0706.4312/ms.tex
1: %%
2: %% Created 2005 10 24
3: %%
4: %% This is a manuscript file creates using AASTeX v5.x LaTeX 2e macros
5: %%
6: 
7: \documentclass[12pt]{emulateapj}
8: %\usepackage{amsmath}
9: %\usepackage{natbib}
10: \usepackage{ulem}
11: \bibliographystyle{apj}
12: 
13: \slugcomment{To be submitted to the Astrophysical Journal}
14: 
15: \begin{document}
16: 
17: \title{Interacting Binaries with Eccentric Orbits.  Secular 
18: Orbital Evolution Due To Conservative Mass Transfer} 
19: 
20: \author{J. F. Sepinsky, B. Willems, V. Kalogera, F. A. Rasio}
21: \affil{Department of Physics and Astronomy, Northwestern University,
22: 2145 Sheridan Road, Evanston, IL 60208}
23: \slugcomment{j-sepinsky, b-willems, vicky@northwestern.edu, and
24: rasio@northwestern.edu}
25: \shortauthors{Sepinsky, Willems, Kalogera, \& Rasio}
26: \shorttitle{Interacting Binaries with Eccentric Orbits}
27: 
28: \begin{abstract}
29: 
30: We investigate the secular evolution of the orbital semi-major axis and
31: eccentricity due to mass transfer in eccentric binaries, assuming
32: conservation of total system mass and orbital angular momentum. Assuming
33: a delta function mass transfer rate centered at periastron, we find
34: rates of secular change of the orbital semi-major axis and eccentricity
35: which are linearly proportional to the magnitude of the mass transfer
36: rate at periastron. The rates can be positive as well as negative, so
37: that the semi-major axis and eccentricity can increase as well as
38: decrease in time. Adopting a delta-function mass-transfer rate of
39: $10^{-9} M_\sun\,{\rm yr}^{-1}$ at periastron yields orbital evolution
40: timescales ranging from a few Myr to a Hubble time or more, depending on
41: the binary mass ratio and orbital eccentricity. Comparison with orbital
42: evolution timescales due to dissipative tides furthermore shows that
43: tides cannot, in all cases, circularize the orbit rapidly enough to
44: justify the often adopted assumption of instantaneous circularization at
45: the onset of mass transfer.  The formalism presented can be incorporated
46: in binary evolution and population synthesis codes to create a
47: self-consistent treatment of mass transfer in eccentric binaries.
48: 
49: \end{abstract}
50: 
51: \keywords{Celestial mechanics, Stars: Binaries: Close, Stars: Mass Loss}
52: 
53: \section{Introduction}
54: \label{sec-intro}
55: Mass transfer between components of close binaries is a common
56: evolutionary phase for many astrophysically interesting binary
57: systems.  Indeed, mass ejection and/or accretion is responsible for
58: many of the most recognizable phenomena associated with close
59: binaries, such as persistent or transient X-ray emission, neutron star
60: spin-up, and orbital contraction or expansion.  Theoretical
61: considerations of these and other associated phenomena in the
62: literature probe these systems quite effectively, yet they often do
63: not consider the effects of any eccentricity associated with the
64: binary orbit.  This can be of particular importance for binaries
65: containing a neutron star or a black hole, where mass loss and natal
66: kicks occurring during compact object formation may induce a
67: significant eccentricity to the binary \citep[e.g.][]{H83, BP95,
68:   K96}. After the formation of the compact object, tides tend to
69: circularize the orbit on a timescale which strongly depends on the
70: ratio of the radius of the compact object's companion to the orbital
71: semi-major axis.  Because of this, orbits are usually assumed to
72: circularize instantaneously when a binary approaches or begins a mass
73: transfer phase. 
74: 
75: Despite our generally well developed understanding of tidal
76: interactions in close binaries, quantitative uncertainties in tidal
77: dissipation mechanisms propagate into the determination of
78: circularization timescales. For example, \citet{MM05} have shown
79: that current theories of tidal circularization cannot explain observed
80: degrees of circularization of solar-type binaries in open clusters. 
81: Circularization of high-mass binaries, on the other hand, is currently
82: thought to be driven predominantly by resonances between dynamic tides
83: and free oscillation modes, but initial conditions play an
84: important role and an extensive computational survey of relevant parts
85: of the initial parameter space has yet to be undertaken \citep{WS99,
86: WS01, WVS03}.
87: 
88: Furthermore, assumptions of instantaneous circularization immediately
89: before or at the onset of mass transfer are in clear contrast with
90: observations of eccentric mass transferring systems. In the most recent
91: catalog of eccentric binaries with known apsidal-motion rates compiled
92: by \citet{PO99}, 26 out of the 128 listed systems are semi-detached or
93: contact binaries. Among these mass-transferring systems, $9$ have
94: measured eccentricities greater than $0.1$. In addition, many
95: high-mass X-ray binaries are known to have considerable orbital
96: eccentricities \citep{Rag05}. While mass transfer in these systems is
97: generally thought to be driven by the stellar wind of a massive O- or
98: B-star, it has been suggested that some of them may also be subjected
99: to atmospheric Roche-lobe overflow at each periastron passage of the
100: massive donor \citep[e.g.][]{P78}. 
101: 
102: \citet{H56}, \citet{K64-2}, and \citet{P64} were the first to study
103: the effects of mass transfer on the orbital elements of eccentric
104: binaries. However, their treatment was restricted to perturbations of
105: the orbital motion caused by the variable component masses.
106: \citet{MW83,MW84} extended these early pioneering studies to include
107: the effects of linear momentum transport from one star to the other,
108: as well as any other possible perturbations caused by the mass
109: transfer stream in the system. However, these authors derived the
110: equations governing the motion of the binary components with respect
111: to a reference frame with origin at the mass center of the binary,
112: which is not an inertial frame. Their equations therefore do not
113: account for the accelerations of the binary mass center caused by the
114: mass transfer (see \S~\ref{sec-appb}). 
115: 
116: More recent work on mass transfer in eccentric binaries has mainly
117: focused on smoothed particle hydrodynamics calculations of the mass
118: transfer stream over the course of a few orbits, without any
119: consideration of the long-term evolution of the binary \citep{L98,
120: Rea05}.
121: 
122: Hence, there is ample observational and theoretical motivation to
123: revisit the study of eccentric mass-transferring binaries.  In this
124: paper, our aim is to derive the equations governing the evolution of
125: the orbital semi-major axis and eccentricity in eccentric
126: mass-transferring binaries, assuming conservation of total system mass
127: and orbital angular momentum. In a subsequent paper, we will
128: incorporate the effects of mass and orbital angular momentum losses
129: from the system.
130: 
131: Our analysis is based on the seminal work of \citet{H69} who was the
132: first to derive the equations of motion of the components of eccentric
133: mass-transferring binaries while properly accounting for the effects
134: of the variable component masses on the stars' mutual gravitational
135: attraction, the transport of linear momentum from one star to the
136: other, the accelerations of the binary mass center due to the
137: redistribution of mass in the system, and the perturbations of the
138: orbital motion caused by the mass-transfer stream.  While the
139: equations of motion derived by \citet{H69} are valid for orbits of
140: arbitrary eccentricity, the author restricted the derivation of the
141: equations governing the evolution of the semi-major axis and
142: eccentricity to orbits with small initial eccentricities.
143: 
144: The paper is organized as follows. In \S\,2 and \S\,3 we present the
145: basic assumptions relevant to the investigation and derive the equations
146: governing the motion of the components of an eccentric mass-transferring
147: binary under the assumption of conservative mass transfer.  The
148: associated equations governing the rates of change of the semi-major
149: axis and the orbital eccentricity are derived in \S\,4, while numerical
150: results for the timescales of orbital evolution due to mass transfer as
151: a function of the initial binary mass ratio and orbital eccentricity are
152: presented in \S\,5. For comparison, timescales of orbital evolution due
153: to dissipative tidal interactions between the binary components are
154: presented in \S\,6.  \S\,7 is devoted to a summary of our main results
155: and a discussion of future work. In the appendices, lastly, we derive
156: an equation for the position of the inner Lagrangian point in eccentric
157: binaries with non-synchronously rotating component stars
158: (Appendix~\ref{sec-appa}), and present an
159: alternative derivation for the equations governing the secular evolution
160: of the orbital semi-major axis and eccentricity assuming instantaneous
161: mass transfer between two point masses (Appendix~\ref{sec-appc}).
162: 
163: \section{Basic Assumptions}
164: 
165: We consider a binary system consisting of two stars in an eccentric
166: orbit with period $P_{\rm orb}$, semi-major axis $a$, and
167: eccentricity $e$. We let the component stars rotate with angular
168: velocities $\vec{\Omega}_1$ and $\vec{\Omega}_2$ parallel to the
169: orbital angular velocity $\vec{\Omega}_{\rm orb}$, and assume the
170: rotation rates to be uniform throughout the stars. We also note that
171: the magnitude of $\vec{\Omega}_{\rm orb}$ varies periodically in time
172: for eccentric binaries, but its direction remains fixed in
173: space. Because of this, the stars cannot be synchronized with the
174: orbital motion at all times.
175: 
176: At some time $t$, one of the stars is assumed to fill its Roche lobe
177: and begins transferring mass to its companion through the inner
178: Lagrangian point $L_1$. We assume this point to lie on the line
179: connecting the mass centers of the stars, even though non-synchronous
180: rotation may cause it to oscillate in the direction perpendicular to
181: the orbital plane with an amplitude proportional to the degree of
182: asynchronism \citep{MW83}. Since the donor's rotation axis is assumed
183: to be parallel to the orbital angular velocity, we can safely assume 
184: that the transferred mass remains confined to the orbital plane.
185: 
186: We furthermore assume that all mass lost from the donor is accreted by
187: its companion, and that any orbital angular momentum transported by
188: the transferred mass is immediately returned to the orbit. The mass
189: transfer thus conserves both the total system mass and the orbital
190: angular momentum.
191: 
192: We also neglect any perturbations to the orbital motion other than
193: those due to mass transfer. At the lowest order of approximation,
194: these additional perturbations (e.g., due to tides, magnetic breaking,
195: or gravitational radiation) are decoupled from those due to mass
196: transfer, and can thus simply be added to obtain the total rates of
197: secular change of the orbital elements.
198: 
199: \section{Equations of Motion}
200: \label{sec-eom}
201: 
202: \subsection{Absolute Motion of the Binary Components}
203: 
204: \label{sec-pert}
205: 
206: Following \citet{H69}, we derive the equations of motion of the
207: components of an eccentric mass-transferring binary with respect to a
208: right handed inertial frame of reference $OXYZ$ which has an arbitrary
209: position and orientation in space (see Fig.~\ref{fig-coords}). We let
210: $M_i$ be the mass of star $i$ at some time $t$ at which mass is
211: transferred from the donor to the accretor, and $M_i + \delta M_i$ the
212: mass of the same star at some time $t+\delta t$, where $\delta t > 0$ is
213: a small time interval. With these notations, $\delta M_i < 0$
214: corresponds to mass loss, and $\delta M_i > 0$ to mass accretion. We
215: furthermore denote the point on the stellar surface at which mass is
216: lost or accreted by $A_i$. For the donor star, $A_i$ corresponds to the
217: inner Lagrangian point $L_1$, while for the accretor, $A_i$ can be any
218: point on the star's equator. For the remainder of the paper, we let
219: $i=1$ correspond to the donor and $i=2$ to the accretor.
220: 
221: \begin{figure*}
222: \epsscale{.9}
223: \plotone{f1.eps}
224: \caption{
225: Schematic representation of the reference frames and position vectors
226: adopted in the derivation of the equations of motion of the 
227: components of an eccentric mass-transferring binary.  The $Z$- and
228: $Z_1$-axes of the $OXYZ$ and $O_1X_1Y_1Z_1$ frames are perpendicular
229: to the plane of the page and are therefore not drawn. The geometry of
230: the system at time $t$ is shown in black, while the geometry at time
231: $t + \delta t$ is shown in gray.  The solid line connecting the origin
232: of the $O_1X_1Y_1Z_1$ frame at time $t$ to the origin of the frame at
233: time $t+\delta t$ represents the path the donor would have taken had
234: no mass transfer occurred.  The dashed line, on the other hand,
235: represents the perturbed motion of the donor's mass center (small
236: solid circles) due to the mass transfer. A similar perturbation is
237: imparted to the motion of the accretor. For clarity, this perturbed
238: motion and the $O_2X_2Y_2Z_2$ reference frame connected to the
239: accretor (see text) are omitted from the figure. The dotted line 
240: illustrates a possible path of a mass element transferred from the
241: donor to the accretor.  The element leaves the donor at the inner
242: Lagrangian point $L_1$ (asterisk) and accretes onto the companion at
243: the point $A_2$ (open star).
244: }
245: \label{fig-coords}
246: \end{figure*}
247: 
248: Because of the mass loss/gain, the center of mass of star $i$ at time
249: $t + \delta t$ is shifted from where it would have been had no mass
250: transfer taken place.  To describe this perturbation, we introduce an
251: additional right-handed coordinate frame $O_i X_i Y_i Z_i$ with a
252: spatial velocity such that its origin follows the unperturbed orbit of
253: star $i$, i.e., the origin of $O_iX_iY_iZ_i$ follows the path the
254: center of mass of star $i$ would have taken had no mass transfer
255: occurred.  Thus, at time $t$, the center of mass of star $i$ lies at the
256: origin of $O_i$, while at time $t+\delta t$ it has a non-zero position
257: vector with respect to $O_i$.  We furthermore let the $Z_i$-axis of the
258: $O_i X_i Y_i Z_i$ frame point in the direction of the orbital angular
259: momentum vector, and let the frame rotate synchronously with the 
260: unperturbed orbital angular velocity of the binary in the absence of
261: mass transfer. The direction and orientation of the $X_i$-axes are
262: then chosen such that at time $t$ the $X_i$-axis points along the
263: direction from the mass center of star $i$ to the mass center of its
264: companion.
265: 
266: To describe the shift in the mass center of star $i$ due to the mass
267: loss/gain, we denote the position vector of $O_i$ at times $t$ and $t
268: + \delta t$ with respect to the inertial frame by $\vec{R}_{_i}$ and
269: $\vec{R}_i^\prime$, respectively.  The position vector of the center
270: of mass of star $i$ at times $t$ and $t + \delta t$ is then given by
271: $\vec{R}_i$ and $\vec{R}_i^\prime + \delta \vec{r}_i^{\,\prime}$, where 
272: $\delta \vec{r}_i^{\,\prime}$ is the position vector of the
273: center of mass of star $i$ at time $t+\delta t$ with respect to
274: $O_i$.  Moreover, we denote by $\vec{r}_{A_i}$ 
275: and $\vec{r}_{A_i}^{\,\prime}$
276: the vectors from $O_i$ to the position where the point $A_i$ on the
277: stellar surface would be at times $t$ and $t + \delta t$, respectively,
278: had no mass been lost/accreted. The various position vectors at 
279: time $t + \delta t$ are related by
280: \begin{equation}
281: \left( M_i+\delta M_i \right) \left( \vec{R}_i^\prime 
282:   + \delta \vec{r}_i^{\,\prime} \right) = M_i\, \vec{R}_i^\prime 
283:   + \delta M_i\,\left( \vec{R}_i^\prime + \vec{r}_{A_i}^{\,\prime} 
284:   \right),
285: \label{eq-cm}
286: \end{equation}
287: which, at the lowest order of approximation in $\delta M_i$ and
288: $\delta \vec{r}_i^{\,\prime}$, yields
289: \begin{equation}
290: \delta \vec{r}_i^{\,\prime} = \frac{\delta M_i}{M_i}\, 
291:   \vec{r}_{A_i}^{\,\prime}.
292: \label{eq-DelR}
293: \end{equation}
294: As expected, the displacement of the center of mass
295: of star $i$ due to the mass loss/gain is directed along the line
296: connecting the center of mass of the star and the mass
297: ejection/accretion point.  
298: 
299: We furthermore denote with $\vec{\rho}_{A_i}^{\, \prime}$ the vector
300: from the center of mass of star $i$ at time $t+\delta t$ to the
301: position where the point $A_i$ would be at time $t+\delta t$, had no
302: mass been transferred between the binary components, and with 
303: $\delta\vec{\rho}^{\,\prime}_{A_i}$ the perturbation of this vector 
304: caused by the mass transfer.  It then follows
305: that $\vec{\rho}_{A_i}^{\, \prime} = \vec{r}_{A_i}^{\,\prime} - \delta
306: \vec{r}_i^{\,\prime}$ and thus, by definition, $\delta
307: \vec{\rho}_{A_i}^{\, \prime} = \delta \vec{r}_i^{\,\prime}$.  At the
308: lowest order of approximation in $\delta M_i$ and $\delta
309: \vec{r}_i^{\,\prime}$, equation~(\ref{eq-DelR}) therefore also yields
310: \begin{equation}
311: \delta \vec{\rho}_{A_i}^{\, \prime} = \frac{\delta
312:   M_i}{M_i}\, \vec{r}_{A_i}^{\, \prime}.
313: \label{eq-Delrho}
314: \end{equation}
315: The definitions of and the relations between these various position
316: vectors are illustrated schematically in Fig.~\ref{fig-coords}.
317: 
318: Next, we denote the absolute velocity of the center of mass of star
319: $i$ with respect to the inertial frame of reference at times $t$ and
320: $t + \delta t$ by $\vec{V}_i$ and $\vec{V}_i^\prime$, respectively,
321: and the absolute velocity of the ejected/accreted mass element by 
322: $\vec{W}_{\delta M_i}$.  The linear momentum $\vec{Q}_1$ of star~$1$ at 
323: time $t$ is then given by
324: \begin{equation}
325: \vec{Q}_1 = M_1\,\vec{V}_1,
326: \label{eq-Q1-1}
327: \end{equation}
328: and the total linear momentum $\vec{Q}_1^\prime$ of star~1 and the
329: ejected mass element at time $t+ \delta t$ by
330: \begin{equation}
331: \vec{Q}_1^\prime = (M_1 + \delta M_1) \vec{V}_1^\prime
332:   - \delta M_1\,\vec{W}_{\delta M_1}.
333: \label{eq-Q1-2}
334: \end{equation}
335: Similarly, the total linear momentum $\vec{Q}_2$ of star~2 and 
336: the mass element to be accreted at time $t$ is given by
337: \begin{equation}
338: \label{eq-Q2-1}
339: \vec{Q}_2 = M_2\,\vec{V}_2 + \delta M_i\,\vec{W}_{\delta M_i},
340: \end{equation}
341: and the linear momentum $\vec{Q}_2^\prime$ of star~2 at time $t+\delta 
342: t$ by
343: \begin{equation}
344: \label{eq-Q2-2}
345: \vec{Q}_2^\prime = (M_2 + \delta M_2) \vec{V}_2^\prime.
346: \end{equation}
347: 
348: At time $t+\delta t$ the velocity of the center of mass of star $i$
349: can be written as
350: \begin{equation}
351: \vec{V}_i^\prime = \vec{V}_{O_i}^\prime
352:    + \left( \vec{\Omega}_{\rm orb}^\prime 
353:    + \delta \vec{\Omega}_{\rm orb}^\prime
354:    \right) \times \delta \vec{r}_i^{\,\prime},  \label{eq-Vtdt}
355: \end{equation}
356: where $\vec{V}_{O_i}^\prime$ is the absolute velocity of the origin of
357: $O_i X_i Y_i Z_i$ at time $t + \delta t$, $\vec{\Omega}_{\rm
358: orb}^\prime$ is the orbital angular velocity of the binary at time
359: $t+\delta t$ in the absence of mass transfer, and $\delta
360: \vec{\Omega}_{\rm orb}^\prime$ is the perturbation of the orbital
361: angular velocity at time $t + \delta t$ due to the mass loss/gain of
362: the binary components.
363: 
364: In the limit of small $\delta t$, taking the difference between the 
365: linear momenta $\vec{Q}_i^\prime$ and $\vec{Q}_i$, dividing the 
366: resulting equation by $\delta t$, and noting that the absolute velocity
367: $\vec{V}_{O_i}$ of the origin of $O_i X_i Y_i Z_i$ at time $t$ is
368: equal to $\vec{V}_i$, yields
369: \begin{equation}
370: M_i \frac{d\vec{V}_{O_i}}{dt} = \vec{F}_i +
371:   \dot{M}_i\, \vec{U}_{\delta M_i}. \label{eq-dVi}  
372: \end{equation}
373: Here, $\vec{F}_i = d\vec{Q}_i/dt$ is the sum of all external forces
374: acting on star $i$, $\dot{M_i} = dM_i/dt$ is the mass loss/accretion
375: rate of star $i$, and  
376: \begin{equation}
377: \vec{U}_{\delta M_i} = \vec{W}_{\delta M_i} - \vec{V}_{O_i} 
378:   - \vec{\Omega}_{\rm orb} \times \vec{r}_{A_i}
379: \label{eq-Ui}
380: \end{equation}
381: is the relative velocity of the ejected/accreted mass element with
382: respect to the ejection/accretion point $A_i$. In the derivation of 
383: equation~(\ref{eq-Ui}), we have made use of equations (\ref{eq-DelR}) and 
384: (\ref{eq-Vtdt}) and restricted ourselves to first-order terms in 
385: the small quantities $\delta M_i$ and $\delta \vec{\Omega}_{\rm 
386: orb}^\prime$.
387: 
388: The absolute acceleration of the center of mass of star $i$ with
389: respect to the inertial frame $OXYZ$ is given by 
390: \begin{equation}
391: \frac{d^2 \vec{R}_i}{dt^2} = \vec{\gamma}_{O_i}
392:   + \vec{\gamma}_{\rm rel,i} + \vec{\gamma}_{\rm cor,i},
393: \end{equation}
394: where $\vec{\gamma}_{O_i} = d\vec{V}_{O_i}/dt$ is the
395: acceleration of the origin of $O_i X_i Y_i Z_i$ with respect to
396: $OXYZ$, $\vec{\gamma}_{\rm rel,i} = (\ddot{M}_i/M_i)\, \vec{r}_{A_i}$
397: is the relative acceleration of the center of mass of star $i$ with
398: respect to $O_i$, and $\vec{\gamma}_{\rm cor,i} = 2\,
399: (\dot{M}_i/M_i)\, (\vec{\Omega}_{\rm orb} \times \vec{r}_{A_i})$ is
400: the Coriolis acceleration of the center of mass of star $i$ with
401: respect to $O_i$\footnote{The centrifugal acceleration does not play a
402: role since it is proportional to $\delta \vec{r}_i^{\,\prime}$ which
403: vanishes for small $\delta t$.}. The expressions for
404: $\vec{\gamma}_{\rm rel,i}$ and $\vec{\gamma}_{\rm cor,i}$ follow from
405: the observation that $d\vec{\rho}^{\,\prime}_{A_i}/dt = 
406: (\dot{M}_i/M_i)\,
407: \vec{r}_{A_i}$, which one obtains by dividing equation~(\ref{eq-Delrho}) by
408: $\delta t$ in the limiting case of small $\delta t$.  The equation of
409: motion for the mass center of star $i$ with respect to the inertial
410: frame $OXYZ$ then becomes
411: \begin{eqnarray}
412: M_i \frac{d^2 \vec{R}_i}{dt^2} &=& \vec{F}_i 
413:   + \dot{M}_i \left( \vec{U}_{\delta M_i} + 2\,
414:   \vec{\Omega}_{\rm orb} \times \vec{r}_{A_i} \right)
415:   \nonumber \\ 
416:  &+& \ddot{M}_i\, \vec{r}_{A_i}.
417: \label{eq-mtn}
418: \end{eqnarray}
419: 
420: \subsection{Relative Motion of the Binary Components}
421: \label{relmot}
422: 
423: We can now obtain the equation describing the relative motion of the
424: accretor (star~2) with respect to the donor (star~1) by taking the
425: difference of the equations of motion of the stars with respect to
426: the inertial frame of reference. For convenience, we first decompose
427: the sum of the external forces acting on each star as
428: \begin{equation}
429: \vec{F}_i = - \frac{G\,M_1\,M_2}{\left| \vec{r} \right|^2}\,
430:   \frac{\vec{R}_i}{| \vec{R}_i |} + \vec{f}_i,
431: \end{equation}
432: where $G$ is the Newtonian constant of gravitation, and $\vec{f}_i$
433: the total gravitational force exerted on star $i$ by the particles in
434: the mass-transfer stream. It follows that
435: \begin{eqnarray}
436: \label{eq-R21}
437: \frac{d^2\vec{r}}{dt^2} &=& -\frac{G \left( M_1+M_2
438:     \right)}{\left| \vec{r} \right|^3}\, \vec{r} 
439:     + \frac{\vec{f}_2}{M_2}
440:     - \frac{\vec{f}_1}{M_1}  \nonumber \\
441: &+& \frac{\dot{M}_2}{M_2} \left( \vec{v}_{\delta M_2} +
442:     \vec{\Omega}_{\rm orb} \times \vec{r}_{A_2} \right) \nonumber \\
443: &-& \frac{\dot{M}_1}{M_1} \left( \vec{v}_{\delta M_1} +
444:     \vec{\Omega}_{\rm orb} \times \vec{r_{A_1}} \right) \\
445: &+& \frac{\ddot{M_2}}{M_2}\, \vec{r}_{A_2} 
446:   - \frac{\ddot{M_1}}{M_1}\, \vec{r}_{A_1}, \nonumber 
447: \end{eqnarray}
448: where $\vec{r} = \vec{R}_2 - \vec{R}_1$ is the position vector of the
449: accretor with respect to the donor, and $\vec{v}_{\delta M_i} =
450: \vec{W}_{\delta M_i} - \vec{V}_{i}$ is the velocity of the
451: ejected/accreted mass element with respect to the mass center of the
452: mass losing/gaining star. 
453: 
454: Equation~(\ref{eq-R21}) can be written in the form of a perturbed
455: two-body problem as
456: \begin{equation}
457: \frac{d^2\vec{r}}{dt^2} = -\frac{G \left( M_1+M_2 \right)}
458:   {\left| \vec{r} \right|^3}\,
459:   \vec{r} + S\, \hat{x} + T\, \hat{y} + W\, \hat{z},
460: \end{equation}
461: where $\hat{x}$ is a unit vector in the direction of $\vec{r}$,
462: $\hat{y}$ is a unit vector in the orbital plane perpendicular to
463: $\vec{r}$ in the direction of the orbital motion, and $\hat{z}$ is a
464: unit vector perpendicular to the orbital plane parallel to and in the
465: same direction as $\vec{\Omega}_{\rm orb}$.  The functions $S$, $T$,
466: and $W$ are found by taking the dot product of the perturbing force 
467: arising from the mass transfer between the binary components and the 
468: unit vector in the $\hat{x}$, $\hat{y}$, and $\hat{z}$ directions, 
469: respectively.  These vector components are
470: \begin{eqnarray}
471: S &=& \frac{f_{2,x}}{M_2} - \frac{f_{1,x}}{M_1}  
472: + \frac{\dot{M}_2}{M_2} \left( v_{\delta M_2,x} - |\vec{\Omega}_{\rm orb}|
473: |\vec{r}_{A_2}| \sin{\phi} \right) \nonumber \\
474: &-& \frac{\dot{M}_1}{M_1} v_{\delta M_1,x} 
475: + \frac{\ddot{M}_2}{M_2} |\vec{r}_{A_2}| \cos{\phi}
476: - \frac{\ddot{M}_1}{M_1} |\vec{r}_{A_1}|, \label{eq-S} \\
477: T &=& \frac{f_{2,y}}{M_2} - \frac{f_{1,y}}{M_1}  
478: + \frac{\dot{M}_2}{M_2} \left( v_{\delta M_2,y} 
479: + |\vec{\Omega}_{\rm orb}| |\vec{r}_{A_2}| \cos{\phi} \right) \nonumber \\ 
480: &-&\frac{\dot{M}_1}{M_1} \left( v_{\delta M_1,y}
481: + |\vec{\Omega}_{\rm orb}||\vec{r}_{A_1}| \right) +
482: \frac{\ddot{M}_2}{M_2} |\vec{r}_{A_2}| \sin{\phi}, \label{eq-T} \\
483: W &=& \frac{f_{2,z}}{M_2} - \frac{f_{1,z}}{M_1}, \label{eq-W}
484: \end{eqnarray}
485: where $\phi$ is the angle between $\hat{x}$ and the vector from the 
486: center of mass of the accretor to the mass
487: accretion point $A_2$, and the subscripts $x$, $y$, and $z$ 
488: denote vector components in the $\hat{x}$, $\hat{y}$, and $\hat{z}$ 
489: directions, respectively. In working out the vector products
490: $\vec{\Omega}_{\rm orb} \times \vec{r}_{A_2}$, we assumed that $A_2$ is 
491: located on the equator of the accreting star\footnote{For brevity, we refer to the point $A_2$ as lying on the 
492: stellar surface. Though, in practice, it can lie at any point near the 
493: star where the 
494: transferred mass can be considered to be part of the accretor. 
495: For instance, if an accretion disk has formed around the 
496: accretor, it would be equally valid to write $A_2$ as the point 
497: where the transferred mass impacts the outer edge of the accretion 
498: disk.}.  The terms contributing to the perturbed orbital motion can be 
499: categorized as follows: (i) term proportional to $\vec{f}_i$ represent 
500: gravitational perturbation on the binary components caused by mass 
501: elements in the mass-transfer stream; (ii) terms proportional to 
502: $\dot{M}_i$ represent linear momentum exchange between the mass donor 
503: and accretor; and (iii) terms proportional to $\ddot{M}_i$ represent 
504: shifts in the position of the mass centers of the mass donor and 
505: accretor due to the non-spherical symmetry of the mass loss or gain.  In 
506: the limiting case where both stars are treated as point masses
507: ($|\vec{r}_{A_1}| \rightarrow 0$ and $|\vec{r}_{A_2}| \rightarrow 0$), 
508: the only non-zero terms in the perturbed equations of motion are those 
509: due to gravitational perturbations of the mass transfer stream and the 
510: transport of linear momentum.
511: 
512: \subsection{Comparison with Previous Work}
513: \label{sec-appb}
514: 
515: The most recent study on the orbital evolution of eccentric
516: mass-transferring binaries has been presented by \citet[][hereafter MW83
517: and MW84, respectively]{MW83, MW84}. These authors extended the work
518: of \citet{H56}, \citet{K64-2}, and \citet{P64} by accounting for the
519: effects of linear momentum transport between the binary components, as
520: well as possible perturbations to the orbital motion caused by the
521: mass transfer stream. However, they also derived the equations
522: describing the motion of the binary components with respect to a frame
523: of reference with origin at the mass center of the binary, which, for
524: mass transferring systems, is not an inertial frame of reference. The
525: equations therefore do not account for the accelerations of the binary
526: mass center caused by the mass transfer. Here, we
527: demonstrate that if the procedure adopted by Matese \& Whitmire is
528: developed with respect to an inertial frame of reference that is not
529: connected to the binary, the resulting equations are in agreement with
530: those derived in \S\,\ref{relmot}.
531: 
532: The core of Matese \& Whitmire's derivation is presented in Section II
533: of MW83. While the authors choose to adopt a reference frame with
534: origin at the binary mass center early on in the investigation, the
535: choice of the frame does not affect the derivation of the equations of
536: motion up to and including their equation~(24). In particular, equation~(13) in
537: MW83, which, in our notation, reads
538: \begin{equation}
539: \label{eq-apb-p}
540: \vec{p}_i = M_i \dot{\vec{R}}_i - \dot{M}_i\vec{r}_{A_i}, 
541: \end{equation}
542: is valid with respect to any inertial frame of reference with
543: arbitrary position and orientation in space. The same applies to equation (1)
544: in MW84:
545: \begin{equation}
546: \label{eq-apb-pdot}
547: \dot{\vec{p}}_i = -G M_1 M_2\,
548:   \frac{\vec{R}_i-\vec{R}_{3-i}}{|\vec{R}_i-\vec{R}_{3-i}|^3}
549:   + \vec{f}_i + \vec{\Psi}_i.
550: \end{equation}
551: In these equations, $\vec{p}_i$ is the linear momentum of star $i$,
552: and $\vec{\Psi}_i=\dot{M}_i (\dot{\vec{R}}_i+\vec{v}_{\delta M_i} )$
553: is the amount of linear momentum transported by the transferred mass
554: per unit time (see equation~(3) of MW84). Substitution of
555: equation~(\ref{eq-apb-p}) into equation~(\ref{eq-apb-pdot}) then yields
556: \begin{eqnarray}
557: M_i\, \ddot{\vec{R}}_i = &-& GM_{3-i}\,
558:   \frac{\vec{R}_i-\vec{R}_{3-i}}{|\vec{R}_i-\vec{R}_{3-i}|^3} \nonumber \\
559:   &+& \frac{\vec{f}_i}{M_i} + \frac{\dot{M}_i}{M_i} 
560:   \left( \vec{v}_{\delta M_i} +
561:   \dot{\vec{r}}_{A_i} \right) + \frac{\ddot{M}_i}{M_i}\vec{R}_{A_i},
562: \label{eq-apb-mrddot}
563: \end{eqnarray}
564: and thus
565: \begin{eqnarray}
566: \label{eq-apb-eqm}
567: \frac{d^2\vec{r}}{dt^2} = &-&\frac{G \left( M_1+M_2
568:   \right)}{\left| \vec{r} \right|^3}\, \vec{r}
569:   + \frac{\vec{f}_2}{M_2} - \frac{\vec{f}_1}{M_1} \nonumber \\
570:   &+& \frac{\dot{M}_2}{M_2} \left( \vec{v}_{\delta M_2} +
571:     \dot{\vec{r}}_{A_2} \right)
572:   - \frac{\dot{M}_1}{M_1} \left( \vec{v}_{\delta M_1} +
573:     \dot{\vec{r}}_{A_1} \right) \nonumber \\
574:   &+& \frac{\ddot{M_2}}{M_2}\, \vec{r}_{A_2} 
575:   - \frac{\ddot{M_1}}{M_1}\, \vec{r}_{A_1}, 
576: \end{eqnarray}
577: where $\vec{r} = \vec{R}_2 - \vec{R}_1$.  Setting $\dot{\vec{r}}_{A_i}
578: = \vec{\Omega}_{\rm orb} \times \vec{r}_{A_i}$, this equation is in
579: perfect agreement with equation~(\ref{eq-R21}) derived in
580: \S\,\ref{relmot}.
581: 
582: In MW83 and MW84, the authors incorrectly set $\vec{R}_1=-M_2\,
583: \vec{r}/(M_1+M_2)$ and $\vec{R}_2=M_1\, \vec{r}/(M_1+M_2)$ in
584: equation~(\ref{eq-apb-mrddot}), which is valid only when the origin of the
585: frame of reference coincide with the mass center of the binary. For a
586: mass-transferring binary, such a frame is, however, not an inertial
587: frame and can therefore not be used for the derivation of the
588: equations of motion of the binary components. Instead of
589: equation~(\ref{eq-apb-eqm}), Matese \& Whitmire therefore find
590: equations~(7)--(8) in MW84, which lack the terms associated with the
591: acceleration of the binary mass center due to the mass transfer.
592: 
593: \section{Orbital Evolution Equations}
594: \label{sec-sec}
595: 
596: \subsection{Secular Variation of the Orbital Elements}
597: 
598: In the classical framework of the theory of osculating elements, the
599: equations governing the rate of change of the orbital semi-major axis 
600: $a$ and eccentricity $e$ due to mass transfer are obtained from the 
601: perturbing functions $S$ and $T$ as \citep[see, e.g.,][]{S60, BC61, D62, 
602: F70}
603: \begin{equation}
604: \frac{da}{dt} = \frac{2}{n(1-e^2)^{1/2}} [ S e \sin{\nu} + T ( 1
605:   + e \cos{\nu} )],
606: \label{eq-dadt}
607: \end{equation}
608: \begin{eqnarray}
609: \lefteqn{\frac{de}{dt} = \frac{(1-e^2)^{1/2}}{na}} 
610:   \nonumber \\
611:  & & \times \left\{ S \sin{\nu} +
612:   T \left[ \frac{2\cos{\nu} +e \left(1+\cos^2{\nu}
613:   \right)}{1+e\cos{\nu}} \right] \right\},
614: \label{eq-dedt}
615: \end{eqnarray}
616: where $n=2\pi/P_{\rm orb}$ is the mean motion and $\nu$ the true
617: anomaly. These equations are independent of the perturbing function
618: $W$ which solely appears in the equations governing the rates of
619: change of the orbital inclination, the longitude of the ascending
620: node, and the longitude of the periastron. 
621: 
622: After substitution of equations~(\ref{eq-S}) and (\ref{eq-T}) for $S$ and
623: $T$ into equations~(\ref{eq-dadt}) and~(\ref{eq-dedt}), the equations
624: governing the rates of change of the semi-major axis and eccentricity
625: contain periodic as well as secular terms. Here we are mainly
626: interested in the long-term secular evolution of the orbit, and so we
627: remove the periodic terms by averaging the equations over one orbital
628: period:
629: \begin{equation}
630: \left< {\frac{da}{dt}} \right>_{\rm sec} \equiv 
631:   \frac{1}{P_{\rm orb}} \int_{-P_{\rm orb}/2}^{P_{\rm orb}/2}
632:   {\frac{da}{dt}}\, dt,
633: \label{eq-dadtsec}
634: \end{equation}
635: \begin{equation}
636: \left< {\frac{de}{dt}} \right>_{\rm sec} \equiv 
637:   \frac{1}{P_{\rm orb}} \int_{-P_{\rm orb}/2}^{P_{\rm orb}/2}
638:   {\frac{de}{dt}}\, dt.
639: \label{eq-dedtsec}
640: \end{equation}
641: The integrals in these definitions are most conveniently computed in
642: terms of the true  anomaly, $\nu$. We therefore make a change of 
643: variables using
644: \begin{equation}
645: dt=\frac{(1-e^2)^{3/2}}{n(1+e\cos{\nu})^2}\,d\nu.
646: \label{eq-dnudt}
647: \end{equation}
648: For binaries with eccentric orbits, the resulting integrals can be
649: calculated analytically only for very specific functional
650: prescriptions of the mass-transfer rate $\dot{M}_1$ (e.g., when
651: $\dot{M}_1$ is approximated by a Dirac delta function centered on the
652: periastron, see \S\,\ref{sec-delta}). In general, the
653: integrals must be computed numerically.
654: 
655: \subsection{Conservation of Orbital Angular Momentum}
656: 
657: \label{sec-angmom}
658: 
659: Since the perturbing functions $S$ and $T$ depend on the properties
660: of the mass transfer stream, calculation of the rates 
661: of secular
662: change of the orbital semi-major axis and eccentricity, in principle,
663: requires the calculation of the trajectories of the particles in the
664: stream \citep[cf.][]{H69b}. As long as no mass is lost from the
665: system, such a calculation automatically incorporates the conservation
666: of total angular momentum in the system. Special cases of angular
667: momentum conservation can, however, be used to bypass the
668: calculation of detailed particle trajectories. Here, we adopt such a
669: special case and assume that any orbital angular momentum carried by
670: the particles in the mass-transfer stream is always immediately
671: returned to the orbit, so that the orbital angular momentum of the
672: binary is conserved.
673: 
674: The orbital angular momentum of a binary with a semi-major axis $a$ and 
675: eccentricity $e$ is given by
676: \begin{equation}
677: J_{\rm orb} = M_1M_2 \left[ \frac{Ga(1-e^2)}{M_1+M_2} \right]^{1/2},
678: \label{eq-J}
679: \end{equation}
680: so that
681: \begin{equation}
682: \frac{\dot{J}_{\rm orb}}{J_{\rm orb}} = \frac{\dot{M_1}}{M_1} +
683: \frac{\dot{M_2}}{M_2}  
684: - \frac{1}{2}\frac{\dot{M_1}+\dot{M_2}}{M_1+M_2}
685: + \frac{1}{2}\frac{\dot{a}}{a} - \frac{e\,\dot{e}}{1-e^2},
686: \label{eq-Jdot}
687: \end{equation}
688: where a dot indicates the time derivative.
689: 
690: In the case of eccentric orbits, substitution of 
691: equations~(\ref{eq-dadt}) and (\ref{eq-dedt}) into equation 
692: (\ref{eq-Jdot}) leads to 
693: \begin{eqnarray}
694: \frac{\dot{J}_{\rm orb}}{J_{\rm orb}} &=& 
695:   \frac{\dot{M_1}}{M_1} + \frac{\dot{M_2}}{M_2} 
696:   - \frac{1}{2}\frac{\dot{M_1}+\dot{M_2}}{M_1+M_2} \nonumber \\ 
697: &+& \frac{(1-e^2)^{1/2}}{n\,a \left(1+e\cos{\nu} \right)}\,T.
698: \label{eq-JT}
699: \end{eqnarray}
700: As we shall see in the next section, by setting $\dot{M}_1 + \dot{M}_2 =
701: 0$ and $< \dot{J}_{\rm orb}/J_{\rm orb} >_{\rm sec} =0$ and substituting
702: equation~(\ref{eq-T}) for $T$, equation~(\ref{eq-JT}) allows us to
703: calculate the $\hat{y}$-component of the final velocities of the 
704: accreting
705: particles as a function of their initial velocities without resorting to
706: the computation of the ballistic trajectories of the mass transfer
707: stream.  
708: 
709: In the limiting case of a circular orbit, equation~(\ref{eq-Jdot}) is 
710: usually used to derive the rate of change of the semi-major axis of 
711: circular binaries under the assumption of conservation of both total 
712: mass ($\dot{M}_1=-\dot{M}_2$) and orbital angular momentum ($\dot{J}_{\rm 
713: orb}=0$):
714: \begin{equation}
715: \frac{da}{dt} = 2a(\frac{M_1}{M_2}-1)\frac{\dot{M}_1}{M_1}.
716: \label{circ-da}
717: \end{equation}
718: 
719: The assumption of orbital angular momentum conservation over secular
720: timescales ($< \dot{J}_{\rm orb}/J_{\rm orb} >_{\rm sec} =0$) is a
721: standard assumption in nearly all investigations of conservative mass
722: transfer in binary systems \citep[e.g., ][]{SPH97, P98}, which is valid
723: over long timescales provided there is no significant storage of angular
724: momentum in the spins of the components stars, the accretion flow,
725: and/or the accretion disk.  In future work, we will investigate the
726: consequences of both mass and orbital angular momentum losses from the
727: binary on the evolution of the orbital elements.
728: 
729: \section{Orbital Evolution Timescales}
730: \label{orbevtim}
731: 
732: In order to assess the timescales of orbital evolution due to mass
733: transfer in eccentric binaries, we observe that, for eccentric binaries, 
734: mass transfer is expected to occur first at the periastron of the 
735: relative orbit, where the component stars are closest to each other.  We 
736: therefore explore the order of magnitude of the timescales assuming a 
737: delta function mass transfer profile centered at the periastron of the 
738: binary orbit
739: \begin{equation}
740: \dot{M}_1 = \dot{M}_0\, \delta \left( \nu \right),
741: \label{eq-del}
742: \end{equation}
743: where $\dot{M}_0 < 0 $ is the instantaneous mass transfer rate, and 
744: $\delta(\nu)$ is the Dirac delta function.
745: 
746: We calculate the rates of secular change of the orbital semi-major axis
747: and eccentricity from equations~(\ref{eq-dadt})--(\ref{eq-dnudt}), and 
748: neglect any gravitational attractions exerted by the particles in the
749: mass-transfer stream on the component stars. Hence, we set
750: \begin{eqnarray}
751: f_{1,x} &=& f_{2,x} = 0, \label{eq-f12x} \\
752: f_{1,y} &=& f_{2,y} = 0. \label{eq-f12y}
753: \end{eqnarray}
754: 
755: \label{sec-delta}
756: 
757: Substituting equations~(\ref{eq-T}), (\ref{eq-dnudt}), and 
758: (\ref{eq-del}) -- (\ref{eq-f12y}) into equation~(\ref{eq-JT}) for $< 
759: \dot{J}_{\rm orb}/J_{\rm orb} >_{\rm sec}=0$ then yields a relationship 
760: between the initial and final $\hat{y}$-component of the velocities of 
761: the transferred mass and the initial and final positions of the 
762: transferred mass given by
763: \begin{eqnarray}
764: qv_{\delta M_2,y} &+& v_{\delta M_1,y} = 
765:  na(1-q)\left(\frac{1+e}{1-e}\right)^{1/2}
766: - |\vec{\Omega}_{{\rm orb}, P}||\vec{r}_{A_1,P}| \nonumber \\
767: &-& q|\vec{\Omega}_{{\rm orb}, P}||\vec{r}_{A_2}|\cos{\phi_P} 
768: \left( \left.1- \frac{d\phi}{d\nu}\right|_{\nu=0} \right),
769: \label{eq-vr}
770: \end{eqnarray}
771: where the subscript $P$ indicates quantities evaluated at the periastron
772: of the binary orbit, $q=M_1/M_2$ is the binary mass ration, and we have 
773: used the relation
774: $d\delta(\nu)/d\nu = -\delta(\nu)/\nu$.  Assuming the transferred mass
775: elements are ejected by star~1 at the $L_1$ point with a velocity
776: $\vec{v}_{\delta M_1}$ equal to the star's rotational velocity at $L_1$,
777: we write
778: \begin{eqnarray}
779: \label{eq-vdm1x}
780: v_{\delta M_1, x} = 0,\\
781: \label{eq-vdm1y}
782: v_{\delta M_1,y}=-|\vec{\Omega}_{{\rm orb},P}||\vec{r}_{A_1, P}|.
783: \end{eqnarray}
784: Moreover, under the assumption that each periastron passage of the
785: binary components give rise to an extremum of $\phi(\nu)$, the
786: derivative $d\phi/d\nu|_{\nu=0}$ is equal to zero in
787: equation~(\ref{eq-vr}), so that
788: \begin{equation}
789: v_{\delta M_2, y} = |\vec{\Omega}_{{\rm orb},P}|\left[ |\vec{r}_{A_1,P}| 
790: \frac{(1-q)}{q} + |\vec{r}_{A_2}|\cos{\phi_P} \right].
791: \label{eq-vrel}
792: \end{equation}
793: 
794: For a binary with orbital period $P_{\rm orb} = 1\,{\rm day}$,
795: eccentricity $e=0.2$, and component masses $M_1=2\,M_\sun$ and
796: $M_2=1.44\,M_\sun$, with $|\vec{r}_{A_1,P}|$ the distance from star~1 to
797: $L_1$ (See Appendix A), and $|\vec{r}_{A_2}|\cos{\phi_P}\approx
798: 2.9\times10^5\,{\rm km}$ (the circularization radius around a compact
799: object for these binary parameters; see \citet{FKR}), the accreting
800: matter has a $\hat{y}$-velocity component of the order of $\sim-45\,{\rm
801: km}\,{\rm s}^{-1}$.  
802: 
803: 
804: After substitution of equation~(\ref{eq-dnudt}) and 
805: equations~(\ref{eq-f12x})--(\ref{eq-vr}), the integrals in
806: equations~(\ref{eq-dadtsec}) and (\ref{eq-dedtsec}) for the rates of
807: secular change of the orbital semi-major axis and eccentricity can be
808: solved analytically to obtain
809: \begin{eqnarray}
810: \left< \frac{da}{dt} \right>_{\rm sec} &=& 
811: \frac{a}{\pi}\frac{\dot{M}_0}{M_1}\frac{1}{(1-e^2)^{1/2}}\left[
812: qe\frac{|\vec{r}_{A_2}|}{a}\cos{\phi_P} \right. \nonumber \\
813:  &+& \left. e\frac{|\vec{r}_{A_1,P}|}{a} +
814: (q-1)(1-e^2) \right],
815: \label{eq-deltaa} 
816: \end{eqnarray}
817: \begin{eqnarray}
818: \left< \frac{de}{dt} \right>_{\rm sec} &=& 
819: \frac{(1-e^2)^{1/2}}{2\pi} \frac{\dot{M}_0}{M_1}\left[ 
820: q\frac{|\vec{r}_{A_2}|}{a}\cos{\phi_P} \right. \nonumber \\
821: &+& \left. \frac{|\vec{r}_{A_1,P}|}{a} + 2(1-e)(q-1) \right].
822: \label{eq-deltae}
823: \end{eqnarray}
824: We note that for a delta-function mass transfer
825: rate given by equation~(\ref{eq-del}) the $\hat{x}$-component of the
826: velocities $\vec{v}_{\delta M_1}$ and $\vec{v}_{\delta M_2}$ does not
827: enter into the derivation of theses equations due to the $\sin{\nu}$ 
828: term in equations~(\ref{eq-dadt}) and (\ref{eq-dedt}).  Furthermore, in 
829: the limiting case of a circular orbit,
830: equation~(\ref{eq-deltaa}) reduces to equation~(\ref{circ-da}), provided that
831: $\dot{M_1}$ in that equation is interpreted as the secular mean mass
832: transfer rate $\left<\right.\!\dot{M_1}\!\left.\right>_{\rm
833: sec}=\dot{M}_0/(2\pi)$.  In Appendix~B, we present an alternative
834: derivations to equations~(\ref{eq-deltaa}) and (\ref{eq-deltae}) in the
835: limiting case where the stars are treated as point masses.
836: 
837: The rates of secular change of the semi-major axis and orbital
838: eccentricity are thus linearly proportional to the magnitude of the
839: mass transfer rate at periastron. Besides the obvious dependencies on
840: $a$, $e$, $q$, and $M_1$, the rates also depend on the ratio of the
841: donor's rotational angular velocity $\Omega_1$ to the orbital angular
842: velocity $\Omega_{{\rm orb},P}$ at periastron through the position of 
843: the $L_1$
844: point, $\vec{r}_{A_1}$. A fitting formula for the position of the
845: $L_1$ point accurate to better than $4\,\%$ over a wide range of $q$,
846: $e$, and $\Omega_1/\Omega_{{\rm orb},P}$ is given by equation~(\ref{XL1fit}) in
847: Appendix~\ref{sec-appa}. While the fitting formula can be used to
848: obtain fully analytical rates of secular change of the semi-major axis
849: and eccentricity, we here use the exact solutions for the
850: position of $L_1$ obtained by numerically solving equation~(\ref{XL}) in
851: Appendix~\ref{sec-appa}.  For a detailed discussion of the 
852: properties of the $L_1$ point in eccentric binaries, we refer the 
853: interested reader to \citet{SWK07}.
854: 
855: \begin{figure*}
856: \plottwo{f2a.eps}{f2b.eps}
857: \caption{Orbital evolution timescales for a delta function mass
858:   transfer profile centered at the periastron of the binary orbit with 
859:   an instantaneous mass transfer rate of $\dot{M}_0 = - 
860:   10^{-9}\,M_\sun\,{\rm yr}^{-1}$. The timescales are calculated under
861:   the assumption that the donor rotates synchronously with the orbital
862:   angular velocity at periastron, and that the accretor is a
863:   $1.44\,M_\odot$ neutron star.  Shown at left (right) are the
864:   timescales for the evolution of the semi-major axis (eccentricity) as
865:   a function of the mass ratio, $q$, for a range of eccentricities, $e$.
866:   Regimes where the timescale is negative correspond to a decrease of
867:   the semi-major axis (eccentricity), while regimes where the
868:   timescale is positive correspond to an increase of the semi-major
869:   axis (eccentricity).
870: }
871: \label{delta1}
872: \end{figure*}
873: 
874: To explore the effects of mass transfer on the orbital elements of
875: eccentric binaries, we calculate the rates of secular change of the
876: semi-major axis and eccentricity and determine the characteristic
877: timescales $\tau_a = a/\dot{a}$ and $\tau_e = e/\dot{e}$. While the
878: actual timescales are given by the absolute values of $\tau_a$
879: and $\tau_e$, we here allow the timescales to be negative as well as
880: positive in order to distinguish negative from positive rates of
881: secular change of the orbital elements. We also note that since
882: $|\vec{r}_{A_1, P}| \propto a$ (see Appendix A), the timescales do not
883: explicitly depend on the orbital semi-major axis $a$ except through 
884: the ratio $|\vec{r}_{A_2}|/a$ of the radius of the accretor to the 
885: semi-major axis.  For convenience, we therefore assume the accretor to 
886: be a compact object with radius $|\vec{r}_{A_2}| << a$.  The timescales 
887: are found to be insensitive to terms containing $|\vec{r}_{A_2}|/a$ in 
888: equation (\ref{eq-deltaa}) and (\ref{eq-deltae}).  Varying 
889: $|\vec{r}_{A_2}|$ from $0$ to $0.01a$ changes the timescales by less 
890: that 10\%.  In what follows, we therefore set $|\vec{r}_{A_2}|=0$.  An
891: implicit dependence on $a$ may then still occur through the amplitude 
892: $\dot{M}_0$ of the mass transfer rate at periastron. Since incorporating 
893: such a dependence in the analysis requires detailed modeling of the 
894: evolution of the donor star, which is beyond the scope of this 
895: investigation, we here restrict ourselves to exploring the timescales of 
896: orbital evolution for a constant $\dot{M}_0$. The linear dependence of 
897: $\left <\dot{a} \right >_{\rm sec}$ and $\left < \dot{e} \right >_{\rm 
898: sec}$ on $\dot{M}_0$ in any case allows for any easy rescaling of 
899: our results to different mass transfer rates.  
900: 
901: In Fig.~\ref{delta1}, we show the variations of $\tau_a$ and $\tau_e$ as
902: functions of $q$ for $\dot{M}_0 = -10^{-9}\,M_\sun\,{\rm yr^{-1}}$ and
903: $e=0.0,0.1, \ldots, 0.9$. In all cases, the donor is assumed to rotate
904: synchronously with the orbital angular velocity at the periastron, and
905: the accretor is assumed to be a neutron star of mass
906: $M_2=1.44\,M_\odot$.  The timescales of the secular evolution of the
907: semi-major axis show a strong dependence on $q$, and a milder dependence
908: on $e$, unless $e \ga 0.7$. The timescales for the secular evolution of
909: the orbital eccentricity always depend strongly on both $q$ and $e$. 
910: These timescales can furthermore be positive as well as negative, so
911: that the semi-major axis and eccentricity can increase as well as
912: decrease under the influence of mass transfer at the periastron of the 
913: binary orbit.
914: 
915: From Fig.~\ref{delta1}, as well as equations~(\ref{eq-deltaa}) and
916: (\ref{eq-deltae}), it can be seen that, for a given ratio of the donor's
917: rotational angular velocity to the orbital angular velocity at
918: periastron, the line dividing positive from negative rates of secular
919: change of the orbital elements is a function of $q$ and $e$. This is
920: illustrated further in Fig.~\ref{contours} where the timescales of
921: orbital evolution are displayed as contour plots in the $(q,e)$-plane.
922: The thick black line near the center of the plots marks the transition
923: values of $q$ and $e$ where the rates of secular change of $a$ and $e$
924: transition from being positive (to the left of the thick black line) to
925: negative (to the right of the thick black line). Varying
926: $\Omega_1/\Omega_{{\rm orb},P}$ between $0.5$ and $1.5$ changes the
927: position of the transition line by less than $10\%$ in comparison to the
928: $\Omega_1/\Omega_{{\rm orb},P}=1$ case displayed in Figures~\ref{delta1}
929: and \ref{contours}.
930: 
931: \begin{figure*}
932: \plottwo{f3a.eps}{f3b.eps}
933: \caption{Contour plot of the orbital evolution timescales for the 
934:   semi-major axis and eccentricity in the
935:   $(q,e)$-plane for the same set of assumptions as adopted in 
936:   figure~\ref{delta1}.  Timescales for the evolution of the 
937:   semi-major axis are shown
938:   on the left; timescales for the evolution of the orbital
939:   eccentricity on the right. The different shades of gray designate
940:   regions of the $(q,e)$ parameter space with timescales (in Gyr) in the 
941:   ranges labeled in the plots.  The thick black line near the center
942:   of each plot designates the transition point where the rate of
943:   change of the semi-major axis or orbital eccentricity changes from
944:   positive (to the left of the thick black line) to negative (to 
945:   the right of the thick black line).}
946: \label{contours}
947: \end{figure*}
948: 
949: In the limiting case of a circular orbit, the orbit expands when
950: $q<1$ and shrinks when $q>1$, in agreement with the classical result
951: obtained from equation~(\ref{circ-da}). For non-zero eccentricities, the
952: critical mass ratio separating positive from negative values of $\left
953: < \dot{a} \right >_{\rm sec}$ decreases with increasing orbital
954: eccentricities. This behavior can be understood by substituting the
955: fitting formula for the position of the $L_1$ point given by
956: equation~(\ref{XL1fit}) in Appendix~\ref{appL1} into equation~(\ref{eq-deltaa})
957: and setting $\left< \dot{a} \right>_{\rm sec} =0$.  However, we can fit 
958: the critical mass ratio separating expanding from shrinking orbits with 
959: a simpler formula given by
960: \begin{equation}
961: q_{\rm crit} \simeq 1 - 0.4e + 0.18e^2.
962: \label{eq-dadttran}
963: \end{equation}
964: 
965: The critical mass ratio separating positive from negative values of
966: $\left < \dot{e} \right >_{\rm sec}$ is largely independent of
967: $e$. Proceeding in a similar fashion as for the derivation of
968: equation~(\ref{eq-dadttran}), we derive the critical mass ratio separating
969: increasing from decreasing eccentricities to be approximately given by
970: \begin{equation}
971: q_{\rm crit} \simeq 0.76 + 0.012e.
972: \label{eq-dedttran}
973: \end{equation}
974: 
975: Last, we note that a more quantitative numerical comparison between
976: the above approximation formulae for $q_{\rm crit}$ and the exact
977: numerical solutions shows that equations~(\ref{eq-dadttran}) and
978: (\ref{eq-dedttran}) are accurate to better than 1\%.
979: 
980: \section{Tidal Evolution Timescales}
981: 
982: A crucial question for assessing the relevance of the work
983: presented here is how the derived orbital evolution timescales compare
984: to the corresponding timescales associated with other orbital
985: evolution mechanisms such as tides. In Fig.~\ref{fig-tid-lin}, we show
986: the secular evolution timescales of the semi-major axis and orbital
987: eccentricity of a mass-transferring binary due to tidal dissipation in
988: the donor star as a function of $q$, for different values of the 
989: eccentricity, $e$. The timescales are strong functions of 
990: $|\vec{r}_{A_1}|/a$ and are determined as in \citet{HTP02}\footnote{Note 
991: that there is a typo in equation~(42) of \citet{HTP02}. The correct equation 
992: for $k/T$ for stars with radiative envelopes is (J. Hurley, Private 
993: Communication)
994: \[
995: \left( k/T\right)_r =
996:   1.9782 \times 10^4\left( MR^2/a^5 \right)^{1/2}
997:   \left(1+q_2\right)^{5/6}E_2\,{\rm yr}^{-1}. \nonumber
998: \]}
999: \citep[see also][]{Z77,Z78,H81}.  The radius $|\vec{r}_{A_1}|$ is
1000: determined by assuming the donor is on the zero-age
1001: main sequence and that the orbital separation is then obtained by 
1002: equating the radius of the donor \citep[given by][]{TEA96} to the 
1003: volume-equivalent radius of its Roche lobe at the periastron of the 
1004: binary orbit \citep[see][]{SWK07}.  As before, we assume the donor
1005: rotates synchronously with the orbital motion at periastron and that the 
1006: accretor is a $1.44\,{\rm M}_\sun$ neutron star.
1007: 
1008: \begin{figure*}
1009: \plottwo{f4a.eps}{f4b.eps}
1010: \caption{Timescales of orbital evolution due to tidal dissipation in a
1011:   Roche-lobe filling component of a close binary under the assumption
1012:   that the donor is a zero-age main-sequence star rotating
1013:   synchronously with the orbital angular velocity at the periastron,
1014:   and the accretor is a $1.44\,M_\odot$ neutron star.  Shown at left
1015:   (right) are the timescales for the evolution of the semi-major axis
1016:   (eccentricity) as a function of the mass ratio, $q$, for 
1017:   different orbital eccentricities $e$.  Regimes
1018:   where the timescales are negative correspond to a decrease of the
1019:   semi-major axis (eccentricity), while regimes where the timescales
1020:   are positive correspond to an increase of the semi-major axis
1021:   (eccentricity). The discontinuity at $q \simeq 0.87$ corresponds to
1022:   the transition from donor stars with convective envelopes ($M_1 \la
1023:   1.25\,M_\odot$) to donor stars with radiative envelopes ($M_1 \ga
1024:   1.25\,M_\odot$).}
1025: \label{fig-tid-lin}
1026: \end{figure*}
1027: 
1028: The timescales of orbital evolution due to tides range from a few Myr to
1029: more than a Hubble time, depending on the binary mass ratio and the
1030: orbital eccentricity. The discontinuity in the timescales at $q \simeq
1031: 0.87$ corresponds to the transition from donor stars with convective 
1032: envelopes ($M_1 \la 1.25\,M_\odot$) to donor stars with radiative 
1033: envelopes ($M_1  \ga 1.25\,M_\odot$) which are subject to different 
1034: tidal dissipation mechanisms.  It follows that tides do not 
1035: necessarily lead to rapid circularization during the early stages of 
1036: mass transfer, especially for orbital eccentricities $e \ga 0.3$. 
1037: Furthermore, for the adopted system parameters, the orbital eccentricity 
1038: always decreases, while the orbital semi-major axis can either increase 
1039: or decrease. Hence, in some regions of the parameter space, the effects 
1040: of tides and mass transfer are additive, while in other regions they are
1041: competitive.  This is illustrated in more detail in 
1042: figure~\ref{fig-tot_timescale} where we show the orbital evolution 
1043: timescales due to the combined effect of tides and mass transfer.  In 
1044: the calculations of the timescales, we have assumed that, at the lowest 
1045: order of approximation, the effects of tides and mass transfer are 
1046: decoupled.  The total rate of change of the orbital elements is then
1047: given by the sum of the rate of change of the orbital elements due to 
1048: tides and mass transfer.
1049: 
1050: \begin{figure*}
1051: \plottwo{f5a.eps}{f5b.eps}
1052: \caption{Orbital evolution timescales due to the combined effects of 
1053: tidal dissipation in a Roche Lobe filling component of a close binary 
1054: system and a delta function mass transfer with an amplitude 
1055: $\dot{M}_0=-10^{-9}\,M_\sun\,{\rm yr}^{-2}$ centered at the 
1056: periastron of the orbit under the assumption that 
1057: the donor is rotating synchronously with 
1058: the orbital angular velocity at periastron, and the accretor is a 
1059: $1.44\,M_\sun$ neutron star.  The contribution to the orbital 
1060: evolution timescales due to tides is determined under the 
1061: assumption that the donor is zero-age main-sequence star.  Shown at left 
1062: (right) are the timescales 
1063: for the evolution of the semi-major axis (eccentricity) as a function 
1064: of the mass ratio, $q$, for a range of eccentricities, $e$.  Regimes 
1065: where 
1066: the timescale is negative correspond to a decrease in the semi-major 
1067: axis (eccentricity), while regimes where the timescales is positive 
1068: correspond to an increase in the semi-major axis (eccentricity).  The 
1069: discontinuity at $q\simeq0.87$ corresponds to a transition of the 
1070: dominant tidal dissipation mechanism which is different for donor 
1071: stars with convective envelopes ($M_1\lesssim1.25\,M_\sun$) than for 
1072: donor stars with radiative envelopes ($M_1 \gtrsim1.25\,M_\sun$).}
1073: \label{fig-tot_timescale}
1074: \end{figure*}
1075: 
1076: When $q \la 0.87$ and $e \ga 0.4$, the effects of tides and mass 
1077: transfer on the orbital semi-major axis are always opposed, with the 
1078: orbital expansion due to mass transfer dominating the orbital shrinkage
1079: due to tides.  In the case of the orbital eccentricity, the 
1080: increase of the eccentricity due to mass transfer dominates the 
1081: decrease due to tides for mass ratios smaller than some critical 
1082: mass ratio which depends strongly on the orbital eccentricity.  Since 
1083: the timescales of orbital evolution
1084: due to mass transfer are inversely proportional to the magnitude
1085: $\dot{M}_0$ of the mass-transfer rate at periastron, the parameter space
1086: where and the extent to which mass transfer dominates increases with the
1087: rate of mass transfer at periastron.
1088: 
1089: \begin{figure*}
1090: \plottwo{f6a.eps}{f6b.eps}
1091: \caption{Contour plots of the total orbital evolution timescales in the 
1092: $(q,e$)-plane due to the combined effects of tidal dissipation and 
1093: mass transfer for the same set of assumptions adopted in 
1094: figure~~\ref{fig-tot_timescale}.  Timescales for the evolution of the 
1095: semi-major axis are shown at left; timescales for the evolution of the 
1096: orbital eccentricity are shown at right.  The different shades of gray 
1097: designate regions of the ($q,e$) parameter space with timescales (in 
1098: Gyr) in the ranges labeled in the plots.  The thick black line designates 
1099: the transition point where the rate of change of the semi-major axis or 
1100: orbital eccentricity changes from positive (to the left of the thick 
1101: black line) to negative (to the right of the thick black line).  The 
1102: vertical white line at $q\simeq0.87$ corresponds to a transition of the 
1103: dominant tidal dissipation mechanism which is different for donor 
1104: stars with convective envelopes ($M_1\lesssim1.25\,M_\sun$) than for 
1105: donor stars with radiative envelopes ($M_1 \gtrsim1.25\,M_\sun$).}
1106: \label{fig-tot_contour}
1107: \end{figure*}
1108: 
1109: In figure~\ref{fig-tot_contour}, we show the total orbital evolution
1110: timescale due to the sum of tidal and mass transfer effects as a contour
1111: plot in the $(q,e)$-plane.  The thick black lines indicate the
1112: transitions from positive (left of the thick black line) to negative
1113: (right of the thick black line) rates of change of the semi-major axis
1114: and eccentricity.  The white dividing line near $q\approx 0.87$
1115: corresponds to the transition between tidal dissipation mechanisms in
1116: stars with convective envelopes ($M_1\lesssim1.25\,{\rm M}_\sun$) and
1117: stars with radiative envelopes ($M_1\gtrsim1.25\,{\rm M}_\sun$).  It
1118: follows that there are large regions of parameter space where the
1119: combined effects of mass transfer and tidal evolution do not rapidly
1120: circularize the orbit.  In particular, for $q > 0.87$ and $e \ga 0.75$
1121: orbital circularization always takes longer than 10\,Gyr, while for $q$
1122: to the left of the thick black line the orbital eccentricity grows
1123: rather than shrinks.  For a given $q$ left of the thick black line, the
1124: timescales for eccentricity growth increase with increasing $e$ though,
1125: so that there is no runaway eccentricity growth.  Hence, for small $q$,
1126: mass transfer at the periastron of eccentric orbits may provide a means
1127: for inducing non-negligible eccentricities in low-mass binary or
1128: planetary systems.  The orbital semi-major axis, on the other hand,
1129: always increases when $e \la 0.55$, but can increase as well as decrease
1130: when $e \ga 0.55$, depending on the binary mass ratio $q$.  We recall
1131: that both the tidal and mass transfer orbital evolution time scales
1132: depend on the ratio of the donor's rotational angular velocity
1133: $\Omega_1$ to the orbital angular velocity $\Omega_{{\rm orb},P}$ and that we have
1134: set $\Omega_1/\Omega_{{\rm orb},P}=1$ in all figures shown.
1135: 
1136: \section{Concluding Remarks}
1137: 
1138: We developed a formalism to calculate the evolution of the
1139: semi-major axis and orbital eccentricity due to mass transfer in
1140: eccentric binaries, assuming conservation of total system mass and
1141: orbital angular momentum. Adopting a delta-function mass-transfer
1142: profile centered at the periastron of the binary orbit yields rates of
1143: secular change of the orbital elements that are linearly proportional to
1144: the magnitude $\dot{M}_0$ of the mass-transfer rate at the periastron.
1145: For $\dot{M}_0 = 10^{-9}\,M_\sun\,{\rm yr}^{-1}$, this yields timescales
1146: of orbital evolution ranging from a few Myr to a Hubble time or longer.
1147: Depending on the initial binary mass ratio and orbital eccentricity, the
1148: rates of secular change of the orbital semi-major axis and eccentricity
1149: can be positive as well as negative, so these orbital elements can 
1150: increase as well as decrease with time. 
1151: 
1152: Comparison of the timescales of orbital evolution due to mass transfer
1153: with the timescales of orbital evolution due to tidal dissipation shows
1154: that the effects can either be additive or competitive, depending on the
1155: binary mass ratio, the orbital eccentricity, and the magnitude of the
1156: mass-transfer rate at the periastron. Contrary to what is often assumed
1157: in even the most state-of-the-art binary evolution and population
1158: synthesis codes, tides do not always lead to rapid circularization
1159: during the early stages of mass transfer. Thus, phases of episodic mass
1160: transfer may occur at successive periastron passages and may persist for
1161: long periods of time.  As a first approximation, the evolution of the
1162: orbital semi-major axis and eccentricity due to mass transfer in
1163: eccentric binaries can be incorporated into binary evolution and
1164: population synthesis codes by means of equations~(\ref{eq-deltaa}) and
1165: (\ref{eq-deltae}) in which the mass-transfer rate is approximated by a
1166: delta-function of amplitude $\dot{M}_0$ centered at the periastron of
1167: the binary orbit.
1168: 
1169: In future papers, we will relax the assumption of conservation of total
1170: system mass and orbital angular momentum, and examine the effects of
1171: non-conservative mass transfer on the orbital elements of eccentric
1172: binaries. We also intend to study the onset of mass transfer in
1173: eccentric binaries in more detail, adopting realistic mass-transfer
1174: rates appropriate for atmospheric Roche-lobe overflow in interacting
1175: binaries as discussed by \citet{R88}.  We will consider individual 
1176: binary systems that are known to be eccentric and transferring mass 
1177: during periastron passage, as well as populations of eccentric 
1178: mass-transferring binaries and their descendants.
1179: 
1180: 
1181: %--------------------------------------------------------------------
1182: %--------------------------- BEGIN APPENDIX -------------------------
1183: %--------------------------------------------------------------------
1184: 
1185: \appendix
1186: 
1187: \section{Equipotential Surfaces and the Inner Lagrangian point in
1188:   Eccentric Binaries}
1189: \label{sec-appa}
1190: \label{appL1}
1191: %\setcounter{figure}{0}
1192: 
1193: \begin{figure}
1194: \plotone{f7.eps}
1195: \caption{Schematic diagram showing the vectors pertinent to the
1196:   derivation of the equipotential surfaces for a non-synchronous,
1197:   eccentric binary.  The center of mass of the system is shown
1198:   as a cross near the center of the diagram and the center of mass of
1199:   each star is shown as a filled circle at the star's center.  The 
1200:   vectors $\vec{R}_1$ and $\vec{R_2}$ begin at the center of mass of 
1201:   the system.}
1202: \end{figure}
1203: 
1204: A crucial element in the description of mass transfer in any binary
1205: system is the location of the inner Lagrangian point ($L_1$) through
1206: which matter flows from the donor to the accretor.  While a solution
1207: for the location of $L_1$ is not analytic, the case of circular orbits
1208: with synchronized components can be approximated by the formula
1209: \begin{equation}
1210: X_{L_1} = 0.5 + 0.22 \log{q},  \label{eq-XL1cs}
1211: \end{equation}
1212: where $X_{L_1}$ is the distance of the $L_1$ point from the mass
1213: center of the donor star in units of the distance between the stars,
1214: and  $q=M_1/M_2$ is the mass ratio of the binary defined as the ratio
1215: of the donor mass $M_1$ to the accretor mass $M_2$. For mass ratios in
1216: the range $0.1 < q < 15$, this formula is valid to an accuracy of
1217: better than 2\% \citep[e.g.][]{DR74}.  The above
1218: formula has been generalized by \citet{PS76} to
1219: include the effect of a non-synchronously rotating donor star: 
1220: \begin{equation}
1221: X_{L_1} = \left[ 0.53 - 0.03 \left( \frac{\Omega_1}{\Omega_{\rm orb}}
1222:   \right)^2 \right] \left( 1 + \frac{4}{9}\log{q} \right). 
1223:   \label{PS}
1224: \end{equation}
1225: Here, $\Omega_1$ is the rotational angular velocity of the donor, and
1226: $\Omega_{\rm orb}$ the orbital angular velocity of the binary. For $0
1227: < \Omega_1/\Omega_{\rm orb} < 1$ and $0.2 < q < 10$, the formula is
1228: valid to an accuracy of better than 3\%. It is to be noted though that
1229: \citet{PS76} derived the position of the $L_1$ point under the
1230: assumption that the orbital and rotational periods of the binary and
1231: its component stars are much longer than the dynamical timescale of
1232: the donor. Despite the better than 3\% accuracy of the fit given by
1233: equation~(\ref{PS}), the formula may therefore still break down because the
1234: underlying formalism is no longer valid.
1235: 
1236: In this paper, it is necessary to generalize the formula for the
1237: position of the $L_1$ point even further to account for a non-zero
1238: eccentricity as well as non-synchronous rotation. For this purpose, we
1239: determine the equipotential surfaces describing the shape of the
1240: components of a non-synchronous, eccentric binary. Our procedure is a
1241: generalization of the steps outlined by \citet{L63} for the
1242: derivation of the equipotentials of a non-synchronous, circular binary.
1243: 
1244: Here, we consider an eccentric binary system where the stars are
1245: considered to be centrally condensed and spherically symmetric, and
1246: thus can be well described by Roche models of masses $M_1$ and $M_2$.
1247: Their orbit is assumed to be Keplerian with semi-major axis $a$ and
1248: eccentricity $e$. Star~1 is furthermore assumed to rotate uniformly
1249: with a constant angular velocity $\vec{\Omega}_1$ parallel to the
1250: orbital angular velocity $\vec{\Omega}_{\rm orb}$. We note that for an
1251: unperturbed Keplerian orbit, the magnitude of $\vec{\Omega}_{\rm orb}$
1252: is a function of time, while the direction and orientation remain
1253: fixed in space.
1254: 
1255: As is customary, we determine the equipotential surfaces with respect
1256: to a Cartesian coordinate frame $OXYZ$ with origin $O$ at the mass
1257: center of star 1, and with the $Z$-axis pointing along
1258: $\vec{\Omega}_1$. The $X$- and $Y$-axes are co-rotating with the star
1259: at angular velocity $\Omega_1$. Since non-synchronously rotating
1260: binary components are inevitably subjected to time-dependent tides
1261: invoked by their companion, the mass elements will oscillate with
1262: frequencies determined by the difference between the rotational and
1263: orbital frequencies. Here, we neglect these tidally induced
1264: oscillations as well as any other type of bulk motion of matter due
1265: to, e.g., convection. This approximation is valid as long as the
1266: characteristic timescale associated with the motion of the mass
1267: elements is sufficiently long compared to the star's dynamical
1268: timescale (more extended discussions on the validity of the
1269: approximation can be found in \citet{L63}, \citet{S78}, and 
1270: \citet{SWK07}).  Star 1 is then completely stationary with respect to 
1271: the co-rotating frame of reference. 
1272: 
1273: With respect to the $OXYZ$ frame, the equation of
1274: motion of a mass element at the surface of star 1 is given by
1275: \begin{equation}
1276: \ddot{\vec{r}}_1 = \ddot{\vec{r}}_0-\ddot{\vec{R}}_1 - \vec{\Omega}_1
1277: \times \left( \vec{\Omega}_1 \times \vec{r}_1 \right) - 2\,
1278: \vec{\Omega}_1 \times \dot{\vec{r}}_1,  
1279: \label{eq-r1}
1280: \end{equation}
1281: where $\vec{r}_0$ and $\vec{r}_1$ are the position vectors of the mass
1282: element with respect to the mass center of the binary and 
1283: star 1, respectively, and $\vec{R}_1$ is the position vector of the center
1284: of mass of star 1 with respect to the center of mass of the binary, and  
1285: $-\vec{\Omega}_1 \times ( \vec{\Omega}_1 \times \vec{r}_1 )$ and $-
1286: 2\, \vec{\Omega}_1 \times \dot{\vec{r}}_1$ are the centrifugal and
1287: Coriolis acceleration with respect to the rotating coordinate frame. 
1288: 
1289: Under the assumption that, in an inertial frame of reference, the only
1290: forces acting on the considered mass element are those resulting from
1291: the pressure gradients in star 1 and the gravitational attractions of
1292: star 1 and its companion, the acceleration of the mass element with
1293: respect to the binary's center of mass is given by
1294: \begin{equation}
1295: \ddot{\vec{r}}_0 = -\frac{1}{\rho}\, \vec{\nabla} P - \vec{\nabla} 
1296: \left( -G\, \frac{M_1}{|\vec{r}_1|} - G\, \frac{M_2}{|\vec{r}_2|} 
1297: \right),
1298: \label{eq-r0}
1299: \end{equation}
1300: where $G$ is the Newtonian constant of gravitation, $\rho$ the mass
1301: density, $P$ the pressure, and $\vec{r}_2$ the position vector of the
1302: mass element with respect to the mass center of star 2.  The gradient in 
1303: equation~(\ref{eq-r0}) and subsequent equations in this Appendix are 
1304: taken with respect to $X$, $Y$, and $Z$.
1305: 
1306: The acceleration of the center of mass of star 1 with respect to the
1307: binary mass center, on the other hand, is given by
1308: \begin{equation}
1309: \ddot{\vec{R}}_1 = -\frac{GM_2}{D^2} \frac{\vec{R}_1}{|\vec{R}_1|}, 
1310: \label{eq-R}
1311: \end{equation}
1312: where $D(t)$ (simplified to $D$ in what follows) is the time-dependent
1313: distance between star 1 and star 2.  
1314: 
1315: Substituting equations~(\ref{eq-r0}) and~(\ref{eq-R}) into
1316: equation~(\ref{eq-r1}), we obtain 
1317: \begin{equation}
1318: \ddot{\vec{r}}_1 = -\frac{1}{\rho} \vec{\nabla} P - \vec{\nabla} \left( 
1319: -G\,
1320: \frac{M_1}{|\vec{r}_1|} - G\, \frac{M_2}{|\vec{r}_2|} \right) +
1321: \frac{GM_2}{D^2} \frac{\vec{R}_1}{|\vec{R}_1|} - \vec{\Omega}_1
1322: \times ( \vec{\Omega}_1 \times \vec{r}_1 ) - 2\, \vec{\Omega}_1 \times
1323: \dot{\vec{r}}_1.
1324: \label{eq-r1.2}
1325: \end{equation}
1326: 
1327: When the $X$-axis coincides with the line connecting the mass centers
1328: of the two stars, the third and fourth term in the right-hand member
1329: of equation~(\ref{eq-r1.2}) can be written as the gradient of a potential
1330: function as  
1331: \begin{equation}
1332: \frac{GM_2}{D^2} \frac{\vec{R}_1}{|\vec{R}_1|} = 
1333:  - \vec{\nabla} \left( \frac{GM_2}{D^2} X \right),
1334: \end{equation}
1335: \begin{equation}
1336: \vec{\Omega}_1\times ( \vec{\Omega}_1 \times \vec{r}_1 ) = -\vec{\nabla}
1337: \left[ \frac{1}{2} |\vec{\Omega}_1|^2 \left( X^2 + Y^2 \right) \right],
1338: \end{equation}
1339: where $X$ and $Y$ are the Cartesian coordinates of the mass element
1340: under consideration. With these transformations, equation~(\ref{eq-r1.2})
1341: becomes 
1342: \begin{equation}
1343: \ddot{\vec{r}}_1 = -\frac{1}{\rho} \vec{\nabla} P - \vec{\nabla} V_1 - 
1344: 2\, \vec{\Omega}_1 \times \dot{\vec{r}}_1,  \label{a8}
1345: \end{equation}
1346: where
1347: \begin{equation}
1348: V_1 = -G \frac{M_1}{|\vec{r}_1|} - G \frac{M_2}{|\vec{r}_2|} -
1349: \frac{1}{2} |\vec{\Omega}_1|^2 (X^2 + Y^2) + \frac{GM_2}{D^2} X.
1350: \end{equation}
1351: 
1352: Since star 1 is assumed to be static in the rotating frame,
1353: $\dot{\vec{r}}_1 = \ddot{\vec{r}}_1 = 0$, so that equation~(\ref{a8})
1354: reduces to
1355: \begin{equation}
1356: \vec{\nabla} P = - \rho \vec{\nabla} V_1.
1357: \end{equation}
1358: This equation governs the instantaneous shape of the equipotential 
1359: surfaces at the instant when the $X$-axis coincides with
1360: the line connecting the mass centers of the two stars. However, since
1361: this instant is entirely arbitrary, the equation is generally valid
1362: and can be used to determine the shape of the star at any phase of the
1363: orbit, with appropriate re-definition of the $X$-axis at every
1364: instant. A similar equation was derived by \citet{A76} and
1365: \citet{W79}.
1366: 
1367: Next, we look for the critical equipotential surface for
1368: which matter starts to flow from star 1 to star 2 through the inner
1369: Lagrangian point $L_1$. For circular binaries with synchronously
1370: rotating component stars, the $L_1$ point is always located on the
1371: line connecting the mass centers of the two stars. \citet{MW83} have
1372: shown that non-synchronous rotation combined with a misalignment 
1373: between the spin axis of star 1 and the orbital plane may cause the 
1374: $L_1$ point to oscillate in the $Z$-direction.  As a first 
1375: approximation, we here neglect
1376: these oscillations and assume the $L_1$ point to be located on the
1377: line connecting the mass centers of the two stars. The position of the
1378: $L_1$ point is then obtained by setting $dV_1/dX=0$. The resulting
1379: equation for the position of $L_1$ is
1380: \begin{equation}
1381: G \frac{M_1}{X^2} - G \frac{M_2}{(X-D)^2} -
1382:   |\vec{\Omega}_1|^2 X + G \frac{M_2}{D^2} = 0,
1383: \end{equation}
1384: which can be rewritten in dimensionless form as
1385: \begin{equation}
1386: \frac{q}{X_{L_1}^2} - \frac{1}{(X_{L_1}-1)^2} 
1387:   - f^2 X_{L_1} (1+q)(1+e) + 1 = 0. 
1388:   \label{XL}
1389: \end{equation}
1390: Here, $X_{L_1}=X/D$ is the position of the $L_1$ point on the $X$-axis
1391: in units of the distance between the two stars, $q=M_1/M_2$ is the
1392: binary mass ratio, $f=\Omega_1/\Omega_{{\rm orb},P}$ is the ratio of the star 1's
1393: rotational angular velocity to the orbital angular velocity at
1394: periastron, and $\Omega_{{\rm orb},P}^2=G(M_1+M_2)(1+e)/D_p^3$. In the latter
1395: expression, $D_P=a(1-e)$ is the distance between the stars at
1396: periastron. Note that in the particular case of a circular orbit,
1397: equation~(\ref{XL}) reduces to the equation for the position of the $L_1$
1398: point derived by \citet{K63} and \citet{PS76}.
1399: 
1400: Equation~(\ref{XL}) must be solved numerically for the location of
1401: $L_1$.  Since we have made no explicit assumptions about the relative
1402: location of the two star in the orbit, and we have let the distance
1403: between the stars $D=a(1-e^2)/(1+e\cos{\nu})$, with $\nu$ the true
1404: anomaly, be an explicit function of time, this equation characterizes
1405: the position of the $L_1$ point at any position in the orbit.
1406: Figure~\ref{fig-l1} shows the relative change
1407: \begin{equation}
1408: \Delta_{X_{L_1}}= \frac{X_{L_1}(\nu)-X_{L_1}(\nu=0)}{X_{L_1}(\nu=0)}
1409: \end{equation}
1410: in the position of the
1411: $L_1$ point as a function of the true anomaly for $q=1$, $f=1$,
1412: and a variety of eccentricities.  The position of the $L_1$ point over
1413: the course of a single orbit varies by less than 15\% when $e \la
1414: 0.1$, and by 30--40\% when $e \ga 0.2$. At apastron, $\Delta_{X_{L1}}$
1415: increases with increasing eccentricity when $e \la 0.5$, and decreases
1416: with increasing eccentricity when $e \ga 0.5$. This turn-around at $e
1417: \simeq 0.5$ is caused by the increasing deviations from synchronous
1418: rotation at apastron with increasing orbital eccentricity.
1419: 
1420: \begin{figure}
1421: \plotone{f8.eps}
1422: \caption{The relative change in the position of the $L_1$ point as a
1423:   function of true anomaly $\nu$, for $q=1$, $f=1$, 
1424: and a range of
1425:   eccentricities.}
1426: \label{fig-l1}
1427: \end{figure}
1428: 
1429: We have also derived a fitting formula for the position of the $L_1$
1430: point at the periastron of the binary orbit as a function of $q$, $e$,
1431: and $f$:
1432: \begin{equation}
1433: X_{L_1} = 0.529+0.231\log{q} - f^2(0.031 + 0.025 e)(1+0.4\log{q}).
1434: \label{XL1fit}
1435: \end{equation}
1436: This fitting formula is accurate to better than $4\,\%$ for mass
1437: ratios $0.1 \le q \le 10$, eccentricities $0 \le e \le 0.9$, and
1438: ratios of stellar rotational angular velocities to orbital angular
1439: velocities at periastron $0 \le f \le 1.5$. In the limiting case of a
1440: circular, synchronized binary, equation~(\ref{XL1fit}) reduces to
1441: equation~(\ref{eq-XL1cs}), while for a circular, non-synchronized binary,
1442: equation~(\ref{XL1fit}) reduces to equation~(\ref{PS}).  A detailed discussion of 
1443: the properties of all five Lagrangian points and the volume of the 
1444: critical Roche Lobe in eccentric binaries is presented in \citet{SWK07}
1445: 
1446: %******************
1447: %*** Appendix C ***
1448: %******************
1449: 
1450: \section{Orbital Evolution Due to Instantaneous Mass Transfer Between 
1451: Two Point Masses}
1452: \setcounter{figure}{0}
1453: \label{sec-appc}
1454: 
1455: When the components of a mass-transferring binary are treated as point
1456: masses and mass transfer is assumed to be instantaneous, an alternative
1457: derivation of the equations governing the secular change of the orbital
1458: semi-major axis and eccentricity may be obtained from the equations for
1459: the specific orbital energy and angular momentum.  This method is
1460: commonly used to relate the post- to pre-supernova orbital parameters of
1461: binaries in which one of the components undergoes an instantaneous
1462: supernova explosion \citep[e.g.][]{H83, BP95, K96}. As in
1463: Section~\ref{orbevtim}, we assume conservation of total system mass and
1464: describe the mass transfer by a delta function mass-transfer rate
1465: centered at the periastron of the binary orbit [see equation~(\ref{eq-del})].
1466: 
1467: The relative orbital velocity, $V_{\rm rel}$, of two stars with masses
1468: $M_1$ (the donor star) and $M_2$ (the accreting star) in a binary with
1469: orbital semi-major axis $a$ and eccentricity $e$ is given by
1470: \begin{equation}
1471: V_{\rm rel}^2 = G(M_1+M_2) \left( \frac{2}{|\vec{r}|} - \frac{1}{a} 
1472: \right),
1473: \label{eq-C-V}
1474: \end{equation}
1475: where $G$ is the Newtonian constant of gravitation, and $\vec{r}$ is the
1476: relative position vector of the accretor with respect to the donor. 
1477: Under the assumption that mass elements are transferred instantaneously
1478: between the two stars, the distance $|\vec{r}|$ is the same right before
1479: and right after mass transfer, so that the rate of change of the orbital
1480: semi-major axis due to instantaneous mass transfer is given by
1481: \begin{equation}
1482: \frac{da}{dt} = \frac{2a^2V_{\rm rel}}{G(M_1+M_2)}\frac{dV_{\rm rel}}{dt}.
1483: \label{eq-C-dadt}
1484: \end{equation}
1485: 
1486: Next, the specific orbital angular momentum of the binary written in 
1487: terms of the binary component masses and orbital elements is given by
1488: \begin{equation}
1489: |\vec{r}\times\vec{V}_{\rm rel}|^2 = G(M_1+M_2)a(1-e^2).
1490: \label{eq-C-J}
1491: \end{equation}
1492: Noting that $|\vec{r}\times\vec{V}_{\rm rel}|=|\vec{r}||\vec{V}_{\rm
1493: rel}|$ at periastron, and proceeding in a similar way as for the
1494: derivation of equation~(\ref{eq-C-dadt}), we derive the rate of change
1495: of the orbital eccentricity due to instantaneous mass transfer to be
1496: \begin{equation}
1497: \frac{de}{dt} = \frac{2a(1-e)|\vec{V}_{\rm 
1498: rel}|}{G(M_1+M_2)}\frac{d|\vec{V}_{\rm rel}|}{dt}. 
1499: \label{eq-C-dedt}
1500: \end{equation}
1501: 
1502: Finally, we eliminate $V_{\rm rel}$ from equations~(\ref{eq-C-dadt})
1503: and~(\ref{eq-C-dedt}) to obtain equations for the rates of changes of
1504: the orbital semi-major axis and eccentricity in terms of the binary
1505: orbital elements and component masses. For this purpose, we use the
1506: conservation of linear momentum to write the rate of change of the
1507: relative orbital velocity due to mass-transfer as
1508: \begin{equation}
1509: \frac{d\vec{V}_{\rm rel}}{dt} = \frac{\dot{M}_2}{M_2}\, \vec{v}_{\delta 
1510: M_2} - \frac{\dot{M}_1}{M_1}\, \vec{v}_{\delta M_1},
1511: \label{eq-C-vrelvec}
1512: \end{equation}
1513: where $\vec{v}_{\delta M_1}$ and $\vec{v}_{\delta M_2}$ are the relative
1514: velocities of the transferred mass elements with respect to the mass
1515: center of the donor and accretor, respectively.  To obtain an equation 
1516: for the change in the magnitude of $\vec{V}_{\rm rel}$, we write the 
1517: vector components of $\vec{V}_{\rm rel}$, $\vec{v}_{\delta M_1}$, and 
1518: $\vec{v}_{\delta M_2}$ with respect to the $\hat{x}$, $\hat{y}$, and 
1519: $\hat{z}$ unit vectors introduced in Section~\ref{relmot}, and take the 
1520: dot product of equation~(\ref{eq-C-vrelvec}) with $\vec{V}_{\rm rel}$. 
1521: Since, $\vec{V}_{\rm rel} = V_{\rm rel}\, \hat{y}$ at the periastron of 
1522: the binary orbit, it follows that
1523: \begin{equation}
1524: \frac{dV_{\rm rel}}{dt} = -\frac{\dot{M}_1}{M_1} \left( qv_{\delta 
1525: M_{2,y}} + v_{\delta M_{1,y}} \right).
1526: \label{eq-C-vrel}
1527: \end{equation}
1528: Assuming conservation of orbital angular momentum it then follows from
1529: equations~(\ref{eq-Jdot}) and~(\ref{eq-C-dadt})--(\ref{eq-C-vrel}) that
1530: \begin{equation}
1531: qv_{\delta M_2,y} + v_{\delta M_1,y}= na(1-q)\left(\frac{1+e}{1-e} 
1532: \right)^{1/2},
1533: \label{eq-C-vdm2}
1534: \end{equation}
1535: where we have used Kepler's 3rd law and that $|\vec{V}_{\rm rel}| = 
1536: \left(G(M_1+M_2)/a\right)^{1/2}\left((1+e)/(1-e)\right)^{1/2}$ at 
1537: the periastron of the binary orbit.  This equation is identical to 
1538: equation~(\ref{eq-vr}) in the limiting
1539: case where $|\vec{r}_{A_1}| \rightarrow 0$ and $|\vec{r}_{A_2}| 
1540: \rightarrow 0$.  Substituting equations~(eq-C-vdm2) and (eq-C-vrel) into 
1541: equations~(\ref{eq-C-dadt}) and
1542: (\ref{eq-C-dedt}) and averaging over the orbital period, we find
1543: \begin{equation}
1544: \left< \frac{da}{dt} \right>_{\rm sec} = 
1545: \frac{a}{\pi}\frac{\dot{M}_0}{M_1}\left( 1-e^2 
1546: \right)^{1/2}(q-1)
1547: \label{eq-C-dadtsec}
1548: \end{equation}
1549: \begin{equation}
1550: \left< \frac{de}{dt} \right>_{\rm sec}= 
1551: \frac{1}{\pi}\frac{\dot{M}_0}{M_1}(1-e^2)^{1/2}(1-e)(q-1).
1552: \label{eq-C-dedtsec}
1553: \end{equation}
1554: These equations are identical to equations~(\ref{eq-dadtsec})
1555: and(\ref{eq-dedtsec}) in the limiting case where
1556: $|\vec{r}_{A_1}|\rightarrow0$ and $|\vec{r}_{A_2}|\rightarrow0$.
1557: 
1558: \acknowledgements
1559: 
1560: We are grateful to an anonymous referee for insightful comments which
1561: helped to improve the paper, as well as to Ronald Taam for useful
1562: discussions, and to Richard O'Shaughnessy for discussions regarding
1563: methods used to fit the position of the $L_1$ point in eccentric,
1564: non-synchronous binaries. This work is partially supported by a NASA
1565: Graduate Fellowship (NNG04GP04H/S1) to J.S., and a NSF CAREER Award
1566: (AST-0449558), a Packard Fellowship in Science and Engineering, and a
1567: NASA ATP Award (NAG5-13236) to V.K.
1568: 
1569: %\bibliography{MT}
1570: \begin{thebibliography}{37}
1571: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1572: 
1573: \bibitem[{{Avni}(1976)}]{A76}
1574: {Avni}, Y. 1976, \apj, 209, 574
1575: 
1576: \bibitem[{{Brandt} \& {Podsiadlowski}(1995)}]{BP95}
1577: {Brandt}, N., \& {Podsiadlowski}, P. 1995, \mnras, 274, 461
1578: 
1579: \bibitem[{{Brouwer} \& {Clemence}(1961)}]{BC61}
1580: {Brouwer}, D., \& {Clemence}, G.~M. 1961, {Methods of celestial mechanics} (New
1581:   York: Academic Press, 1961)
1582: 
1583: \bibitem[{{Danby}(1962)}]{D62}
1584: {Danby}, J. 1962, {Fundamentals of celestial mechanics} (New York: Macmillan,
1585:   1962)
1586: 
1587: \bibitem[{{Drobyshevski} \& {Reznikov}(1974)}]{DR74}
1588: {Drobyshevski}, E.~M., \& {Reznikov}, B.~I. 1974, Acta Astronomica, 24, 189
1589: 
1590: \bibitem[{{Eggleton}(1983)}]{Egg83}
1591: {Eggleton}, P.~P. 1983, \apj, 268, 368
1592: 
1593: \bibitem[{{Fitzpatrick}(1970)}]{F70}
1594: {Fitzpatrick}, P.~M. 1970, {Principles of celestial mechanics} (New York,
1595:   Academic Press [1970])
1596: 
1597: \bibitem[{{Frank, King, \& Raine}(2002)}]{FKR}
1598: {Frank}, J., {King}, A., \& {Raine}, D. 2002, {Accretion Power in 
1599: Astrophysics Third Edition} (Cambridge University 
1600: Press [2002])
1601: 
1602: \bibitem[{{Hadjidemetriou}(1969{\natexlab{a}})}]{H69b}
1603: {Hadjidemetriou}, J.~D. 1969{\natexlab{a}}, \apss, 3, 330
1604: 
1605: \bibitem[{{Hadjidemetriou}(1969{\natexlab{b}})}]{H69}
1606: ---. 1969{\natexlab{b}}, \apss, 3, 31
1607: 
1608: \bibitem[{{Hills}(1983)}]{H83}
1609: {Hills}, J.~G. 1983, \apj, 267, 322
1610: 
1611: \bibitem[{{Huang}(1956)}]{H56}
1612: {Huang}, S.~S. 1956, \aj, 61, 49
1613: 
1614: \bibitem[{{Hurley} {et~al.}(2002){Hurley}, {Tout}, \& {Pols}}]{HTP02}
1615: {Hurley}, J.~R., {Tout}, C.~A., \& {Pols}, O.~R. 2002, \mnras, 329, 897
1616: 
1617: \bibitem[{{Hut}(1981)}]{H81}
1618: {Hut}, P. 1981, \aap, 99, 126
1619: 
1620: \bibitem[{{Kalogera}(1996)}]{K96}
1621: {Kalogera}, V. 1996, \apj, 471, 352
1622: 
1623: \bibitem[{{Kruszewski}(1963)}]{K63}
1624: {Kruszewski}, A. 1963, Acta Astronomica, 13, 106
1625: 
1626: \bibitem[{{Kruszewski}(1964)}]{K64-2}
1627: ---. 1964, Acta Astronomica, 14, 241
1628: 
1629: \bibitem[{{Layton} {et~al.}(1998){Layton}, {Blondin}, {Owen}, \&
1630:   {Stevens}}]{L98}
1631: {Layton}, J.~T., {Blondin}, J.~M., {Owen}, M.~P., \& {Stevens}, I.~R. 1998, New
1632:   Astronomy, 3, 111
1633: 
1634: \bibitem[{{Limber}(1963)}]{L63}
1635: {Limber}, D.~N. 1963, \apj, 138, 1112
1636: 
1637: \bibitem[{{Matese} \& {Whitmire}(1983)}]{MW83}
1638: {Matese}, J.~J., \& {Whitmire}, D.~P. 1983, \apj, 266, 776
1639: 
1640: \bibitem[{{Matese} \& {Whitmire}(1984)}]{MW84}
1641: ---. 1984, \apj, 282, 522
1642: 
1643: \bibitem[{{Meibom} \& {Mathieu}(2005)}]{MM05}
1644: {Meibom}, S., \& {Mathieu}, R.~D. 2005, \apj, 620, 970
1645: 
1646: \bibitem[{{Petrova} \& {Orlov}(1999)}]{PO99}
1647: {Petrova}, A.~V., \& {Orlov}, V.~V. 1999, \aj, 117, 587
1648: 
1649: \bibitem[{{Petterson}(1978)}]{P78}
1650: {Petterson}, J.~A. 1978, \apj, 224, 625
1651: 
1652: \bibitem[{{Piotrowski}(1964)}]{P64}
1653: {Piotrowski}, S.~L. 1964, Acta Astronomica, 14, 251
1654: 
1655: \bibitem[{{Pratt} \& {Strittmatter}(1976)}]{PS76}
1656: {Pratt}, J.~P., \& {Strittmatter}, P.~A. 1976, \apjl, 204, L29
1657: 
1658: \bibitem[{{Pribulla}(1998)}]{P98}
1659: {Pribulla}, T. 1998, Contrib. Astron. Obs. Skalnat\'e Pleso, 28, 101
1660: 
1661: \bibitem[{{Raguzova} \& {Popov}(2005)}]{Rag05}
1662: {Raguzova}, N.~V., \& {Popov}, S.~B. 2005, ArXiv Astrophysics e-prints
1663: 
1664: \bibitem[{{Reg{\"o}s} {et~al.}(2005){Reg{\"o}s}, {Bailey}, \&
1665:   {Mardling}}]{Rea05}
1666: {Reg{\"o}s}, E., {Bailey}, V.~C., \& {Mardling}, R. 2005, \mnras, 358, 544
1667: 
1668: \bibitem[{{Ritter}(1988)}]{R88}
1669: {Ritter}, H. 1988, \aap, 202, 93
1670: 
1671: \bibitem[{{Savonije}(1978)}]{S78}
1672: {Savonije}, G.~J. 1978, \aap, 62, 317
1673: 
1674: \bibitem[{{Sepinsky} {et~al.}(2007){Sepinsky}, {Willems}, \& 
1675: {Kalogera}}]{SWK07}
1676: {Sepinsky}, J.~F., {Willems}, B., \& {Kalogera}, V. 2007, \apj, in press
1677: 
1678: \bibitem[{{Soberman} {et~al.}(1997){Soberman}, {Phinney}, \& {van den 
1679: Heuvel}}]{SPH97}
1680: {Soberman}, G.~E., {Phinney}, E.~S., \& {van den Heuvel}, E.~P.~J. 1997, 
1681: \aap, 327, 620
1682: 
1683: \bibitem[{{Sterne}(1960)}]{S60}
1684: {Sterne}, T.~E. 1960, {An introduction to celestial mechanics} (Interscience
1685:   Tracts on Physics and Astronomy, New York: Interscience Publication, 1960)
1686: 
1687: \bibitem[{{Tout} {et~al.}(1996){Tout}, {Pols}, {Eggleton}, \& {Han}}]{TEA96}
1688: {Tout}, C.~A., {Pols}, O.~R., {Eggleton}, P.~P., \& {Han}, Z. 1996, \mnras,
1689:   281, 257
1690: 
1691: \bibitem[{{Willems} {et~al.}(2003){Willems}, {van Hoolst}, \&
1692:   {Smeyers}}]{WVS03}
1693: {Willems}, B., {van Hoolst}, T., \& {Smeyers}, P. 2003, \aap, 397, 973
1694: 
1695: \bibitem[{{Wilson}(1979)}]{W79}
1696: {Wilson}, R.~E. 1979, \apj, 234, 1054
1697: 
1698: \bibitem[{{Witte} \& {Savonije}(1999)}]{WS99}
1699: {Witte}, M.~G., \& {Savonije}, G.~J. 1999, \aap, 350, 129
1700: 
1701: \bibitem[{{Witte} \& {Savonije}(2001)}]{WS01}
1702: ---. 2001, \aap, 366, 840
1703: 
1704: \bibitem[{{Zahn}(1977)}]{Z77}
1705: {Zahn}, J.-P. 1977, \aap, 57, 383
1706: 
1707: \bibitem[{{Zahn}(1978)}]{Z78}
1708: {Zahn}, J.-R. 1978, \aap, 67, 162
1709: 
1710: \end{thebibliography}
1711: 
1712: \end{document}
1713: