gr-qc0110038/ms.tex
1: \documentclass[12pt,preprint]{aastex}
2: 
3: \received{}
4: \accepted{}
5: \journalid{VOL}{JOURNAL DATE}
6: \articleid{START PAGE}{END PAGE}
7: \paperid{MANUSCRIPT ID}
8: 
9: \lefthead{Yoshida \& Lee}
10: \righthead{Relativistic $r$-modes in Slowly Rotating Neutron Stars}
11: 
12: \begin{document}
13: 
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
15: %% O. FRONT MATTER                                          %%
16: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
17: 
18: 
19: \title{Relativistic $r$-modes in Slowly Rotating Neutron Stars: \\
20:            Numerical Analysis in the Cowling Approximation}
21: %
22: \author{Shijun Yoshida\altaffilmark{1} and Umin Lee}
23: %
24: \affil{Astronomical Institute, Graduate School of Science, 
25: Tohoku University, Sendai 980-8578, 
26: Japan \\ yoshida@astr.tohoku.ac.jp, lee@astr.tohoku.ac.jp}
27: 
28: \altaffiltext{1}{Research Fellow of the Japan Society for 
29: the Promotion of Science.}
30: 
31: \begin{abstract}
32: 
33: We investigate the properties of relativistic $r$-modes of slowly rotating 
34: neutron stars by using a relativistic version of the Cowling 
35: approximation. 
36: In our formalism, we take into account the influence of the Coriolis like 
37: force on the stellar oscillations, but ignore the effects of 
38: the centrifugal like force. For three neutron star models,
39: we calculated the fundamental $r$-modes with $l'=m=2$ and $3$.
40: We found that the oscillation frequency $\bar\sigma$ of
41: the fundamental $r$-mode is in a good approximation given by 
42: $\bar\sigma\approx \kappa_0\,\Omega$, where 
43: $\bar\sigma$ is defined in the 
44: corotating frame at the spatial infinity, and
45: $\Omega$ is the angular frequency of rotation of the star.
46: The proportional coefficient $\kappa_0$ is only weakly dependent on $\Omega$, but
47: it strongly depends on the relativistic parameter $GM/c^2R$, where $M$ and $R$ are 
48: the mass and the radius of the star.
49: All the fundamental $r$-modes with $l'=m$ computed in this study
50: are discrete modes with distinct regular eigenfunctions, and
51: they all fall 
52: in the continuous part of the frequency spectrum associated with
53: Kojima's equation (Kojima 1998).
54: These relativistic $r$-modes are
55: obtained by including the effects of rotation higher
56: than the first order of $\Omega$ so that the buoyant force plays a role, 
57: the situation of which is quite similar to that
58: for the Newtonian $r$-modes.
59: 
60: \end{abstract}
61: 
62: 
63: \keywords{instabilities --- stars: neutron --- 
64: stars: oscillations --- stars: rotation}
65: 
66: 
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68: %%  INTRODUCTION                                            %%
69: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70: 
71: 
72: \section{Introduction}
73: 
74: 
75: 
76: It is Andersson (1998) and Friedman \& Morsink (1998) who realized 
77: that the $r$-modes in rotating stars are unstable against 
78: the gravitational radiation reaction.
79: Since then, a large number of papers have been published to explore
80: the possible importance of the instability in neutron stars.
81: The $r$-modes in rotating stars are restored by the Coriolis force.
82: They have the dominant toroidal component 
83: of the displacement vector and their oscillation frequencies are
84: comparable to the rotation frequency $\Omega$ of the star
85: (see, e.g., Bryan 1889; Papaloizou \& Pringle 1978;
86: Provost, Berthomieu, \& Roca 1981; Saio 1982; Unno et al. 1989).
87: For the $r$-mode instability,
88: see reviews by, e.g., Friedman \& Lockitch (1999),
89: Andersson \& Kokkotas (2001), Lindblom (2001), and
90: Friedman \& Lockitch (2001).
91: Most of the studies on the $r$-mode instability in neutron stars, however,
92: have been done within the framework of Newtonian dynamics.
93: Since the relativistic factor can be as large as $GM/c^2R\sim 0.2$
94: for neutron stars where $M$ and $R$ are respectively the mass and the radius, 
95: the relativistic effects on the $r$-modes are essential.
96: 
97: From time to time the effects of general relativity on the $r$-modes
98: in neutron stars have been discussed in the context of the $r$-mode instability.
99: Kojima (1998) is the first who investigated the $r$-modes in neutron stars
100: within the framework of general relativity.
101: In the slow rotation approximation, 
102: Kojima (1998) derived a second order ordinary 
103: differential equation governing the relativistic $r$-modes, expanding the 
104: linearized Einstein equation to the first order of $\Omega$, and assuming that 
105: the toroidal component of the displacement vector is dominant and 
106: the oscillation frequency is comparable to $\Omega$.
107: Kojima (1998) showed that this equation has a singular 
108: property and allows a continuous part in the frequency spectrum
109: of the $r$-modes
110: (see, also, Beyer \& Kokkotas 1998). 
111: Recently, Lockitch, Andersson, \& Friedman (2001) showed that Kojima's equation 
112: is appropriate for non-barotropic stars but not for barotropic ones. 
113: They found that both discrete regular $r$-modes and continuous singular 
114: $r$-modes are allowed in Kojima's equation for uniform density stars, and 
115: suggested that the discrete regular $r$-modes are a relativistic 
116: counterpart of the Newtonian $r$-modes. 
117: Yoshida (2001) and Ruoff \& Kokkotas (2001a) 
118: showed that discrete regular $r$-mode solutions to Kojima's equation
119: exist only for some restricted ranges of the polytropic index and the relativistic
120: factor for polytropic models, and 
121: that regular $r$-mode solutions do not exist for the typical ranges of the
122: parameters appropriate for neutron stars. 
123: 
124: 
125: The appearance of continuous singular $r$-mode solutions in
126: Kojima's equation (Kojima 1998) caused a stir in the community of 
127: people who are interested in the $r$-mode instability.
128: Lockitch et al. (2001) (see also Beyer \& Kokkotas 1999) suggested that 
129: the singular property in Kojima's equation could be avoided if 
130: the energy dissipation associated with the gravitational radiation is
131: properly included in the
132: eigenvalue problem because the eigenfrequencies become complex due to the
133: dissipations
134: and the singular point can be detoured when integrated along the real axis.
135: Very recently, this possibility of avoiding the singular property in 
136: Kojima's equation has been examined by Yoshida \& Futamase 
137: (2001) and Ruoff \& Kokkotas (2001b), who showed that the basic 
138: properties of Kojima's equation do not change even if the gravitational 
139: radiation reaction effects are approximately included into the equation. 
140: Quite interestingly, 
141: as shown by Kojima \& Hosonuma (2000), if the third order rotational effects are added
142: to the original Kojima's equation (Kojima 1998),
143: the equation for the $r$-modes becomes 
144: a fourth order ordinary linear differential equation, which has no singular properties 
145: if the Schwarzschild discriminant associated with the buoyant force does 
146: not vanish inside the star.
147: By solving a simplified version of the extended Kojima's equation for a simple
148: toy model,
149: Lockitch \& Andersson (2001) showed that because of the higher order rotational terms  
150: the singular solution in the original Kojima's equation can be avoided.
151: Obviously, we need to solve the complete version of the extended Kojima's equation to
152: obtain a definite conclusion concerning the continuous singular $r$-mode solutions.
153: 
154:  
155: It may be instructive to turn our attention to 
156: a difference in mathematical property between
157: the Newtonian $r$-modes 
158: and the relativistic $r$-modes associated with Kojima's equation. 
159: It is usually assumed that in the lowest order of $\Omega$
160: the eigenfunction as well as the eigenfrequency of the $r$-modes is proportional to $\Omega$.
161: In the case of the Newtonian $r$-modes, 
162: if we employ a perturbative method for $r$-modes in which the angular frequency
163: $\Omega$ is regarded as a small expanding parameter,
164: the radial eigenfunctions of order of $\Omega$ can be determined 
165: by solving a differential equation derived from the terms of order of $\Omega^3$,  
166: which bring about the couplings between the oscillations and the buoyant force
167: in the interior
168: (e.g., Provost et al 1981, Saio 1982).
169: In other words, there is no differential equation, 
170: of order of $\Omega$, which determines the
171: radial eigenfunction of the Newtonian $r$-modes.
172: On the other hand,
173: in the case of the general relativistic $r$-modes derived from Kojima's equation, 
174: we do not have to take account of the rotational effects of order of $\Omega^3$
175: to obtain the radial eigenfunctions of order of $\Omega$.
176: That is, the eigenfrequency and eigenfunction of the $r$-modes are both
177: determined by a differential equation (i.e., Kojima's equation) derived from 
178: the terms of order of $\Omega$.
179: This remains true even if we take the Newtonian limit of Kojima's equation
180: to calculate the $r$-modes (Lockitch et al. 2001; Yoshida 2001).
181: We think that this is an essential difference between the Newtonian $r$-modes
182: and the relativistic $r$-modes associated with Kojima's equation.
183: Considering that Kojima's equation can give no relativistic counterpart of the
184: Newtonian $r$-mode,
185: it is tempting to assume that some terms representing certain physical processes
186: are missing in the original Kojima' equation.
187: On the analogy of the $r$-modes in Newtonian dynamics, 
188: we think that the buoyant force in the interior plays an essential role to obtain
189: a relativistic counterpart of the Newtonian $r$-modes and 
190: that the terms due to the buoyant force
191: will appear when the rotational effects higher than the first order of $\Omega$
192: are included.
193: This is consistent with the suggestions made by
194: Kojima \& Hosonuma (2000) and Lockitch \& Andersson (2001).
195: In this paper, 
196: we calculate relativistic $r$-modes by taking account of the effects of the buoyant force
197: in a relativistic version of the Cowling approximation, in which all the 
198: metric perturbations are omitted.
199: In our formulation, all the terms associated with the Coriolis force are included but 
200: the terms from the centrifugal force are all ignored.
201: Our formulation can take account of the rotational contributions, due to the Coriolis force,
202: higher than the first order of $\Omega$.
203: Note that our method of solution is not a perturbation theory in which
204: $\Omega$ is regarded as a small expanding parameter for the eigenfrequency and eigenfunction.
205: Similar treatment has been employed in Newtonian stellar pulsations in rotating stars
206: (see, e.g., Lee \& Saio 1986; Unno et al. 1989;
207: Bildsten, Ushomirsky \& Cutler 1996; Yoshida \& Lee 2001). 
208: This treatment is justified
209: for low frequency modes because the Coriolis force 
210: dominates the centrifugal force in the equations of motion. 
211: The plan of this paper is as follows. In \S 2, we describe the formulation 
212: of relativistic stellar pulsations in the relativistic Cowling 
213: approximation. In \S 3, we show the modal properties of the $r$-modes in
214: neutron star models. In \S 4, we discuss about our results, specially the effect of 
215: the buoyancy in the fluid core on the modes. \S 5 is devoted to conclusions. 
216: In this paper, we use units in which $c=G=1$, where $c$ and $G$ denote the 
217: velocity of light and the gravitational constant, respectively. 
218: 
219: 
220: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
221: %%  Formulation                                             %%
222: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
223: 
224: \section{Formulation}
225: 
226: \subsection{Equilibrium State}
227: 
228: We consider slowly and uniformly rotating relativistic stars in equilibrium.
229: If we take account of the rotational effects up to first order of $\Omega$,
230: the geometry in the stars can be described by the following line element 
231: (see, e.g. Thorne 1971): 
232: %
233: \begin{eqnarray}
234: ds^2 = g_{\alpha\beta}dx^\alpha dx^\beta = 
235:      - e^{2 \nu(r)} dt^2 + e^{2 \lambda(r)} dr^2 + r^2 d\theta^2 + 
236:            r^2 \sin^2 \theta d\varphi^2 - 
237:            2 \omega(r) \, r^2 \sin^2 \theta dt d\varphi  \, . 
238: \label{metric}
239: \end{eqnarray}
240: %
241: The fluid four-velocity in a rotating star is given by 
242: %
243: \begin{equation}
244: u^\alpha = \gamma(r,\theta) \, (t^\alpha + \Omega\,\varphi^\alpha) \, , 
245: \end{equation}
246: %
247: where $t^\alpha$ and $\varphi^\alpha$ stand for the timelike and rotational 
248: Killing vectors, respectively. 
249: Here the function $\gamma$ is chosen to satisfy 
250: the normalization condition $u^\alpha u_\alpha = -1$. 
251: If we consider the accuracy up to order of $\Omega$, the function $\gamma$ reduces to:
252: %
253: \begin{equation}
254: \gamma = e^{-\nu(r)} \, .  
255: \end{equation}
256: %
257: Once the
258: physical quantities of the star such as the pressure, $p(r)$, 
259: the mass-energy density, $\rho(r)$, and the metric function, $\nu(r)$, are given, 
260: the rotational effect on the metric, $\omega(r)$ can be obtained from a 
261: well-known numerical procedure (see, e.g., Thorne 1971).  
262: 
263: 
264: \subsection{Pulsation Equations in the Cowling Approximation}
265: 
266: General relativistic pulsation equations are usually obtained by linearizing
267: Einstein's field equation.
268: The linearized Einstein equation contains perturbations associated with
269: the metric fluctuations and the fluid motions.
270: In this paper, to simplify the problem, we employ
271: a relativistic version of the Cowling approximation, 
272: in which all the metric perturbations are omitted in the 
273: pulsation equations (see McDermott, Van Horn, \& Scholl 1983, and Finn 1988).
274: The relativistic Cowling approximation is accurate enough 
275: for the $f$- and $p$-modes in non-rotating stars (Lindblom \& Splinter 1992).
276: It is also the case for the modes in slowly rotating stars (Yoshida \& Kojima 1997). 
277: The relativistic Cowling approximation is a good approximation 
278: for oscillation modes in which
279: the fluid motions are dominating over the metric fluctuations 
280: to determine the oscillation frequency.
281: Therefore, it is justified to employ the relativistic
282: Cowling approximation for the $r$-modes, for which the fluid motion is dominating.
283: 
284: 
285: 
286: If we employ the Cowling approximation, we can obtain our basic equations for pulsations
287: from the perturbed energy and momentum conservation laws:
288: %
289: \begin{eqnarray}
290: &&\delta\,(u^\alpha \nabla_\beta T^\beta_\alpha) = 0 \, , \ \ \ \ 
291: {\rm (energy\ conservation\ law)} 
292: \label{ena-eq} \\
293: &&\delta\,(q^\alpha_\gamma \nabla_\beta T^\beta_\alpha) = 0 \, , \ \ \ \
294: {\rm (momentum\ conservation\ law)} 
295: \label{mom-eq}
296: \end{eqnarray}
297: %
298: where $\nabla_\alpha$ is the covariant derivative associated with the metric, 
299: $T^\alpha_\beta$ is the energy-momentum tensor, and $q^\alpha_\beta$ is 
300: the projection tensor with respect to the fluid four-velocity. 
301: Here, $\delta Q$ denotes the Eulerian change in the physical quantity $Q$. 
302: In this paper, we adapt the adiabatic condition for the pulsation: 
303: %
304: \begin{eqnarray}
305: \Delta p = \frac{p\,\Gamma}{\rho+p}\,\Delta \rho \, ,
306: \label{ad-rel}
307: \end{eqnarray}
308: % 
309: where $\Gamma$ is the adiabatic index defined as 
310: %
311: \begin{eqnarray}
312: \Gamma = \frac{\rho+p}{p}\,\left(\frac{\partial p}{\partial \rho}\right)_{ad}
313:  \, ,
314: \end{eqnarray}
315: % 
316: and $\Delta Q$ stands for the Lagrangian change in the physical quantity 
317: $Q$. 
318: The relation between the Lagrangian and the Eulerian changes is given 
319: by the equation: 
320: %
321: \begin{eqnarray}
322: \Delta Q = \delta Q + \pounds_\zeta Q \, , 
323: \end{eqnarray}
324: % 
325: where $\pounds_\zeta$ is the Lie derivative along the Lagrangian displacement vector
326: $\zeta^\alpha$, which is defined by the relation:
327: %
328: \begin{eqnarray}
329: \delta \hat{u}^{\alpha} = q^\alpha_\beta \delta u^\beta = 
330: q^\alpha_\beta (\pounds_u \zeta)^\beta \, .
331: \end{eqnarray}
332: % 
333: Notice that we have 
334: $\delta \hat{u}^{\alpha} = \delta u^\alpha$ in the Cowling 
335: approximation. 
336: Because we are interested in pulsations of stationary rotating 
337: stars, we can assume that all the perturbed quantities have 
338: time and azimuthal dependence given by $e^{i\sigma t+i m \varphi}$, where $m$ is 
339: a constant integer and $\sigma$ is a constant frequency measured by an inertial 
340: observer at the spatial infinity. 
341: Because of this assumption, the relation 
342: between the Lagrangian displacement $\zeta^\alpha$ and the velocity 
343: perturbation $\delta \hat{u}^\alpha$ reduces to an algebraic relationship: 
344: %
345: \begin{eqnarray}
346: \delta \hat{u}^\alpha = i \gamma \bar{\sigma} \zeta^\alpha \, , 
347: \label{def-disp}
348: \end{eqnarray}
349: % 
350: where $\bar{\sigma}$ is the frequency defined in the corotating frame defined as 
351: $\bar{\sigma}=\sigma + m \Omega$. 
352: Note that the gauge freedom in 
353: $\zeta^\alpha$ has been used to demand the relation 
354: $u_\alpha \zeta^\alpha =0$. 
355: By substituting equations (\ref{ad-rel}) and 
356: (\ref{def-disp}) into equations (\ref{ena-eq}) and (\ref{mom-eq}), we can 
357: obtain the perturbed energy equation,     
358: %
359: \begin{eqnarray}
360: \frac{1}{\gamma}\,\nabla_\alpha (\gamma \zeta^\alpha) + 
361: \frac{1}{p \Gamma}\,(\delta p + \zeta^\alpha \nabla_\alpha p) = 0 \, ,
362: \label{ena-eq2}
363: \end{eqnarray}
364: %
365: and the perturbed momentum equation, 
366: %
367: \begin{eqnarray}
368: - \gamma^2\bar{\sigma}^2 g_{\alpha\beta}\zeta^\beta +
369: 2 i \gamma \bar{\sigma} \zeta^\beta \nabla_\beta u_\alpha + 
370: q_\alpha^\beta \nabla_\beta\,\left( \frac{\delta p}{\rho+p}\right) + 
371: \left( \frac{\delta p}{\rho+p}\,q_\alpha^\beta + \zeta^\beta\, 
372: \frac{\nabla_\alpha p}{\rho+p} \right) \, A_\beta = 0 \, ,
373: \label{mom-eq2} 
374: \end{eqnarray}
375: % 
376: where $A_\alpha$ is the relativistic Schwarzschild discriminant defined by 
377: %
378: \begin{eqnarray}
379: A_\alpha = \frac{1}{\rho+p}\,\nabla_\alpha \rho - 
380:            \frac{1}{\Gamma p}\,\nabla_\alpha p \, .
381: \end{eqnarray}
382: %
383: Notice that equations (\ref{ena-eq2}) and (\ref{mom-eq2}) have been derived 
384: without the assumption of slow rotation. 
385: Physically 
386: acceptable solutions of equations (\ref{ena-eq2}) and (\ref{mom-eq2}) must 
387: satisfy boundary conditions at the center and the 
388: surface of the star. 
389: The surface boundary condition at $r=R$ is given by
390: %
391: \begin{eqnarray}
392: \Delta p = \delta p + \zeta^\alpha \nabla_\alpha p = 0 \, ,
393: \end{eqnarray}
394: %
395: and the inner boundary condition is 
396: that all the eigenfunctions are regular at the center ($r=0$). 
397: 
398: 
399: On the analogy between general relativity and Newtonian gravity, 
400: the second term on the left hand side of equation (\ref{mom-eq2}) is 
401: interpreted as a relativistic counterpart of the Coriolis force. 
402: In our formulation, the terms due to the Coriolis 
403: like force are included in the perturbation equations, but the terms due to the 
404: centrifugal like force, which are proportional to $\Omega^2/(GM/R^3)$, are 
405: all ignored. 
406: In Newtonian theory of oscillations, this approximation is justified for low frequency modes 
407: satisfying the conditions $\left|2\Omega/\bar{\sigma}\right|\ge 1$ and 
408: $\Omega^2/(GM/R^3)\ll1$ (Lee \& Saio 1986; Unno et al. 1989; 
409: Bildsten, Ushomirsky \& Cutler 1996; Yoshida \& Lee 2001). 
410: In general relativity, we note that it is difficult to make a clear 
411: distinction between inertial forces such as the Coriolis force and  
412: the centrifugal force. 
413: From a physical point of view, however, it is also acceptable to use the approximation 
414: for low frequency oscillations in the background spacetime described by 
415: the metric (\ref{metric}).  
416: 
417: 
418: The eigenfunctions are expanded in terms of spherical harmonic
419: functions $Y^m_l(\theta,\varphi)$ with different values of $l$ for a given 
420: $m$.  
421: The Lagrangian displacement, $\zeta^k$ and the pressure perturbation, 
422: $\delta p/(\rho+p)$ are expanded as
423: %
424: \begin{equation}
425: \zeta^r = r \sum_{l\geq\vert m \vert}^{\infty}  \, S_l(r)  
426: Y_l^m (\theta,\varphi) \, e^{i \sigma t} \,  , 
427: \label{xi-r}
428: \end{equation}
429: %
430: \begin{equation}
431: \zeta^\theta = \sum_{l,l'\geq\vert m \vert}^{\infty} 
432: \left\{ H_l (r) {\partial {Y_l^m (\theta,\varphi)}\over\partial\theta}  
433: - T_{l'} (r) \frac{1}{\sin \theta} \, 
434: {\partial{Y_{l'}^m (\theta,\varphi)}\over\partial\varphi} \right\} \, 
435: e^{i \sigma t} \, ,
436: \label{xi-th}
437: \end{equation}
438: %
439: \begin{equation}
440: \zeta^\varphi = {1\over\sin^2\theta}\,  
441: \sum_{l,l'\geq\vert m \vert}^{\infty} \left\{ H_l (r) 
442: {\partial {Y_l^m (\theta,\varphi)}\over\partial\varphi}
443:         + T_{l'} (r) \sin\theta\,{\partial {Y_{l'}^m (\theta,\varphi) }
444:             \over\partial\theta} \right\} \, e^{i \sigma t} \, , 
445: \label{xi-ph}
446: \end{equation}
447: %
448: \begin{equation}
449: \frac{\delta p}{\rho+p} = \sum_{l \geq\vert m \vert}^{\infty} \delta U_l (r) 
450: Y_l^m (\theta,\varphi) \, e^{i \sigma t} \, , 
451: \label{del-p}
452: \end{equation}
453: %
454: where $l=|m|+2k$ and $l'=l+1$ for even modes and $l=|m|+2k+1$ and $l'=l-1$
455: for odd modes where $k=0,~1,~2~\cdots$ (Regge \& Wheeler 1957; Thorne 1980).   
456: Here, even and odd modes are, respectively, characterized 
457: by their symmetry and antisymmetry of the eigenfunction with respect to the 
458: equatorial plane. 
459: Substituting the perturbed quantities 
460: (\ref{xi-r})--(\ref{del-p}) into linearized equations (\ref{ena-eq2}) and 
461: (\ref{mom-eq2}), we obtain an infinite system of coupled ordinary differential 
462: equations for the expanded coefficients. 
463: The details of our basic equations are given in the Appendix. 
464: Note that non-linear terms of
465: $q\equiv 2\bar\omega/\bar{\sigma}$ where $\bar\omega\equiv \Omega-\omega$
466: are kept in our basic equations.  
467: For numerical calculations, the infinite set of ordinary differential equations are 
468: truncated to be a finite set by discarding all the expanding coefficients 
469: associated with $l$ larger than $l_{\rm max}$, the value of which is determined 
470: so that the eigenfrequency and the eigenfunctions are well converged
471: as $l_{\rm max}$ increases (Yoshida \& Lee 2000a). 
472: 
473: 
474: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
475: %%  r-Modes of Neutron Star Models                          %%
476: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
477: 
478: \section{$r$-Modes of Neutron Star Models}
479: 
480: 
481: The neutron star models that we use in this paper are the same as those
482: used in the modal analysis by McDermott, Van Horn, \& Hansen (1988).
483: The models are taken from the evolutionary sequences 
484: for cooling neutron stars calculated by Richardson 
485: et al. (1982), where the envelope structure is constructed 
486: by following Gudmundsson, Pethick \& Epstein (1983). 
487: These models are composed of a fluid core, a solid crust and a surface 
488: fluid ocean, and the interior temperature is finite and is not 
489: constant as a function of the radial distance $r$. 
490: The models are not barotropic and the Schwarzschild discriminant 
491: $\vert A\vert$ has finite values in the interior of the star. 
492: In order to avoid the complexity in the modal properties of relativistic $r$-modes
493: brought about by the existence of the solid crust in the models
494: (see Yoshida \& Lee 2001), we treat the whole
495: interior of the models as a fluid in the following modal analysis.
496: 
497: 
498: We computed frequency spectra of $r$-modes for the neutron star models called
499: NS05T7, NS05T8, and NS13T8 (see, McDermott et al. 1988). 
500: The physical properties such as
501: the total mass $M$, the radius $R$, the central 
502: density $\rho_c$, the central temperature $T_c$ and the relativistic factor 
503: $GM/c^2 R$ are summarized in Table 1 
504: (for other quantities, see McDermott et al. 1988).  
505: In Figures 1 and 2, scaled eigenfrequencies $\kappa\equiv\bar{\sigma}/\Omega$ 
506: of the $r$-modes of the three neutron star models are 
507: given as functions of 
508: $\hat\Omega\equiv\Omega/\sqrt{GM/R^3}$ for $m=2$ and $3$ cases, respectively. 
509: Here only the fundamental $r$-modes with $l'=m$ are considered because they 
510: are most important for the $r$-mode instability of neutron stars. 
511: We note that it is practically impossible to correctly calculate 
512: rotationally induced modes at $\hat{\Omega} \sim 0$ because of their coupling with 
513: high overtone $g$-modes having extremely low frequencies. 
514: From these figures, we can see that the scaled eigenfrequency $\kappa$ is 
515: almost constant as $\hat\Omega$ varies. 
516: In other words, the relation $\bar\sigma \sim \kappa_0 \Omega$ 
517: is a good approximation for the fundamental $r$-modes with $l'=m$, 
518: where $\kappa_0$ is a constant. 
519: Comparing the two frequency curves, which nearly overlap each other, 
520: for the models NS05T7 and NS05T8,
521: it is found that the detailed interior structure of the stars such 
522: as the temperature distribution $T(r)$ does not strongly affect
523: the frequency of the fundamental $r$-modes with $l'=m$. 
524: This modal property is the same as that found for the fundamental $l'=m$ $r$-modes 
525: in Newtonian dynamics (see, Yoshida \& Lee 2000b).
526: On the other hand, comparing the frequency curves for the models NS05T7 (NS05T8)
527: and NS13T8, we note that
528: the $r$-mode frequency of relativistic stars is strongly dependent
529: on the relativistic factor $GM/c^2R$ of the models. 
530: This is because the values of the effective rotation frequency 
531: $\bar\omega\equiv\Omega-\omega$ 
532: in the interior is strongly influenced by the relativistic factor. 
533: Similar behavior of the $GM/c^2R$ 
534: dependence of the $r$-mode frequency has been found in the analysis of Kojima's 
535: equation (Yoshida 2001; Ruoff \& Kokkotas 2001a).   
536: In Table 2, the values of $\kappa_0$ for the $r$-modes shown in Figures 1 and 2 are 
537: tabulated, where $\kappa_0$ are evaluated at $\hat\Omega=0.1$. 
538: The boundary values for the continuous part of the frequency spectrum 
539: for the $l'=m=2$ $r$-modes 
540: derived from Kojima's equation (Kojima 1998) are also listed in the same table. 
541: As shown by Kojima (1998) (see, also, Beyer \& Kokkotas 1999; Lockitch et al. 2001), 
542: if an $r$-mode falls in the frequency region bounded by the boundary values, 
543: Kojima's equation becomes singular and yields a continuous frequency spectrum and 
544: singular eigenfunctions as solutions. 
545: Although all the $r$-modes obtained in the present study are in the bounded 
546: frequency region, 
547: our numerical procedure shows that the $r$-modes obtained here are  
548: isolated and discrete eigenmodes.
549: In fact, no sign of continuous frequency spectrum appears in the present numerical
550: analysis (for a sign 
551: of the appearance of a continuous frequency spectrum in a numerical analysis, see
552: Schutz \& Verdaguer 1983). 
553: 
554: 
555: In Figures 3 and 4, the eigenfunctions $i\,T_2$ 
556: for the $l'=m=2$ fundamental $r$-modes 
557: in the neutron star models NS05T8 and NS13T8 at $\hat{\Omega}=0.1$ are shown. 
558: We can confirm from these figures that the
559: eigenfunctions show no singular property, even though the  
560: frequencies are in the continuous part of the spectrum associated with Kojima's equation. 
561: These figures also show
562: that the fluid motion due to the $r$-modes is more confined 
563: near the stellar surface for NS13T8 than for NS05T8. 
564: This suggests that the eigenfunctions $i\,T_m$ tend to be confined to the stellar surface 
565: as the relativistic factor of the star increases. 
566: The same property appears in regular $r$-mode solutions derived 
567: from Kojima's equation (Yoshida 2001; Ruoff \& Kokkotas 2001a; 
568: Yoshida \& Futamase 2001). 
569: But, in the present case, the confinement of the eigenfunction for NS13T8 
570: is not so strong, even though the model NS13T8 is highly 
571: relativistic in the sense that the relativistic factor is as large as
572: $GM/c^2R=0.249$.  
573:  
574: 
575: \section{Discussion}
576: 
577: 
578: In the neutron star 
579: models analyzed in the last section, the buoyant force is produced by 
580: thermal stratification in the star. 
581: However, as suggested by 
582: Reisenegger \& Goldreich (1992), the buoyant force in the core of neutron 
583: stars might become even stronger if the effect of 
584: the smooth change of the chemical composition 
585: of charged particles (protons and electrons) in the core is taken into account.
586: We thus think it legitimate to examine the effects of the enhanced
587: buoyant force on the modal properties of the relativistic $r$-modes.
588: Although our neutron star models do not provide 
589: enough information regarding the composition gradient, 
590: we may employ for this experiment an approximation formula for the 
591: Schwarzschild discriminant due to the composition gradient in the core
592: given by Reisenegger \& Goldreich (1992):
593: %
594: \begin{equation}
595: r A_r = 3.0\times 10^{-3}\,\left(\frac{\rho}{\rho_{nuc}}\right)
596: \frac{r}{\rho}\frac{d\rho}{dr} \, , 
597: \label{cma}
598: \end{equation} 
599: %
600: where $\rho_{nuc}$ denotes the nuclear density $\rho_{nuc}=2.8\times10^{14} 
601: \, {\rm g \, cm^{-3}}$. 
602: In Figure 5, Brunt-V\"ais\"al\"a 
603: frequency due to the thermal stratification (solid line) and that due to the composition
604: gradient (dashed line) are given as a function of $\log(1-r/R)$ 
605: for model NS13T8, where the frequencies are normalized by $(GM/R^3)^{1/2}$.
606: In this paper, the relativistic Brunt-V\"ais\"al\"a frequency $N$ is defined as  
607: %
608: \begin{equation}
609: N^2 = \frac{A_r}{\rho+p}\, \frac{d p}{dr} \, .
610: \end{equation}
611: %
612: This figure shows that in the core
613: the Brunt-V\"ais\"al\"a frequency due to the composition gradient
614: is by several orders of magnitude larger and hence
615: the corresponding buoyant force is much stronger than those produced by
616: the thermal stratification.
617: Note that the Brunt-V\"ais\"al\"a frequency due to the composition gradient
618: becomes as large as $\sim(GM/R^3)^{1/2}$ in the core (see also, e.g., Lee 1995).
619: 
620: Replacing the Schwarzschild discriminant in the original neutron star models
621: with that due to the composition gradient calculated by using equation (19),
622: we computed the fundamental $r$-modes with $l'=m=2$ for the models NS05T8 and NS13T8,
623: and plotted $\kappa=\bar\sigma/\hat \Omega$
624: against rotation frequency $\hat{\Omega}$ in Figure 6, where
625: $\kappa$'s for the original models with the thermal stratification were
626: also plotted for convenience. 
627: As shown by the figure, $\kappa$ slightly increases
628: as $\hat\Omega\rightarrow 0$ for the models with the enhanced buoyant force.
629: This is remarkable because such behavior of $\kappa$ for the fundamental $r$-modes
630: with $l'=m$
631: is not found in the case of Newtonian 
632: $r$-modes in non-barotropic stars (Yoshida \& Lee 2000b).
633: The slight increase of $\kappa$ with decreasing $\hat\Omega$
634: may be explained in terms of the
635: property of the eigenfunctions of the $r$-modes.
636: We depict
637: the eigenfunctions $i\,T_2$ of the $l'=m=2$ fundamental $r$-modes 
638: in the neutron star model NS13T8 at $\hat{\Omega}=0.2$ and $.02$
639: in Figures 7 and 8,
640: where the solid line and dashed line in each figure denote the $r$-modes 
641: in the models with the compositional stratification
642: and with the thermal stratification, respectively.
643: As shown by Figure 7, in the case of rapid rotation,
644: the buoyant force in the core does not strongly
645: affect the properties of the eigenfunctions.
646: On the other hand, in the case of slow rotation, Figure 8 shows that
647: the amplitude of the eigenfunction is strongly confined to the region near
648: the surface for the model with the enhanced buoyant force.
649: For relativistic $r$-modes, as a result of the general relativistic frame dragging effect,
650: the effective rotation frequency $\bar\omega=\Omega-\omega\propto\Omega$ 
651: acting on local fluid elements 
652: is an increasing function of the distance $r$ from the stellar center, and
653: thus fluid elements near the stellar surface are rotating faster 
654: than those near the stellar center. 
655: The oscillation frequency of the $r$-modes may be determined by
656: the mean rotation frequency obtained by averaging local 
657: rotation frequencies over the whole interior of the star with
658: a certain weighting function associated with the eigenfunction.
659: In this case, it is reasonable to expect that 
660: the relativistic $r$-modes that have large amplitudes only in the regions
661: near the surface
662: get larger values of $\kappa$ than those that have large amplitudes deep in the core.
663: 
664: 
665: For the models with the enhanced buoyant force,
666: the current quadrupole moment of the $r$
667: modes that determines the strength of the instability will be largely reduced
668: when $\hat\Omega$ is small as a result of the strong confinement of the amplitudes.
669: But, since the $r$-mode instability is believed to operate at rapid rotation rates,
670: the effect of the amplitude confinement at small rotation rates
671: may be irrelevant to the instability.
672: 
673: 
674: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
675: %%       DISCUSSIONS AND CONCLUSIONS                        %%
676: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
677: 
678: 
679: \section{Conclusion}
680: 
681: 
682: In this paper, we have investigated the properties of relativistic
683: $r$-modes in slowly rotating neutron stars 
684: in the relativistic Cowling approximation by taking account of 
685: higher order effects of rotation than the first oder of $\Omega$.
686: In our formalism, only the influence of the Coriolis like 
687: force on the oscillations are taken into account, and no effects of 
688: the centrifugal like force are considered. 
689: We obtain the fundamental $r$-modes associated with 
690: $l'=m=2$ and $3$ for three neutron star models. 
691: We find that the fundamental $r$-mode frequencies are in a good approximation 
692: given by $\bar\sigma\approx \kappa_0\,\Omega$.
693: The proportional coefficient $\kappa_0$ 
694: is only weakly dependent on $\Omega$, but strongly
695: depends on the relativistic parameter $GM/c^2R$.
696: For the fundamental $r$-modes with $l'=m$, we find that
697: the buoyant force in the core is more influential to the relativistic
698: $r$-modes than to the Newtonian $r$-modes.
699: All the $r$-modes obtained in this paper are discrete modes with distinct regular 
700: eigenfunctions, and they all fall in 
701: the frequency range of the continuous spectrum of Kojima's equation. 
702: We may conclude that the fundamental $r$-modes obtained in this paper are
703: the relativistic counterpart of the Newtonian $r$-modes.
704: 
705: 
706:     
707: Here, it is legitimate to mention the relation between the present work and 
708: the study by Kojima \& Hosonuma (1999), who studied relativistic $r$-modes 
709: in slowly rotating stars in the Cowling approximation. 
710: By applying the Laplace transformation to linearized equations 
711: derived for the $r$-modes, Kojima \& Hosonuma (1999) examined 
712: how a single component of initial perturbations with axial parity
713: evolves as time goes, and showed
714: that the perturbations cannot oscillate with a single frequency. 
715: In this paper, we have presented a formulation for small
716: amplitude relativistic oscillations in rotating stars 
717: in the relativistic Cowling approximation,
718: and solved the oscillation equations as a boundary-eigenvalue problem for the $r$-modes.
719: In our formulation, we assume neither axially dominant eigenfunctions nor
720: low frequencies to calculate the $r$-modes.
721: Considering these differences in the treatment of the $r$-mode oscillations
722: between the two studies,
723: it is not surprising that our results are not necessarily consistent with
724: those by Kojima \& Hosonuma (1999).
725: 
726: 
727: 
728: Our results suggest that the appearance of singular $r$-mode solutions
729: can be avoided by extending the original Kojima's equation so that
730: terms due to the buoyant force in the stellar interior are included. 
731: As discussed recently by Yoshida \& Futamase (2001) and Lockitch \& 
732: Andersson (2001), we believe that the answer to the question whether 
733: $r$-mode oscillations in uniformly rotating relativistic stars show true 
734: singular behavior may be given by solving the forth order 
735: ordinary differential equation derived by Kojima \& 
736: Hosonuma (2000) for $r$-modes. 
737: Verification of this possibility remains as a future study. 
738: 
739: 
740: 
741: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
742: %%         END OF MAIN BODY OF PAPER                        %%
743: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
744: 
745: \acknowledgements
746: 
747: We would like to thank H. Saio for useful comments. We are grateful to the 
748: anonymous referee for useful suggestions regarding the buoyancy due to the 
749: composition gradient. S.Y. would like to thank Y. Eriguchi and T. Futamase for 
750: fruitful discussions and continuous encouragement. 
751: 
752: 
753: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
754: %% APPENDIX                                                 %%
755: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
756: 
757: \appendix
758: 
759: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
760: %% Basic equations for ...                                  %%
761: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
762: 
763: \section{Basic Equations}
764: 
765: We introduce column vectors $\bf{y}_1$, $\bf{y}_2$, $\bf{h}$, and $\bf{t}$, 
766: whose components are defined by 
767: %
768: \begin{equation}
769: y_{1, k} = S_l (r) \, ,
770: \end{equation}
771: %
772: \begin{equation}
773: y_{2, k} = \frac{1}{r \frac{d\nu}{dr}} \,  \delta U_l(r) \, , 
774: \end{equation}
775: %
776: \begin{equation}
777: h_{,k} = H_l (r) \, ,
778: \end{equation}
779: %
780: and
781: %
782: \begin{equation}
783: t_{,k} = T_{l'} (r) \, ,
784: \end{equation}
785: %
786: where $l=\vert m \vert + 2 k -2 $ and $l'=l+1$ for ``even'' modes, and 
787: $l=\vert m \vert + 2 k -1 $ and $l'=l-1$ for ``odd'' modes, and $k = 1, 2, 3, \dots$. 
788: 
789: 
790: In vector notation, equations for the adiabatic nonradial pulsation
791: in a slowly rotating star are written as follows:  
792: 
793: 
794: 
795: \noindent
796: The perturbed energy equation (\ref{ena-eq2}) reduces to 
797: %
798: \begin{eqnarray}
799: r \, \frac{d {\bf y}_1}{dr} + 
800: \left( 3 - \frac{V}{\Gamma} + r\frac{d \lambda}{dr} \right) \, {\bf y}_1
801: + \frac{V}{\Gamma}\,{\bf y}_2 - {\bf \Lambda}_0 {\bf h}  
802: +c_2 \hat{\bar{\omega}} \hat{\sigma} \, (-m {\bf h}+{\bf C}_0 i {\bf t}) 
803: = 0 \, .
804: \label{differ-1}
805: \end{eqnarray}
806: %
807: The $r$ component of the perturbed momentum equations (\ref{mom-eq2}) 
808: reduces to 
809: %
810: \begin{eqnarray}
811: r \, \frac{d{\bf y}_2}{dr} - 
812: ( e^{2\lambda}\,c_1 \hat{\bar{\sigma}}^2 + r A_r ) \, {\bf y}_1 
813: +(U+r A_r) {\bf y}_2 - c_1 \hat{\bar{\sigma}}^2 \chi
814: (-m {\bf h}+{\bf C}_0 i {\bf t}) = 0 \, .
815: \label{differ-2}
816: \end{eqnarray}
817: %
818: Here, 
819: %
820: \begin{equation}
821: U = \frac{r\frac{d}{dr}(r\frac{d\nu}{dr})}{\frac{d\nu}{dr}} \, , 
822: \hspace{0.5in}
823: V = - \frac{d\ln p}{d\ln r} \, ,
824: \end{equation}
825: %
826: \begin{equation}
827: \chi = \frac{e^{2\nu}}{r}\frac{d}{dr} 
828: \left( r^2 e^{-2 \nu} \frac{\bar{\omega}}{\bar{\sigma}}\right)\, ,
829: \end{equation}
830: %
831: \begin{equation}
832: c_1 = \frac{r^2 e^{-2 \nu}}{r \frac{d\nu}{dr}}\, \frac{M}{R^3} \, , 
833: \hspace{0.5in}
834: c_2 = r^2 e^{-2 \nu} \frac{M}{R^3} \, ,
835: \end{equation}
836: %
837: and
838: $\hat{\bar{\omega}} \equiv \bar{\omega}/(GM/R^3)^{1/2}$, 
839: $\hat{\bar{\sigma}} \equiv \bar{\sigma}/(GM/R^3)^{1/2}$, and 
840: $\hat{\sigma} \equiv \sigma/(GM/R^3)^{1/2}$ are frequencies in the unit of 
841: the Kepler frequency at the stellar surface, where $\bar\omega\equiv\Omega-\omega$
842: and $\bar\sigma\equiv\sigma+m\Omega$. 
843: 
844: \noindent
845: The $\theta$ and $\varphi$ components of the perturbed momentum equations 
846: (\ref{mom-eq2}) reduce to 
847: %
848: \begin{equation}
849: {\bf L}_0 \, {\bf h} + {\bf M}_1 \, i {\bf t}
850:  = \frac{1}{c_1 \hat{\bar{\sigma}}^2} \, {\bf y}_2 + \lbrace 
851: {\bf O}+{\bf M}_1{\bf L}_1^{-1}{\bf K}
852:  \rbrace \, \left(\chi \,{\bf y}_1+
853:  \frac{c_2 \hat{\bar{\omega}}\hat{\sigma}}{c_1 \hat{\bar{\sigma}}^2}\,
854:  {\bf y_2} \right) \, , \label{alge-1}
855: \end{equation}
856: %
857: \begin{equation}
858: {\bf L}_1 \, i {\bf t} + {\bf M}_0 \, {\bf h}
859:  = {\bf K} \, \left(\chi \,{\bf y}_1+
860:  \frac{c_2 \hat{\bar{\omega}}\hat{\sigma}}{c_1 \hat{\bar{\sigma}}^2}\,
861:  {\bf y_2} \right) \, , \label{alge-2}
862: \end{equation}
863: %
864: where 
865: \begin{equation}
866: {\bf O} = m {\bf \Lambda}_0^{-1} - {\bf M}_1{\bf L}_1^{-1}{\bf K} \, . 
867: \end{equation}
868: 
869: The quantities ${\bf C}_0$, ${\bf K}$, ${\bf L}_0$, 
870: ${\bf L}_1$, ${\bf \Lambda}_0$, ${\bf \Lambda}_1$, ${\bf M}_0$, 
871: ${\bf M}_1$ are matrices written as follows:  
872: %
873: 
874: \noindent
875: For even modes, 
876: %
877: \[
878: ({\bf C}_0)_{i,i} = - (l+2) J^m_{l+1} \, , \hspace{.3in}
879: ({\bf C}_0)_{i+1,i} = (l+1) J^m_{l+2} \, ,
880: \]
881: %
882: \[
883: ({\bf K})_{i,i} = \frac{J^m_{l+1}}{l+1} \, , \hspace{.3in}
884: ({\bf K})_{i,i+1} = - \frac{J^m_{l+2}}{l+2} \, ,
885: \]
886: %
887: \[
888: ({\bf L}_0)_{i,i} = 1 - \frac{m q}{l(l+1)} \, , \hspace{.3in}
889: ({\bf L}_1)_{i,i} = 1 - \frac{m q}{(l+1)(l+2)} \, ,
890: \]
891: %
892: \[
893: ({\bf \Lambda}_0)_{i,i} = l(l+1) \, , \hspace{.3in}
894: ({\bf \Lambda}_1)_{i,i} = (l+1)(l+2) \, ,
895: \]
896: %
897: \[
898: ({\bf M}_0)_{i,i} = q \frac{l}{l+1} \, J^m_{l+1} \, , \hspace{.3in}
899: ({\bf M}_0)_{i,i+1} = q \frac{l+3}{l+2} \, J^m_{l+2} \, ,
900: \]
901: %
902: \[
903: ({\bf M}_1)_{i,i} = q \frac{l+2}{l+1} \, J^m_{l+1} \, , \hspace{.3in}
904: ({\bf M}_1)_{i+1,i} = q \frac{l+1}{l+2} \, J^m_{l+2} \, ,
905: \]
906: %
907: where $l=\vert m \vert + 2 i -2 $ for $i = 1,2,3,\dots $, and
908: $q \equiv 2 \bar{\omega} / \bar{\sigma}$, and
909: \begin{equation}
910: J^m_l \equiv \left[ \frac{(l+m)(l-m)}{(2l-1)(2l+1)} \right]^{1/2} 
911: \, .  \label{J_l}
912: \end{equation}
913: %
914: 
915: 
916: \noindent
917: For odd modes, 
918: %
919: \[
920: ({\bf C}_0)_{i,i} = (l-1) J^m_{l} \, , \hspace{.3in}
921: ({\bf C}_0)_{i,i+1} = -(l+2) J^m_{l+1} \, ,
922: \]
923: %
924: \[
925: ({\bf K})_{i,i} = - \frac{J^m_l}{l} \, , \hspace{.3in}
926: ({\bf K})_{i+1,i} = \frac{J^m_{l+1}}{l+1} \, ,
927: \]
928: %
929: \[
930: ({\bf L}_0)_{i,i} = 1 - \frac{m q}{l(l+1)} \, , \hspace{.3in}
931: ({\bf L}_1)_{i,i} = 1 - \frac{m q}{l(l-1)} \, ,
932: \]
933: %
934: \[
935: ({\bf \Lambda}_0)_{i,i} = l(l+1) \, , \hspace{.3in}
936: ({\bf \Lambda}_1)_{i,i} = l(l-1) \, ,
937: \]
938: %
939: \[
940: ({\bf M}_0)_{i,i} = q \frac{l+1}{l} \, J^m_l \, , \hspace{.3in}
941: ({\bf M}_0)_{i+1,i} = q \frac{l}{l+1} \, J^m_{l+1} \, ,
942: \]
943: %
944: \[
945: ({\bf M}_1)_{i,i} = q \frac{l-1}{l} \, J^m_l \, , \hspace{.3in}
946: ({\bf M}_1)_{i,i+1} = q \frac{l+2}{l+1} \, J^m_{l+1} \, ,
947: \]
948: %
949: where $l=\vert m \vert + 2 i -1 $ for $i = 1,2,3,\dots $.
950: 
951: Eliminating $\bf h$ and 
952: $\it i \bf t$ from equations (\ref{differ-1}) and (\ref{differ-2}) by 
953: using equations (\ref{alge-1}) and (\ref{alge-2}), equations 
954: (\ref{differ-1}) and (\ref{differ-2}) reduce to a set of first-order 
955: linear ordinary  differential equations for ${\bf y}_1$ and ${\bf y}_2$ 
956: as follows:
957: %  
958: \begin{eqnarray}
959: r\,\frac{d {\bf y}_1}{dr} &=& 
960:  \left\lbrace \left( \frac{V}{\Gamma} -3 -r\,\frac{d\lambda}{dr}\right) \, 
961:  {\bf 1} + \chi \,\left({\bf{\cal F}}_{11} + 
962:  c_2\hat{\bar{\omega}}\hat{\sigma} 
963:  {\bf{\cal F}}_{21}\right) \right\rbrace \, {\bf y}_1 \nonumber \\
964: &+& \left\lbrace 
965: \frac{1}{c_1 \hat{\bar{\sigma}}^2}\left({\bf{\cal F}}_{12} + 
966:  c_2\hat{\bar{\omega}}\hat{\sigma}
967:  {\bf{\cal F}}_{22}\right) + 
968:  \frac{c_2 \hat{\bar{\omega}}\hat{\sigma}}{c_1 \hat{\bar{\sigma}}^2}
969:  \left({\bf{\cal F}}_{11} + c_2\hat{\bar{\omega}}\hat{\sigma}
970:  {\bf{\cal F}}_{21}\right) - \frac{V}{\Gamma}{\bf 1} 
971: \right\rbrace 
972: \, {\bf y}_2 \, , \label{basic-eq1}
973: \end{eqnarray}
974: %
975: \begin{eqnarray}
976: r\,\frac{d {\bf y}_2}{dr} &=& \Bigl\lbrace 
977: (e^{2\lambda} c_1 \hat{\bar{\sigma}}^2 + r A_r) \, {\bf 1} - 
978: c_1 \hat{\bar{\sigma}}^2 q^2 {\bf{\cal F}}_{21} \Bigr\rbrace \, 
979: {\bf y}_1 \nonumber \\ 
980: &+& \Bigl\lbrace -(r A_r+U){\bf 1} - \chi \,( {\bf{\cal F}}_{22} + 
981:  c_2\hat{\bar{\omega}}\hat{\sigma} {\bf{\cal F}}_{21} )\Bigr\rbrace 
982:  \, {\bf y}_2 \, , \label{basic-eq2}
983: \end{eqnarray}
984: %
985: where 
986: %
987: \begin{equation}
988: {\bf{\cal F}}_{11} = {\bf W} {\bf O} \, ,
989: \end{equation}
990: %
991: \begin{equation}
992: {\bf{\cal F}}_{12} = {\bf W} \, ,
993: \end{equation}
994: %
995: \begin{eqnarray}
996: {\bf{\cal F}}_{21} = {\bf{\cal R}} {\bf W} {\bf O} 
997:  - {\bf C}_0 {\bf L}_1^{-1} {\bf K} \, ,
998: \end{eqnarray}
999: %
1000: \begin{equation}
1001: {\bf{\cal F}}_{22} = {\bf{\cal R}} \, {\bf W} \, ,
1002: \end{equation}
1003: %
1004: \begin{eqnarray}
1005: {\bf{\cal R}} = m{\bf \Lambda}_0^{-1} 
1006:  + {\bf C}_0 {\bf L}_1^{-1} {\bf M}_0 {\bf \Lambda}_0^{-1} \, ,
1007: \end{eqnarray}
1008: %
1009: \begin{equation}
1010: {\bf W} = {\bf \Lambda}_0 
1011: ({\bf L}_0 - {\bf M}_1 {\bf L}_1^{-1} {\bf M}_0 )^{-1} \, .
1012: \end{equation}
1013: %
1014: The surface boundary conditions $\Delta p (r=R)  = 0$ are given by 
1015: %
1016: \begin{equation}
1017: {\bf y}_1 - {\bf y}_2  = 0 \, . \label{boundary-1}
1018: \end{equation}
1019: %
1020: The inner boundary conditions at the stellar center are the regularity 
1021: conditions of the eigenfunctions.
1022: 
1023: 
1024: 
1025: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1026: %%  BIBLIOGRAPHY                                            %%
1027: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1028: 
1029: \begin{thebibliography}{}
1030: 
1031: \bibitem[Andersson 1998]{nils97}
1032:  Andersson, N. 1998, \apj, 502, 708
1033: %
1034: \bibitem[Andersson \& Kokkotas 2001]{ak01}
1035:  Andersson, N., \& Kokkotas, K. D. 2001, Int. J. Mod. Phys., D 10, 381 
1036: %
1037: \bibitem[Beyer \& Kokkotas 1999]{bk99}
1038:  Beyer, H. R., \& Kokkotas, K. D. 1999, \mnras, 308, 745
1039: %
1040: \bibitem[Bildsten, Ushomirsky \& Cutler 1996]{buc96}
1041:  Bildsten, L., Ushomirsky, G., \& Cutler, C. 1996, \apj, 460, 827 
1042: %
1043: \bibitem[Bryan 1889]{b1889}
1044:  Bryan, G. H. 1889, Phil. Trans. R. Soc. London, A 180, 187 
1045: %
1046: \bibitem[Finn 1988]{f88}
1047:  Finn, L. S. 1988, \mnras, 101, 367
1048: %
1049: \bibitem[Friedman \& Lockitch 1999]{fl99} 
1050:  Friedman, J. L., \& Lockitch, K. H. 1999, Prog. Theor. Phys. Suppl., 136, 121
1051: %
1052: \bibitem[Friedman \& Lockitch 2001]{fl01} 
1053:  Friedman, J. L., \& Lockitch, K. H. 2001, preprint (gr-qc/0102114)
1054: %
1055: \bibitem[Friedman \& Morsink 1998]{jfs97} 
1056:  Friedman, J. L., \& Morsink, S. M. 1998, \apj, 502, 714
1057: %
1058: \bibitem[Gudmundsson, Pethick \& Epstein 1983]{gpe83} 
1059:  Gudmundsson, E. H., Pethick, C. J., \& Epstein, R. I. 1983, \apj, 272, 286
1060: %
1061: \bibitem[Kojima 1998]{k98}
1062:  Kojima, Y. 1998, \mnras, 293, 49
1063: %
1064: \bibitem[Kojima \& Hoshonuma 1999]{kh99}
1065:  Kojima, Y., \& Hosonuma, M. 1999, \apj, 520, 788
1066: %
1067: \bibitem[Kojima \& Hoshonuma 2000]{kh00}
1068:  Kojima, Y., \& Hosonuma, M. 2000, \prd, 62, 044006
1069: %
1070: \bibitem[Lee 1995]{l95}
1071:  Lee, U. 1995, \aap, 303, 515
1072: %
1073: \bibitem[Lee \& Saio 1986]{ls86}
1074:  Lee, U., \& Saio, H. 1986, \mnras, 221, 365
1075: %
1076: \bibitem[Lindblom 2001]{l01}
1077:  Lindblom, L. 2001, preprint (astro-ph/0101136) 
1078: %
1079: \bibitem[Lindblom \& Splinter 1990]{ls90}
1080:  Lindblom, L., \& Splinter, R. J. 1990, \apj, 348, 198
1081: %
1082: \bibitem[Lockitch \& Andersson 2001]{la01}
1083:  Lockitch, K. H., \& Andersson, N. 2001, preprint (gr-qc/0106088)
1084: %
1085: \bibitem[Lockitch et al. 2001]{laf01}
1086:  Lockitch, K. H., Andersson, N.,\& Friedman, J. L. 2001, \prd, 63, 024019
1087: %
1088: \bibitem[McDermott, Van Horn, \& Hansen 1988]{mvh88}
1089:  McDermott, P. N., Van Horn, H. M, \& Hansen, C. J. 1988, \apj, 325, 725
1090: %
1091: \bibitem[McDermott, Van Horn, \& Scholl 1983]{mvs83}
1092:  McDermott, P. N., Van Horn, H. M, \& Scholl, J. F. 1983, \apj, 268, 837
1093: %
1094: \bibitem[Papaloizou \& Pringle 1978]{pp78}
1095:  Papaloizou, J., \& Pringle, J. E. 1978, \mnras, 182, 423
1096: %
1097: \bibitem[Provost et al. 1981]{pbr81} 
1098:  Provost, J., Berthomieu, G., \& Rocca, A. 1981, \aap, 94, 126
1099: %
1100: \bibitem[Regge \& Wheeler 1957]{rw57}
1101:  Regge, T., \& Wheeler, J. A. 1957, Phys. Rev., 108, 1063  
1102: %
1103: \bibitem[Reisenegger \& Goldreich 1992]{rg92}
1104:  Reisenegger, A. \& Goldreich, P. 1992, \apj, 395, 240
1105: %
1106: \bibitem[Richardson et al. 1982]{rvrm82}
1107:  Richardson, M. B., Van Horn, H. M., Ratcliff, K. F., \& Malone, R. C. 1982, 
1108: \apj, 255, 624
1109: %
1110: \bibitem[Ruoff \& Kokkotas 2001a]{rk01a}
1111:  Ruoff, J., \& Kokkotas, K. D. 2001a, preprint (gr-qc/0101105)
1112: %
1113: \bibitem[Ruoff \& Kokkotas 2001b]{rk01b}
1114:  Ruoff, J., \& Kokkotas, K. D. 2001b, preprint (gr-qc/0106073)
1115: %
1116: \bibitem[Saio 1982]{s82}
1117:  Saio, H. 1982, \apj, 256, 717
1118: %
1119: \bibitem[Schutz \& Verdaguer 1983]{sv83}
1120:  Schutz, B. F., \& Verdaguer, E. 1983, \mnras, 202, 881 
1121: %
1122: \bibitem[Thorne 1971]{t71}
1123:  Thorne, K. S. 1971, in General Relativity and Cosmology, ed. R. K. Sachs 
1124:  (Academic Press), 237
1125: %
1126: \bibitem[Thorne 1980]{t80}
1127:  Thorne, K. S. 1980, Rev. Mod. Phys., 52, 299
1128: %
1129: \bibitem[Unno et al.\ 1989]{u89}
1130:  Unno, W., Osaki, Y., Ando, H., Saio, H., \& Shibahashi, H. 1989, 
1131:  Nonradial Oscillations of Stars, Second Edition  (Tokyo: Univ. Tokyo Press) 
1132: %
1133: \bibitem[Yoshida 2001]{y01}
1134:  Yoshida, S. 2001, \apj, 558, 263
1135: %
1136: \bibitem[Yoshida \& Futamase 2001]{yf01}
1137:  Yoshida, S., \& Futamase, T. 2001, \prd, 64, 123001 
1138: %
1139: \bibitem[Yoshida \& Kojima 1997]{yk97}
1140:  Yoshida, S., \& Kojima, Y. 1997, \mnras, 289, 117
1141: %
1142: \bibitem[Yoshida \& Lee 2000a]{yl00a}
1143:  Yoshida, S., \& Lee, U. 2000a, \apj, 529, 997
1144: %
1145: \bibitem[Yoshida \& Lee 2000b]{yl00b}
1146:  Yoshida, S., \& Lee, U. 2000b, \apjs, 129, 353
1147: %
1148: \bibitem[Yoshida \& Lee 2001]{yl01}
1149:  Yoshida, S., \& Lee, U. 2001, \apj, 546, 1121
1150: %
1151: \end{thebibliography}
1152: %
1153: 
1154: 
1155: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1156: %% TABLES                                                   %%
1157: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1158: 
1159: \newpage
1160: 
1161: \begin{deluxetable}{cccccc}
1162: \footnotesize
1163: \tablecaption{Neutron Star Models}
1164: \tablewidth{0pt}
1165: \tablehead{ \colhead{Model} & \colhead{$M\ (M_{\sun})$}  
1166:  & \colhead{$R$ (km)} & \colhead{$\rho_c$ (g $\rm cm^3$)}    
1167:  & \colhead{$T_c$ (K)} & \colhead{$GM/(c^2 R)$}
1168: } 
1169: \startdata
1170: NS05T7&$0.503$&$9.839$&$9.44\times 10^{14}$&$1.03\times 10^7$
1171: &$7.54\times 10^{-2}$\nl
1172: NS05T8&$0.503$&$9.785$&$9.44\times 10^{14}$&$9.76\times 10^7$
1173: &$7.59\times 10^{-2}$\nl 
1174: NS13T8&$1.326$&$7.853$&$3.63\times 10^{15}$&$1.05\times 10^8$
1175: &$2.49\times 10^{-1}$\nl
1176: \enddata
1177: \label{ns-model}
1178: \end{deluxetable}
1179: 
1180: \begin{deluxetable}{ccccc}
1181: \footnotesize
1182: \tablecaption{Scaled Eigenfrequencies $\kappa_0$ of Fundamental 
1183: $r$-modes}
1184: \tablewidth{0pt}
1185: \tablehead{
1186:  \colhead{Model}&\colhead{$\kappa_0 (l=m=2)$}
1187:                 &\colhead{$\kappa_0 (l=m=3)$}
1188:                 &\colhead{$2/3\times\bar{\omega}(0)/\Omega$}
1189:                 &\colhead{$2/3\times\bar{\omega}(R)/\Omega$}
1190: } 
1191: \startdata
1192: NS05T7 &$ 0.600 $&$ 0.455 $&$ 0.523 $&$ 0.645 $ \nl
1193: NS05T8 &$ 0.601 $&$ 0.456 $&$ 0.524 $&$ 0.642 $ \nl
1194: NS13T8 &$ 0.393 $&$ 0.309 $&$ 0.208 $&$ 0.489 $ \nl
1195: \enddata
1196: \label{eigen-f}
1197: \end{deluxetable}
1198: 
1199: 
1200: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1201: %% FIGURES                                                  %%
1202: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1203: 
1204: 
1205: \newpage
1206: 
1207: \begin{figure}
1208: \epsscale{.5}
1209: \plotone{f1.eps}
1210: \caption{Scaled frequencies $\kappa=\bar{\sigma}/\Omega$ of 
1211: the fundamental $r$-modes in the neutron star models NS05T7, NS05T8, and NS13T8 
1212: are plotted as functions of $\hat{\Omega}=\Omega/(GM/R^3)^{1/2}$ for 
1213: $l'=m=2$. 
1214: Note that the two $r$-mode frequency curves for the models NS05T7 and NS05T8
1215: overlap each other almost completely.}
1216: \end{figure}
1217: 
1218: %\newpage
1219: 
1220: \begin{figure}
1221: \epsscale{.5}
1222: \plotone{f2.eps}
1223: \caption{Same as Figure 1 but for the case of $m=3$.}
1224: \end{figure}
1225: 
1226: \newpage
1227: 
1228: \begin{figure}
1229: \epsscale{.5}
1230: \plotone{f3.eps}
1231: \caption{Eigenfunction $i\ T_2$ of the 
1232: $r$-mode with $l'=m=2$ for the model NS05T8 at $\hat{\Omega}=0.1$ 
1233: is given as a function of $r/R$. Here, normalization of the eigenfunction 
1234: is chosen as $i\ T_2(R)=1$.} 
1235: \end{figure}
1236: 
1237: \begin{figure}
1238: \epsscale{.5}
1239: \plotone{f4.eps}
1240: \caption{Same as Figure 4 but for the model NS13T8.}
1241: \end{figure}
1242: 
1243: \begin{figure}
1244: \epsscale{.5}
1245: \plotone{f5.eps}
1246: \caption{Brunt-V\"ais\"al\"a frequency due to the thermal stratification (solid line) and that 
1247: due to the composition gradient (dashed line) are given as a function of $\log(1-r/R)$ 
1248: for model NS13T8, where the frequencies are normalized by $(GM/R^3)^{1/2}$.}
1249: \end{figure}
1250: 
1251: \begin{figure}
1252: \epsscale{.5}
1253: \plotone{f6.eps}
1254: \caption{Scaled frequencies $\kappa=\bar{\sigma}/\Omega$ of 
1255: the $l'=m=2$ fundamental $r$-modes in the neutron star models NS05T8 and NS13T8 
1256: are plotted as functions of $\hat{\Omega}=\Omega/(GM/R^3)^{1/2}$. 
1257: Labels "comp." and "temp." stand for the frequencies of the models with 
1258: the buoyancy due to the composition gradient and due to the temperature gradient, 
1259: respectively.}
1260: \end{figure}
1261: 
1262: \begin{figure}
1263: \epsscale{.5}
1264: \plotone{f7.eps}
1265: \caption{Eigenfunctions $i\,T_2$ of the $l'=m=2$ fundamental $r$-modes 
1266: in the neutron star model NS13T8 at $\hat{\Omega}=0.2$ are given as a function 
1267: of $r/R$. Here, normalization of the eigenfunction is chosen as $i\ T_2(R)=1$.
1268: The solid line and dashed line denote the $r$-modes in the models with the compositional 
1269: stratification and with the thermal stratification, respectively.}  
1270: \end{figure}
1271: 
1272: \begin{figure}
1273: \epsscale{.5}
1274: \plotone{f8.eps}
1275: \caption{Same as Figure 7 but for the model at $\hat{\Omega}=0.02$.}
1276: \end{figure}
1277: 
1278: 
1279: \end{document}
1280: 
1281: 
1282: