0805.3128/ms.tex
1: \documentclass[usenatbib]{mn2e}
2: 
3: \usepackage{epsf}
4: 
5: \newcommand\be{{\bmath e}}
6: \newcommand\bu{{\bmath u}}
7: \newcommand\bcdot{{\bmath\cdot}}
8: \newcommand\bnabla{{\bmath\nabla}}
9: \newcommand\bfT{\mathbf{T}}
10: \newcommand\bfone{\mathbf{1}}
11: \newcommand\real{\mathrm{Re}}
12: \newcommand\imag{\mathrm{Im}}
13: \newcommand\half{{\textstyle\frac{1}{2}}}
14: \newcommand\rmD{\mathrm{D}}
15: \newcommand\rmb{\mathrm{b}}
16: \newcommand\rmd{\mathrm{d}}
17: \newcommand\rme{\mathrm{e}}
18: \newcommand\rmi{\mathrm{i}}
19: \newcommand\rmK{\mathrm{K}}
20: \newcommand\rmm{\mathrm{m}}
21: \newcommand\rmp{\mathrm{p}}
22: \newcommand\rmr{\mathrm{r}}
23: \newcommand\rms{\mathrm{s}}
24: \newcommand\rmT{\mathrm{T}}
25: \newcommand\f{\frac}
26: \newcommand\p{\partial}
27: \newcommand\cst{\mathrm{constant}}
28: \newcommand\he{\mathrm{He}}
29: 
30: \title[Three-dimensional eccentric discs around Be stars]
31: {Three-dimensional eccentric discs around Be stars}
32: 
33: \author[Gordon I. Ogilvie]{Gordon I. Ogilvie\\
34: Department of Applied Mathematics and Theoretical Physics,
35: University of Cambridge, Centre for Mathematical Sciences,\\
36: Wilberforce Road, Cambridge CB3 0WA
37: }
38: 
39: \begin{document}
40: 
41: \maketitle
42: 
43: \label{firstpage}
44:  
45: \begin{abstract}
46:   One-armed oscillation modes in the circumstellar discs of Be~stars
47:   may explain the cyclical variations in their emission lines.  We
48:   show that a three-dimensional effect, involving vertical
49:   motion and neglected in previous treatments, profoundly influences
50:   the dynamics.  Using a secular theory of eccentric discs that
51:   reduces the problem to a second-order differential equation, we show
52:   that confined prograde modes are obtained for all reasonable disc
53:   temperatures and stellar rotation rates.  We confirm these results
54:   using a numerical analysis of the full set of linearized equations
55:   for three-dimensional isothermal discs including viscous
56:   terms that couple the horizontal motions at different altitudes.  In
57:   order to make these modes grow, viscous damping must be overcome by
58:   an excitation mechanism such as viscous overstability.
59: \end{abstract}
60: 
61: \begin{keywords}
62:   accretion, accretion discs --- circumstellar matter --- hydrodynamics ---
63:   stars: emission-line, Be
64: \end{keywords}
65: 
66: \section{Introduction}
67: 
68: Classical Be~stars \citep{2003PASP..115.1153P} are rapidly rotating
69: early-type stars that exhibit Balmer emission lines.  It is widely
70: agreed that these lines, which are generally double-peaked, originate
71: in a relatively thin circumstellar disc that is in approximately
72: Keplerian rotation \citep[and references
73: therein]{2007ASPC..361..230O}.  While the precise mechanism by which
74: the disc is formed remains controversial, it is likely to resemble a
75: viscous decretion disc that is expelled by the action of a torque at
76: its inner boundary \citep{1991MNRAS.250..432L,2003PASP..115.1153P}.
77: 
78: Many Be~stars show cyclical variations in their double-peaked
79: emission lines over years or decades, with the red and blue
80: peaks alternately becoming more prominent \citep[and references
81: therein]{1997A&A...318..548O}.  An explanation of this phenomenon was
82: given by \citet{1991PASJ...43...75O}, who proposed that a
83: low-frequency, one-armed oscillation mode \citep{1983PASJ...35..249K}
84: occurs in the disc.  This is equivalently to saying that the disc
85: becomes eccentric and the slow precession of its elliptical shape
86: gives rise to the cyclical changes in the observed emission lines.
87: 
88: Eccentric discs, in which fluid elements, solid particles or stars
89: follow elliptical orbits of variable eccentricity around a central
90: mass, have further applications in systems as diverse as planetary
91: rings, protoplanetary systems, close binary stars and galactic nuclei.
92: A detailed understanding of the origin of eccentricity and the rates
93: of precession in the circumstellar discs of Be~stars would therefore
94: be of general interest.
95: 
96: \citet{1991PASJ...43...75O} originally considered a disc that orbits
97: in a point-mass potential and obtained a sequence of retrograde modes
98: in which the precession of the disc is in a direction opposite to its
99: rotation.  Retrograde precession is a natural consequence of the
100: pressure forces in the disc, which cause a small departure from
101: Keplerian rotation and allow the eccentricity to propagate in a
102: wavelike manner.  The global modes found by
103: \citet{1991PASJ...43...75O} are weighted towards the outer part of the
104: disc and their periods become extremely long as the outer radius of
105: the disc is increased to realistic values.
106: 
107: \citet{1992A&A...265L..45P} and \citet{1993A&A...276..409S} considered
108: the effect of the quadrupole gravitational potential associated with
109: the rotational deformation of the star, which tends to cause a
110: prograde precession of elliptical orbits.  They showed that, when the
111: quadrupole effect is taken into account, prograde modes can be
112: obtained that are naturally confined in the inner part of the disc and
113: are insensitive to the outer boundary condition.  Subsequent
114: observations confirmed that the precession is indeed prograde
115: \citep{1994A&A...288..558T}.
116: 
117: However, \citet{1997A&A...318..548O} found that confined prograde
118: modes can be obtained only when the disc is sufficiently cool, so that
119: the quadrupole effect dominates over the tendency of pressure to
120: produce retrograde precession and extended modes.  He concluded that a
121: different mechanism is required in the hotter discs of early-type
122: Be~stars such as the prototype $\gamma$~Cas, and proposed
123: that radiative line forces could explain the required prograde
124: precession.  Unfortunately the modelling of radiative forces is
125: subject to considerable uncertainty.  \citet{2006A&A...447..277F}
126: found that the resulting model has little predictive power owing to
127: the sensitivity of the results to the parameters.
128: 
129: More recently, \citet{2006A&A...456.1097P} investigated an alternative
130: way to obtain prograde modes in hotter discs.  This involves replacing
131: the rigid inner boundary condition at the stellar surface with a free
132: boundary condition, on the basis that a gap is formed between the star
133: and the disc.  Such a gap is not generally expected in the scenario of
134: the viscous decretion disc but might be possible in alternative
135: models.
136: 
137: All of the treatments described so far are based on two-dimensional
138: models of the disc that neglect aspects of its vertical structure and
139: motion.  At first sight, such an approach seems reasonable for
140: studying eccentric modes in thin discs, where the motion might be
141: assumed to be purely horizontal and independent of height.  However,
142: in presenting a three-dimensional, non-linear theory of eccentric
143: discs, we have previously argued that three-dimensional effects are of
144: considerable importance \citep{2001MNRAS.325..231O}.  In particular,
145: the variation of the vertical gravitational acceleration around an
146: elliptical orbit excites an oscillatory vertical motion in an
147: eccentric disc that should not be neglected.
148: 
149: In this paper we show that, in fact, three-dimensional effects are
150: essential to understanding the precession of eccentric discs around
151: Be~stars.  They allow confined prograde modes to be obtained even when
152: the stellar quadrupole moment is negligible and when the inner
153: boundary is rigid.  This property allows us to give a unified
154: description of eccentric discs of Be~stars of all stellar types
155: without introducing uncertain radiative forces or modifying the inner
156: boundary condition.  While we do not deny that in some cases radiative
157: forces might be important, or that the inner boundary condition might
158: differ from a rigid one, we show that these innovations are
159: unnecessary.
160: 
161: Most previous analyses have not discussed the processes that
162: could cause eccentric modes to grow or decay, and which are therefore
163: relevant to explaining the occurrence of eccentric discs around
164: Be~stars.  Viscous overstability
165: \citep{1978MNRAS.185..629K,2006MNRAS.372.1829L} provides a possible
166: explanation; \citet{2001A&A...369..117N} have applied this idea to
167: Be~stars and estimated the associated growth rate.  We defer to a
168: future investigation a detailed analysis of the effects of viscous or
169: turbulent stresses.  A non-linear treatment, which could address the
170: saturation of the growth mechanism and attempt to predict the observed
171: amplitudes and precession rates of the eccentric modes, also remains
172: to be undertaken.
173: 
174: The structure of this paper is as follows.  In Section~\ref{s:basic}
175: we describe the basic state of the disc.  We then discuss an
176: approximate, secular theory of three-dimensional eccentric discs in
177: Section~\ref{s:secular}, comparing it with the two-dimensional
178: theories used by other authors.  In Section~\ref{s:full} we solve the
179: full linearized equations accurately using a spectral method,
180: including some effects of viscosity, and compare the results with
181: those of the secular theory.  Conclusions are given in
182: Section~\ref{s:conc}.
183: 
184: \section{Basic state of the disc}
185: \label{s:basic}
186: 
187: We consider a basic state consisting of a steady, axisymmetric disc
188: around a rotating star.  Adopting cylindrical polar coordinates
189: $(r,\phi,z)$, we write the gravitational potential of the star as
190: \begin{equation}
191:   \Phi=-GM(r^2+z^2)^{-1/2}\left[1+
192:   \f{Q}{3}\f{(r^2-2z^2)R^2}{(r^2+z^2)^2}\right],
193: \end{equation}
194: where $M$ and $R$ are its mass and (equatorial) radius.  This
195: potential consists of monopole and quadrupole components, the latter
196: being of dimensionless strength $Q$ and arising from the rotational
197: deformation of the star.  (The parameter $Q$ is related to the
198: gravitational moment $J_2$ used in planetary science by $Q=3J_2/2$.)
199: For a star with uniform angular velocity $\Omega_*$, we have
200: $Q=k_2\Omega_*^2R^3/GM$, where $k_2$ is the apsidal motion constant.
201: We neglect higher-order multipole components as well as the
202: self-gravitation of the disc.
203: 
204: In common with most other treatments, we assume that the stellar
205: radiation maintains the disc at a constant temperature $T_\rmd$,
206: slightly less than the effective temperature $T_\rme$ of the star.
207: The isothermal sound speed $c_\rms=(\mathcal{R}T_\rmd/\mu)^{1/2}$ is
208: therefore constant, both in the equilibrium state and for the
209: perturbations.  We neglect any viscous or turbulent stresses and any
210: resulting meridional motions in the disc.
211: 
212: The basic state then has velocity $\bu=r\Omega(r)\,\be_\phi$.  The
213: angular velocity $\Omega$ is independent of $z$ because the basic
214: state is isothermal and therefore barotropic.  The balance of forces
215: requires $r\Omega^2\,\be_r=\bnabla(\Phi+h)$, where $h=c_\rms^2\ln\rho$
216: is the (pseudo-)enthalpy of an isothermal gas.
217: 
218: The potential and enthalpy in the midplane are
219: $\Phi_\rmm(r)=\Phi(r,0)=-(GM/r)[1+(Q/3)(R/r)^2]$ and
220: $h_\rmm(r)=h(r,0)$, and the radial force balance implies
221: \begin{equation}
222:   \Omega^2=\f{GM}{r^3}\left[1+Q\left(\f{R}{r}\right)^2\right]+\f{1}{r}\p_rh_\rmm.
223: \label{omsq}
224: \end{equation}
225: The associated epicyclic frequency $\kappa(r)$ is given by
226: \begin{eqnarray}
227:   \kappa^2&=&\f{1}{r^3}\p_r(r^4\Omega^2)\nonumber\\
228:   &=&\f{GM}{r^3}\left[1-Q\left(\f{R}{r}\right)^2\right]+\f{1}{r^3}\p_r(r^3\p_rh_\rmm).
229: \label{kasq}
230: \end{eqnarray}
231: We also refer to the Keplerian angular velocity $\Omega_\rmK(r)$
232: defined by
233: \begin{equation}
234:   \Omega_\rmK^2=\f{GM}{r^3}.
235: \end{equation}
236: 
237: For a thin disc we may expand the potential about the midplane to
238: obtain $\Phi\approx\Phi_\rmm+\half\Omega_z^2z^2$, where the vertical
239: frequency $\Omega_z(r)$ is given by
240: \begin{equation}
241:   \Omega_z^2=\f{GM}{r^3}\left[1+3Q\left(\f{R}{r}\right)^2\right].
242: \end{equation}
243: The vertical force balance then implies $h=h_\rmm-\half\Omega_z^2z^2$
244: and so $\rho=\rho_\rmm\exp[-(z^2/2H^2)]$, where $\rho_\rmm(r)$ is the
245: midplane density and $H(r)=c_\rms/\Omega_z$ is the scaleheight.  The
246: surface density is $\Sigma(r)=(2\pi)^{1/2}\rho_\rmm H$.
247: 
248: Since the basic state is steady and axisymmetric, wave modes may be
249: considered in which the dependence on azimuth and time is of the form
250: $\exp(\rmi m\phi-\rmi\omega t)$, where $m$ is the azimuthal wavenumber
251: and $\omega$ is the wave frequency.  In the case $m=1$ that is of
252: interest here, $\omega$ is also the angular pattern speed of the mode
253: in an inertial frame of reference.  This can be identified with the
254: precession rate of the eccentric disc, which is positive if the
255: precession is prograde (i.e.\ in the same direction as the rotation of
256: the disc).
257: 
258: \section{Secular treatment}
259: \label{s:secular}
260: 
261: \subsection{Comparison of two- and three-dimensional theories}
262: 
263: An approximate theory can be derived in which the disc is assumed to
264: be nearly Keplerian and the frequency $\omega$ of the wave is much
265: less than the orbital frequency $\Omega$ of the disc.  This type of
266: approximation, which is related to the secular theory of celestial
267: mechanics, has been discussed in the two-dimensional linear case by
268: \citet{2001AJ....121.1776T}, \citet{2002A&A...388..615P} and
269: \citet{2006MNRAS.368.1123G}, and in the three-dimensional non-linear
270: case by \citet{2001MNRAS.325..231O}.  The derivation of this theory in
271: the case of a three-dimensional isothermal disc around a Be~star is
272: presented in Appendix~\ref{s:appendix}, which draws on the analysis of
273: Section~\ref{s:full} below.  The two-dimensional approximation is also
274: discussed there.
275: 
276: In the secular theory the radial velocity $u(r)$ satisfies the
277: equation
278: \begin{equation}
279:   (\omega-f)u=\f{1}{2r^2}\p_r\left[\f{c_\rms^2r^3}{\Sigma}\p_r\left(\f{\Sigma u}{r\Omega}\right)\right],
280: \label{secular}
281: \end{equation}
282: where $f(r)$ is given by
283: \begin{equation}
284:   f=\f{\Omega^2-\kappa^2}{2\Omega}
285: \label{f2d}
286: \end{equation}
287: in a two-dimensional disc, but
288: \begin{equation}
289:   f=\f{\Omega^2-\kappa^2}{2\Omega}+\f{9c_\rms^2}{4r^2\Omega}
290: \label{f3d}
291: \end{equation}
292: in a three-dimensional disc.  With the dependence
293: $\rme^{\rmi\phi-\rmi\omega t}$ assumed above, the radial velocity is
294: related to the complex eccentricity $E(r)=e\,\rme^{\rmi\varpi}$, where
295: $e(r)$ is the eccentricity and $\varpi(r)$ is the longitude of
296: periastron measured in a frame of reference that rotates with the
297: angular pattern speed $\omega$ of the mode, by $u^*=\rmi r\Omega E$
298: \citep{2001MNRAS.325..231O}.  It can be seen from
299: equation~(\ref{secular}) that $f$ is a local contribution to the
300: global precession rate $\omega$ of the eccentric mode.  Indeed, an
301: integral expression for the mode frequency in the case of rigid
302: boundary conditions\footnote{Although there is no physical
303: justification for a rigid outer boundary condition, a confined mode
304: decays sufficiently fast with increasing $r$ that the same integral
305: relation is recovered in the limit of large $r_\mathrm{out}$.  Such
306: modes are in fact insensitive to the outer boundary condition.} ($u=0$
307: at $r=r_\mathrm{in}$ and $r=r_\mathrm{out}$) is
308: \begin{eqnarray}
309:   \lefteqn{\omega\int_{r_\mathrm{in}}^{r_\mathrm{out}}\f{\Sigma r|u|^2}{\Omega}\,\rmd r=\int_{r_\mathrm{in}}^{r_\mathrm{out}}f\,\f{\Sigma r|u|^2}{\Omega}\,\rmd r}&\nonumber\\
310:   &&-\int_{r_\mathrm{in}}^{r_\mathrm{out}}\f{c_\rms^2r^3}{2\Sigma}\left|\p_r\left(\f{\Sigma u}{r\Omega}\right)\right|^2\,\rmd r,
311: \label{vp}
312: \end{eqnarray}
313: which follows from equation~(\ref{secular}) after multiplication by
314: $\Sigma ru^*/\Omega$ and an integration by parts.  The first term on
315: the right-hand side shows the contribution to the precession rate from
316: $f$ in the form of an integral weighted by the structure of the mode,
317: while the second term shows a retrograde contribution associated with
318: pressure.  (Note that the pressure also contributes indirectly through
319: its effect on $f$.)
320: 
321: In a two-dimensional disc $f$ is given by equation~(\ref{f2d}) and
322: corresponds to the expected expression for the local apsidal
323: precession rate.  Note that, in this case, $f\approx\Omega-\kappa$ if,
324: as assumed, the disc is nearly Keplerian
325: ($|\Omega^2-\kappa^2|\ll\Omega^2$).  There is, in fact, more than one
326: version of the two-dimensional theory.  \citet{1993A&A...276..409S}
327: work throughout with vertically averaged equations and find
328: \begin{equation}
329:   f=Q\f{GMR^2}{r^5\Omega}-\f{c_\rms^2}{2r^2\Omega}\p_r(r^2\p_r\ln\Sigma).
330: \label{2ds}
331: \end{equation}
332: \citet{1997A&A...318..548O} calculates $\Omega$ and $\kappa$ using a
333: three-dimensional equilibrium disc but then applies vertically
334: averaged equations for the perturbations.  He therefore uses
335: \begin{equation}
336:   f=Q\f{GMR^2}{r^5\Omega}-\f{c_\rms^2}{2r^2\Omega}\p_r(r^2\p_r\ln\rho_\rmm),
337: \label{2do}
338: \end{equation}
339: which agrees with equations~(\ref{omsq}) and~(\ref{kasq}) above.
340: 
341: In a three-dimensional disc, however, equation~(\ref{f3d}) gives the
342: relevant expression for $f$ as
343: \begin{equation}
344:   f=Q\f{GMR^2}{r^5\Omega}-\f{c_\rms^2}{2r^2\Omega}\p_r(r^2\p_r\ln\rho_\rmm)+
345:   \f{9c_\rms^2}{4r^2\Omega}.
346: \label{3d}
347: \end{equation}
348: The last term represents an additional local contribution to the
349: prograde precession of the disc that arises from the three-dimensional
350: dynamics including the vertical motion; it is discussed further in
351: Section~\ref{s:interp} below.  For the parameters relevant to
352: Be~stars, this term is never negligible (it corresponds to a
353: precession period of order $1\,\mathrm{yr}$ at the stellar surface)
354: and declines much more slowly with radius than the quadrupole term.
355: 
356: We refer to the three theories described above as `2DS'
357: \citep[equation~\ref{2ds}
358: above;][]{1993A&A...276..409S,2006A&A...456.1097P}, `2DO'
359: \citep[equation~\ref{2do}
360: above;][]{1991PASJ...43...75O,1997A&A...318..548O} and `3D'
361: (equation~\ref{3d}).  Note, however, that the secular approximations
362: were usually not employed in the cited papers.  The 3D~secular theory
363: is based on similar assumptions and approximations to the non-linear
364: analysis of \citet{2001MNRAS.325..231O}.  In particular, some viscous
365: or other stress is required to couple different layers in the disc so
366: that they tend to adopt the same eccentricity.  In the absence of such
367: stresses the eccentricity may vary significantly with $z$ and the
368: results can differ.  This complication is discussed in
369: Section~\ref{s:full}, where the full linearized equations are solved.
370: 
371: \subsection{Application to power-law discs}
372: \label{s:powerlaw}
373: 
374: It is consistent with the spirit of the secular approximation to
375: neglect the differences between $\Omega$, $\kappa$, $\Omega_z$ and
376: $\Omega_\rmK$ except where essential (i.e.\ in the quantity
377: $\Omega^2-\kappa^2$).  Accordingly, we may take $\Omega\propto
378: r^{-3/2}$ and $H\propto r^{3/2}$.  For a surface density profile
379: $\Sigma\propto r^{-\sigma}$ the density in the midplane varies as
380: $\rho_\rmm\propto r^{-\sigma-3/2}$.  We introduce the dimensionless
381: radial coordinate
382: \begin{equation}
383:   x=\f{r}{R}
384: \end{equation} 
385: and the small parameter
386: \begin{equation}
387:   \epsilon=c_\rms\left(\f{R}{GM}\right)^{1/2},
388: \end{equation}
389: which is a measure of the angular semithickness $H/r$ of the disc at the
390: stellar surface.  The three theories then give
391: \begin{equation}
392:   \mathrm{2DS:\phantom{O3D}}\f{f}{\Omega}=Qx^{-2}+\half\sigma\epsilon^2x,
393: \end{equation}
394: \begin{equation}
395:   \mathrm{2DO:\phantom{S3D}}\f{f}{\Omega}=Qx^{-2}+\half(\sigma+{\textstyle\f{3}{2}})\epsilon^2x,
396: \end{equation}
397: \begin{equation}
398:   \mathrm{3D:\phantom{S2DO}}\f{f}{\Omega}=Qx^{-2}+\half(\sigma+6)\epsilon^2x.
399: \end{equation}
400: We therefore consider the generic form
401: \begin{equation}
402:   \f{f}{\Omega}=Qx^{-2}+s\epsilon^2x,
403: \end{equation}
404: where $s$ is a constant.  It is convenient to work with dimensionless,
405: rescaled values of the mode frequency and quadrupole strength,
406: $\tilde\omega$ and $\tilde Q$, defined by
407: \begin{equation}
408:   \omega=\epsilon^2\left(\f{GM}{R^3}\right)^{1/2}\tilde\omega,
409: \end{equation}
410: \begin{equation}
411:   Q=\epsilon^2\tilde Q.
412: \end{equation}
413: Equation~(\ref{secular}) then becomes
414: \begin{eqnarray}
415:   \lefteqn{\left[\tilde\omega-(\tilde Qx^{-7/2}+sx^{-1/2})\right]u}&\nonumber\\
416:   &&=\f{1}{2x^2}\p_x\left[x^{\sigma+3}\p_x(x^{-\sigma+1/2}u)\right],
417: \label{secular2}
418: \end{eqnarray}
419: which is an eigenvalue problem of Sturm--Liouville form, when
420: considered together with appropriate boundary conditions.  Note that
421: the parameter $\epsilon$ drops out of the equation under the above
422: rescalings.  For prograde modes ($\tilde\omega>0$), one solution of
423: this equation decays exponentially as $x\to\infty$, while the other
424: solution grows exponentially.  We therefore consider the equation in
425: the domain $1<x<\infty$, requiring solutions to satisfy a rigid
426: boundary condition $u=0$ at the stellar surface $x=1$ and to decay as
427: $x\to\infty$.
428: 
429: Reasonable values of $\tilde Q$ for Be~stars can be estimated
430: following \citet{2006A&A...447..277F}.  Their Table~4 is based on
431: stellar models of intermediate main-sequence age.  There is
432: considerable uncertainty in the applicability of these models because
433: of the rapid rotation of Be~stars.  Nevertheless, using these data, we
434: estimate that $\tilde Q$ ranges from approximately~$8$ at the lower
435: end ($M=2.51\,M_\odot$) to approximately~12 at the upper end
436: ($M=15.85\,M_\odot$).  These values assume, somewhat arbitrarily, a
437: disc temperature of $T_\rmd=(2/3)T_\rme$ and a stellar rotation rate
438: that is $95\%$ of the critical value.  Similarly, for the stellar
439: models of $\gamma$~Cas and $59$~Cyg described in Tables~2 and~3 of
440: \citet{2006A&A...447..277F}, we estimate $\tilde Q\approx9$ and~$13$,
441: respectively.  It is clear that significantly larger values of $\tilde
442: Q$ cannot be obtained by further increasing the stellar rotation rate,
443: and in fact smaller values may be more appropriate.  The values of
444: $\epsilon$ for all these models are in the range $0.021$--$0.026$, again
445: assuming that $T_\rmd=(2/3)T_\rme$.
446: 
447: \subsection{Solutions in the absence of a quadrupole}
448: 
449: Consider first the case without a quadrupole term, $\tilde Q=0$.  The
450: solution that decays as $x\to\infty$ is then
451: \begin{equation}
452:   u=x^{(\sigma-3)/2}K_\nu(z),
453: \label{knu}
454: \end{equation}
455: with
456: \begin{equation}
457:   \nu=2[(\sigma+2)^2-8s]^{1/2},\qquad z=4(2\tilde\omega)^{1/2}x^{1/4},
458: \end{equation}
459: where $K_\nu$ is the modified Bessel function of the second kind of
460: order $\nu$ \citep{1965hmfw.book.....A}.  Note that $z$ is real and
461: positive for the prograde modes of interest.  This solution has no
462: zeros in $z>0$ if $\nu$ is real, but does so if $\nu$ is imaginary.
463: To match a rigid boundary condition at $x=1$ we therefore require
464: $\nu^2<0$.  The 2DS~theory gives $8s=4\sigma$, so
465: $\nu^2=4(\sigma^2+4)$ and modes are never confined.  The 2DO~theory
466: gives $8s=4\sigma+6$, so $\nu^2=4(\sigma^2-2)$ and modes may be
467: confined for $\sigma^2<2$.  The 3D~theory gives $8s=4\sigma+24$, so
468: $\nu^2=4(\sigma^2-20)$ and modes may be confined for $\sigma^2<20$.
469: 
470: This analysis is somewhat misleading because the modes may not be
471: adequately confined to be applicable to the discs of Be~stars.  In
472: Fig.~\ref{f:knu} we plot $K_\nu(z)$ versus $x$ for $\nu=8\rmi$ and
473: $\nu=\sqrt{8}\rmi$, choosing $\tilde\omega$ such that $K_\nu(z)=0$ at
474: $x=1$.  Only the former case provides an adequately confined mode.
475: This shows that the 2DO~theory cannot in fact produce confined
476: prograde modes in the absence of a quadrupole, even for the most
477: optimistic choice of surface density profile, because $|\nu|$ is never
478: large enough.  However, sufficiently large imaginary values of $\nu$
479: are obtained in the 3D~theory.
480: 
481: \begin{figure}
482: \centerline{\epsfysize8cm\epsfbox{fig1.eps}}
483: \caption{Confinement of prograde modes in the absence of a quadrupole.
484: The modified Bessel function $K_\nu(z)$ (multiplied by a constant for
485: convenient normalization) is plotted against $x=r/R$ for the cases
486: $\nu=8\rmi$ (solid line) and $\nu=\sqrt{8}\rmi$ (dashed line).
487: According to equation~(\ref{knu}), the modified Bessel function, which
488: decays exponentially as $x\to\infty$ if $\nu$ is imaginary, describes
489: the confinement of the eigenfunction.  The solid line is relevant to
490: the 3D~theory for a surface density index $\sigma=2$, and shows a good
491: confinement of the mode.  The dashed line is relevant to the
492: 2DO~theory in the (optimal) case $\sigma=0$, but here the confinement
493: is very poor.}
494: \label{f:knu}
495: \end{figure}
496: 
497: When $\nu$ is imaginary, confined solutions are obtained, in
498: principle, when $\tilde\omega=z_n^2/32$ where $z_{n,\nu}$ is the $n$th
499: zero of $K_\nu(z)$ in $z>0$.  For example, consider the 3D~theory with
500: $\sigma=2$, which is the value expected for a steady decretion disc
501: with constant alpha viscosity parameter, far inside its outer radius.
502: Then $s=4$ and $\nu=8\rmi$, and zeros occur at $z=4.802$, $3.067$,
503: $2.029$, $\dots$\ Therefore, in principle, $\tilde\omega=0.7207$,
504: $0.2940$, $0.1286$, $\dots$ are the scaled frequencies of confined
505: modes.  However, only the first of these, corresponding to an
506: eigenfunction with no nodes (Fig.~\ref{f:knu}, solid line), gives rise
507: to a mode that is adequately confined in a disc of modest radial
508: extent.
509: 
510: \subsection{Critical quadrupole strength}
511: \label{s:qcrit}
512: 
513: The integral expression for the dimensionless frequency eigenvalue of
514: a mode that satisfies a rigid boundary condition at the stellar
515: surface and decays as $r\to\infty$ is (cf.\ equation~\ref{vp})
516: \begin{eqnarray}
517:   \lefteqn{\tilde\omega\int_1^\infty x^{-\sigma+5/2}u^2\,\rmd x=
518:   \int_1^\infty(\tilde Qx^{-\sigma-1}+sx^{-\sigma+2})u^2\,\rmd x}&\nonumber\\
519:   &&-\f{1}{2}\int_1^\infty x^{\sigma+3}\left[\p_x(x^{-\sigma+1/2}u)\right]^2\,\rmd x,
520: \end{eqnarray}
521: where we take $u$ to be real.  Following \citet{2006A&A...456.1097P},
522: we note that this expression has the usual variational property
523: associated with self-adjoint eigenvalue problems.  In this case, a
524: confined prograde mode exists if and only the right-hand side of this
525: equation can be made positive by a trial function $u(x)$ satisfying
526: the appropriate boundary conditions.  This is clearly possible if
527: either $\tilde Q$ or $s$ is large enough.  What is the minimum value
528: of $\tilde Q$ (for a given value of $s$) such that the right-hand side
529: can just be made to vanish for a non-trivial $u$?  This happens when
530: \begin{eqnarray}
531:   \lefteqn{\tilde Q\int_1^\infty x^{-\sigma-1}u^2\,\rmd x=
532:   \f{1}{2}\int_1^\infty
533:   x^{\sigma+3}\left[\p_x(x^{-\sigma+1/2}u)\right]^2\,\rmd
534:   x}&\nonumber\\ &&-\int_1^\infty sx^{-\sigma+2}u^2\,\rmd x.
535: \end{eqnarray}
536: The minimum value of $\tilde Q$ for which this equation can be
537: satisfied is given by the corresponding Euler--Lagrange equation,
538: which is just equation~(\ref{secular2}) with $\tilde\omega$ set to
539: zero.  The solutions are
540: \begin{equation}
541:   u=x^{(\sigma-3)/2}J_{\pm\mu}(w),
542: \end{equation}
543: with
544: \begin{equation}
545:   \mu=\f{\nu}{6}=\f{1}{3}[(\sigma+2)^2-8s]^{1/2},\qquad w=\f{\tilde Q^{1/2}}{3}\left(\f{x}{2}\right)^{-3/2}.
546: \end{equation}
547: If $\mu^2<0$ we have seen that confined modes can be found, in
548: principle, even if $\tilde Q=0$.  Consider then the case $\mu^2>0$.
549: The solution that decays\footnote{Since $J_{\mu}(w)\propto w^\mu$ for
550: small $w$, $u\propto x^{[\sigma-3(1+\mu)]/2}$ for large $x$.  This
551: solution satisfies the condition that
552: $x^{7/2}u\p_x(x^{-\sigma+1/2}u)\to0$ as $x\to\infty$, which is
553: required to carry out the integration by parts and derive the
554: variational principle and Euler--Lagrange equation, while $J_{-\mu}$
555: does not.} as $x\to\infty$ is $J_{\mu}(w)$.  The critical condition
556: for a rigid inner boundary condition is therefore $\tilde
557: Q=9w_\mu^2/8$, where $w_\mu$ is the first zero of $J_\mu(w)$ in $w>0$.
558: (This is an increasing function of $\mu$ and therefore a decreasing
559: function of $s$.)
560: 
561: For example, if $\sigma=2$, we require $\tilde Q>10.8$ for confinement
562: in the 2DO~theory, or $\tilde Q>15.9$ in the 2DS~theory.  These values
563: are not very sensitive to $\sigma$; \citet{2006A&A...456.1097P} quote
564: $\tilde Q>17.3$ for $\sigma=5/2$ and a rigid inner boundary condition,
565: which agrees with this analysis.
566: 
567: \subsection{Schr\"odinger analogy}
568: 
569: A useful description of confined modes uses an analogy with bound
570: states in quantum mechanics.  If $y=x^{1/4}$ and
571: $u=x^{\sigma/2-13/8}\psi(y)$, then $\psi(y)$ satisfies the
572: Schr\"odinger equation
573: \begin{equation}
574:   -\f{\rmd^2\psi}{\rmd y^2}+[V(y)-E]\psi=0
575: \end{equation}
576: with an effective potential
577: \begin{equation}
578:   V=\f{1}{4}(16\sigma^2+64\sigma+63-128s)y^{-2}-32\tilde Qy^{-14}
579: \label{v}
580: \end{equation}
581: and an effective energy eigenvalue $E=-32\tilde\omega$.  The
582: coefficient of $1/4y^2$ in $V$ is $(16\sigma^2+63)$ for the
583: 2DS~theory, $(16\sigma^2-33)$ for 2DO and $(16\sigma^2-321)$ for 3D.
584: It is therefore likely to be negative only in the 3D~theory.  Note
585: that the disc occupies the region $y>1$.  A bound state of negative
586: energy, equivalent to a confined prograde mode, can be obtained if
587: there is a sufficiently deep and wide potential well.  The $\tilde Q$
588: term tends to create a deep well close to the stellar surface.  In the
589: 2D~theories it competes with the $y^{-2}$ term, which contributes a
590: repulsive potential.  In the 3D~theory, however, the $y^{-2}$ term
591: creates a much wider well, so allowing broader modes with slower
592: precession rates, even if $\tilde Q=0$.  These potentials are plotted
593: in Fig.~\ref{f:v} for the case $\sigma=2$ and $\tilde Q=10$.  For
594: these parameters the 3D~potential is deep and wide enough to support a
595: bound state in the case of a rigid inner boundary condition,
596: while the others are not.
597: 
598: \begin{figure}
599: \centerline{\epsfysize8cm\epsfbox{fig2.eps}}
600: \caption{Effective potentials $V(y)$ (equation~\ref{v}) describing the
601: confinement of prograde modes by analogy with bound states in quantum
602: mechanics.  Potentials are shown for the three theories for the case
603: $\sigma=2$ and $\tilde Q=10$.  The bottom of the potential well at
604: $y=1$ (i.e.\ at the stellar surface) in each case is considerably
605: deeper than can be shown on the graph.  However, only in the 3D~theory
606: is the well deep and wide enough in this case to support a bound state
607: in the case of a rigid inner boundary condition.}
608: \label{f:v}
609: \end{figure}
610: 
611: \subsection{Precession rates and mode shapes}
612: 
613: \begin{figure*}
614: \centerline{\epsfysize8.5cm\epsfbox{fig3a.eps}\qquad\epsfysize8.5cm\epsfbox{fig3b.eps}}
615: \caption{Dimensionless precession rates of confined prograde modes in
616: the three theories, for the case $\sigma=2$.  The left and
617: right panels are for rigid and free inner boundary conditions,
618: respectively.  These results are insensitive to the outer boundary
619: condition provided that the disc is reasonably large.  In the
620: 2D~theories, confined prograde modes are found only for sufficiently
621: large values of $\tilde Q$.  The dotted line indicates a
622: second confined mode, with one radial node in its eigenfunction, found
623: in the 3D~theory; this and higher-order modes are less well confined
624: and therefore more sensitive to the outer boundary condition.}
625: \label{f:ot}
626: \end{figure*}
627: 
628: We now compute the eigenvalues $\tilde\omega$ of
629: equation~(\ref{secular2}) by a shooting method, adopting a rigid inner
630: boundary condition at the stellar surface and seeking prograde
631: confined modes that decay as $x\to\infty$.  (In practice this is done
632: by imposing a rigid outer boundary condition and verifying that the
633: eigenvalue is completely insensitive to the value of the outer radius
634: provided it is sufficiently large.)  The results are shown in the left
635: panel of Fig.~\ref{f:ot}.  Here we confirm that confined prograde
636: modes exist in the 2D~theories only for sufficiently large values of
637: $\tilde Q$, as described in Section~\ref{s:qcrit}.  In contrast, the
638: 3D~theory allows such modes to be obtained for any value of $\tilde
639: Q$.  Furthermore, the precession rates are much larger in the
640: 3D~theory.  Since the realistic values of $\tilde Q$ are probably in
641: the vicinity of $10$ or smaller, it is clear that the
642: three-dimensional effects are of essential importance in the
643: case of a rigid inner boundary condition.
644: 
645: In the right panel of Fig.~\ref{f:ot} we show comparable
646: results for the free inner boundary condition considered by
647: \citet{2006A&A...456.1097P}, meaning that the Lagrangian pressure
648: perturbation is zero at $r=R$.  In the secular theory this is
649: equivalent to $\p_rE=0$, or $u+2r\p_r u=0$.  As described by
650: \citet{2006A&A...456.1097P}, a free inner boundary condition allows
651: prograde modes to be found for smaller values of $\tilde Q$ in the
652: 2DS~theory.  It is still true, however, that the three-dimensional
653: effects have an important effect on the results.  Since the
654: eigenfunctions obtained with a free inner boundary condition are
655: generally peaked at the inner radius, there may also be a conflict
656: between the nonlinear development of such an eccentric mode and the
657: existence of a stellar surface, unless there is a wide gap between the
658: star and the disc.
659: 
660: To convert these eigenvalues into physical units, we again make use of
661: the stellar models in \citet{2006A&A...447..277F}.  We then find that
662: the precession period is
663: \begin{equation}
664:   P=\f{C_P}{\tilde\omega}\,\mathrm{yr},
665: \end{equation}
666: where $C_P$ ranges from approximately $1.6$ at the lower end to
667: approximately $2.3$ at the upper end (or $2.4$ for $\gamma$~Cas and
668: $1.8$ for $59$~Cyg).  This conversion factor is independent of
669: assumptions regarding the stellar rotation rate, except inasmuch as
670: the rotation affects the equatorial radius, but $C_P$ is inversely
671: proportional to the assumed disc temperature.
672: 
673: The shapes of the confined modes in the 3D~theory are illustrated in
674: Fig.~\ref{f:ef}.  For larger values of $\tilde Q$, the mode is
675: increasingly confined in the inner part of the disc.  Although
676: higher-order modes may exist, one of which is referred to in
677: Fig.~\ref{f:ot}, we focus here on the fundamental confined mode with
678: the simplest radial structure.
679: 
680: \begin{figure}
681: \centerline{\epsfysize8cm\epsfbox{fig4.eps}}
682: \caption{Eigenfunctions of confined prograde modes in the 3D~secular
683: theory with a rigid inner boundary condition.  The eccentricity
684: (proportional to $u/r\Omega$, and with arbitrary normalization) is
685: plotted versus radius for the case $\sigma=2$, and for relative
686: quadrupole strengths $\tilde Q=0$ (solid line), $5$ (dotted line),
687: $10$ (dashed line) and $20$ (dot-dashed line).  The first case
688: corresponds to the solid line in Fig.~\ref{f:knu}.}
689: \label{f:ef}
690: \end{figure}
691: 
692: Since $\tilde\omega$ is close to~$1$ in the 3D~theory for reasonable
693: values of~$\tilde Q$ and with a rigid inner boundary
694: condition, precession rates in the vicinity of $1$--$3\,\mathrm{yr}$
695: are obtained.  Observed cycle times, which can vary significantly even
696: for the same star, are more usually in the range
697: $5$--$10\,\mathrm{yr}$ \citep{1997A&A...318..548O}.  This discrepancy
698: might occur because the discs have different temperature or density
699: profiles from those assumed here, either of which would affect the
700: precession period.  Non-linearity of the eccentric mode may also
701: increase the precession period by altering the distribution of
702: eccentricity so that it is less peaked in the inner part of the disc.
703: Furthermore, non-isothermal effects or radiative forces might be
704: important.  In some cases, including $\gamma$~Cas and $59$~Cyg, the
705: presence of a relative close binary companion may also affect the
706: dynamics of the eccentric mode, although it would seem most
707: likely to contribute to the prograde precession and therefore to
708: decrease the precession period.
709: 
710: \subsection{Physical interpretation of the three-dimensional dynamics}
711: \label{s:interp}
712: 
713: What is the origin of the additional prograde precession that occurs
714: in a three-dimensional disc?  As described in
715: Appendix~\ref{s:appendix} and Section~\ref{s:full} below, two
716: different types of motion are involved in an eccentric disc.  One
717: (corresponding to the $n=0$ mode in later sections) consists of
718: horizontal velocities and enthalpy perturbations that are independent
719: of $z$; this describes the eccentric orbital motion of the gas.  The
720: other (corresponding to $n=2$) involves a vertical velocity
721: proportional to $z$ and an enthalpy perturbation proportional to
722: $z^2-H^2$; this is a vertical `breathing' mode of the disc.  Coupling
723: of these motions occurs because of the variation of the vertical
724: gravitational acceleration (or, equivalently, the vertical frequency
725: $\Omega_z$, or the scaleheight $H$) with radius.  Vertical hydrostatic
726: equilibrium cannot be maintained in an eccentric disc because a fluid
727: element in an elliptical orbit experiences a vertical gravitational
728: acceleration that oscillates with the orbital frequency.  The
729: breathing mode is excited and the associated enthalpy perturbation
730: affects the horizontal dynamics, contributing to the precession of the
731: eccentric mode.  This contribution is found to be always
732: positive.  In the more general situation of a non-isothermal disc
733: undergoing adiabatic perturbations, the three-dimensional contribution
734: to $f$ (in which we compare a 3D-type theory with a 2DS-type theory)
735: is
736: \begin{equation}
737:   \f{3(\gamma+1)}{2\gamma}\f{P}{\Sigma r^2\Omega},
738: \end{equation}
739: where $\gamma$ is the adiabatic index and $P$ is the
740: vertically integrated pressure (Ogilvie \& Goodchild, in preparation),
741: but the derivation of this expression is beyond the scope of the
742: present paper.
743: 
744: \section{Full treatment}
745: \label{s:full}
746: 
747: \subsection{Inviscid dynamics}
748: 
749: The dynamical equations governing an isothermal, inviscid disc are
750: \begin{equation}
751:   (\p_t+\bu\bcdot\bnabla)\bu=-\bnabla(\Phi+h),
752: \label{motion}
753: \end{equation}
754: \begin{equation}
755:   (\p_t+\bu\bcdot\bnabla)h=-c_\rms^2\bnabla\bcdot\bu.
756: \end{equation}
757: When these are linearized about the basic state described in
758: Section~\ref{s:basic}, we obtain
759: \begin{equation}
760:   -\rmi\hat\omega u_r'-2\Omega u_\phi'=-\p_r h',
761: \end{equation}
762: \begin{equation}
763:   -\rmi\hat\omega u_\phi'+\f{\kappa^2}{2\Omega}u_r'=-\f{\rmi mh'}{r},
764: \end{equation}
765: \begin{equation}
766:   -\rmi\hat\omega u_z'=-\p_z h',
767: \end{equation}
768: \begin{eqnarray}
769:   \lefteqn{-\rmi\hat\omega h'+u_r'\p_r h+u_z'\p_z h}&\nonumber\\
770:   &&=-c_\rms^2\left[\f{1}{r}\p_r(ru_r')+\f{\rmi mu_\phi'}{r}+\p_zu_z'\right],
771: \end{eqnarray}
772: where $\hat\omega=\omega-m\Omega$ is the Doppler-shifted wave
773: frequency and the perturbations, denoted by primed quantities, have
774: the dependence $\exp(-\rmi\omega t+\rmi m\phi)$.
775: 
776: Following \citet{1987PASJ...39..457O}, we decompose the vertical
777: structure of the mode into the basis of Hermite polynomials defined by
778: \begin{equation}
779:   \he_n(\zeta)=\rme^{\zeta^2/2}\left(-\f{\rmd}{\rmd \zeta}\right)^n\rme^{-\zeta^2/2},
780: \end{equation}
781: where $\zeta=z/H$ is a dimensionless vertical coordinate and
782: $n=0,1,2,\dots$\  These polynomials satisfy the differential equation
783: \begin{equation}
784:   \he_n''(\zeta)-\zeta\,\he_n'(\zeta)+n\,\he_n(\zeta)=0,
785: \end{equation}
786: the recurrence relations
787: \begin{equation}
788:   \he_n'(\zeta)=n\,\he_{n-1}(\zeta),
789: \end{equation}
790: \begin{equation}
791:   \zeta\,\he_n(\zeta)=\he_{n+1}(\zeta)+n\,\he_{n-1}(\zeta)
792: \end{equation}
793: and the orthogonality relation
794: \begin{equation}
795:   \int_{-\infty}^\infty\rme^{-\zeta^2/2}\,\he_m(\zeta)\,\he_n(\zeta)\,\rmd \zeta=(2\pi)^{1/2}n!\,\delta_{mn}.
796: \end{equation}
797: The first three are $\he_0(\zeta)=1$, $\he_1(\zeta)=\zeta$ and
798: $\he_2(\zeta)=\zeta^2-1$.
799: 
800: We therefore expand
801: \begin{eqnarray}
802:   u_r'(r,z)&=&\sum_nu_n(r)\,\he_n(\zeta),\\
803:   u_\phi'(r,z)&=&\sum_nv_n(r)\,\he_n(\zeta),\\
804:   u_z'(r,z)&=&\sum_nw_n(r)\,\he_{n-1}(\zeta),\\
805:   h'(r,z)&=&\sum_nh_n(r)\,\he_n(\zeta).
806: \end{eqnarray}
807: with $u_n=v_n=h_n=0$ for $n<0$ and $w_n=0$ for $n<1$.  Bearing in mind
808: that $H$ depends on $r$, we have
809: \begin{equation}
810:   \p_r\,\he_n(\zeta)=-(\p_r\ln H)[n\,\he_n(\zeta)+n(n-1)\,\he_{n-2}(\zeta)].
811: \end{equation}
812: The projected equations are then
813: \begin{eqnarray}
814:   \lefteqn{-\rmi\hat\omega u_n-2\Omega v_n=-\p_r h_n}&\nonumber\\
815:   &&+(\p_r\ln H)[nh_n+(n+1)(n+2)h_{n+2}],
816: \label{un}
817: \end{eqnarray}
818: \begin{equation}
819:   -\rmi\hat\omega v_n+\f{\kappa^2}{2\Omega}u_n=-\f{\rmi mh_n}{r},
820: \label{vn}
821: \end{equation}
822: \begin{equation}
823:   -\rmi\hat\omega w_n=-\f{nh_n}{H},
824: \label{wn}
825: \end{equation}
826: \begin{eqnarray}
827:   \lefteqn{-\rmi\hat\omega\f{h_n}{c_\rms^2}+\f{1}{r\Sigma}\p_r(r\Sigma u_n)+\f{\rmi mv_n}{r}-\f{w_n}{H}}&\nonumber\\
828:   &&+(\p_r\ln H)(nu_n+u_{n-2})=0,
829: \label{hn}
830: \end{eqnarray}
831: \citep[cf.][]{2002ApJ...565.1257T,2006MNRAS.368..917Z}.
832: 
833: This approach corresponds to a (Galerkin) spectral treatment of the
834: partial differential equations governing the linearized dynamics,
835: which is much preferable to a finite-difference treatment.  In
836: practice this system of equations must be truncated by setting $u_n$,
837: etc., to zero for $n>N$ for some integer $N$.  The 2DO~theory is
838: obtained, in fact, by considering a radical truncation, $N=0$, of the
839: equations.
840: 
841: \subsection{Selected viscous effects}
842: \label{s:viscous}
843: 
844: To include viscosity, a term
845: \begin{equation}
846:   \f{1}{\rho}\bnabla\bcdot\bfT
847: \end{equation}
848: should be added to the right-hand side of the equation of
849: motion~(\ref{motion}), where
850: \begin{equation}
851:   \bfT=\rho\nu[\bnabla\bu+(\bnabla\bu)^\rmT]+\rho(\nu_\rmb-{\textstyle\f{2}{3}}\nu)(\bnabla\bcdot\bu)\bfone
852: \end{equation}
853: is the viscous stress tensor.  In the context of an isothermal disc it
854: is reasonable to assume that the kinematic shear and bulk viscosities
855: $\nu$ and $\nu_\rmb$ depend only on $r$.  We parametrize them as
856: \begin{equation}
857:   \nu=\alpha c_\rms H,\qquad\nu_\rmb=\alpha_\rmb c_\rms H.
858: \end{equation}
859: 
860: A full treatment of the effects of viscosity is complicated, not only
861: because the above expression for the viscous force must be evaluated
862: in cylindrical polar coordinates and then projected on to the basis of
863: Hermite polynomials, but also because the basic state is modified to
864: include a meridional flow driven by viscous stresses, which should be
865: considered in the linearized equations.  This problem is therefore
866: deferred to a future investigation.
867: 
868: In the present paper we adopt a simpler approach in which only
869: selected viscous effects are included.  We consider what might
870: be assumed to be the dominant viscous terms, i.e.\ those involving two
871: derivatives with respect to~$z$.  Since
872: \begin{equation}
873:   \f{1}{\rho}\p_z[\rho\p_z\,\he_n(\zeta)]=-\f{n}{H^2}\,\he_n(\zeta),
874: \end{equation}
875: the inviscid perturbation equations (\ref{un})--(\ref{wn}) are
876: modified by the addition of the viscous terms
877: \begin{equation}
878:   -\rmi\hat\omega u_n=\cdots-\f{\nu}{H^2}nu_n,
879: \label{unviscous}
880: \end{equation}
881: \begin{equation}
882:   -\rmi\hat\omega v_n=\cdots-\f{\nu}{H^2}nv_n,
883: \end{equation}
884: \begin{equation}
885:   -\rmi\hat\omega w_n=\cdots-\f{(\nu_\rmb+{\textstyle\f{4}{3}}\nu)}{H^2}(n-1)w_n,
886: \label{wnviscous}
887: \end{equation}
888: while equation~(\ref{hn}) is unchanged.  These terms act to damp the
889: mode, but have most effect on components of large $n$.  They have no
890: effect on $u_0$ and $v_0$, which represent horizontal motions
891: independent of $z$.  These viscous terms can also be thought of as
892: providing a coupling between different layers of the disc and thereby
893: encouraging it to adopt a horizontal motion independent of $z$.  We
894: show below that this effect is of considerable importance.
895: 
896: \subsection{Numerical solutions}
897: 
898: We solve the system of ordinary differential equations in $r$ for
899: modes with $m=1$ using a Chebyshev collocation (i.e.\ pseudospectral)
900: method.  This approach converts the differential equations and
901: boundary conditions into an algebraic generalized eigenvalue problem
902: for the frequency $\omega$, which we solve using a standard direct
903: method.  Specifically, equations~(\ref{un})--(\ref{hn}),
904: supplemented by the viscous terms
905: (\ref{unviscous})--(\ref{wnviscous}), are solved for $n=0,2,4,\dots,N$
906: with $u_n$, etc., set to zero for $n>N$.  Rigid boundary conditions
907: $u_n=0$ are adopted at both inner and outer boundaries, but it is
908: ensured that the modes obtained are completely insensitive to the
909: value of the outer radius and therefore to the choice of outer
910: boundary condition.
911: 
912: For comparison with the results in Section~\ref{s:powerlaw}, we
913: consider a disc with a midplane density profile $\rho_\rmm\propto
914: r^{-\sigma-3/2}$.  We also include a shear viscosity corresponding to
915: a constant $\alpha$ parameter, but no bulk viscosity.
916: 
917: Sample results are shown in Table~\ref{t:ot}.  The convergence of the
918: eigenfrequency with increasing truncation order $N$ of the Hermite
919: polynomial basis is remarkable.  The case $N=0$ corresponds exactly to
920: the two-dimensional theory considered by \citet{1991PASJ...43...75O},
921: and therefore agrees well with the 2DO~secular approximation.  Here
922: $\tilde Q=20$ is large enough to support a confined prograde mode.
923: The precession rate is much larger in the case $N=2$ and hardly varies
924: as further Hermite polynomials are included.  It agrees reasonably
925: well with the 3D~secular theory for an inviscid disc.  The slight
926: offset of the precession frequency is attributable partly to errors in
927: the secular approximation, which is valid only to leading order in
928: $\epsilon$, and partly to the effects of viscosity.  As described in
929: Appendix~\ref{s:appendix}, the viscous damping of vertical motions
930: considered in the full model can be represented within the 3D~secular
931: theory by multiplying the coefficient $9/4$ in the three-dimensional
932: expression~(\ref{f3d}) for $f$ by $(1-\rmi\beta)/(1+\rmi\beta)$, where
933: $\beta=\alpha_\rmb+{\textstyle\f{4}{3}}\alpha$.  Table~\ref{t:ot}
934: shows that this viscous secular theory gives good agreement with the
935: full model for $\alpha=0.1$.
936: 
937: \begin{table}
938: \caption{Scaled frequency eigenvalues obtained from the full
939: linearized equations for a disc with $\epsilon=0.02$, $\alpha=0.1$,
940: $\tilde Q=20$ and $\sigma=2$.  Rapid convergence is seen with
941: increasing values of the vertical truncation number $N$.  Comparable
942: results from the secular theories are shown below.  A negative value
943: of $\mathrm{Im}(\tilde\omega)$ represents the (scaled) exponential
944: damping rate of the mode.}
945: \begin{tabular}{lll}
946: Version&$\mathrm{Re}(\tilde\omega)$&$\mathrm{Im}(\tilde\omega)$\\
947: \hline
948: Full, $N=0$&1.189356&\\
949: Full, $N=2$&2.782638&-0.449908\\
950: Full, $N=4$&2.782585&-0.449510\\
951: Full, $N=6$&2.782585&-0.449510\\
952: \\
953: 2DS secular, inviscid&0.636435&\\
954: 2DO secular, inviscid&1.193322&\\
955: 3D secular, inviscid&2.890299&\\
956: 3D secular, viscous&2.829523&-0.448547\\
957: \hline
958: \end{tabular}
959: \label{t:ot}
960: \end{table}
961: 
962: The viscous damping rate of the modes is considerable.  Although the
963: dominant motion is horizontal and independent of $z$, and therefore
964: does not incur any viscous forces in our approximation, the
965: accompanying vertical motion is damped.  To excite eccentric modes in
966: a three-dimensional disc, this damping must be overcome.  Viscous
967: overstability may be able to do this, but detailed calculations are
968: required and there is uncertainty in the applicability of a
969: Navier--Stokes viscosity to turbulent stresses in the disc.  A
970: simple estimate can be made as follows.  In a two-dimensional,
971: isothermal, Keplerian shearing sheet with constant kinematic shear
972: viscosity $\nu$ and no bulk viscosity, the maximum local growth rate
973: of the overstability is $\approx0.034\,\alpha\Omega$ and occurs for a
974: radial wavelength $\approx13\,H$ \citep{2006MNRAS.372.1829L}.
975: Although our eigenfunctions do not have an obviously wavelike form,
976: this suggests that overstability may be able to compensate for the
977: damping rate found in Table~\ref{t:ot}, which corresponds to only
978: $\approx0.0018\,\alpha\Omega$ for the parameters adopted there.
979: 
980: A sufficiently large viscosity is required to couple different layers
981: in the disc effectively.  If the viscosity is reduced to
982: $\alpha=0.01$, with other parameters as in Table~\ref{t:ot}, a
983: frequency eigenvalue of $\tilde\omega=2.7831-0.1793\,\rmi$ is
984: obtained, and a slightly larger value of $N$ is required to obtain the
985: same convergence.  While the precession rate agrees well with the case
986: of $\alpha=0.1$, the damping rate is now larger than predicted by the
987: viscous secular 3D~theory.  This happens because of the increasing
988: $z$-dependence of the horizontal motion in the absence of a strong
989: coupling between layers; although $u_2$ is still much smaller than
990: $u_0$ for $\alpha=0.01$, the ratio $u_2/u_0$ is several times larger
991: than in the case $\alpha=0.1$.  The viscous damping of this shearing
992: motion is more important, for $\alpha=0.01$, than that of the vertical
993: motion considered in the viscous secular theory.
994: 
995: If $\alpha$ is reduced further, the precession rate starts to
996: deviate from the 3D~secular theory and more vertical structure
997: develops in the eigenfunction, requiring a larger value of $N$ for
998: convergence.  Similar behaviour was found by
999: \citet{2006MNRAS.372.1829L}.  The behaviour of a three-dimensional
1000: inviscid eccentric disc could be very difficult to describe.
1001: 
1002: We also note that the stresses associated with tangled
1003: magnetic fields in a disc in which the magnetorotational instability
1004: occurs may provide an elastic, or viscoelastic, coupling between
1005: different layers \citep{2001MNRAS.325..231O}.  The associated damping
1006: rate may be smaller than that estimated on the basis of a
1007: Navier--Stokes viscosity.
1008: 
1009: \section{Conclusions}
1010: \label{s:conc}
1011: 
1012: In this paper we have examined the linear dynamics of one-armed
1013: oscillation modes in the circumstellar discs of Be~stars.  A
1014: three-dimensional effect, first identified by
1015: \citet{2001MNRAS.325..231O} but neglected in previous treatments of
1016: Be~stars, makes a crucial positive contribution to the precession
1017: rates of such modes.  It allows confined prograde modes to be obtained
1018: for all reasonable disc temperatures and stellar rotation rates.  This
1019: property allows us to give a unified description of eccentric discs of
1020: Be~stars of all stellar types without introducing uncertain radiative
1021: forces or modifying the inner boundary condition.  While we do not
1022: deny that in some cases radiative forces might be important, or that
1023: the inner boundary condition might differ from a rigid one, we have
1024: shown that these innovations are unnecessary.  We obtained these
1025: results using a secular theory of eccentric discs and confirmed them
1026: using a spectral treatment of the full linearized equations for
1027: three-dimensional isothermal discs including viscous terms that couple
1028: the horizontal motions at different altitudes.  In order to
1029: make these modes grow, viscous damping must be overcome by an
1030: excitation mechanism such as viscous overstability, which will be
1031: investigated in a subsequent paper.
1032: 
1033: The three-dimensional dynamics that we have described may also
1034: have important consequences for the behaviour of eccentric discs in
1035: other circumstances.  For example, in cataclysmic variable stars
1036: exhibiting superhumps, the relation between the precession rate of the
1037: disc and the binary mass ratio is an important observational property
1038: that is not fully explained by current theoretical models
1039: \citep{2006MNRAS.368.1123G,2007MNRAS.378..785S}.
1040: 
1041: The physical model adopted in this paper is idealized.  To explain in
1042: detail the observed cyclical behaviour of the emission lines of
1043: Be~stars within this theoretical framework is likely to require a
1044: treatment of non-isothermal and non-linear effects as well as a better
1045: understanding of the time-dependent behaviour of the density
1046: distribution of the circumstellar disc.  In addition, future work
1047: should attempt to identify and assess mechanisms, such as viscous
1048: overstability, by which one-armed oscillation modes may be excited in
1049: circumstellar discs.  Nevertheless, the effect investigated in this
1050: paper enhances the credibility of the one-armed oscillation model by
1051: providing a natural explanation of confined prograde modes.
1052: 
1053: \section*{acknowledgments}
1054: 
1055: I thank Atsuo Okazaki, John Papaloizou and an anonymous
1056: referee for useful comments.
1057: 
1058: \appendix
1059: 
1060: \section{Derivation of the secular theory}
1061: \label{s:appendix}
1062: 
1063: We consider equations~(\ref{un})--(\ref{hn}) together with the viscous
1064: effects described in Section~\ref{s:viscous}.  When $v_n$ and $w_n$
1065: are eliminated, we have
1066: \begin{eqnarray}
1067:   \lefteqn{\f{(\hat\omega_{\mathrm{h}n}^2-\kappa^2)}{\rmi\hat\omega_{\mathrm{h}n}}u_n=-\p_rh_n+\f{2m\Omega}{r\hat\omega_{\mathrm{h}n}}h_n}&\nonumber\\
1068:   &&+(\p_r\ln H)[nh_n+(n+1)(n+2)h_{n+2}],
1069: \label{unelim}
1070: \end{eqnarray}
1071: \begin{eqnarray}
1072:   \lefteqn{\left(-\rmi\hat\omega-\f{n\Omega_z^2}{\rmi\hat\omega_{\mathrm{v}n}}-\f{m^2c_\rms^2}{r^2}\f{1}{\rmi\hat\omega_{\mathrm{h}n}}\right)\f{h_n}{c_\rms^2}+\f{m\kappa^2}{2r\Omega\hat\omega_{\mathrm{h}n}}u_n}&\nonumber\\
1073:   &&+\f{1}{r\Sigma}\p_r(r\Sigma u_n)+(\p_r\ln H)(nu_n+u_{n-2})=0,
1074: \label{hnelim}
1075: \end{eqnarray}
1076: where $\hat\omega_{\mathrm{h}n}=\hat\omega+\rmi\alpha n\Omega_z$ and
1077: $\hat\omega_{\mathrm{v}n}=\hat\omega+\rmi(\alpha_\rmb+{\textstyle\f{4}{3}}\alpha)(n-1)\Omega_z$
1078: are the Doppler-shifted wave frequency modified to allow for the
1079: viscous damping of horizontal and vertical motions, respectively.  We
1080: are interested in the case $m=1$.
1081: 
1082: This system of equations is not closed because the $u_0$ equation
1083: refers to $h_2$, which depends on $u_2$.  In turn, the $u_2$ equation
1084: refers to $h_4$, which depends on $u_4$, and so on.  However, suppose
1085: for the time being that $u_2$ can be neglected in the $h_2$ equation
1086: (i.e.\ equation~\ref{hnelim} with $n=2$).  Then we have a
1087: closed system
1088: \begin{equation}
1089:   \f{(\hat\omega^2-\kappa^2)}{\rmi\hat\omega}u_0=-\p_rh_0+\f{2\Omega}{r\hat\omega}h_0+2(\p_r\ln H)h_2,
1090: \label{closed1}
1091: \end{equation}
1092: \begin{equation}
1093:   \left(-\rmi\hat\omega-\f{c_\rms^2}{r^2}\f{1}{\rmi\hat\omega}\right)\f{h_0}{c_\rms^2}+\f{\kappa^2}{2r\Omega\hat\omega}u_0+\f{1}{r\Sigma}\p_r(r\Sigma u_0)=0,
1094: \label{closed2}
1095: \end{equation}
1096: \begin{equation}
1097:   \left(-\rmi\hat\omega-\f{2\Omega_z^2}{\rmi\hat\omega_{\mathrm{v}2}}-\f{c_\rms^2}{r^2}\f{1}{\rmi\hat\omega_{\mathrm{h}2}}\right)\f{h_2}{c_\rms^2}+(\p_r\ln H)u_0\approx0.
1098: \label{closed3}
1099: \end{equation}
1100: These equations agree with a two-dimensional theory except for the
1101: additional $h_2$ term in equation~(\ref{closed1}), which can be
1102: related to $u_0$ through the approximate equation~(\ref{closed3}).
1103: 
1104: In the secular approximation we neglect the differences between
1105: $\Omega$, $\kappa$, $\Omega_z$ and $\Omega_\rmK$ except where
1106: essential, so that $\Omega\propto r^{-3/2}$ and $H\propto r^{3/2}$,
1107: and we assume a low frequency $|\omega|\ll\Omega$.  The leading
1108: approximations to the above equations in the inviscid case are then
1109: \begin{equation}
1110:   2\rmi\left(\f{\Omega^2-\kappa^2}{2\Omega}-\omega\right)u_0=-\p_rh_0-\f{2h_0}{r}+\f{3h_2}{r},
1111: \end{equation}
1112: \begin{equation}
1113:   \rmi\Omega\f{h_0}{c_\rms^2}-\f{u_0}{2r}+\f{1}{r\Sigma}\p_r(r\Sigma u_0)=0,
1114: \end{equation}
1115: \begin{equation}
1116:   -\rmi\Omega\f{h_2}{c_\rms^2}+\f{3u_0}{2r}=0.
1117: \label{h2}
1118: \end{equation}
1119: These combine to give equation~(\ref{secular}) for $u=u_0$, with the
1120: three-dimensional expression~(\ref{f3d}) for $f$.  If $h_2$ is
1121: neglected altogether, equation~(\ref{secular}) is obtained, but with
1122: the two-dimensional expression~(\ref{f2d}) for $f$.  It can be
1123: seen from the above equations that $h_2$ always makes a positive
1124: contribution to the precession rate $\omega$.
1125: 
1126: Neglecting $u_2$ compared to $u_0$ is equivalent to assuming that the
1127: eccentricity is independent of $z$, since $\he_0(\zeta)=1$ and
1128: $\he_2(\zeta)=\zeta^2-1$.  Under what conditions is this assumption
1129: reasonable?  Rough estimates based on equations~(\ref{unelim})
1130: and~(\ref{hnelim}) show that $u_2$ can be neglected in the $h_2$
1131: equation if $\alpha\Omega$ is much larger than the precession
1132: frequencies $\omega$ or $f$, meaning that the shear viscosity prevents
1133: the significant development of a $z$-dependent horizontal motion.
1134: This condition is readily satisfied in the circumstellar discs of
1135: Be~stars for reasonable values of $\alpha$.  Ultimately, however, the
1136: validation of this approximation comes from the numerical solution of
1137: the full system of linearized equations.
1138: 
1139: If viscosity is retained in this analysis, within the secular
1140: approximation, then the coefficient of $h_2$ in equation~(\ref{h2}) is
1141: multiplied by $(1+\rmi\beta)/(1-\rmi\beta)$, where
1142: $\beta=\alpha_\rmb+{\textstyle\f{4}{3}}\alpha$.  In this case,
1143: equation~(\ref{secular}) is again obtained, but in
1144: expression~(\ref{f3d}) for $f$ the coefficient $9/4$ of the
1145: three-dimensional term is multiplied by $(1-\rmi\beta)/(1+\rmi\beta)$.
1146: In this case the mode decays because of the viscous damping of the
1147: vertical motion.  For sufficiently small $\alpha$, however,
1148: the decay rate is enhanced by the viscous damping of the $z$-dependent
1149: horizontal motion.
1150: 
1151: \begin{thebibliography}{}
1152: \bibitem[Abramowitz \& Stegun(1965)]{1965hmfw.book.....A}
1153: Abramowitz, M.\ \& Stegun, I.~A., 1965, Handbook of Mathematical Functions, Dover, New York
1154: \bibitem[Fi{\v r}t \& Harmanec(2006)]{2006A&A...447..277F}
1155: Fi{\v r}t, R.\ \& Harmanec, P., 2006, A\&A, 447, 277 
1156: \bibitem[Goodchild \& Ogilvie(2006)]{2006MNRAS.368.1123G}
1157: Goodchild, S.\ \& Ogilvie, G., 2006, MNRAS, 368, 1123 
1158: \bibitem[Kato(1978)]{1978MNRAS.185..629K}
1159: Kato, S., 1978, MNRAS, 185, 629
1160: \bibitem[Kato(1983)]{1983PASJ...35..249K}
1161: Kato, S., 1983, PASJ, 35, 249
1162: \bibitem[Latter \& Ogilvie(2006)]{2006MNRAS.372.1829L}
1163: Latter, H.~N.\ \& Ogilvie, G.~I., 2006, MNRAS, 372, 1829
1164: \bibitem[Lee et al.(1991)]{1991MNRAS.250..432L}
1165: Lee, U., Osaki, Y.\ \& Saio, H., 1991, MNRAS, 250, 432 
1166: \bibitem[Negueruela et al.(2001)]{2001A&A...369..117N}
1167: Negueruela, I., Okazaki, A.~T., Fabregat, J., Coe, M.~J., Munari, U.\ \& Tomov, T., 2001, A\&A, 369, 117
1168: \bibitem[Ogilvie(2001)]{2001MNRAS.325..231O}
1169: Ogilvie, G.~I., 2001, MNRAS, 325, 231 
1170: \bibitem[Okazaki(1991)]{1991PASJ...43...75O}
1171: Okazaki, A.~T., 1991, PASJ, 43, 75
1172: \bibitem[Okazaki(1997)]{1997A&A...318..548O}
1173: Okazaki, A.~T., 1997, A\&A, 318, 548 
1174: \bibitem[Okazaki(2007)]{2007ASPC..361..230O}
1175: Okazaki, A.~T., 2007, in Active OB-Stars: Laboratories for Stellar and Circumstellar Physics, eds S.~Stefl, S.~P.~Owocki \& A.~T.~Okazaki, ASP Conf.\ Ser., 361, 230
1176: \bibitem[Okazaki et al.(1987)]{1987PASJ...39..457O}
1177: Okazaki, A.~T., Kato, S.\ \& Fukue, J., 1987, PASJ, 39, 457 
1178: \bibitem[Papaloizou(2002)]{2002A&A...388..615P}
1179: Papaloizou, J.~C.~B., 2002, A\&A, 388, 615 
1180: \bibitem[Papaloizou \& Savonije(2006)]{2006A&A...456.1097P}
1181: Papaloizou, J.~C.~B.\ \& Savonije, G.~J., 2006, A\&A, 456, 1097
1182: \bibitem[Papaloizou et al.(1992)]{1992A&A...265L..45P}
1183: Papaloizou, J.~C., Savonije, G.~J.\ \& Henrichs, H.~F., 1992, A\&A, 265, L45 
1184: \bibitem[Porter \& Rivinius(2003)]{2003PASP..115.1153P}
1185: Porter, J.~M.\ \& Rivinius, T., 2003, PASP, 115, 1153 
1186: \bibitem[Savonije \& Heemskerk(1993)]{1993A&A...276..409S}
1187: Savonije, G.~J.\ \& Heemskerk, M.~H.~M., 1993, A\&A, 276, 409 
1188: \bibitem[Smith et al.(2007)]{2007MNRAS.378..785S}
1189: Smith, A.~J., Haswell, C.~A., Murray, J.~R., Truss, M.~R.\ \& Foulkes, S.~B.\ 2007, MNRAS, 378, 785
1190: \bibitem[Tanaka et al.(2002)]{2002ApJ...565.1257T}
1191: Tanaka, H., Takeuchi, T.\ \& Ward, W.~R., 2002, ApJ, 565, 1257 
1192: \bibitem[Telting et al.(1994)]{1994A&A...288..558T}
1193: Telting, J.~H., Heemskerk, M.~H.~M., Henrichs, H.~F.\ \& Savonije, G.~J., 1994, A\&A, 288, 558 
1194: \bibitem[Tremaine(2001)]{2001AJ....121.1776T}
1195: Tremaine, S., 2001, AJ, 121, 1776 
1196: \bibitem[Zhang \& Lai(2006)]{2006MNRAS.368..917Z}
1197: Zhang, H.\ \& Lai, D., 2006, MNRAS, 368, 917
1198: \end{thebibliography}
1199: 
1200: \label{lastpage}
1201: 
1202: \end{document}
1203: