\alpha)$ 
1372: are defined as
1373: %
1374: \begin{equation}
1375: \chi_1(\alpha)=\alpha+a{d\alpha\over da}, \quad
1376: \chi_2(\alpha)=2\alpha+a{d\alpha\over da}, \quad
1377: \chi_3(\alpha)=3\alpha+a{d\alpha\over da},
1378: \end{equation}
1379: %
1380: and
1381: %
1382: \begin{equation}
1383: \chi_0(\alpha)=-a{d\alpha\over da}+a{d\over da}\left(a{d\alpha\over da}\right).
1384: \end{equation}
1385: %
1386: The matrices $E_0$, $E_1$, $F_0$, $F_1$, $I$, $\tilde J$, $\tilde J^+$, $\tilde K$, $\tilde K^+$,
1387: $\tilde L_0$, $\tilde L_0^+$, $\tilde L_1$, $\tilde L_1^+$, 
1388: $\tilde M_0$, $\tilde M_0^+$, $\tilde M_1$, and $\tilde M_1^+$ are defined as
1389: %
1390: \begin{equation}
1391: E_0=\left(1+2\chi_1(\alpha)\right)\pmb{1}+2\chi_1(\beta)A_0, \quad
1392: E_1=(1+\chi_2(\alpha))\pmb{1}+\chi_2(\beta)A_0,
1393: \end{equation}
1394: %
1395: \begin{equation}
1396: F_0=\nu\Lambda_0^{-1}\left[(1+2\alpha+2\beta)\pmb{1}+12\beta A_0\right], \quad
1397: F_1=\nu\Lambda_1^{-1}\left[(1+2\alpha+2\beta)\pmb{1}+12\beta A_1\right],
1398: \end{equation}
1399: %
1400: %
1401: \begin{equation}
1402: I=\pmb{1}-2\chi_1(\alpha)\pmb{1}-2\chi_1(\beta)A_0,
1403: \end{equation}
1404: %
1405: \begin{equation}
1406: \tilde J=\tilde J^+-3\beta\Lambda_0^{-1}(2A_0+B_0), 
1407: \end{equation}
1408: %
1409: \begin{equation}
1410: \tilde J^+=m\nu\Lambda_0^{-1}\left[(1+\chi_2(\alpha))\pmb{1}+\chi_2(\beta)A_0\right],
1411: \end{equation}
1412: %
1413: \begin{equation}
1414: \tilde K =\tilde K^+-3m\beta\Lambda_1^{-1}Q_1,
1415: \end{equation}
1416: %
1417: \begin{equation}
1418: \tilde K^+
1419: =\nu\Lambda_1^{-1}\left[((1+\chi_2(\alpha))\pmb{1}+\chi_2(\beta)A_1)C_1
1420: +2((1+\chi_2(\alpha)-\chi_2(\beta))\pmb{1}+2\chi_2(\beta)A_1)Q_1\right],
1421: \end{equation}
1422: %
1423: \begin{equation}
1424: \tilde L_0=(1+2\alpha)\pmb{1}-mF_0+\beta\Lambda_0^{-1}(2A_0\Lambda_0+6B_0), 
1425: \end{equation}
1426: %
1427: \begin{equation}
1428: \tilde L_1=(1+2\alpha)\pmb{1}-mF_1+\beta\Lambda_1^{-1}(2A_1\Lambda_1+6B_1), 
1429: \end{equation}
1430: %
1431: \begin{equation}
1432: \tilde M_0=\tilde M_0^+-6m\beta\Lambda_1^{-1}Q_1, 
1433: \end{equation}
1434: %
1435: \begin{equation}
1436: \tilde M_0^+=\nu\Lambda_1^{-1}
1437: [(1+2\alpha-2\beta)\pmb{1}+4\beta A_1]Q_1\Lambda_0+F_1C_1, \quad
1438: \end{equation}
1439: %
1440: \begin{equation}
1441: \tilde M_1=\tilde M_1^+-6m\beta\Lambda_0^{-1}Q_0, 
1442: \end{equation}
1443: %
1444: \begin{equation}
1445: \tilde M_1^+=\nu\Lambda_0^{-1}
1446: [(1+2\alpha-2\beta)\pmb{1}+4\beta A_0]Q_0\Lambda_1+F_0C_0,
1447: \end{equation}
1448: %
1449: where $\pmb{1}$ denotes the unit matrix,
1450: and the matrices $A_0$, $A_1$, $B_0$, $B_1$, $C_0$, $C_1$,
1451: $Q_0$, $Q_1$, $\Lambda_0$, and $\Lambda_1$ as well as the quantities $U$ and $c_1$
1452: are defined in Yoshida \& Lee (2000a).
1453: 
1454: The oscillation equations in the superfluid core are given as a set of linear ordinary
1455: differential equations for the variables $\pmb{y}_j$ with $j=1$ to 4, which
1456: are obtained by eliminating the vectors
1457: $\pmb{h}$ and $i\pmb{t}$ in equations (A3) to (A8) 
1458: using the algebraic equations (A9) to (A12).
1459: 
1460: 
1461: \newpage
1462: 
1463: 
1464: \begin{deluxetable}{cccccccc}
1465: \footnotesize
1466: \tablecaption{$\kappa_0$, $\kappa_2$, and 
1467: $\tau^0_i$'s for the $r^o$- and $r^s$-modes of $l^\prime=m=2$.}
1468: \tablewidth{0pt}
1469: \tablehead{ \colhead{Modes}&\colhead{$\eta$}& \colhead{$\kappa_0$} & \colhead{$\kappa_2$}
1470: & \colhead{$\tau^0_{GJ}$(s)} & \colhead{$\tau^0_{GD}$(s)} &
1471: \colhead{$\tau^0_{S}$(s)} & \colhead{$\tau^0_{MF}$(s)} }
1472: \startdata
1473:  $r^o$& 0 & 0.6667 & 0.408 & $3.31$ & $3.83\times 10^2$ & $-7.55\times10^7$ &  $-\infty$ \\
1474: $r^s$&  & 0.6667 & 0.511 & $2.12\times10^4$ & $2.18\times 10^6$  & $-2.99 \times 10^6$ 
1475: & $-\infty$ \\
1476: $r^o$& 0.02 & 0.6667 & 0.408 & $3.31$ & $3.83\times 10^2$ & $-6.76\times10^7$ 
1477: &  $-2.52\times10^2$ \\
1478: $r^s$ & & 0.8532 & 0.652 & $4.76\times10^{8}$ & $1.21\times 10^6$  & $-1.99 \times 10^6$ 
1479: & $-2.94\times 10^{-1}$ \\
1480:  $r^o$&0.04 & 0.6667 & 0.408 & $3.31$ & $3.83\times 10^2$ & $-6.76\times10^7$ 
1481: &  $-1.70\times10^3$ \\
1482:  $r^s$& & 1.0408 & 0.787 & $7.89\times10^{8}$ & $2.55\times 10^6$  & $-1.14 \times 10^6$ 
1483: & $-7.99\times 10^{-2}$ \\
1484:  $r^o$&0.06 & 0.6667 & 0.408 & $3.31$ & $3.83\times 10^2$ & $-6.76\times10^7$ 
1485: &  $-2.62\times10^3$ \\
1486:  $r^s$ && 1.2296 & 0.821 & $3.64\times10^{9}$ & $1.03\times 10^7$  & $-4.67 \times 10^4$ 
1487: & $-5.20\times 10^{-2}$ \\
1488: \enddata
1489: \label{coefficients0}
1490: \end{deluxetable}
1491: 
1492: 
1493: \begin{deluxetable}{ccccc}
1494: \footnotesize
1495: \tablecaption{$\kappa_0$ for the $i^o$- and $i^s$-modes for $m=2$ and $\eta=0$.}
1496: \tablewidth{0pt}
1497: \tablehead{ \colhead{$l_0-|m|$}& \colhead{$i^o$} & \colhead{$i^s$}}
1498: \startdata
1499: 2 & 1.0978 & 1.1190 \\
1500:  & -0.5683 & -0.5109 \\
1501: 3 & 1.3564 & 1.3869 \\
1502:   & 0.5180 & 0.5077 \\
1503:   & -1.0293 & -0.9723 \\
1504: 4 & 1.5191 & 1.5526 \\
1505:  & 0.8639 & 0.8612 \\
1506:  & -0.2738 & -0.2417\\
1507:  & -1.2738 & -1.2238 \\
1508: 5 & 1.6273 & 1.6600 \\
1509:   & 1.1046 & 1.1134 \\
1510:   & 0.4215 & 0.4060 \\
1511:   & -0.7028 & -0.6524 \\
1512:  & -1.4340 & -1.3968 \\
1513: \enddata
1514: \label{coefficients0}
1515: \end{deluxetable}
1516: 
1517: 
1518: \begin{deluxetable}{ccccccc}
1519: \footnotesize
1520: \tablecaption{$\kappa_0$, $\kappa_2$, and 
1521: $\tau^0_i$'s for the $r^o$- and $r^s$-modes of $l^\prime=m=2$ for $\eta=0.04$ without the drag force.}
1522: \tablewidth{0pt}
1523: \tablehead{ \colhead{Modes}& \colhead{$\kappa_0$} & \colhead{$\kappa_2$}
1524: & \colhead{$\tau^0_{GJ}$(s)} & \colhead{$\tau^0_{GD}$(s)} &
1525: \colhead{$\tau^0_{S}$(s)} & \colhead{$\tau^0_{MF}$(s)} }
1526: \startdata
1527: $r^o$ & 0.6667 & 0.408 & $3.31$ & $3.83\times 10^2$ & $-7.49\times10^7$ &  $-9.10\times10^2$ \\
1528: $r^s$ & 0.6667 & 0.529 & $4.33\times10^4$ & $2.63\times 10^6$  & $-1.13 \times 10^6$ & $-7.30\times10^{-2}$ \\
1529: \enddata
1530: \label{coefficients0}
1531: \end{deluxetable}
1532: 
1533: 
1534: 
1535: \newpage
1536: 
1537: 
1538: 
1539: \begin{thebibliography}{}
1540: 
1541: \bibitem[]{} Akmal, A., Pandharipande, V.R., \& Ravenhall, D.G., 1998, Phys. Rev. C., 58, 1804
1542: 
1543: \bibitem[]{} Alpar, M.A., Langer, S.A., \& Sauls, J.A., 1984, ApJ, 282, 533
1544: 
1545: \bibitem[]{} Andersson, N. 1998, ApJ, 502, 708
1546: 
1547: \bibitem[]{} Andersson, N., \& Comer, G.L. 2001, MNRAS, 328, 1129
1548: 
1549: \bibitem[]{} Andersson, N., Kokkotas, K.D., \& Stergioulas, N. 1999, ApJ, 516, 307
1550: 
1551: \bibitem[]{} Andreev, A.F., \& Bashkin, E.P. 1976, Sov. Phys.-JETP, 42, 164
1552: 
1553: \bibitem[]{} Bildsten, L., \& Ushomirsky, G. 2000, ApJ, 529, L33
1554: 
1555: \bibitem[]{} Borumand, M., Joynt, R., \& Klu\'zniak, W. 1996, Phys. Rev. C., 54, 2745
1556: 
1557: \bibitem[]{} Chandrasekhar, S. 1933a, MNRAS, 93, 390
1558: 
1559: \bibitem[]{} Chandrasekhar, S. 1933b, MNRAS, 93, 462
1560: 
1561: \bibitem[]{} Cutler, L., \& Lindblom, L. 1987, ApJ, 314, 234
1562: 
1563: \bibitem[]{} Epstein, R.I. 1988, ApJ, 333, 880
1564: 
1565: \bibitem[]{} Feynman, R.P. 1972, Statistical Mechanics, Addison Wesley, Reading
1566: 
1567: \bibitem[]{} Flowers, E., \& Itoh, N. 1979, ApJ, 230, 847
1568: 
1569: \bibitem[]{} Friedman, J.L., \& Morsink, S.M. 1998, ApJ, 502, 714
1570: 
1571: \bibitem[]{} Friedman, J.L., \& Schutz, B.F. 1978, ApJ, 222, 281
1572: 
1573: \bibitem[]{} Haensel, P., Levenfish, K.P., \& Yakovlev, D.G. 2002, A\&A, 381, 1080
1574: 
1575: \bibitem[]{} Jones, P.B. 2001a, Phys. Rev. Lett., 86, 1384
1576: 
1577: \bibitem[]{} Jones, P.B. 2001b, Phys. Rev. D, 64, 084003
1578: 
1579: \bibitem[]{} Kaminker, A.D., Haensel, P., \& Yakovlev, D.G. 2001, A\&A, 373, L17
1580: 
1581: \bibitem[]{} Kaminker, A.D., Yakovlev, D.G., \& Gnedin, O.Y. 2002, A\&A, 383, 1076
1582: 
1583: \bibitem[]{} Khalatnikov, I.M. 1965, An Introduction to the Theory of Superfluidity,
1584: Benjamin, New York
1585: 
1586: \bibitem[]{} Kinney, J.B., \& Mendell, G. 2002, gr-qc/0206001
1587: 
1588: \bibitem[]{} Landau, L.D., \& Lifshitz, E.M. 1987, Fluid Mechanics, 2nd edn, Pergamon Press, Oxford
1589: 
1590: \bibitem[]{} Lee, U. 1993, ApJ, 405, 359
1591: 
1592: \bibitem[]{} Lee, U. 1995, A\&A, 303, 515
1593: 
1594: \bibitem[]{} Lee, U., \& Saio, H. 1986, MNRAS, 221, 365
1595: 
1596: \bibitem[]{} Lee, U., \& Strohmayer, T.E. 1996, A\&A, 311, 155
1597: 
1598: \bibitem[]{} Levin, Y., \& Ushomirsky, G. 2001, MNRAS, 324, 917
1599: 
1600: \bibitem[]{} Lindblom, L., \& Mendell, G. 1994, ApJ, 421, 689
1601: 
1602: \bibitem[]{} Lindblom, L., \& Mendell, G. 2000, Phys. Rev. D., 61, 104003
1603: 
1604: \bibitem[]{} Lindblom, L., Owen, B.J., \& Morsink, S.M. 1998, Phys. Rev. Lett, 80, 4843
1605: 
1606: \bibitem[]{} Lindblom, L., \& Owen, B.J. 2002, Phys. Rev. D, 65, 084039 
1607: 
1608: \bibitem[]{} Lockitch, K.H., \& Friedman, J.L. 1999, ApJ, 521, 764
1609: 
1610: \bibitem[]{} McDermott, P.N., Van Horn, H.M., \& Hansen, C.J. 1988, ApJ, 325, 725
1611: 
1612: \bibitem[]{} Mendell, G. 1991a, ApJ, 380, 515
1613: 
1614: \bibitem[]{} Mendell, G. 1991b, ApJ, 380, 530
1615: 
1616: \bibitem[]{} Mendell, G. 1998, MNRAS, 296, 903
1617: 
1618: %\bibitem[]{} Mendell, G. 2001, Phys. Rev. D, 64, 44009
1619: 
1620: \bibitem[]{} Morsink, S.M. 2002, ApJ, 571, 435
1621: 
1622: \bibitem[]{} Phinney, E.S., \& Kulkarni, S.R. 1994, ARA\&A, 32, 591
1623: 
1624: \bibitem[]{} Reisenegger, A., \& Goldreich, P. 1992, ApJ, 395, 240
1625: 
1626: \bibitem[]{} Saio, H. 1981, ApJ, 244, 299
1627: 
1628: \bibitem[]{} Sawyer, R.F. 1989, Phys. Rev. D., 39, 3804
1629: 
1630: %\bibitem[]{} Serot, B.D. 1979, Phys. Lett. B., 86, 146
1631: 
1632: \bibitem[]{} Shapiro, S.L., \& Tuekolwsky, S.A. 1983, 
1633: Black Holes, White Dwarfs, and Neutron Stars, 
1634: Wiley, New York
1635: 
1636: \bibitem[]{} Tassoul, J.L. 1978, Theory of Rotating Stars, Princeton Univ. Press, Princeton
1637: 
1638: \bibitem[]{} Thorne, K. 1980, Rev. Mod. Phys., 52, 299
1639: 
1640: \bibitem[]{} Tilley, D.R., \& Tilley, J. 1990, Superfluidity and Superconductivity, 3rd edn, 
1641: IOP Publishing Ltd, Bristol
1642: 
1643: \bibitem[]{} Unno, W., Osaki, Y., Ando, H., Saio, H., \& Shibahashi, H. 1989, 
1644: Nonradial Oscillations of Stars, 2nd ed, University of Tokyo Press, Tokyo
1645: 
1646: \bibitem[]{} van der Klis, M. 2000, ARA\&A, 38, 717
1647: 
1648: \bibitem[]{} Yoshida, S. 2001, ApJ, 558, 263
1649: 
1650: \bibitem[]{} Yoshida, S., \& Lee, U. 2000a, ApJ, 529, 997
1651: 
1652: \bibitem[]{} Yoshida, S., \& Lee, U. 2000b, ApJS, 129, 353
1653: 
1654: \bibitem[]{} Yoshida, S., \& Lee, U. 2001, ApJ, 546, 1121
1655: 
1656: \bibitem[]{} Yoshida, S., \& Lee, U. 2002a, ApJ, 567, 1112
1657: 
1658: \bibitem[]{} Yoshida, S., \& Lee, U. 2002b, in preparation.
1659: 
1660: \end{thebibliography}
1661: 
1662: 
1663: \begin{figure}
1664: \epsscale{.5}
1665: %\plotone{f1.eps}
1666: 
1667: \caption{Toroidal components $iT_m$ and $iT_{m+2}$ of the $r$-modes of
1668: $l^\prime=m=2$ are plotted 
1669: as a function of $a/R$ for $\bar\Omega=0.01$ and $\eta=0.04$, 
1670: where the solid, dotted, dashed, and dash-dotted lines
1671: are used to indicate, respectively, $iT_m^n$, $iT_{m+2}^n$, $iT_m^p$, and $iT_{m+2}^p$
1672: in the superfluid core.
1673: The amplitudes are normalized by the maximum value.
1674: Panels (a) and (b) are for the $r^o$- and $r^s$-modes, respectively.}
1675: 
1676: \end{figure}
1677: 
1678: \begin{figure}
1679: \epsscale{.5}
1680: %\plotone{f2.eps}
1681: 
1682: \caption{Difference $iT_m^p-iT_m^n$ versus $a/R$ for the $l^\prime=m=2$ $r^o$-mode
1683: for $\bar\Omega=0.01$ and $\eta=0.04$, where 
1684: the amplitude normalization $\max(|iT_m|)=1$ have been used.}
1685: 
1686: \end{figure}
1687: 
1688: \begin{figure}
1689: \epsscale{.5}
1690: %\plotone{f3.eps}
1691: 
1692: \caption{Same as Figure 1 but for the case of $\eta=0$.}
1693: 
1694: \end{figure}
1695: 
1696: \begin{figure}
1697: \epsscale{.5}
1698: %\plotone{f4.eps}
1699: 
1700: \caption{Toroidal components $iT_m$ and $iT_{m+2}$ of inertial modes of
1701: $m=2$ and $l_0-|m|=3$ are plotted 
1702: as a function of $a/R$ for $\bar\Omega=0.01$ and $\eta=0$, 
1703: where the solid, dotted, dashed, and dash-dotted lines
1704: are used to indicate, respectively, $iT_m^n$, $iT_{m+2}^n$, $iT_m^p$, and $iT_{m+2}^p$
1705: in the superfluid core.
1706: The amplitudes are normalized by the maximum value.
1707: Panels (a) and (b) are for the $i^o$-mode of $\kappa_0=0.5180$ and the $i^s$-mode of
1708: $\kappa_0=0.5077$, respectively.}
1709: 
1710: \end{figure}
1711: 
1712: \begin{figure}
1713: \epsscale{.5}
1714: %\plotone{f5.eps}
1715: 
1716: \caption{Same as Figure 4 but for $l_0-|m|=5$.
1717: Panels (a) and (b) are for the $i^o$-mode of $\kappa_0=0.4215$ and the $i^s$-mode of
1718: $\kappa_0=0.4060$, respectively.}
1719: 
1720: \end{figure}
1721: 
1722: \begin{figure}
1723: \epsscale{.5}
1724: %\plotone{f6.eps}
1725: 
1726: \caption{Same as Figure 4 but for $l_0-|m|=5$.
1727: Panels (a) and (b) are for the $i^o$-mode of $\kappa_0=1.1046$ and 
1728: the $i^s$-mode of $\kappa_0=1.1134$, respectively.}
1729: 
1730: \end{figure}
1731: 
1732: \begin{figure}
1733: \epsscale{.5}
1734: %\plotone{f7.eps}
1735: 
1736: \caption{Avoided crossing as a function of $\eta$
1737: between the $l^\prime=m$ $r^s$-mode and the $i^o$-mode that tends to $\kappa_0=1.1046$ as
1738: $\eta\rightarrow0$, where $m=2$ and $\bar\Omega=0.01$ have been assumed.
1739: The dashed line indicates $\kappa_0$ given by equation (49) for the $r^s$-mode.}
1740: 
1741: \end{figure}
1742: 
1743: \begin{figure}
1744: \epsscale{.5}
1745: %\plotone{f8.eps}
1746: 
1747: \caption{Mode crossings as a function of $\eta$
1748: between the $l^\prime=m$ $r^o$-mode (dashed line) and the $i^s$-modes (solid lines)
1749: that tend to $\kappa_0=0.5077$ and $\kappa_0=0.4060$ as $\eta\rightarrow0$, 
1750: where $m=2$ and $\bar\Omega=0.01$ have been assumed.}
1751: 
1752: \end{figure}
1753: 
1754: 
1755: \begin{figure}
1756: \epsscale{.5}
1757: %\plotone{f9.eps}
1758: 
1759: \caption{Differences $iT_m^p-iT_m^n$ for the $l^\prime=m=2$ $r^o$-mode at $\bar\Omega=0.01$
1760: are plotted  as a function of $a/R$ for the cases of $\eta=0.023$ (panel a)
1761: and $\eta=0.0484$ (panel b).
1762: The amplitude normalization is given by $\max(|iT_m|)=1$.}
1763: 
1764: \end{figure}
1765: 
1766: \begin{figure}
1767: \epsscale{.5}
1768: %\plotone{f10.eps}
1769: 
1770: \caption{Toroidal components $iT_m$ and $iT_{m+2}$ of inertial modes of
1771: $m=2$ are plotted as a function of $a/R$ for $\bar\Omega=0.01$, 
1772: where the solid, dotted, dashed, and dash-dotted lines
1773: are used to indicate, respectively, $iT_m^n$, $iT_{m+2}^n$, $iT_m^p$, and $iT_{m+2}^p$
1774: in the superfluid core.
1775: The amplitudes are normalized by the maximum value.
1776: Panel (a) shows the $i^s$-mode of $l_0-|m|=3$ at
1777: $\eta=0.023$, and panel (b) the $i^s$-mode of $l_0-|m|=5$ at $\eta=0.0484$}
1778: 
1779: \end{figure}
1780: 
1781: 
1782: \begin{figure}
1783: \epsscale{.5}
1784: %\plotone{f11.eps}
1785: 
1786: \caption{ $NJ^\prime_{mm}(a)$ versus $a/R$ for the $l^\prime=m=2$ $r^s$-mode
1787: for the cases of $\eta=0.04$ (solid line)
1788: and $\eta=0$ (dashed line), where $N\equiv \sqrt{|\omega\sigma^{2l+1}|N_l/2E}$
1789: and $J^\prime_{mm}(a)=\int_0^adadJ^\prime_{mm}/da$.}
1790: 
1791: \end{figure}
1792: 
1793: \begin{figure}
1794: \epsscale{.5}
1795: %\plotone{f12.eps}
1796: 
1797: \caption{ $-\tau^0_{MF}$ versus $\eta$ for the $l^\prime=m=2$
1798: $r^o$-mode, where $\rho_s=2.8\times10^{14}$g/cm$^3$ has been used, and
1799: $-\tau^0_{MF}$ is given in second.}
1800: 
1801: \end{figure}
1802: 
1803: \begin{figure}
1804: \epsscale{.5}
1805: %\plotone{f13.eps}
1806: 
1807: \caption{ Eigenfunction $\delta\beta_{2,k}$ as a function of $r/R$ for the $r^o$-mode of
1808: $l^\prime=m=2$ for $\bar\Omega=0.01$ and $\eta=0.04$, where amplitude normalizaation has been given
1809: by $y_{2,k=1}(r=R)=f_{m+1,m}\bar\Omega^2$. See text for definition of the factor $f_{lm}$ and 
1810: the function $y_{2,k}$. The solid, dotted, and dashed lines are for
1811: $\delta\beta_{2,1}$, $100\times \delta\beta_{2,2}$, and $100\times \delta\beta_{2,3}$, respectively.}
1812: 
1813: \end{figure}
1814: 
1815: \begin{figure}
1816: \epsscale{.5}
1817: %\plotone{f14.eps}
1818: 
1819: \caption{Same as Figure 1 but for that
1820: we have ignored the drag force terms on the right hand side of the velocity equations
1821: (18) and (19) to calculate the $r$-modes.}
1822: 
1823: \end{figure}
1824: 
1825: 
1826: 
1827: \end{document}
1828: 
1829: