1: %\documentclass[12pt,preprint]{aastex}
2: \documentclass[apj]{emulateapj}
3: \usepackage{apjfonts}
4: \usepackage{epsfig}
5: \usepackage{amsmath}
6: \usepackage{mathrsfs}
7:
8: % math commands
9: \newcommand{\be}{\begin{eqnarray}}
10: \newcommand{\ee}{\end{eqnarray}}
11: \newcommand{\beq}{\begin{equation}}
12: \newcommand{\eeq}{\end{equation}}
13: \def\simless{\mathbin{\lower 3pt\hbox
14: {$\rlap{\raise 5pt\hbox{$\char'074$}}\mathchar"7218$}}}
15: \def\simgreat{\mathbin{\lower 3pt\hbox
16: {$\rlap{\raise 5pt\hbox{$\char'076$}}\mathchar"7218$}}} %> or of order
17:
18: % variables
19: %\renewcommand{\vec}[1]{\mbox{\boldmath $\displaystyle #1$}}
20: \renewcommand{\vec}[1]{\mathbf{#1}}
21: \newcommand{\xivec}{{\mbox{\boldmath $\xi$}}}
22:
23: \newcommand{\grad}{\boldsymbol{\nabla}}
24: \newcommand{\yr}{\,{\rm yr}}
25: \newcommand{\myr}{\,{\rm Myr}}
26: \newcommand{\gyr}{\,{\rm Gyr}}
27: \newcommand{\cm}{\,{\rm cm}}
28: \newcommand{\km}{\,{\rm km}}
29: \newcommand{\pc}{\,{\rm pc}}
30: \newcommand{\kpc}{\,{\rm kpc}}
31: \newcommand{\kev}{\,{\rm keV}}
32: \newcommand{\ev}{\,{\rm eV}}
33: \newcommand{\ergs}{\,{\rm erg\,s^{-1}}}
34: \newcommand{\s}{\,{\rm s}}
35: \newcommand{\msun}{\,M_\odot}
36: \newcommand{\rsun}{\,R_\odot}
37: \newcommand{\lsun}{\,L_{\rm \odot}}
38: \newcommand{\lx}{L_{\rm X}}
39: \newcommand{\fx}{F_{\rm X}}
40: \newcommand{\mdot}{\,M_\odot\,{\rm yr}^{-1}}
41:
42: \newcommand{\fdisk}{\mathscr{F}}
43: \newcommand{\pert}{\varepsilon}
44:
45: % -----------------------------------------------------------
46: % -----------------------------------------------------------
47:
48: \begin{document}
49:
50: % -----------------------------------------------------------
51: % -----------------------------------------------------------
52:
53: \shorttitle{ELLIPSOIDAL OSCILLATIONS}
54: \shortauthors{PFAHL, ARRAS, \& PAXTON}
55:
56: % -----------------------------------------------------------
57: % -----------------------------------------------------------
58:
59: \title{Ellipsoidal Oscillations Induced by Substellar Companions: \\
60: A Prospect for the {\em Kepler} Mission}
61:
62: \author{Eric Pfahl\altaffilmark{1}, Phil Arras\altaffilmark{1,2},
63: and Bill Paxton\altaffilmark{1}}
64:
65: \altaffiltext{1}{Kavli Institute for Theoretical Physics, University
66: of California, Santa Barbara, CA 93106; pfahl@kitp.ucsb.edu,
67: paxton@kitp.ucsb.edu}
68:
69: \altaffiltext{2}{Department of Astronomy, University of Virginia,
70: P.O. Box 400325, Charlottesville, VA 22904-4325; arras@virginia.edu}
71:
72: % -----------------------------------------------------------
73: % -----------------------------------------------------------
74:
75: \begin{abstract}
76:
77: Hundreds of substellar companions to solar-type stars will be
78: discovered with the {\em Kepler} satellite. {\em Kepler}'s extreme
79: photometric precision gives access to low-amplitude stellar
80: variability contributed by a variety of physical processes. We
81: discuss in detail the periodic flux modulations arising from the
82: tidal force on the star due to a substellar companion. An analytic
83: expression for the variability is derived in the equilibrium-tide
84: approximation. We demonstrate analytically and through numerical
85: solutions of the linear, nonadiabatic stellar oscillation equations
86: that the equilibrium-tide formula works extremely well for stars of
87: mass $<$$1.4\msun$ with thick surface convection zones. More
88: massive stars with largely radiative envelopes do not conform to the
89: equilibrium-tide approximation and can exhibit flux variations
90: $\ga$10 times larger than naive estimates. Over the full range of
91: stellar masses considered, we treat the oscillatory response of the
92: convection zone by adapting a prescription that A. J. Brickhill
93: developed for pulsating white dwarfs. Compared to other sources of
94: periodic variability, the ellipsoidal lightcurve has a distinct
95: dependence on time and system parameters. We suggest that
96: ellipsoidal oscillations induced by giant planets may be detectable
97: from as many as $\sim$100 of the $10^5$ {\em Kepler} target stars.
98: For the subset of these stars that show transits and have
99: radial-velocity measurements, all system parameters are well
100: constrained, and measurement of ellipsoidal variation provides a
101: consistency check, as well as a test of the theory of forced stellar
102: oscillations in a challenging regime.
103:
104: \end{abstract}
105:
106: \keywords{planetary systems --- stars: oscillations ---
107: techniques: photometric}
108:
109: % -----------------------------------------------------------
110: % -----------------------------------------------------------
111:
112: \section{INTRODUCTION}
113: \label{sec:intro}
114:
115: The upcoming {\em Kepler}\footnote{\url{http://kepler.nasa.gov}}
116: satellite will continuously monitor $\sim$$10^5$ main-sequence stars
117: of mass $\simeq$0.5--$1.5\msun$ over 4--6 years with fractional
118: photometric precisions of $\sim$$10^{-5}$. Such high sensitivity,
119: which is unattainable from the ground, will allow for the robust
120: detection of Earth-size planets that transit their host stars, and the
121: measurement of asteroseismic oscillations as a probe of stellar
122: structure \citep[e.g.,][]{Borucki2004,Basri2005}. These missions will
123: also discover hundreds of ``hot Jupiters'' with orbital periods of
124: $<$10 days, revealed by their transits or reflected starlight
125: \citep[e.g.,][]{Jenkins2003}. Continuous observations of these
126: systems are likely to show a myriad of novel physical effects,
127: including Doppler flux variability of the host stars \citep{Loeb2003},
128: photometric dips due to moons or rings around the planets
129: \citep{Sartoretti1999,Brown2001}, and the impact of additional
130: perturbing planets on transit timing
131: \citep{Miralda-Escude2002,Agol2005,Holman2005}. The same ideas apply
132: if the companion is a more massive brown dwarf, but these are rarely
133: found in close orbits around solar-type stars
134: \citep[e.g.,][]{Grether2006}.
135:
136: Here we scrutinize another mechanism for generating periodic
137: variability of a star closely orbited by a giant planet or brown
138: dwarf. A star subject to the tidal gravity of a binary companion has a
139: nonspherical shape and surface-brightness distribution. In the
140: simplest approximation, the stellar surface is a prolate ellipsoid
141: with its long axis on the line connecting the two objects. As the
142: tidal bulge tracks the orbital motion, differing amounts of light
143: reach the observer. For a solar-type star orbited by a perturbing
144: companion of mass $M_p$ with period $P_{\rm orb}$, the expected
145: fractional amplitude of this ellipsoidal variability is
146: $\sim$$10^{-2}(M_p/M_\odot)(1\,{\rm day}/P_{\rm orb})^2$. This effect
147: has a long history in the study of eclipsing binary stars \citep[see
148: the review by][]{Wilson1994}, but was mentioned only recently in the
149: exoplanet context.
150:
151: \citet{Udalski2002}, \citet{Drake2003}, and \citet{Sirko2003} noted
152: that if ellipsoidal light variations are detected from the ground,
153: where the fractional photometric precision is $\ga$$10^{-3}$, then the
154: perturber must be fairly massive (e.g., $\ga$$0.1\msun$). They
155: offered this idea as a test to distinguish between planetary transits
156: and eclipses by low-mass stars. The superior sensitivity of {\em
157: Kepler} offers the possibility of measuring ellipsoidal variability
158: induced by giant planets ($M_p \sim 10^{-3}$--$10^{-2}\msun$) with
159: orbital periods of $\la$10\,days.
160:
161: \citet{Loeb2003} compare the ellipsoidal variability induced by a
162: planetary companion to flux modulations arising from reflected
163: starlight and the Doppler effect. The three amplitudes are similar
164: when the companion has an orbital period of $\la$3 days and an optical
165: albedo of $\la$0.1. In a sufficiently long observation it should be
166: possible to separately extract each of the signals, since their
167: Fourier decompositions are distinct. Precise physical modeling of the
168: ellipsoidal lightcurve could provide an independent constraint on the
169: mass of the companion, as well as important clues regarding stellar
170: tidal interactions.
171:
172: Ellipsoidal variability is typically modeled under the assumption that
173: the distorted star maintains hydrostatic balance and precisely fills a
174: level surface of an appropriate potential (e.g., the Roche potential).
175: The measured flux is then just an integral of the intensity over the
176: visible stellar surface, where the intensity includes the effects of
177: limb darkening and gravity darkening \citep[e.g.,][]{Kopal1942}. This
178: approach is strictly valid only when the orbit is circular and the
179: star rotates at the orbital frequency, so that a stationary
180: configuration exists in the coorbital frame. These conditions may not
181: be satisfied when the companion has a low mass or long period, because
182: of the weak tidal interaction. In fact, a state of tidal equilibrium
183: may not be attainable in the case of a planetary companion
184: \citep[e.g.,][]{Rasio1996}. Equilibrium models of ellipsoidal
185: lightcurves do have a realm of validity for noncircular orbits and
186: asynchronously rototating stars, and have been applied successfully to
187: somewhat eccentric binaries \citep[e.g.,][]{Soszynski2004}. However,
188: by construction, such models ignore fluid inertia and the possibility
189: exciting normal modes of oscillation, effects that may be of critical
190: importance in a wide range of observationally relevant circumstances.
191: Here we apply the machinery of linear stellar oscillation theory to
192: the weak tidal forcing of stars by substellar companions.
193: Conceptually, our investigation bridges {\em Kepler}'s planetary and
194: astroseismology programs.
195:
196: Section~2 describes the geometry of the problem, provides quantitative
197: measures for the strength of the tidal interaction, discusses our
198: simplifying assumptions, and presents the mathematical framework for
199: calculating ellipsoidal variability. In \S~\ref{sec:eqtide}, we
200: consider the equilibrium-tide approximation and derive an analytic
201: expression for the ellipsoidal lightcurve. A brief review of von
202: Zeipel's theorem and its limitations is given in
203: \S~\ref{sec:vonzeipel}. Tidally forced, nonadiabatic stellar
204: oscillations are addressed in \S~\ref{sec:nonad}, where we argue for a
205: simple treatment of perturbed surface convection zones, use this
206: prescription to calculate the ellipsoidal variability of deeply
207: convective stars, estimate analytically the surface flux perturbation
208: in mainly radiative stars, and show select numerical results. Our
209: main conclusions are summarized in \S~\ref{sec:summary}. We conclude
210: in \S~\ref{sec:detection} with remarks on the measurement of
211: ellipsoidal oscillations in the presence of other sources of periodic
212: variability.
213:
214: % -----------------------------------------------------------
215:
216: \section{PRELIMINARIES}
217: \label{sec:prelim}
218:
219: Consider a star of mass $M$ and radius $R$ is orbited by a substellar
220: companion of mass $M_p$ and radius $R_p$. We work in spherical
221: coordinates $(r, \theta, \phi)$ with the origin at the star's center
222: and the pole direction ($\theta = 0$) parallel to the orbital angular
223: momentum vector. The orbit is then described by $(d,
224: \pi/2, \phi_p)$, where $d$ and $\phi_p$ are, respectively, the
225: time-dependent orbital separation and true anomaly; $\phi_p = 0$ marks
226: the phase of periastron. We assume that the orbit is strictly
227: Keplerian with fixed semimajor axis $a$ and eccentricity $e$, such
228: that $d = a(1 - e^2)/(1 + e\cos \phi_p)$. The direction to the
229: observer from the center of the star is $(\theta_o, \phi_o)$, so that
230: the conventional orbital inclination is $I = \pi - \theta_o$.
231:
232: We imagine that the gravity of the companion raises nearly symmetrical
233: tidal bulges on opposite sides of the star that rotate at the orbital
234: frequency. A rough measure of both the height of the tides relative
235: to the unperturbed stellar radius and the fractional amplitude of the
236: ellipsoidal variability is given by the ratio of the tidal
237: acceleration to the star's surface gravity:
238: %
239: \be\label{eq:tidescale}
240: \pert \equiv \frac{M_p}{M}\left( \frac{R}{a} \right)^3 \sim
241: 10^{-5}\, \frac{M_p}{M_J} \frac{M_\odot}{M}
242: \left( \frac{P_*}{2.8\ {\rm hr}}\frac{1\ {\rm day}}{P_{\rm orb}} \right)^2
243: ~,
244: \ee
245: %
246: where $M_J \simeq 10^{-3}\msun$ is the mass of Jupiter,
247: $P_*=2\pi(R^3/GM)^{1/2}=2.8\,[(R/R_\odot)^{3}(M_\odot/M)]^{1/2}\,{\rm
248: hr}$ is the dynamical time of the star. For main-sequence stars
249: with $R/R_\odot \simeq M/M_\odot$, we see that $\pert \propto M_p M
250: P_{\rm orb}^{-2}$. The maximum value of $\pert$ is attained when the
251: companion fills its Roche lobe at an orbital separation of $a \simeq 2
252: R_p(M/M_p)^{1/3}$, which gives
253: %
254: \be
255: \pert_{\rm max} &\simeq&
256: \left(\frac{M_p}{M}\right)^2
257: \left(\frac{R}{2R_p}\right)^3 \nonumber \\
258: &\simeq&
259: 10^{-4}\,\left(\frac{M_p}{M_J}\right)^2
260: \frac{M}{M_\odot}
261: \left(\frac{0.1\rsun}{R_p}\right)^3
262: ~,
263: \ee
264: %
265: where we have applied a fixed value of $R_p = 0.1\rsun$, appropriate
266: for both giant planets and old brown dwarfs. Note that $\pert_{\rm
267: max}\sim 1$ for massive brown dwarfs ($M_p/M_J \sim 80$).
268: Hereafter, we consider only cases with $\pert \ll 1$.
269:
270: For orbital periods as short as $\simeq$1 day, tidal torques on the
271: star from a planetary companion are rather ineffective at altering the
272: stellar rotation rate \citep[e.g.,][]{Rasio1996}. Therefore, as
273: already mentioned in \S~\ref{sec:intro}, we should not generally
274: expect the star to rotate synchronously with the orbit, and so there
275: is no frame in which the star appears static. This holds when the
276: orbit is circular, and is obviously true when the there is a finite
277: eccentricity. In fact, $\simeq$30\% of the known
278: exoplanets\footnote{\url{http://vo.obspm.fr/exoplanetes/encyclo/encycl.html}}
279: with $P_{\rm orb} < 10$\,days have eccentricities of $>$0.1. Small
280: variable distortions of the star from its equilibrium state, due to a
281: combination of asynchronous rotation and orbital eccentricity, should
282: be viewed as waves excited by the tidal force of the companion. Our
283: task is to study such tidally forced stellar oscillations in the
284: linear domain in order to understand the corresponding lightcurves.
285:
286: In order to greatly simplify the mathematical description of the
287: stellar oscillations, we assume that the star is nonrotating in the
288: inertial frame. When the stellar rotation frequency is nonzero, but
289: much smaller than the tidal forcing frequency, the effect of rotation
290: is to introduce fine structure into the oscillation frequency
291: spectrum, and cause the oscillation eigenfunctions to be slightly
292: modified as a result of the Coriolis force \citep[for a
293: discussion, see][]{Unno1989}. Tidal pumping of a slowly rotating star
294: by an orbiting companion has a dominant period of $P_{\rm orb}/2$---a
295: few days in the cases of interest. By contrast, single solar-type
296: stars with ages $>$1\,Gyr tend to have rotation periods of $>$10 days
297: \citep[e.g.,][]{Skumanich1972,Pace2004}; the Sun has an equatorial
298: rotation period of $\simeq$25 days. Slowly rotating stars with masses
299: of $\simeq$$1\msun$ are prime targets for {\em Kepler}, since they
300: exhibit low intrinsic variability. Based on this selection effect, and
301: the inability of tidal torques to spin up the star, our assumption of
302: vanishing stellar rotation seems generally justified.
303:
304: The general framework for calculating the measurable flux modulations
305: associated with ellipsoidal stellar oscillations is as follows. We
306: consider small perturbations to a spherical, nonrotating background
307: stellar model, such that fluid elements at equilibrium position
308: $\vec{x}$ are displaced in a Lagrangian fashion to position $\vec{x} +
309: \xivec$. Variations in the measured flux from an oscillating star
310: arise from two physically distinct contributions
311: \citep[e.g.,][]{Dziembowski1977}: (1) changes in the shape of the star
312: due to radial fluid displacements $\xi_r=\xivec\cdot \vec{e}_r$, where
313: $\vec{e}_r$ is the radial unit vector, and (2) hot and cold spots
314: generated by local Lagrangian perturbations $\Delta F$ to the heat
315: flux. Our main task in \S\S~\ref{sec:eqtide} and \ref{sec:nonad} is
316: to compute $\xi_r$ and $\Delta F$ according to the relevant physics.
317:
318: Given the dependences of $\xi_r$ and $\Delta F$ on $(r,\theta,\phi)$,
319: it is straightforward to compute the time varying component of the
320: measured flux. The flux\footnote{Our calculations concern the
321: bolometric flux, although is relatively straightforward to modify
322: the analysis for narrow-band measurements.} received from a star at
323: distance $D$ is \citep[e.g.,][]{Robinson1982}
324: %
325: \be
326: \fdisk & = &
327: \frac{1}{D^2} \int dS\
328: \vec{n} \cdot \vec{n}_o \ F \ h(\vec{n} \cdot \vec{n}_o)
329: \ee
330: %
331: where $dS$ is an area element at the stellar photosphere, $F$ is the
332: net flux of radiation out of the surface element, $h$ is the
333: limb-darkening function, $\vec{n}$ and $\vec{n}_o$ are unit vectors
334: normal to the surface and toward the observer, respectively, and the
335: integration is over the visible stellar disk. Vertical displacement
336: at the surface yields changes in $\fdisk$ through changes in surface
337: area and $\vec{n}\cdot \vec{n}_o$. Following \citet{Dziembowski1977},
338: we expand $\xi_r$ and $\Delta F$ in spherical harmonics,
339: %
340: \be
341: \xi_r(r,\theta,\phi,t) & = & \sum_{\ell = 0}^\infty \sum_{m = -\ell}^\ell
342: \xi_{r, \ell m}(r,t) Y_{\ell m}(\theta,\phi)~, \\
343: \Delta F(r,\theta,\phi,t) & = & \sum_{\ell = 0}^\infty \sum_{m = -\ell}^\ell
344: \Delta F_{\ell m}(r,t) Y_{\ell m}(\theta,\phi)~,
345: \ee
346: %
347: and carry out the appropriate linear expansions to obtain the
348: fractional variability
349: %
350: \be\label{eq:fdiskgen}
351: \frac{\delta\fdisk}{\fdisk} =
352: \sum_{\ell=0}^\infty
353: \left[ (2b_\ell - c_\ell) \frac{\xi_{r,\ell}^o}{R} +
354: b_\ell \frac{\Delta F_\ell^o}{F} \right]~.
355: \label{eq:dflux}
356: \ee
357: %
358: Here $\xi_{r, \ell}^o$ and $\Delta F_\ell^o$ are components evaluated
359: at the surface ($r = R$) and in the direction of the observer:
360: %
361: \be\label{eq:xirdelf_obs}
362: \frac{\xi_{r, \ell}^o}{R} & = &
363: \sum_{m = -\ell}^\ell \frac{\xi_{r, \ell m}(R,t)}{R}
364: Y_{\ell m}(\theta_o,\phi_o)~, \\
365: \frac{\Delta F_\ell^o}{F} & = &
366: \sum_{m = -\ell}^\ell \frac{\Delta F_{\ell m}(R,t)}{F}
367: Y_{\ell m}(\theta_o,\phi_o)~.
368: \ee
369: %
370: The terms $b_\ell$ and $c_\ell$ are given by
371: %
372: \be
373: b_\ell =
374: \int_0^1 d\mu \mu\, P_\ell\, h~,~~
375: c_\ell =
376: \int_0^1 d\mu (1-\mu^2)\frac{dP_\ell}{d\mu}
377: \left( h + \mu \frac{dh}{d\mu} \right)\ ,
378: \ee
379: %
380: where $\mu = \vec{n}\cdot\vec{n}_o$, the $P_\ell(\mu)$ are ordinary
381: Legendre polynomials, and $h(\mu)$ is normalized such that $\int_0^1
382: d\mu \mu\,h = 1$. The linear limb-darkening function is
383: %
384: \be\label{eq:limbdark}
385: h(\mu) = \frac{6}{(3-\gamma)}\big[ 1 - \gamma(1 - \mu)\big] ~;
386: \ee
387: %
388: more general nonlinear functions of $\mu$ \citep[e.g.,][]{Claret2000}
389: will not be considered here. The classical Eddington limb-darkening
390: function is $h = 1 + 3\mu/2$ \citep[$\gamma = 3/5$;
391: e.g.,][]{Mihalas1970}. Table~\ref{tab:ldterms} shows shows functional
392: forms and particular values of $b_\ell$ and $c_\ell$ for $\ell = 2$
393: and 3.
394:
395: % -----------------------------------------------------------
396:
397: \section{EQUILIBRIUM TIDE}
398: \label{sec:eqtide}
399:
400: Vertical displacement of the stellar surface is often accurately
401: modeled by assuming that the tidally perturbed fluid remains in
402: hydrostatic balance. The cause and magnitude of the surface flux
403: perturbation is a more complicated affair. In this section, we apply
404: a simple parameterization of the flux perturbation and obtain a
405: complete set of formulae for computing the ellipsoidal lightcurve.
406: Subsequent sections provide more detailed calculations. In
407: particular, we show in \S~\ref{sec:thickconv} that stars with deep
408: convective envelopes (the majority of {\em Kepler} targets) have
409: surface flux variations that conform to the equilibrium-tide
410: approximation.
411:
412: %
413: \input{tab1.tex}
414: %
415:
416: When the tidal forces on the stellar fluid change sufficiently slowly,
417: the star can stay very nearly in hydrostatic equilibrium. If the net
418: acceleration required to balance the pressure gradient is derivable
419: from a potential, then equilibrium implies that a fluid element
420: remains on an equipotential surface. Since we neglect stellar
421: rotation, there is no centrifugal force, and the total potential is
422: the sum of the gravitational potential $\varphi$ from the spherical
423: background stellar model and the perturbing tidal potential $U \sim
424: \pert\varphi \ll \varphi$. For our analytic work, we neglect the
425: modification of $\varphi$ due to the tide. In general, the Eulerian
426: variation $\delta\varphi$ should be added to $U$, as we do in our
427: numerical models (see \S~\ref{sec:numerical} and the Appendix); we
428: find that $|\delta\varphi|/|U| \sim 10^{-2}$.
429:
430: In the absence of tidal forces, a given fluid element sits at
431: equilibrium position $\vec{x}$ with total potential
432: $\varphi(\vec{x})$. Gentle inclusion of the tidal potential causes
433: the fluid element to move to position $\vec{x} + \xivec$ while
434: preserving the value of the total potential. This is expressed
435: mathematically by
436: %
437: \be
438: \varphi({\vec{x}}) & = &
439: \varphi(\vec{x} + \xivec) + U(\vec{x} + \xivec,t) \nonumber \\
440: & = & \varphi(\vec{x}) + \xivec\cdot\grad\varphi + U(\vec{x},t) +
441: {\cal O}(\xi^2, \xi U)~. \ee
442: %
443: We see that $\xivec\cdot\grad\varphi = \xi_r g$, where $g = GM_r/r^2$
444: is the background gravitational acceleration at mass coordinate $M_r$.
445: To first order, the radial displacement of the equilibrium tide is
446: \citep[see also][]{Goldreich1989}
447: %
448: \be\label{eq:eqtide}
449: \xi_r(\vec{x},t) & \simeq & - U(\vec{x},t)/g~,
450: \ee
451: %
452: which tells us the geometry of the star as a function of time.
453:
454: The tidal potential within the star can be expanded as
455: %
456: \be\label{eq:U}
457: U(r,\theta,\phi,t) = - \frac{GM_p}{d}
458: \sum_{\ell = 2}^\infty
459: \left(\frac{r}{d}\right)^{\ell} P_\ell(\cos\psi)~,
460: \ee
461: %
462: where $\cos\psi = \sin\theta \cos(\phi_p-\phi)$. There is no $\ell =
463: 1$ term, since this would give the acceleration of the star's center
464: of mass, which is already incorporated into the orbital dynamics. The
465: angular expansion of $\xi_r$ follows immediately from
466: eq.~(\ref{eq:eqtide}):
467: %
468: \be\label{eq:xireqtide}
469: \frac{\xi_r(r,\theta,\phi,t)}{r} = \frac{M_p}{M_r}
470: \sum_{\ell = 2}^\infty
471: \left(\frac{r}{d}\right)^{\ell+1} P_\ell(\cos\psi)~.
472: \ee
473: %
474: In order to express $U$ and $\xi_r$ in spherical harmonics, we utilize
475: the addition theorem,
476: %
477: \be\label{eq:legendre}
478: P_\ell(\cos\psi) =
479: \frac{4\pi}{2\ell+1}
480: \sum_{m = -\ell}^\ell
481: Y_{\ell m}^*(\pi/2,\phi_p)\
482: Y_{\ell m}(\theta,\phi)~,
483: \ee
484: %
485: where ``$*$'' denotes the complex conjugate. Note that $Y_{\ell
486: m}(\pi/2,\phi_p)$ is nonzero only when $\ell -m$ is even. For the
487: dominant $\ell = 2$ components of $U$ and $\xi_r$, the surface values
488: of $U/\varphi$ and $\xi_r/R$ are $\sim$$\pert$, as expected.
489:
490: From eqs.~(\ref{eq:xirdelf_obs}), (\ref{eq:xireqtide}), and
491: (\ref{eq:legendre}), the components $\xi_{r,\ell}^o/R$ of the surface
492: radial displacement toward the observer are immediately apparent. As
493: we will see in \S~\ref{sec:nonad}, the computation of $\Delta F/F$ is,
494: in general, rather technical. However, in the special case where the
495: stellar fluid responds adiabatically to a slowly varying tidal
496: potential, $\Delta F_\ell/F$ varies in phase with and in proportion to
497: $\xi_{r,\ell}/r$ in the linear approximation of the equilibrium tide.
498: Making this assumption, we write $\Delta F_\ell/F= -\lambda_\ell
499: \xi_{r,\ell}/R$ at the surface, where the $\lambda_\ell$ are real
500: constants that depend on the stellar structure (see
501: \S~\ref{sec:conv}). We will see in \S~\ref{sec:vonzeipel} that
502: $\lambda_\ell = \ell + 2$ is a good first guess for radiative stars,
503: and so we might generally expect $\lambda_\ell$ to be positive and
504: ${\cal O}(\ell)$.
505:
506: We now have the ingredients for the fractional variability
507: (eq.~[\ref{eq:fdiskgen}]), and we obtain
508: %
509: \be\label{eq:fdiskeq}
510: \frac{\delta \fdisk}{\fdisk} =
511: \pert \sum_{\ell = 2}^\infty
512: \Bigg(\frac{R}{a}\Bigg)^{\ell - 2}
513: \Bigg(\frac{a}{d}\Bigg)^{\ell + 1}
514: f_\ell \ P_\ell (\cos\psi_o)~,
515: \ee
516: %
517: where $f_\ell = (2-\lambda_\ell)b_\ell - c_\ell$, and $\cos\psi_o =
518: \sin\theta_o\cos(\phi_p - \phi_o)$. The $\ell = 2$ and 3 Legendre
519: polynomials can be expanded as
520: %
521: \begin{multline}\label{eq:P2}
522: P_2(\cos\psi_o) =
523: \frac{1}{4}
524: \big[
525: -(3\cos^2 I - 1) \\
526: + 3\sin^2 I\ \cos 2(\phi_p - \phi_o)
527: \big]~,
528: \end{multline}
529: %
530: %
531: \begin{multline}\label{eq:P3}
532: P_3(\cos\psi_o) =
533: \frac{1}{8}\sin I
534: \big[
535: -3(5\cos^2 I - 1)\cos (\phi_p - \phi_o) \\
536: + 5\sin^2 I\ \cos 3(\phi_p - \phi_o)
537: \big]~,
538: \end{multline}
539: %
540: where we have substituted $\theta_o = \pi - I$. The Eddington
541: limb-darkening formula gives (see Table~\ref{tab:ldterms})
542: %
543: \be
544: f_2 = -\frac{13}{10}\left(1 + \frac{\lambda_2}{4}\right)~,~~
545: f_3 = -\frac{5}{8}\left(1 + \frac{\lambda_3}{10}\right)~.
546: \ee
547: %
548: It is important to note that $f_2<0$ when $\lambda_2 \geq 0$ (see
549: below).
550:
551: In eq.~(\ref{eq:fdiskeq}), the orbital dynamics are described by the
552: evolution of $d$ and $\phi_p$ (see \S~\ref{sec:prelim}). For a
553: circular orbit, we have $d = a$ and $\phi_p = \Omega t$, where $\Omega
554: = 2\pi/P_{\rm orb}$, and $t$ is the time since periastron (modulo
555: $P_{\rm orb}$). Example lightcurves with $e = 0$, $\gamma = 3/5$,
556: $\lambda_\ell=0$, and $I = \pi/2$ are shown in
557: Fig.~\ref{fig:eqtide_shape} for $a/R = \{2, 4, 8, 16\}$. When $R/a
558: \ll 1$, the $\ell = 2$ piece of $\delta\fdisk/\fdisk$ is a good
559: approximation, and the temporal flux variation approaches a pure
560: cosine with angular frequency $2\Omega$ (see eq.~[\ref{eq:P2}]).
561: Because $f_2 < 0$, the dominant $\ell = 2$ component of the
562: ellipsoidal variability has {\em minimum} light when tidal bulge is
563: aligned with the direction to the observer. As $R/a$ increases, so
564: does the importance of $\ell >2$ terms and their extra harmonic
565: content, as seen in eq.~(\ref{eq:P3}) and Fig.~\ref{fig:eqtide_shape}.
566:
567: Additional harmonics in $\delta\fdisk/\fdisk$ also result from a
568: finite eccentricity. At the ${\cal O}(e)$ level, signals with
569: frequencies $\Omega$ and $3\Omega$, and amplitudes of $\sim$$\pert e$,
570: are present in the $\ell = 2$ component of $\delta\fdisk/\fdisk$,
571: which compete with the $\ell = 3$ piece when $e \sim R/a$. Notice
572: that when $e > 0$ the flux is variable even when the orbit is viewed
573: face-on ($I = 0$ or $\pi$), by virtue of changes in $d^{-3} = 1+3
574: e\cos\Omega t + {\cal O}(e^2)$. For $I = 0$, we see that
575: $P_3(\cos\psi_o)$ vanishes, leaving the largest contribution
576: $\delta\fdisk/\fdisk \simeq -1.5 \pert e f_2\cos(\Omega t)$.
577:
578:
579: %
580: \begin{figure}
581: \centerline{\epsfig{file = fig1.eps,angle = 0,width = \linewidth}}
582: \caption{ Disk-averaged flux variation (eq.~[\ref{eq:fdiskeq}]) for an
583: edge-on circular orbit under the equilibrium-tide approximation
584: (eqs.~[\ref{eq:dflux}] and [\ref{eq:xireqtide}]) with $\Delta F/F =
585: 0$ at the surface. The four curves correspond to $a/R=2$ (black),
586: $4$ (red), $8$ (blue) and $16$ (green). In order to compare the
587: shapes of the curves, $\delta\fdisk/\fdisk$ has been multiplied by
588: $(a/R)^3(M/M_p)$. As $a/R$ increases, higher harmonics decrease in
589: strength and the lightcurve approaches a pure cosine with frequency
590: $2/P_{\rm orb}$. The tidal bulge closest to the companion points
591: toward the observer at integer values of $t/P_{\rm orb}$.}
592: \label{fig:eqtide_shape}
593: \end{figure}
594: %
595:
596: % -----------------------------------------------------------
597:
598: \section{An Aside on von Zeipel's Theorem}
599: \label{sec:vonzeipel}
600:
601: Our equilibrium calculation in the last section used the simple
602: prescription $\Delta F_\ell/F = - \lambda_\ell \xi_{r,\ell}/R$. There
603: remains the question of what physics determines $\Delta F/F$. A
604: common practice in empirical studies of close eclipsing
605: binaries---systems that tend to be nearly in tidal equilibrium---is to
606: use some variant of the \citet{vonZeipel1924} theorem, which was
607: originally formulated for purely radiative, strictly hydrostatic
608: stars. In equilibrium, all the thermodynamic variables depend only on
609: the local value of the total potential $\Phi$. Thus, the radiative
610: flux can be written as \citep[e.g.,][]{Hansen1994}
611: %
612: \be\label{eq:vonzeipel}
613: \vec{F} & \propto & - \frac{1}{\kappa \rho} \frac{dT^4}{d\Phi}
614: \grad \Phi~,
615: \ee
616: %
617: where $\rho$ is the mass density, $T$ is the effective temperature,
618: and $\kappa(\rho, T)$ is the opacity. Equation~(\ref{eq:vonzeipel})
619: is the essence of von Zeipel's theorem, which says that the magnitude
620: $F$ of the radiative flux is proportional to the magnitude of the net
621: acceleration $A = |\grad\Phi|$. When $\Phi = \varphi + U$ (see
622: \S~\ref{sec:eqtide}), we obtain $A = g + \partial U/\partial r +
623: {\cal O}(\xi^2)$, so that the Lagrangian flux
624: perturbation about equilibrium is
625: %
626: \be\label{eq:fluxvar}
627: \frac{\Delta F}{F} = \frac{\Delta A}{g} =
628: \frac{\Delta g}{g} + \frac{1}{g}\frac{\partial U}{\partial r}~,
629: \ee
630: %
631: where $\Delta g/g = -2\xi_r/r$, due to the change in radius at
632: approximately constant enclosed mass. Substituting the
633: equilibrium-tide result $U = -\xi_r g$ into eq.~(\ref{eq:fluxvar}), we
634: obtain the compact expression $\Delta F/F = -\partial \xi_r/\partial
635: r$. Using eq.~(\ref{eq:xireqtide}), we find
636: %
637: \be\label{eq:fluxvar2}
638: \frac{\Delta F_\ell}{F} = -(\ell + 2) \frac{\xi_{r,\ell}}{r}~,
639: \ee
640: %
641: from which we identify $\lambda_\ell = \ell + 2$.
642:
643: Although the application of von Zeipel's theorem is instructive, the
644: underlying physical assumptions are inaccurate for slowly rotating
645: main-sequence stars of mass 1.0--$1.6\msun$ with tidal forcing
646: periods of days. We are now led to investigate the general problem
647: of forced nonadiabatic stellar oscillations.
648:
649: % -----------------------------------------------------------
650:
651: \section{FORCED NONADIABATIC OSCILLATIONS}
652: \label{sec:nonad}
653:
654: % -----------------------------------------------------------
655:
656: %
657: \begin{figure}
658: \centerline{\epsfig{file = fig2.eps,angle = 0,width = \linewidth}}
659: \caption{Important oscillations frequencies and time scales as a
660: function of pressure for a $1\msun$ main-sequence star. The four
661: curves show the Brunt-V\"ais\"all\"a frequency $N$ (black), Lamb
662: frequency $S_l$ (red; $\ell = 2$ is shown), inverse thermal time
663: $t_{\rm th}^{-1}$ (blue), and inverse eddy turnover time
664: $t^{-1}_{\rm ed}$ (green). Large, real values of $N$ occur in the
665: radiative core and very near the photosphere, while $N^2<0$ in the
666: convective envelope. Gravity waves propagate only where the angular
667: frequency is below both $N$ and $L_\ell$. The two horizontal lines
668: delimit the range of tidal forcing frequencies of interest here.}
669: \label{fig:prop_1.0msun}
670: \end{figure}
671: %
672:
673:
674: The equilibrium analysis ignores fluid inertia and the excitation of
675: the star's natural oscillation modes. While this assumption may be
676: valid near the surface of the star, it does not hold deeper in the
677: interior. Gravity waves ($g$-modes; restored by buoyancy) can
678: propagate in the radiative interiors of Sun-like stars with a range of
679: oscillation periods that includes the tidal forcing periods of
680: interest ($\la$3 days). Tidal forcing of radiative regions may
681: produce substantial deviations from hydrostatic balance, as well as
682: large surface amplitudes of $\Delta F/F$, in particular if resonant
683: oscillations are excited. This is especially relevant for
684: main-sequence stars of mass $M \ga 1.4$--$1.5\msun$ with mainly
685: radiative envelopes. Less massive stars ($M \la 1.3$--$1.4\msun$)
686: have rather deep convective envelopes that can block information about
687: the dynamic interior from being conveyed to the surface. Here we
688: investigate each of these regimes with both analytic estimates and
689: numerical models of oscillating stars.
690:
691: Our calculations employ realistic models of 0.9--$1.6\msun$
692: main-sequence stars, constructed with the EZ stellar evolution code
693: \citep{Paxton2004}, a distilled and rewritten version of the program
694: originally created by Peter Eggleton. We adopt Solar metallicity and
695: a convective mixing length of 1.6 times the pressure scale height.
696: All stars are evolved to an age when the core hydrogen abundance has
697: the Solar value of $X_H=0.35$. Models with 199 radial grid points are
698: interpolated to yield $\ga$$10^4$ points in which the $g$-mode radial
699: wavelength is well resolved in the core.
700:
701: %
702: \begin{figure}
703: \centerline{\epsfig{file = fig3.eps,angle = 0,width = \linewidth}}
704: \caption{ Same as Fig.~\ref{fig:prop_1.0msun}, but for a $1.6M_\odot$
705: main-sequence star. Note two geometrically thin, relatively
706: inefficient ($t_{\rm th} \sim t_{\rm ed}$) convection zones near the
707: surface. The spike in $N$ near the center is at the edge of the
708: convective core, and signals a steep gradient in the mean molecular
709: weight.}
710: \label{fig:prop_1.6msun}
711: \end{figure}
712: %
713:
714:
715: Figures~\ref{fig:prop_1.0msun} and \ref{fig:prop_1.6msun} illustrate
716: some of the differences between $1\msun$ and $1.6\msun$ stars, and
717: serve to introduce several important physical quantities used in the
718: remainder of this section. The Lamb frequency,
719: %
720: \be\label{eq:Lamb}
721: S_\ell = [\ell(\ell + 1)]^{1/2}\frac{c_s}{r}~,
722: \ee
723: %
724: is the inverse of the horizontal sound-crossing time scale, where $c_s$
725: is the sound speed, and $[\ell(\ell + 1)]^{1/2}/r \equiv k_h$ is the
726: horizontal wavenumber of the oscillation. For fixed chemical
727: composition, the squared Brunt-V\"ais\"all\"a frequency is
728: %
729: \be\label{eq:Brunt}
730: N^2 \simeq \frac{g}{H_p}(\nabla _{\rm ad} - \nabla)~,
731: \ee
732: %
733: where $H_p = -(d\ln p/dr)^{-1} = p/(\rho g)$ is the pressure scale height,
734: and $\nabla = d\ln T/d \ln p$ is the temperature gradient\footnote{Do
735: not confuse the temperature gradient $\nabla$ with the spatial
736: gradient $\grad$ used in \S~\ref{sec:eqtide}.} ($\nabla_{\rm ad}$
737: is the adiabatic value). Radiative regions have $\nabla_{\rm ad} - \nabla
738: > 0$ ($N^2 > 0$), and $N$ represents the frequency of buoyancy
739: oscillations. In convection zones, $\nabla_{\rm ad} - \nabla < 0$ and
740: $N^2 < 0$, indicating that $g$-modes are evanescent. When $N^2 < 0$,
741: the time scale
742: %
743: \be\label{eq:teddy}
744: t_{\rm ed} \sim |N|^{-1}~,
745: \ee
746: %
747: approximates the turnover time of convective motions (for details and
748: modifications for radiative losses, see, e.g., Kippenhahn \& Weigert
749: 1990). A shell of radius $r$, thickness $H_p$ (size of the largest
750: convective eddies), and radiative luminosity $L$ cools on the thermal
751: time scale
752: %
753: \be\label{eq:tth}
754: t_{\rm th} = \frac{4\pi r^2 H_p \rho C_p T}{L}~,
755: \ee
756: %
757: where $C_p$ is the specific heat at constant pressure.
758:
759: The $1\msun$ model (Fig.~\ref{fig:prop_1.0msun}) has one deep
760: convection zone with $t_{\rm ed} \ll t_{\rm th}$ over most of the
761: region, indicating that convection very efficiently transports energy
762: and causes the zone to be essentially isentropic. By consrast, the
763: $1.6\msun$ star (Fig.~\ref{fig:prop_1.6msun}) has two thin surface
764: convection zones with $t_{\rm ed} \sim t_{\rm th}$, and thus the
765: radiative and convective fluxes are comparable. Gravity waves with
766: frequency $\omega$ propagate only in radiative regions where $\omega <
767: N$ and $\omega < S_\ell$. For the $1\msun$ star, heat and entropy
768: generated by $g$-modes in the radiative interior may be strongly
769: mitigated owing to the long thermal time at the base of the deep
770: convection zone. On the other hand, $g$-modes in a $1.6\msun$ star
771: can propagate very near the surface, producing qualitatively different
772: results.
773:
774: We now go on to elucidate the physics of the flux perturbations. All
775: the analytic and numerical work that follows assumes that the tidal
776: potential has the generic form $U \propto r^\ell Y_{\ell m}(\theta,
777: \phi) \exp (-i\omega t)$ with forcing frequency $\omega$.
778:
779: %
780: \begin{figure}
781: \centerline{\epsfig{file = fig4.eps,angle = 0,width = \linewidth}}
782: \caption{Eddy turnover time at base of the convective envelope versus
783: stellar age for a range of stellar masses.}
784: \label{fig:ted_vs_age}
785: \end{figure}
786: %
787:
788: % -----------------------------------------------------------
789:
790: \subsection{ Heat Transfer in a Convective Envelope }
791: \label{sec:conv}
792:
793: Calculation of the perturbed convective flux in oscillating stars is a
794: thorny issue. For the purposes of our study, we argue for an
795: especially simple treatment that draws from previous work on this
796: subject. Specifically, we modify the prescription of Brickhill (1983,
797: 1990; see also Goldreich \& Wu 1999a,b), which was originally applied
798: to white-dwarf pulsations, into a form appropriate for the tidal flow
799: problem.
800:
801: In the mixing-length theory of convection, heat is transported by
802: eddies with a spectrum of sizes $l\la H_p$, speeds $v_l$, and turnover
803: times $t_{\rm ed}(l) = l/v_l$. The Kolmogorov scalings for turbulent
804: motions give $v_l \propto l^{1/3}$, $t_{\rm ed} \propto l^{2/3}$, and
805: an energy density per unit mixing length interval $\propto$$l^{-1/3}$.
806: We see that in the unperturbed star most of the convective energy flux
807: ($\propto$$v_l^3$ at scale $l$) is carried by the largest eddies ($l
808: \sim H_p$). Convection efficiently transports energy when the
809: radiative thermal time scale associated with the dominant eddies is
810: much longer than $t_{\rm ed}$. Alternatively, efficient convection
811: implies that the gradient of the specific entropy $s$ is small; i.e.,
812: $d\ln s/d \ln p \ll 1$. If all the convective energy flux $F$ is
813: carried by eddies with mixing length $l$, the flux and entropy
814: gradient are related by \citep[e.g.,][]{Kippenhahn1990}
815: %
816: \be\label{eq:dsdlnp}
817: \frac{1}{C_p}\frac{ds}{d\ln p}(l) & = & \left( \nabla -
818: \nabla_{\rm ad} \right) \sim \left[ \frac{F}{p c_s} \left(
819: \frac{H_p}{l} \right)^2 \right]^{2/3}~.
820: \ee
821: %
822: Efficient convection enforces $\nabla - \nabla_{\rm ad}\ll 1$, which
823: implies $d\ln s/d \ln p \ll 1$, since $s \ga C_p$ in the convective
824: regions of our background models.
825:
826: %
827: \begin{figure}
828: \centerline{\epsfig{file = fig5.eps,angle = 0,width = \linewidth}}
829: \caption{Thermal time at the base of the convection envelope versus
830: stellar age for a range of stellar masses.}
831: \label{fig:tth_vs_age}
832: \end{figure}
833: %
834:
835: Gravity waves with the tidal forcing frequency $\omega$ are excited in
836: the radiative region below the convection zone. Convective eddies can
837: transport heat during a forcing period only if $t_{\rm ed}< 2\pi/\omega$
838: \citep[e.g.,][]{Brickhill1990,Goldreich1999b}. Inspection of Fig.~
839: \ref{fig:prop_1.0msun} shows that in the $1\msun$ model, the largest
840: eddies have $t_{\rm ed} \simeq 20 (p/p_{\rm bcz})^{0.5}$\,days, where
841: $p_{\rm bcz} \simeq 10^{13.5}\ {\rm dyne\ cm^{-2}}$ is the pressure at
842: the base of convection zone. Using the Kolmogorov scaling, the
843: ``resonant'' length $l_{\rm res}$ for which $\omega t_{\rm ed}/2\pi =
844: 1$ is
845: %
846: \be\label{eq:lres}
847: \frac{l_{\rm res}}{H_p} &
848: \sim & 10^{-2}\ \left( \frac{2\pi/\omega}{1\ {\rm day}} \right)^{3/2}
849: \left( \frac{p_{\rm bcz}}{p} \right)^{3/4}~,
850: \ee
851: %
852: which is $>$1 for all periods $2\pi/\omega > 1$\,day when $p \la
853: 10^{12} \ {\rm dyne\ cm^{-2}}$, which still encompasses much of the
854: convection zone. Now imagine the situation where all the
855: convective flux is carried by eddies of size $\la$$l_{\rm res}$. The
856: entropy gradient for this range of mixing lengths is
857: %
858: \be\label{eq:entropyres}
859: \frac{1}{C_p} \frac{ds}{d\ln p}(l_{\rm res})
860: & \sim & 10^{-3}
861: \left(\frac{2\pi/\omega}{1\ {\rm day}}\right)^{-1}
862: \left( \frac{p}{p_{\rm bcz}} \right)~,
863: \ee
864: %
865: where we have adopted $F/pc_s \sim 10^{-8}$ at the base of the
866: convection zone, as indicated by our $1\msun$ stellar model.
867:
868: These arguments suggest that convection is efficient in a $1\msun$
869: star at the forcing periods of interest even if small ``resonant''
870: eddies carry all the energy flux near the base of the convection zone.
871: At larger radii, but not too near the photosphere, convection is both
872: efficient and rapid ($\omega t_{\rm ed}/2\pi < 1$) over the full
873: spectrum of eddies. Rapid convection on all scales $l \la H_p$
874: enforces isentropy in the convection zone, such that $s$ and its
875: Lagrangian perturbation $\Delta s$ are nearly constant, as in the
876: \citet{Brickhill1983,Brickhill1990} picture. While convection at the
877: base is rapid only on small scales, it is still highly efficient,
878: which yields $s \simeq \textrm{constant}$ and further indicates that
879: $\Delta s/C_p$ is small in magnitude, as we demonstrate in
880: \S~\ref{sec:thickconv}.
881:
882: As the stellar mass increases, the convection zone thins and
883: $t_{\rm ed}$ at the base decreases (see Figs.~\ref{fig:prop_1.6msun}
884: and \ref{fig:ted_vs_age}). Rapid convection holds over the
885: bulk of the convection zone for masses
886: $\ga$$1\msun$. However, the assumption that the convection is
887: efficient starts to break down at 1.4--$1.5\msun$, since $t_{\rm th}
888: \sim t_{\rm ed}$ at the base (see
889: Figs.~\ref{fig:prop_1.6msun} and \ref{fig:tth_vs_age}). For the full
890: range of stellar masses considered here, we assume that $s$ and
891: $\Delta s$ are constant in convection zones.
892:
893: % -----------------------------------------------------------
894:
895: \subsection{Analytic Result for Thick Convection Zones}
896: \label{sec:thickconv}
897:
898: In a fully convective star, the emergent luminosity is determined
899: entirely by the surface boundary conditions. Under our assumption
900: that $\Delta s$ is constant in the convection zone, the perturbed
901: luminosity is likewise a function only of the boundary conditions.
902: Stars of mass $\la$1.3--$1.4\msun$ have long thermal times ($\omega
903: t_{\rm th} \gg 1$) at the top of the interior radiative region (see
904: Fig. \ref{fig:tth_vs_age}), so that the flux perturbation $\Delta F$
905: is approximately the ``quasi-adiabatic'' value, derived by ignoring
906: $\Delta s \propto (\omega t_{\rm th})^{-1}$ in eq.~(\ref{eq:y5}). We
907: assume efficient convection continues to just below the photosphere.
908:
909: At the photosphere, we adopt the usual Stefan-Boltzmann relation, $F =
910: \sigma T^4$, and the hydrostatic condition, $p\kappa/A=2/3$, where $A$
911: is the total acceleration defined in \S~\ref{sec:vonzeipel}, and 2/3
912: is the photospheric optical depth. Taking the photosphere to define
913: the stellar surface, we compute the Lagrangian perturbations,
914: %
915: \be\label{eq:dphotf}
916: \frac{\Delta F}{F} = 4\frac{\Delta T}{T}~,
917: \ee
918: %
919: and
920: %
921: \be\label{eq:dphotp}
922: \frac{\Delta p}{p} - \frac{\Delta A}{g} + \frac{\Delta \kappa}{\kappa} = 0~.
923: \ee
924: %
925: Using $s$ and $p$ as our independent thermodynamic variables, we write
926: $\Delta \kappa/\kappa=\kappa_{\rm ad}\Delta p/p + \kappa_s \Delta
927: s/C_p$ and $\Delta T/T=\Delta s/C_p + \nabla_{\rm ad}\Delta p/p$.
928: In our numerical work (see \S~\ref{sec:numerical}), we self-consistently compute the perturbation $\Delta A$ to the effective surface gravity, in order to follow resonant oscillations, where the equilibrium-tide result fails. However, we are now addressing non-resonant forcing, for which
929: we use the equilibrium-tide approximation at the surface, giving
930: $\Delta A/g=-\partial \xi_r/\partial r$ (see \S~\ref{sec:vonzeipel}).
931: We now have
932: %
933: \be
934: \frac{\Delta p}{p} & = & - \left(\frac{\kappa_s \Delta s/C_p +
935: \partial\xi_r/\partial r}{1+\kappa_{\rm ad}} \right)~,
936: \ee
937: %
938: and upon substitution,
939: %
940: \be\label{eq:surfacedf}
941: \frac{\Delta F}{F} =
942: 4 \left( \frac{1 + \kappa_{\rm ad}- \nabla_{\rm ad} \kappa_s}{1 + \kappa_{\rm ad}} \right)
943: \frac{\Delta s}{C_p}
944: - \frac{4\nabla_{\rm ad}}{1 + \kappa_{\rm ad}} \frac{\partial\xi_r}{\partial r}~.
945: \ee
946: %
947: Equation (\ref{eq:surfacedf}) differs from \citet{Goldreich1999a} in
948: that we retain the gravity perturbation in eq.~(\ref{eq:dphotp}),
949: whereas they consider a constant-gravity atmosphere (and no tidal
950: perturbation). For $g$-modes in white dwarfs, the interesting region
951: is near the surface and the motion is mainly horizontal, so that
952: $\Delta A = \Delta g = 0$ is a good approximation. Since the
953: equilibrium tide has large vertical motions, the $\Delta A$ term must
954: be retained.
955:
956: The luminosity change across the convection zone is derived from the
957: entropy equation (eq.~[\ref{eq:y6}]). If we ignore horizontal flux
958: perturbations (set $\ell = 0$ in eq.~[\ref{eq:y6}]) and energy
959: generation, the equation for the luminosity perturbation $\Delta
960: L/L=2\xi_r/r + \Delta F/F$ is
961: %
962: \be\label{eq:entropy}
963: \frac{d(\Delta L/L)}{dM_r} & = & i\omega T \Delta s/L~.
964: \ee
965: %
966: Integrating over the convection zone with constant $\Delta s$, we
967: obtain
968: %
969: \be\label{eq:ldiff}
970: \frac{\Delta L_{\rm ph}}{L} - \frac{\Delta L_{\rm bcz}}{L} =
971: i\omega \Delta s \int_{\rm cz} dM_r T/L~,
972: \ee
973: %
974: where the subscript ``ph'' refers to the photosphere. We define
975: $t_{\rm cz} = C_{p,{\rm ph}} \int_{\rm cz} dM_r T/L$ to be the mean
976: thermal time of the convection zone, so that the right-hand side of
977: eq.~(\ref{eq:ldiff}) is $i\omega t_{\rm cz} \Delta s/C_{p,{\rm ph}}$.
978:
979: Figure~\ref{fig:tth_vs_age} shows that the thermal time at the base of
980: the convection zone (of order $t_{\rm cz}$) for $M \la 1.3\msun$ is
981: orders of magnitude longer than the forcing periods of 1--10\,days.
982: Insofar as $|\Delta L|/L \sim |\xi_r|/r$ at any location in the star
983: (i.e., if resonances are neglected), we see that
984: $(|\xi_r|/r)^{-1}|\Delta s|/C_p \sim (\omega t_{\rm cz})^{-1} \ll 1$
985: in stars with deep convective envelopes. In this limit,
986: eq.~(\ref{eq:surfacedf}) becomes
987: %
988: \be\label{eq:simplesurfacedf}
989: \frac{\Delta F}{F} \simeq
990: - \frac{4\nabla_{\rm ad}}{1 + \kappa_{\rm ad}}
991: \frac{\partial\xi_r}{\partial r}~.
992: \ee
993: %
994: If we had set $\Delta A = 0$, the amplitude of the photospheric flux
995: perturbation would have been $\sim$$|\Delta s|/C_p$ rather than the
996: much larger value $\sim$$|\xi_r|/R$.
997:
998: Photospheric flux perturbations in tidally forced solar-type stars
999: with thick convective envelopes arise mainly from changes in the local
1000: effective gravity. This statement is reminiscent of, but physically
1001: distinct from, von Zeipel's theorem (eqs.~[\ref{eq:fluxvar}] and
1002: [\ref{eq:fluxvar2}]). We have recovered our equilibrium-tide scaling,
1003: $\Delta F_\ell/F = -\lambda_\ell \xi_{r,\ell}/R$, where
1004: eq.~(\ref{eq:simplesurfacedf}) gives
1005: %
1006: \be\label{eq:lambda}
1007: \lambda_\ell = 4(\ell + 2)
1008: \frac{\nabla_{\rm ad}}{1 + \kappa_{\rm ad}}~.
1009: \ee
1010: %
1011: For $M = 1.0$--$1.4\msun$, we find $\lambda_2 \simeq 1.9$--1.1. These
1012: estimates neglect resonant excitation of $g$-modes, a point
1013: addressed in \S~\ref{sec:numerical}.
1014:
1015: % -----------------------------------------------------------
1016:
1017: \subsection{ Analytic Result for Radiative Envelopes }
1018: \label{sec:radiative}
1019:
1020: As the stellar mass increases beyond $1.4\msun$, the outer convective
1021: region thins and sits close to the surface, where $t_{\rm ed} \sim
1022: t_{\rm th}$. Figure~\ref{fig:prop_1.6msun} shows that the $1.6\msun$
1023: model has two thin, inefficient surface convection zones, as well as a
1024: convective core. Radiative energy transport is important throughout
1025: the envelopes of these more massive stars. We now consider the
1026: idealized case of a completely radiative envelope, and obtain an
1027: analytic approximation for $\Delta L/L$ at the surface.
1028:
1029: Near the surface of a radiative star, we have $H_p/r \ll 1$, $4\pi r^3
1030: \rho /M_r \ll 1$, and $\omega^2 r/g \ll 1$ for $2\pi/\omega =
1031: 1$--10\,days. Under these conditions, the quasi-adiabatic luminosity
1032: perturbation becomes \citep[e.g.,][]{Unno1989}
1033: %
1034: \be
1035: \frac{\Delta L_{\rm qad}}{L} \simeq
1036: - \zeta \frac{\Delta p}{p}
1037: + \frac{gk_h^2H_p}{\omega^2}
1038: \left( \frac{\nabla_{\rm ad}}{\nabla} - 1 \right)
1039: \left( \frac{\Delta p}{p}+
1040: \frac{\xi_r-\xi_{r,\rm eq}}{H_p} \right) \nonumber \\
1041: \label{eq:qaddl}
1042: \ee
1043: %
1044: where
1045: %
1046: \be
1047: \zeta=\kappa_{\rm ad} - 4\nabla_{\rm ad} + \frac{\nabla_{\rm ad}}{\nabla}
1048: - \frac{d\nabla_{\rm ad}}{d\ln T}~,
1049: \ee
1050: %
1051: and $\xi_{r,\rm eq}$ is the equilibrium-tide radial displacement
1052: (eq.~[\ref{eq:eqtide}]). Nonzero values of $\Delta p/p$ and $(\xi_r -
1053: \xi_{r,{\rm eq}})/H_p$ indicate deviations from hydrostatic
1054: equilibrium. Care must be taken with these terms, because the
1055: denominators $p$ and $H_p$ become very small close to the surface.
1056:
1057: With the help of the Appendix, we define the variables
1058: %
1059: \be
1060: \alpha & = & y_1-y_2+y_3=
1061: -\frac{H_p}{r}\frac{\Delta p}{p}~, \\
1062: \beta & = & y_2+\frac{U}{gr}=
1063: \frac{H_p}{r}\frac{\Delta p}{p} + \frac{\xi_r-\xi_{r,\rm eq}}{r}~,
1064: \ee
1065: %
1066: which satisfy the differential equations
1067: %
1068: \be\label{eq:dalphadr}
1069: \frac{d\alpha}{d r} & \simeq & - \frac{d\ln \rho}{d r} \alpha
1070: + \frac{gk_h^2}{\omega^2} \beta + (\ell+4)\frac{U}{gr^2}~, \\
1071: \frac{d\beta}{d r} & = & - \frac{N^2}{g} \alpha + \frac{\beta}{r}~.
1072: \ee
1073: %
1074: When $\omega^2 \ll gk_h^2 H_p$, these equations produce the $g$-mode
1075: dispersion relation $k_r^2=k_h^2N^2/\omega^2$ (in the limit
1076: $k_r^2/k_h^2 \ll 1$) for radial wavenumber $k_r$. For these
1077: propagating waves, the surface amplitudes of $\alpha$ and $\beta$ are
1078: determined at the core radiative-convective boundary, where $g$-modes
1079: are driven \citep[e.g.,][]{Goldreich1989}. On the other hand, when
1080: $\omega^2 \gg g k_h^2 H_p$, the $g$-modes are evanescent
1081: \citep[see][]{Unno1989} and we neglect the term $gk_h^2\beta/\omega^2$
1082: in eq.~(\ref{eq:dalphadr}). This limit yields the approximate
1083: solution $\alpha \simeq - (\ell+4)(H_p U/gR^2)$, or $\Delta p/p \simeq
1084: (4+\ell)U/gR$. In this case, $\Delta p/p$ is not small compared to
1085: the fractional fluid displacement, and thus the equilibrium-tide
1086: approximation loses validity.
1087:
1088: From our stellar models, we find that the evanescent regime
1089: corresponds to forcing periods of $\la$4--8\,days for $M =
1090: 1$--$1.6\msun$, most of the range of interest. The high-frequency
1091: limit of eq.~(\ref{eq:qaddl}) is
1092: %
1093: \be\label{eq:dlqadapprox}
1094: \frac{\Delta L_{\rm qad}}{L} \simeq -\zeta (\ell+ 4)\ \frac{U}{gR}~.
1095: \ee
1096: %
1097: This relation should be evaluated at the layer where $\omega t_{\rm
1098: th} \simeq 1$, above which the luminosity effectively ``freezes
1099: out.'' Figure \ref{fig:radiativeflux} shows the quasi-adiabic flux
1100: perturbation $\Delta F_{\rm qad}/F=\Delta L_{\rm qad}/L-2\xi_r/R$,
1101: evaluated where $\omega t_{\rm th}=1$, for a range of forcing periods
1102: and $M = 1.5$--$1.7\msun$. Note that $|\Delta F/F|$ can be an order of
1103: magnitude larger than $|U|/gR$, because of the rather large values of
1104: $|\zeta|(\ell + 4)$ for $\ell \geq 2$. Much larger perturbations are
1105: possible when $g$-modes are resonantly excited in a radiative star, as
1106: we discuss in the next section.
1107:
1108: We must point out that the quasi-adiabatic approximation is
1109: technically inappropriate when $\omega t_{\rm th} \sim 1$.
1110: Equation~(\ref{eq:dlqadapprox}) should be viewed as an estimate of the
1111: modulus of the luminosity perturbation at the surface. If, for
1112: instance, $|\Delta s|/C_p \ga |U|/gR$ where $\omega t_{\rm th} \sim
1113: 1$, then $\Delta L/L$ at the surface will have a substantial imaginary
1114: part (see eq.~[\ref{eq:entropy}]). This is what we find in the
1115: numerical calculations summarized in the next section.
1116:
1117: %
1118: \begin{figure}
1119: \centerline{\epsfig{file = fig6.eps,angle = 0,width = \linewidth}}
1120: \caption{ Ratio of surface Lagrangian flux perturbation $\Delta F/F$
1121: to equilibrium-tide displacement $-U/gR$ for a range of forcing
1122: periods in the limit where surface $g$-modes are evanescent. The
1123: flux is evaluated at the location where $\omega t_{\rm th} = 1$.
1124: Dashed, solid, and dotted curves correspond to $M=1.5$, 1.6, and
1125: $1.7\msun$, respectively.} \label{fig:radiativeflux}
1126: \end{figure}
1127: %
1128:
1129: % -----------------------------------------------------------
1130:
1131: \subsection{ Numerical Examples }
1132: \label{sec:numerical}
1133:
1134: %
1135: \begin{figure*}
1136: \centerline{\epsfig{file = fig7.eps,angle = 0,width = 1.\linewidth}}
1137: \caption{Responses of tidally forced $1\msun$ and $1.6\msun$
1138: main-sequence stars. Black, red, blue, and green curves denote,
1139: respectively, the logarithms of $\xi_r/r$, $(\delta p+\rho \delta
1140: \varphi)/\rho g r$, $\Delta s/C_p$, and $\Delta L/L$. Solid
1141: (dashed) curves show the real (imaginary) parts. The tidal
1142: potential has been scaled so that $\xi_r/R = 1$ corresponds to the
1143: equilibrium-tide value. The four panels show the following: (a)
1144: non-resonant response of a $1\msun$ star tidally forced at a period
1145: of $2\pi/\omega \simeq 2.91\,{\rm day}$; (b) resonant response of a
1146: $1\msun$ star with $2\pi/\omega \simeq 1.00\,{\rm day}$; (c)
1147: non-resonant response of a $1.6\msun$ star with $2\pi/\omega \simeq
1148: 3.00\,{\rm day}$; (d) resonant response of a $1.6\msun$ star with
1149: $2\pi/\omega \simeq 1.02\,{\rm day}$}
1150: \label{fig:allvsdepth}
1151: \end{figure*}
1152: %
1153:
1154:
1155: Here we show solutions of the perturbed mass, momentum, and energy
1156: equations that describe linear, nonadiabatic oscillations of a star
1157: subject to a varying tidal force. The equations listed in the
1158: Appendix are the same as in \citet{Unno1989} for radiative
1159: regions, but augmented to include the tidal acceleration.
1160: In convection zones, we apply the prescription $\Delta s =
1161: \textrm{constant}$ based on our conclusions in \S~\ref{sec:conv}.
1162: Figure~\ref{fig:allvsdepth} summarizes how the interiors of $1\msun$
1163: and $1.6\msun$ stars respond to resonant and non-resonant tidal
1164: forcing. The tidal potential has been scaled so that $\xi_r/R = 1$
1165: corresponds to the equilibrium-tide surface displacement.
1166:
1167: For our $1\msun$ model, the non-resonant response to a forcing period
1168: of $\simeq$3\,days is shown in Fig.~\ref{fig:allvsdepth}a. We see
1169: that $\xi_r/R$ matches the equilibrium-tide result at the surface; the
1170: imaginary piece is completely negligible. We also find that our
1171: approximation for $\Delta F/F$ at the surface
1172: (eq.~[\ref{eq:simplesurfacedf}]) works very well. A factor of
1173: $\sim$10 decay in $|\Delta L|/L$ occurred in order for $|\Delta F|/F
1174: \sim \xi_r/R$ at the surface. Variation of $\Delta s/C_p$ in the
1175: convection zone ($\log [p/(GM^2/R^4)] > -2.5$) is due to changes in
1176: $C_p$. In the radiative interior, the oscillations are caused by most
1177: nearly resonant $g$-modes, whose amplitudes rise rapidly as the core
1178: is approached, due to conservation of wave luminosity. We have
1179: checked that the quasi-adiabatic approximation of $\Delta L/L$ is
1180: valid in the radiative region; the ratio of the real and imaginary
1181: parts is found to be roughly constant for the ingoing gravity-wave
1182: \citep[see also][]{Zahn1977}.
1183:
1184: In order to model the resonant response of a $1\msun$ star, we tuned
1185: the forcing period to $\simeq$1\,day (see Figs.~\ref{fig:allvsdepth}b
1186: and \ref{fig:xirdf_magphase_env_1.0msun}). At the surface, both
1187: $\xi_r$ and $\Delta L$ have dominant imaginary parts, due to the short
1188: radial wavelength of the $g$-mode compared to the equilibrium-tide
1189: fluid displacement. The entropy at the base of the convection zone is
1190: very strongly perturbed in comparison to the non-resonant case, but
1191: $\Delta L$ is still damped by orders of magnitude as the surface is
1192: approached.
1193:
1194: Figure \ref{fig:xirdf_magphase_env_1.0msun} shows the surface values
1195: of the complex modulus and phase of $\xi_r/R$ and $\Delta F/F$ versus
1196: forcing period. The phase is $\tan^{-1}(\textrm{Imaginary/Real}) \in
1197: (-\pi, \pi)$. Solid lines connect points halfway between $g$-mode
1198: resonant periods. We find that the equilibrium-tide approximation
1199: given by eqs.~(\ref{eq:eqtide}) and (\ref{eq:simplesurfacedf}) is
1200: excellent for non-resonant forcing. Dashed curves give the maximum
1201: and minimum values that occur on resonance. One example of a
1202: resonance is shown in the insets. Resonant forcing at periods of
1203: $<$2\,days yields surface values of $\xi_r/R$ and $\Delta F/F$ that
1204: differ substantially from the equilibrium-tide results. However, the
1205: ratio of resonance width to the spacing between adjacent resonances is
1206: $\sim$$10^{-4}$, making resonant forcing very unlikely. It is
1207: noteworthy that at forcing periods of $>$2\,days, the equilibrium-tide
1208: result holds extremely well even when precisely on a resonance. As
1209: explained by Zahn (1975), the resonant response can be considered as
1210: the sum of the equilibrium tide and the most nearly resonant wave. As
1211: the period increases, the $g$-mode radial wavelength decreases,
1212: resulting in a reduction of the overlap integral for the mode and the
1213: tidal force, which in turn gives a decreased amplitude of the wave
1214: component relative to the equilibrium tide.
1215:
1216: The non-resonant response of the $1.6\msun$ star is shown in
1217: Fig.~\ref{fig:allvsdepth}c. We see that the equilibrium-tide result
1218: provides a good match to $\xi_r/R$. Our estimate for the modulus of
1219: the radiative luminosity perturbation in the evanescent limit
1220: (eq.~[\ref{eq:dlqadapprox}]) agrees reasonably well with what is in
1221: Fig.~\ref{fig:allvsdepth}c. We also see that $\Delta L/L$ does
1222: roughly ``freeze-out'' when $\omega t_{\rm th} \simeq 1$, just below
1223: the base of the convection zone at $\log p/(GM^2/R^4) \simeq -9$ (see
1224: Fig.~\ref{fig:prop_1.6msun}). Our expectations in
1225: \S~\ref{sec:radiative} regarding the imaginary part of $\Delta L/L$
1226: are borne out in Fig.~\ref{fig:allvsdepth}c
1227:
1228: A resonantly excited $1.6\msun$ star exhibits huge surface flux
1229: perturbations, radial displacements, and phase lags, as seen in
1230: Fig.~\ref{fig:allvsdepth}d. In
1231: Fig.~\ref{fig:xirdf_mag_point_1.6msun}, surface values of $|\xi_r|/R$
1232: and $|\Delta F|/F$ are plotted as a function of forcing period, where
1233: we have taken care to resolve resonances. Resonant amplitudes vary
1234: non-monotonically with period, in contrast to the smooth behavior of
1235: the $1\msun$ star (Fig.~\ref{fig:xirdf_magphase_env_1.0msun}).
1236: Although we do not show the results here, similar plots for masses
1237: between $1\msun$ and $1.6\msun$ show progressively more structure as
1238: the mass increases. The cause of this irregularity is not clear, but
1239: may have to do with the two thin surface convection zones changing the
1240: overlap of successive $g$-modes with the tidal force.
1241:
1242: %
1243: \begin{figure}
1244: \centerline{\epsfig{file = fig8.eps,angle = 0,width = \linewidth}}
1245: \caption{ Surface radial displacement and Lagrangian flux perturbation
1246: versus forcing period for a $1.0M_\odot$ star. Solid lines connect
1247: points halfway between resonant $g$-mode periods, while dashed
1248: curves give the maximum and minimum values found on resonance. The
1249: equilibrium-tide approximation is extremely good, except when the
1250: forcing period is $<$2\,days and resonant.}
1251: \label{fig:xirdf_magphase_env_1.0msun}
1252: \end{figure}
1253: %
1254:
1255: % -----------------------------------------------------------
1256:
1257: \section{SUMMARY}
1258: \label{sec:summary}
1259:
1260: We have investigated in detail the ellipsoidal oscillations of
1261: 0.9--$1.6\msun$ main-sequence stars induced by substellar companions.
1262: Classical models of ellipsoidal variability
1263: \citep[e.g.,][]{Wilson1994} are built on the assumption of hydrostatic
1264: balance in a frame corotating with the binary orbit. This approach is
1265: justified in the context of short-period ($P_{\rm orb} \la 10$\,days)
1266: binaries containing two stars of comparable mass, where tidal
1267: dissipation circularizes the orbits and synchronizes the stellar spins
1268: with the orbital frequency. However, when the companion has a very
1269: low mass, we cannot assume that the binary is in complete tidal
1270: equilibrium; in fact, this state may be unattainable (see
1271: \S~\ref{sec:prelim}). In this case, one must, in general, appeal to a
1272: dynamical description of the tidal interaction. A substellar
1273: companion with $P_{\rm orb} \ga 1$\,day raises tides on the star that
1274: are a small fraction of the stellar radius (see
1275: eq.~[\ref{eq:tidescale}]), permitting a linear analysis of the stellar
1276: oscillations.
1277:
1278: While the root of our study is a dynamical treatment of stellar tidal
1279: perturbations, the equilibrium-tide approximation does have an
1280: important realm of validity (see below). For this reason, we derived
1281: in \S~\ref{sec:eqtide} a general expression (eq.~[\ref{eq:fdiskeq}])
1282: for the measurable flux variation of a star that remains in
1283: hydrostatic equilibrium under the influence of a small external tidal
1284: force. This formula (1) assumes that the local perturbation to the
1285: energy flux at the stellar surface is proportional to and in phase
1286: with the equilibrium-tide radial fluid displacement at each angular
1287: order $\ell$ (eq.~[\ref{eq:eqtide}]), (2) neglects stellar rotation,
1288: and (3) applies to inclined and eccentric orbits. As expected, the
1289: fractional amplitude of the modulation is $\sim$$\pert \equiv
1290: (M_p/M)(R/a)^3$ for small eccentricities and $I = 90^\circ$, or
1291: $\sim$$10^{-5} (M_p/M_J)(P_{\rm orb}/1\,{\rm day})^{-2}$ for a star
1292: like the Sun (see \S~\ref{sec:prelim}).
1293:
1294: %
1295: \begin{figure}
1296: \centerline{\epsfig{file = fig9.eps,angle = 0,width = \linewidth}}
1297: \caption{ Surface radial displacement and Lagrangian flux perturbation
1298: versus forcing period for a $1.6M_\odot$ star. Curves connect
1299: evenly spaced points away from resonances, with finer spacing near
1300: resonance periods.}
1301: \label{fig:xirdf_mag_point_1.6msun}
1302: \end{figure}
1303: %
1304:
1305: A common practice is to use von Zeipel's theorem when computing the
1306: surface radiative flux from a tidally distorted star (see
1307: \S~\ref{sec:vonzeipel}). The theorem assumes that the star is in
1308: hydrostatic equilibrium and that the energy transport in the outer
1309: layers is purely by radiative diffusion. As already mentioned, the
1310: hydrostatic assumption is technically unjustified for substellar
1311: perturbers. Moreover, the majority of {\em Kepler} targets will be
1312: main-sequence stars with masses of $<$$1.4\msun$, which have
1313: substantial surface convection zones. Evidently, von Zeipel's theorem
1314: is an inappropriate starting point for the conditions of interest.
1315:
1316: Section~\ref{sec:conv} discusses heat transport in perturbed stars
1317: with convective envelopes. Heuristic arguments are used to develop a
1318: simple treatment of the perturbed convection zone in main-sequence
1319: stars of mass $<$$1.6\msun$ with forcing periods of 1--10\,days. We
1320: suggest that both the specific entropy $s$ and its Lagrangian
1321: perturbation $\Delta s$ are spatially constant in convective regions,
1322: a model partly inspired by the ideas of
1323: \citet{Brickhill1983,Brickhill1990}.
1324:
1325: %
1326: \input{tab2.tex}
1327: %
1328:
1329: Using this prescription, we analytically compute in
1330: \S~\ref{sec:thickconv} the perturbed flux at the photosphere of deeply
1331: convective stars ($M\la 1.4\msun$), where the thermal time scale at
1332: the base of the convection zone is much longer than the forcing
1333: period. We find that $\Delta s/C_p$ is negligible near the top of the
1334: convection zone, and that the photospheric flux perturbation is
1335: proportional to changes in the effective surface gravity. Thus, we
1336: recover the equilibrium-tide result, $\Delta F/F = -\lambda_\ell
1337: \xi_r/R$, at the surface, where $\lambda_\ell$ depends on the
1338: adiabatic derivatives of opacity and temperature with respect to
1339: pressure (see eq.~[\ref{eq:lambda}]). Numerical solutions of the
1340: equations of linear, nonadiabatic stellar oscillations (see
1341: \S~\ref{sec:numerical} and Fig.~\ref{fig:allvsdepth}a) corroborate our
1342: analytic estimates in the non-resonant regime. Resonant excitations
1343: of $g$-modes in the radiative stellar interior cause large departures
1344: from the equilibrium-tide approximation when the forcing period is
1345: $<$2\,days (Figs.~\ref{fig:allvsdepth}b and
1346: \ref{fig:xirdf_magphase_env_1.0msun}). However, the likelihood of
1347: being on a resonance is small, and at periods of $>$2\,day the
1348: equilibrium-tide result holds for $M \simeq 1\msun$ even with resonant
1349: forcing.
1350:
1351: Stars of mass $\ga$$1.4\msun$ have thin, relatively inefficient
1352: surface convection zones. Thus, $g$-modes can propagate very close to
1353: the surface and produce large flux perturbations and fluid
1354: displacements. Analytic arguments in \S~\ref{sec:radiative} indicate
1355: that the surface flux perturbations in these stars have non-resonant
1356: amplitudes of $\sim$$10\pert$ (eq.~[\ref{eq:dlqadapprox}] and
1357: Fig.~\ref{fig:radiativeflux}), in rough agreement with our numerical
1358: calculations (Fig.~\ref{fig:allvsdepth}c). As seen in
1359: Figs.~\ref{fig:allvsdepth}d and \ref{fig:xirdf_mag_point_1.6msun}, a
1360: resonantly forced $1.6\msun$ star can exhibit flux perturbation
1361: amplitudes of $>$$100\pert$ at forcing periods of $\simeq$1\,day.
1362: While the amplitudes are not as extreme at longer periods, their
1363: dependence on period is rather erratic
1364: (Fig.~\ref{fig:xirdf_mag_point_1.6msun}), an issue that deserves
1365: further study. It will be difficult to derive physical
1366: interpretations from the ellipsoidal variability of these more massive
1367: stars.
1368:
1369: % -----------------------------------------------------------
1370:
1371: \section{DETECTION PROSPECTS}
1372: \label{sec:detection}
1373:
1374: The dominant sources of periodic variability of a star with a
1375: substellar companion are transit occultations (when $|\cos I| <
1376: [R+R_p]/a$), Doppler flux modulations, reflection of starlight from
1377: the companion, and ellipsoidal oscillations. For each of these
1378: signals, Table ~\ref{tab:fluxmod} lists the characteristic amplitude,
1379: period with the largest power in the Fourier spectrum, and orbital
1380: phase(s) at which the light is a maximum or minimum. The transit
1381: contribution is included for completeness, but its duration is
1382: sufficiently short---a fraction $\simeq$$(R+R_p)/(\pi a)$ of $P_{\rm
1383: orb}$---that it should often be possible to excise it from the data
1384: \citep[see][]{Sirko2003}. Of the remaining signals, the Doppler
1385: variability is the simplest, being purely sinusoidal with period
1386: $P_{\rm orb}$ when the orbit is circular. The dominant $\ell = 2$
1387: piece of the equilibrium-tide approximation to the ellipsoidal
1388: variability (see eqs.~[\ref{eq:fdiskeq}] and [\ref{eq:P2}]) is also
1389: sinusoidal when $e = 0$, but with period $P_{\rm orb}/2$. Reflection
1390: is more problematic, as its time dependence is generally not
1391: sinusoidal and not known a priori.
1392:
1393: If the companion scatters light as a Lambert sphere
1394: \citep[e.g.,][]{Seager2000}, the Fourier spectrum of the reflection
1395: variability has finite amplitude at all harmonics of the orbital
1396: frequency $\Omega$, but the amplitude at $2\Omega$ is roughly 1/5 of
1397: the amplitude at $\Omega$. Therefore, the reflection and ellipsoidal
1398: variability amplitudes may be similar at a frequency of $2\Omega$ when
1399: $\alpha = 0.1$, $M_p \sim M_J$, and $P_{\rm orb} \simeq 1$\,day.
1400: Also, the orbital phase at which the reflected light is a maximum is
1401: distinct from both the Doppler and ellipsoidal cases, further
1402: distinguishing the signals. However, Lambert scattering is probably
1403: never appropriate in real planetary atmospheres. Infrared reemission
1404: of absorbed optical light, multiple photon scattering, and anisotropic
1405: scattering typically conspire to narrow the peak in the reflection
1406: lightcurve and lower the albedo, decreasing the prominence of the
1407: reflection signal. These issues are sensitive to the atmospheric
1408: chemistry and the uncertain details in models of irradiated giant
1409: planets. For reasonable choices regarding the atmospheric
1410: composition, calculated optical albedos of Jovian planets range from
1411: $<$0.01 to $\simeq$0.5 \citep{Seager2000,Sudarsky2000}. Recent
1412: photometric observations of HD~209458, the star hosting the
1413: first-detected transiting giant planet ($P_{\rm orb} \simeq
1414: 3.5$\,days), constrain the planetary albedo to be $<$0.25
1415: \citep{Rowe2006}.
1416:
1417: Detailed lightcurve simulations will be required to say how well the
1418: different periodic signals can be extracted from the data. This is
1419: beyond the scope of the current study. We now do the simpler exercise
1420: of isolating the ellipsoidal modulations and assessing when this
1421: effect alone should be detectable. For a star of apparent visual
1422: magnitude $V$ and an integration time of $T = 6\,{\rm hr}$, {\em
1423: Kepler}'s photon shot noise is\footnote{An integration time of $T =
1424: 6$\,hr is chosen for convenience; {\em Kepler}'s nominal exposure
1425: time is 30\,min. Here we use the $V$-band flux as a reference, but,
1426: in fact, the {\em Kepler} bandpass is 430--890\,nm, which spans $B$,
1427: $V$, and $R$ colors.}
1428: %
1429: \be
1430: \left(\frac{\delta \fdisk}{\fdisk}\right)_{\rm shot}
1431: \sim 10^{-5} 10^{0.2(V - 12)}
1432: \left(\frac{T}{6\,{\rm hr}}\right)^{-1/2}~.
1433: \ee
1434: %
1435: Instrumental noise should contribute at a similar level
1436: \citep[e.g.,][]{Koch2006}. If the data is folded at the orbital
1437: period and binned in time intervals $T \ll P_{\rm orb}$, the shot
1438: noise is suppressed by a factor of $\sim$$n_{\rm orb}^{-1/2}$, where
1439: $n_{\rm orb}$ is the number of folded cycles. After folding 1 year of
1440: continuous photometric data using $T = 6$\,hr, a star with $V <12$
1441: orbited by a giant planet with $P_{\rm orb} \la 3$\,days may have a
1442: fractional shot noise per time bin of $\la$$10^{-6}$. This is less
1443: than the ellipsoidal amplitude, $(\delta\fdisk/\fdisk)_{\rm ell}$,
1444: when $I$ is not too small.
1445:
1446: The actual situation is not so simple when the data spans of weeks or
1447: months, because the intrinsic stochastic variability of the star will
1448: not have a white-noise power spectrum. Over times of $\la$1\,day, the
1449: Sun shows variability of $(\delta \fdisk/\fdisk)_{\rm int} \sim
1450: 10^{-5}$, but the amplitude rises steeply between $\sim$1 and 10 days
1451: to $\sim$$10^{-3}$. Intrinsic variability tends to be large near the
1452: rotation period of the star, due mainly to starspots. Low-frequency
1453: variability may not too damaging for the study of ellipsoidal
1454: oscillations induced by planets with $P_{\rm orb} \la 3$\,days, but
1455: more study is needed.
1456:
1457: {\em Kepler}'s target list will contain $\simeq$$10^5$ main-sequence
1458: FGK stars with $V = 8$--14. The statistics of known exoplanets
1459: indicate that 1--2\% of all such stars host a giant planet ($M_p \ga
1460: M_J$) with $P_{\rm orb} < 10$\,days \citep[e.g.,][]{Marcy2005}. Of
1461: these ``hot Jupiters,'' $\simeq$30\% have $P_{\rm orb} =1$--3\,days.
1462: It seems that a maximum of $\sim$$10^3$ {\em Kepler} stars will have
1463: detectable ellipsoidal modulations. If we neglect intrinsic stellar
1464: variability and consider only shot noise, then many systems with
1465: $P_{\rm orb} \la 3$\,days and $V < 14$ will have signal-to-noise $S/N
1466: > 1$ after $\sim$100 cycles are monitored; this may amount to $>$100
1467: stars. Obviously, the number drops when we place higher demands on
1468: $S/N$ and include the intrinsic variability. The results depend
1469: critically on the distributions of $M_p$ and $P_{\rm orb}$.
1470:
1471: In order to better estimate the number of stars with potentially
1472: detectable ellipsoidal oscillation, we perform a simple population
1473: synthesis calculation. Denote the set of star-planet system
1474: parameters by $\vec{P} = \{ M, M_p, P_{\rm orb}, I\}$, and let
1475: $f(\vec{P})d\vec{P}$ be the probability of having a system in the
1476: 4-dimensional volume $d\vec{P}$. We assume that the planetary orbits
1477: are circular and obtain $(\delta\fdisk/\fdisk)_{\rm ell}$ from the
1478: equilibrium-tide estimate in Table~\ref{tab:fluxmod}. Given the mass
1479: of the star, we compute its absolute $V$ magnitude on the
1480: main-sequence using the approximation \citep[see also][]{Henry1993}
1481: %
1482: \be
1483: \mathscr{M}_V = 4.8 - 10.3 \log (M/M_\odot) ~,
1484: \ee
1485: %
1486: which is in accord with the usual mass-luminosity relation
1487: $\log(L/L_\odot) \simeq 4 \log(M/M_\odot)$ for $M \simeq 1\msun$.
1488: With a maximum apparent magnitude of $V_{\rm max} = 14$ for the {\em
1489: Kepler} targets, the maximum distance of the star is
1490: %
1491: \be
1492: D_{\rm max} = 10^{1 + 0.2(14 - \mathscr{M}_V)}\pc ~.
1493: \ee
1494: %
1495: With a certain signal-to-noise threshold $(S/N)_{\rm min}$, there is a
1496: maximum distance $D_d < D_{\rm max}$ to which the ellipsoidal
1497: variability is detectable. For a spatially uniform population, the
1498: detectable fraction of systems is $(D_d/D_{\rm max})^3$. Thus, the
1499: net detectable fraction among all systems is
1500: %
1501: \be
1502: \mathscr{E} = \int d\vec{P} f(\vec{P})
1503: \left(\frac{D_d}{D_{\rm max}}\right)^3~,
1504: \ee
1505: %
1506: an integral over all relevant $\vec{P}$ space.
1507:
1508: When the only noise is intrinsic to the star, $N =
1509: (\delta\fdisk/\fdisk)_{\rm int}$ and $S/N$ is independent of distance,
1510: so that $D_{d,{\rm int}} = D_{\rm max}$ when $S/N > (S/N)_{\rm min}$,
1511: and $D_{d,{\rm int}} = 0$ otherwise. In the case of pure shot noise,
1512: there is a maximum magnitude $V_d$ for which the ellipsoidal
1513: oscillations are detectable:
1514: %
1515: \be\label{eq:detfrac}
1516: V_d = 5\log \left[\frac{(\delta\fdisk/\fdisk)_{\rm ell}}{\chi\ (S/N)_{\rm min} }\right]~,
1517: \ee
1518: %
1519: where $\chi \sim 10^{-8.4}(T_6 n_{100})^{-1/2}$ is the value of
1520: $(\delta\fdisk/\fdisk)_{\rm shot}$ for $V = 0$, $T = 6 T_6\,{\rm hr}$,
1521: and $n_{\rm orb} = 100n_{100}$. The corresponding distance is given
1522: by $\log [D_{d,{\rm shot}}/10\pc] = 0.2(V_d - \mathscr{M}_V)$ if $V_d
1523: < V_{\rm max}$, and is $D_{d,{\rm shot}} = D_{\rm max}$ when $V_d >
1524: V_{\rm max}$. We take the maximum detectable distance to be $D_d =
1525: \min\{D_{d,{\rm int}}, D_{d,{\rm shot}}\}$.
1526:
1527: %
1528: \input{tab3.tex}
1529: %
1530:
1531: At this point the simplest approach is to assume that the parameters
1532: $\{ M, M_p, P_{\rm orb}, I\}$ are statistically independent and carry
1533: out a Monte Carlo integration to obtain $\mathscr{E}$. To this end,
1534: we draw $M$ from the \citet{Kroupa1993} initial mass function in the
1535: range of 0.5--$1.5\msun$. The planetary mass is chosen from the
1536: distribution $f(M_p)\propto M_p^{-x}$ for $M_p = 1$--$10M_J$.
1537: \citet{Marcy2005} find that $x \simeq 1$ when considering all detected
1538: planets; the shape of $f(M_p)$ is not well constrained at $P_{\rm orb}
1539: < 10$\,days. We let $x = 1$ and 2. We adopt $f(P_{\rm orb}) \propto
1540: P_{\rm orb}^{-y}$ over 1--10\,days. Multiplying the resulting value
1541: of $\mathscr{E}$ by 1000 provides a crude estimate of the actual
1542: number of {\em Kepler} targets with detectable ellipsoidal
1543: variability. No single value of $y$ is consistent with the data, and
1544: so we consider the reasonable range $y = -1$, 0, and $+1$.
1545: Inclinations are chosen under the assumption that the orbits are
1546: randomly oriented, such that $f(\cos I) = 1/2$ for $I\in (0,\pi)$.
1547: Our calculations use fixed values of $(\delta\fdisk/\fdisk)_{\rm int}
1548: = 10^{-5}$ and $T_6 = n_{100} = 1$.
1549:
1550: Results of our Monte Carlo integrations are shown in
1551: Table~\ref{tab:detnum} as actual numbers of {\em Kepler} targets. The
1552: largest number of detectable systems is obtained when $x = y = 1$,
1553: parameters that yield the largest proportions short periods and
1554: massive planets. We expect that $\sim$10--100 {\em Kepler} stars may
1555: exhibit ellipsoidal oscillations with $S/N \ga 5$. A handful of
1556: systems might have $S/N \ga 10$. Higher harmonics from the $\ell = 3$
1557: components of eq.~(\ref{eq:fdiskeq}) or modest eccentricities might be
1558: accessible for at most a few stars.
1559:
1560: Our integrations also check for cases where the planet is transiting.
1561: As $(S/N)_{\rm min}$ increases from 1 to 5, the fraction of systems in
1562: Table~\ref{tab:detnum} with $|\cos I| < (R+R_p)/a$ runs from
1563: $\simeq$30\% to $\simeq$50\%, with a weak dependence on $x$ and $y$.
1564: Such significant fractions stand to reason, since systems with the
1565: shortest periods have the highest ellipsoidal amplitudes and transit
1566: probabilities. Transit measurements directly give $P_{\rm orb}$,
1567: $\sin I \ga 0.95$ (for $P_{\rm orb} \ga 1$\,day), and $(R_p/R)^2$.
1568: The planet mass $M_p$ can be determined with the addition of
1569: spectroscopic radial velocity measurements, which should be possible
1570: for most of the {\em Kepler} targets with detectable ellipsoidal
1571: oscillations. The ellipsoidal amplitude then depends on the
1572: unmeasured stellar mass and radius via $\pert \propto R^3/M^2$
1573: (eq.~[\ref{eq:tidescale}]), as well as the stellar photospheric
1574: conditions (eq.~[\ref{eq:simplesurfacedf}]). If $M$ and $R$ are
1575: obtained from stellar models, ellipsoidal variability may provide an
1576: interesting consistency check on all the system parameters, as well as
1577: test the theory of forced stellar oscillations.
1578:
1579: As a last point, we emphasize that stars of mass $\ga$$1.4\msun$ may
1580: have typical ellipsoidal amplitudes of $\sim$$10\pert$. However, such
1581: stars will also be younger than most {\em Kepler} targets and probably
1582: have intrinsic variability $\gg$$10^{-5}$. We carried out Monte Carlo
1583: integrations with $M = 1.4$--$1.6\msun$, $(\delta\fdisk/\fdisk)_{\rm
1584: ell} = 10\pert\sin^2 I$, and $x = y = 1$. As we vary
1585: $(\delta\fdisk/\fdisk)_{\rm int}$ from $10^{-5}$ to $10^{-4}$,
1586: $\mathscr{E}$ decreases from large values of $\simeq$0.4 to a small
1587: fraction of $\simeq$0.03 for $(S/N)_{\rm min} = 10$. Unfortunately,
1588: we do not know how many such stars will be included in the {\em
1589: Kepler} target list. Also, there has not yet been a discovery of a
1590: giant planet with $P_{\rm orb} < 10$\,days around a star of mass
1591: $\geq$$1.4\msun$, but exoplanet surveys tend to exclude these more
1592: massive stars.
1593:
1594: %--------------------------------------------------------------------------------
1595:
1596: \acknowledgements
1597:
1598: We thank Tim Brown for general discussions and addressing {\em Kepler}
1599: questions, J{\o}rgen Christensen-Dalsgaard for guidance on stellar
1600: luminosity perturbations, and Mike Muno for advice on signal
1601: processing. This work was supported by NSF grant PHY05-51164.
1602:
1603: %--------------------------------------------------------------------------------
1604:
1605: \appendix
1606:
1607: \section{OSCILLATION EQUATIONS}
1608: \label{sec:equations}
1609:
1610: Here we list the nonadiabatic, linearized fluid equations that we
1611: solve numerically. The reader is referred to \cite{Unno1989} for a
1612: complete discussion. Scalar and vector quantities are expanded in
1613: spherical harmonics $Y_{\ell m}$ and poloidal vector harmonics,
1614: respectively. The momentum, mass, and energy equations are written in
1615: terms of the dimensionless variables $y_1=\xi_r/r$, $y_2=(\delta
1616: p/\rho+\delta \varphi)/gr$, $y_3=\delta \varphi/gr$, $y_4=g^{-1}
1617: d\delta \varphi/dr$, $y_5=\Delta s/C_p$, and $y_6=\Delta L/L$. Here
1618: $L$ is the total (radiative plus convective) luminosity. The radial
1619: flux perturbation is $\Delta F/F=\Delta L/L-2\xi_r/r$. In radiative
1620: zones, the nonadiabatic equations are
1621: %
1622: \be
1623: \frac{dy_1}{d\ln r} & = & y_1 \left( \frac{gr}{c_s^2}-3 \right)
1624: + y_2 \left( \frac{gk_h^2r}{\omega^2} - \frac{gr}{c_s^2} \right)
1625: + y_3 \frac{gr}{c_s^2} - y_5 \rho_s + \frac{k_h^2}{\omega^2} U ~,
1626: \\
1627: \frac{dy_2}{d\ln r} & = & y_1 \left(\frac{\omega^2-N^2}{g/r} \right)
1628: + y_2 \left( 1-\eta + \frac{N^2}{g/r} \right)
1629: - y_3 \frac{N^2}{g/r} - \rho_s y_5 - \frac{1}{g} \frac{dU}{dr} ~,
1630: \\
1631: \frac{dy_3}{d \ln r} & = & y_3 \left( 1 - \eta \right)
1632: + y_4 ~,
1633: \\
1634: \frac{dy_4}{d \ln r} & = & y_1 \eta \frac{N^2}{g/r}
1635: + y_2 \eta \frac{gr}{c_s^2}
1636: + y_3 \left[ \ell(\ell+1) - \eta \frac{gr}{c_s^2} \right]
1637: - y_4 \eta + y_5 \rho_s \eta ~,
1638: \\
1639: \frac{dy_5}{d\ln r} & = & y_1 \frac{r}{H_p}
1640: \left[
1641: \nabla_{\rm ad} \left( \eta - \frac{\omega^2}{g/r} \right)
1642: + 4 \left( \nabla - \nabla_{\rm ad} \right) + c_2
1643: \right]
1644: + y_2 \frac{r}{H_p}
1645: \left[
1646: \left( \nabla_{\rm ad} - \nabla \right) \frac{gk_h^2r}{\omega^2} - c_2
1647: \right]
1648: \nonumber \\
1649: && + y_3 \frac{r}{H_p} c_2
1650: + y_4 \frac{r}{H_p} \nabla_{\rm ad}
1651: + y_5 \frac{r}{H_p} \nabla \left( 4 - \kappa_s \right)
1652: - y_6 \frac{r}{H_p} \nabla
1653: + \frac{r}{H_p}
1654: \left[
1655: \nabla_{\rm ad} \left( \frac{dU/dr}{g} + \frac{k_h^2}{\omega^2} U \right)
1656: - \nabla \frac{k_h^2}{\omega^2} U
1657: \right] ~,
1658: \label{eq:y5}
1659: \\
1660: \frac{dy_6}{d\ln r} & = &
1661: y_1 \ell(\ell+1)\left(\frac{\nabla_{\rm ad}}{\nabla} - 1 \right)
1662: - y_2 \ell(\ell+1)\frac{\nabla_{\rm ad}}{\nabla}
1663: + y_3 \ell(\ell+1)\frac{\nabla_{\rm ad}}{\nabla}
1664: + y_5 \left[
1665: i\omega \frac{4\pi r^3 \rho C_p T}{L} - \frac{\ell(\ell+1)}{\nabla}\frac{H_p}{r}
1666: \right] ~,
1667: \label{eq:y6}
1668: \ee
1669: %
1670: where $c_s$ is the sound speed, $\eta = d\ln M_r/d\ln r$,
1671: $c_2=(r/H_p)\nabla (\kappa_{\rm ad}-4\nabla_{\rm ad})+\nabla_{\rm
1672: ad}(d\ln \nabla_{\rm ad}/d\ln r + r/H_p)$, and we have ignored
1673: energy generation terms. Note that the tidal acceleration $-\grad U$
1674: has been added to the momentum equations. In convection zones, we
1675: ignore turbulent viscosity effects and replace the radiative diffusion
1676: equation (eq.~[\ref{eq:y5}]) with the prescription $\Delta s =
1677: \textrm{constant}$ (see \S~\ref{sec:conv}), or more precisely
1678: %
1679: \be
1680: \frac{d}{dr} \left( y_5 C_p \right) = 0~.
1681: \ee
1682: %
1683: Equation~(\ref{eq:y6}) still involves the total (convective plus
1684: radiative) luminosity. We ignore energy generation and horizontal flux
1685: perturbation terms, i.e. we ignore all terms with spherical harmonic
1686: index $\ell$ in eq.~(\ref{eq:y6}) in convection zones.
1687:
1688: At the center of the star, we require the solutions to be finite, and
1689: also set $\Delta s=0$. At the surface, we set $\delta p=\rho g \xi_r$
1690: and we require $\delta \varphi$ to decrease outward. This boundary
1691: condition is only approximate, as $g$-modes may propagate above the
1692: convection zone for wave periods of $\ga$$4\, {\rm days}$ in our
1693: $1M_\odot$ model. The final surface boundary condition is given by
1694: eq.~(\ref{eq:dphotp}). Care must be used in the radiative zone just
1695: below the photosphere, since the entropy perturbation is far from the
1696: quasi-adiabatic value. If we solve the radiative diffusion equation
1697: in this region, we find that the entropy increases by $\sim$10 orders
1698: of magnitude in just a few grid points. However, we regard this
1699: behavior as unphysical, because the region at the top of the
1700: convection zone is optically thin. To eliminate this unphysical
1701: behavior, we set $\Delta s$ to a constant at such low optical depths.
1702:
1703: %--------------------------------------------------------------------------------
1704:
1705: \begin{thebibliography}{}
1706:
1707: \bibitem[Agol et al.(2005)]{Agol2005} Agol, E., Steffen, J.,
1708: Sari, R., \& Clarkson, W.\ 2005, \mnras, 359, 567
1709:
1710: \bibitem[Basri et al.(2005)]{Basri2005} Basri, G., Borucki,
1711: W.~J., \& Koch, D.\ 2005, New Astronomy Review, 49, 478
1712:
1713: %\bibitem[Bord{\'e} et al.(2003)]{Borde2003} Bord{\'e}, P., Rouan, D.,
1714: % \& L{\'e}ger, A.\ 2003, \aap, 405, 1137
1715:
1716: \bibitem[Borucki et al.(2004)]{Borucki2004} Borucki, W., et al.\ 2004,
1717: ESA SP-538: Stellar Structure and Habitable Planet Finding, 177
1718:
1719: \bibitem[Brickhill(1983)]{Brickhill1983} Brickhill, A.~J.\ 1983,
1720: \mnras, 204, 537
1721:
1722: \bibitem[Brickhill(1990)]{Brickhill1990} Brickhill, A.~J.\ 1990,
1723: \mnras, 246, 510
1724:
1725: \bibitem[Brown et al.(2001)]{Brown2001} Brown, T.~M., Charbonneau, D.,
1726: Gilliland, R.~L., Noyes, R.~W., \& Burrows, A.\ 2001, \apj, 552, 699
1727:
1728: \bibitem[Claret(2000)]{Claret2000} Claret, A.\ 2000, \aap, 363,
1729: 1081
1730:
1731: %\bibitem[Clayton(1983)]{Clayton1983} Clayton, D.~D.\ 1983, Chicago:
1732: % University of Chicago Press, 1983,
1733:
1734: %\bibitem[Cowling(1941)]{Cowling1941} Cowling, T.~G.\ 1941, \mnras,
1735: %101, 367
1736:
1737: \bibitem[Drake(2003)]{Drake2003} Drake, A.~J.\ 2003, \apj, 589, 1020
1738:
1739: \bibitem[Dziembowski(1977)]{Dziembowski1977} Dziembowski, W.\ 1977,
1740: Acta Astronomica, 27, 203
1741:
1742: \bibitem[Goldreich \& Nicholson(1989)]{Goldreich1989} Goldreich, P.,
1743: \& Nicholson, P.~D.\ 1989, \apj, 342, 1079
1744:
1745: \bibitem[Goldreich \& Wu(1999a)]{Goldreich1999a} Goldreich, P., \& Wu,
1746: Y.\ 1999, \apj, 511, 904
1747:
1748: \bibitem[Goldreich \& Wu(1999b)]{Goldreich1999b} Goldreich, P., \& Wu,
1749: Y.\ 1999, \apj, 523, 805
1750:
1751: %\bibitem[Goodman \& Dickson(1998)]{Goodman1998} Goodman, J., \&
1752: % Dickson, E.~S.\ 1998, \apj, 507, 938
1753:
1754: \bibitem[Grether \& Lineweaver(2006)]{Grether2006} Grether, D., \&
1755: Lineweaver, C.~H.\ 2006, \apj, 640, 1051
1756:
1757: \bibitem[Hansen \& Kawaler(1994)]{Hansen1994} Hansen, C. J., \& Kawaler, S. D. 1994, Stellar Interiors: Principles, Structure, and Evolution, ( New York: Springer)
1758:
1759: \bibitem[Henry \& McCarthy(1993)]{Henry1993} Henry, T.~J., \&
1760: McCarthy, D.~W., Jr.\ 1993, \aj, 106, 773
1761:
1762: \bibitem[Holman \& Murray(2005)]{Holman2005} Holman, M.~J., \&
1763: Murray, N.~W.\ 2005, Science, 307, 1288
1764:
1765: \bibitem[Jenkins \& Doyle(2003)]{Jenkins2003} Jenkins, J.~M., \&
1766: Doyle, L.~R.\ 2003, \apj, 595, 429
1767:
1768: \bibitem[Kippenhahn \& Weigert(1990)]{Kippenhahn1990}
1769: Kippenhahn, R., \& Weigert, A. 1990, Stellar Structure and Evolution (Berlin: Springer)
1770:
1771: \bibitem[Koch et al.(2006)]{Koch2006} Koch, D., et al.\ 2006,
1772: \apss, 304, 391
1773:
1774: %\bibitem[Kopal(1941)]{Kopal1941} Kopal, Z.\ 1941, \apj, 94, 159
1775:
1776: \bibitem[Kopal(1942)]{Kopal1942} Kopal, Z.\ 1942, \apj, 96, 20
1777:
1778: %\bibitem[Kopal(1959)]{Kopal1959} Kopal, Z.\ 1959, Close Binary Systems
1779: % (New York: Wiley)
1780:
1781: \bibitem[Kroupa et al.(1993)]{Kroupa1993} Kroupa, P., Tout, C.~A.,
1782: \& Gilmore, G.\ 1993, \mnras, 262, 545
1783:
1784: \bibitem[Loeb \& Gaudi(2003)]{Loeb2003} Loeb, A., \& Gaudi,
1785: B.~S.\ 2003, \apjl, 588, L117
1786:
1787: %\bibitem[Loeb(2005)]{Loeb2005} Loeb, A.\ 2005, \apjl, 623, L45
1788:
1789: %\bibitem[Lucy(1967)]{Lucy1967} Lucy, L.~B.\ 1967, Zeitschrift
1790: % fur Astrophysik, 65, 89
1791:
1792: \bibitem[{{Marcy} {et~al.}(2005){Marcy}, {Butler}, {Fischer}, {Vogt}, {Wright},
1793: {Tinney}, \& {Jones}}]{Marcy2005}
1794: {Marcy}, G., {Butler}, R.~P., {Fischer}, D., {Vogt}, S., {Wright}, J.~T.,
1795: {Tinney}, C.~G., \& {Jones}, H.~R.~A. 2005, Prog. Theor. Phys. Supp., 158, 24
1796:
1797: \bibitem[Mihalas(1970)]{Mihalas1970} Mihalas, D. 1970, Stellar Atmospheres (San Francisco: Freeman)
1798:
1799: \bibitem[Miralda-Escud{\'e}(2002)]{Miralda-Escude2002}
1800: Miralda-Escud{\'e}, J.\ 2002, \apj, 564, 1019
1801:
1802: \bibitem[Pace \& Pasquini(2004)]{Pace2004} Pace, G., \&
1803: Pasquini, L.\ 2004, \aap, 426, 1021
1804:
1805: \bibitem[Paxton(2004)]{Paxton2004} Paxton, B.\ 2004, \pasp, 116,
1806: 699
1807:
1808: %\bibitem[Pols et al.(1995)]{Pols1995} Pols, O.~R., Tout, C.~A.,
1809: %Eggleton, P.~P., \& Han, Z.\ 1995, \mnras, 274, 964
1810:
1811: \bibitem[Rasio et al.(1996)]{Rasio1996} Rasio, F.~A., Tout,
1812: C.~A., Lubow, S.~H., \& Livio, M.\ 1996, \apj, 470, 1187
1813:
1814: \bibitem[Robinson et al.(1982)]{Robinson1982} Robinson, E.~L.,
1815: Kepler, S.~O., \& Nather, R.~E.\ 1982, \apj, 259, 219
1816:
1817: \bibitem[Rowe et al.(2006)]{Rowe2006} Rowe, J.~F., et al.\ 2006,
1818: \apj, 646, 1241
1819:
1820: %\bibitem[Russell \& Shapley(1912)]{Russell1912} Russell, H.~N., \&
1821: % Shapley, H.\ 1912, \apj, 36, 385
1822:
1823: \bibitem[Sartoretti \& Schneider(1999)]{Sartoretti1999} Sartoretti,
1824: P., \& Schneider, J.\ 1999, \aaps, 134, 553
1825:
1826: \bibitem[Seager et al.(2000)]{Seager2000} Seager, S., Whitney,
1827: B.~A., \& Sasselov, D.~D.\ 2000, \apj, 540, 504
1828:
1829: %\bibitem[Seager \& Mall{\'e}n-Ornelas(2003)]{Seager2003} Seager,
1830: %S., \& Mall{\'e}n-Ornelas, G.\ 2003, \apj, 585, 1038
1831:
1832: \bibitem[Sirko \& Paczy{\'n}ski(2003)]{Sirko2003} Sirko, E., \&
1833: Paczy{\'n}ski, B.\ 2003, \apj, 592, 1217
1834:
1835: \bibitem[Skumanich(1972)]{Skumanich1972} Skumanich, A.\ 1972, \apj,171, 565
1836:
1837: \bibitem[Soszynski et al.(2004)]{Soszynski2004} Soszynski, I.,
1838: Udalski, A., Kubiak, M., Szymanski, M.~K., Pietrzynski, G., Zebrun,
1839: K., Wyrzykowski, O.~S.~L., \& Dziembowski, W.~A.\ 2004, Acta
1840: Astronomica, 54, 347
1841:
1842: \bibitem[Sudarsky et al.(2000)]{Sudarsky2000} Sudarsky, D., Burrows,
1843: A., \& Pinto, P.\ 2000, \apj, 538, 885
1844:
1845: \bibitem[Udalski et al.(2002)]{Udalski2002} Udalski, A., et al.\
1846: 2002, Acta Astron., 52, 1
1847:
1848: \bibitem[Unno et al.(1989)]{Unno1989} Unno, W., Osaki, Y.,
1849: Ando, H., Saio, H., \& Shibahashi, H.\ 1989, Nonradial Oscillations
1850: of Stars, Second
1851: Edition (Tokyo: Univ. Tokyo Press)
1852:
1853:
1854: \bibitem[von Zeipel(1924)]{vonZeipel1924} von Zeipel, H.\ 1924,
1855: \mnras, 84, 665
1856:
1857: \bibitem[Wilson(1994)]{Wilson1994} Wilson, R.~E.\ 1994, \pasp,
1858: 106, 921
1859:
1860: \bibitem[Zahn(1975)]{Zahn1977} Zahn, J.-P.\ 1975, \aap, 41, 329
1861:
1862: \end{thebibliography}
1863:
1864: \end{document}
1865:
1866:
1867: