1: %\documentclass[10pt,preprint]{aastex}
2: \documentclass[onecolumn]{emulateapj}
3:
4: \usepackage{multirow,graphicx}
5:
6: \renewcommand{\vec}[1]{\mathbf{#1}}
7: \newcommand{\el}{$2003~{\rm EL_{61}}$}
8:
9: \begin{document}
10:
11: \title{L\'{e}vy Flights of Binary Orbits due to Impulsive Encounters}
12: \author{Benjamin F. Collins\altaffilmark{1} and Re'em Sari\altaffilmark{1,2}}
13: \altaffiltext{1}{California Institute of Technology, MC 130-33,
14: Pasadena, CA 91125}
15: \altaffiltext{2}{Racah Institute of Physics, Hebrew University, Jerusalem 91904,
16: Israel}
17: \email{bfc@tapir.caltech.edu}
18:
19:
20: \begin{abstract}
21: We examine the evolution of an almost circular Keplerian orbit
22: interacting with unbound perturbers. We calculate the change in
23: eccentricity and angular momentum that results from a single
24: encounter, assuming the timescale for the interaction is shorter than
25: the orbital period. The orbital perturbations are incorporated into a
26: Boltzmann equation that allows for eccentricity dissipation. We
27: present an analytic solution to the Boltzmann equation that describes
28: the distribution of orbital eccentricity and relative inclination as a
29: function of time. The eccentricity and inclination of the binary do
30: not evolve according to a normal random walk but perform a L\'{e}vy
31: flight. The slope of the mass spectrum of perturbers dictates whether
32: close gravitational scatterings are more important than distant tidal
33: ones. When close scatterings are important, the mass spectrum sets
34: the slope of the eccentricity and inclination distribution functions.
35: We use this general framework to understand the eccentricities of
36: several Kuiper belt systems: Pluto, \el, and Eris. We use the model
37: of \citet{TBGE07} to separate the non-Keplerian components of the
38: orbits of Pluto's outer moons Nix and Hydra from the motion excited by
39: interactions with other Kuiper belt objects. Our distribution is
40: consistent with the observations of Nix, Hydra, and the satellites of
41: \el and Eris. We address applications of this work to objects outside
42: of the solar system, such as extrasolar planets around their stars and
43: millisecond pulsars.
44: \end{abstract}
45:
46: \keywords{Kuiper Belt --- planets and satellites --- minor planets, asteroids}
47:
48: \section{Introduction}
49:
50: Several binary Kuiper belt objects (KBOs) have well-measured small
51: orbital eccentricities \citep{NGC07}. \citet{SBL03} investigate
52: numerically the forcing of the eccentricity of the Pluto-Charon orbit
53: by interloping KBOs. They find that the system almost never possesses
54: an eccentricity as high as the observed value of 0.003 \citep{TBGE07};
55: depending on the model of tidal damping used, they find median values
56: of $10^{-5} - 10^{-4}$. Our goal is to develop an analytic theory that
57: describes the effects of a population of unbound perturbers on a
58: binary orbit and can be applied simply to any binary, in the Kuiper
59: belt or elsewhere.
60:
61: The interaction of a binary system with its environment has been
62: studied extensively in the literature \citep{H75,SS95,Y02,MME07,SHM07}.
63: One interesting context is
64: white dwarf-pulsar binaries, which are expected to be circular. For
65: these objects pulse timing produces very accurate measurements of
66: their orbital motion; such measurements reveal that their
67: eccentricities are typically very small but finite, around
68: $10^{-4}-10^{-5}$ \citep{S04}. \citet{P92} investigated the effects
69: of passing stars on the orbit of such a binary and found that for
70: Galactic pulsars, the perturbations are sub-dominant compared to the
71: effects of atmospheric fluctuations in the companion star. The higher
72: density environment of a globular cluster however can induce an order
73: of magnitude higher eccentricity. \citet{RH95} and \citet{HR96}
74: present a detailed account of the changes in orbital parameters for
75: binaries in a stellar cluster. The work of these authors focuses on
76: the regime where a perturbing body interacts with the binary on
77: timescales longer than the orbital period of the binary. In the
78: Kuiper belt, a single interaction between a binary and an unbound
79: object occurs over a shorter timescale than the orbital period of the
80: binary. We focus on this regime, where the perturbations
81: to the orbital dynamics can be approximated as discrete impulses.
82:
83: The main result of this work is that we have identified the
84: perturbative evolution of the eccentricity and relative inclination of
85: a nearly circular binary orbit as a L\'{e}vy flight, a specific type of
86: random walk through phase space \citep{SZF95}. The entire
87: distribution function of the eccentricity and inclination is then
88: determined by calculating the frequency of perturbations as a function
89: of their magnitude. We find a simple analytic expression for this
90: distribution function.
91:
92: We take the following steps to arrive at
93: our conclusion. In section \ref{secSingleEnc} we
94: calculate the effect of one perturber on a two-body orbit, examining
95: separately the tidal effects of distant scatterings, close encounters
96: with a single binary member, and direct collisions. We describe the
97: effects of many such encounters in section \ref{secBE}, and write a
98: Boltzmann equation that describes the distribution function of the
99: orbital eccentricity and the inclination of the binary relative to its
100: initial plane. The quantitative description of the binary's evolution
101: given by this distribution function reveals its nature as a L\'{e}vy flight.
102: In section \ref{secMassSpec}, we allow for a
103: distribution of perturbing masses and discuss the different L\'{e}vy
104: distributions that result.
105:
106: We then use the analytic theory to examine the orbits of binary KBOs
107: being perturbed by the other members of the Kuiper belt. Section
108: \ref{secKBB} applies our analysis to several specific Kuiper belt
109: binaries. We briefly discuss the relevance of this theory to other
110: astrophysical systems in section \ref{secApp}, and summarize our
111: conclusions in section \ref{secConclusions}.
112:
113: \section{A Single Encounter}
114: \label{secSingleEnc}
115:
116: We use the following terminology to describe the geometry of the
117: encounter between a single perturber and a two-body orbit. We refer
118: to the two bound bodies as ``the binary.'' The members of the
119: binary have masses $m_1$ and $m_2$, with a total mass labeled
120: $m_b=m_1+m_2$ and $m_1 \geq m_2$. The position of body 2 relative to body
121: 1 is given by $\vec r_b$, and the relative velocity by $\vec v_b$. We
122: distinguish between the magnitude and direction of a vector with the
123: notation $\vec r_b = r_b \hat r_b$. We assume
124: $v_b \approx \Omega r_b$, where $\Omega$ is the orbital frequency of
125: the binary. We write the orbital period as $T_{\rm orb}=2 \pi/\Omega$.
126:
127: We label the mass of the perturber $m_p$. The position of the
128: perturber as a function of time, $\vec r_p(t)$, is described by two
129: vectors: $\vec r_p(t) = \vec b + \vec v_p t$. The vector $\vec b$
130: specifies the closest point of the perturber's trajectory to body 1,
131: and $\vec v_p$ is velocity of the perturber relative to body 1. Each
132: encounter geometry is uniquely specified by $\vec b$ and $\vec v_p$
133: under the constraint $\vec b \cdot \vec v_p =0$. Figure
134: \ref{figCartoon} depicts the arrangement of the vectors $\vec r_b,
135: \vec v_b, \vec r_p(t), \vec b,$ and $\vec v_p$. We assume $T_{\rm
136: orb} \gg b/v_p$ so that we may ignore the motion of the binary
137: during the interaction. We further assume that the effects of the
138: gravity of the binary on the perturber are small; the perturber then
139: travels along a straight path with a constant $\vec v_p$. This
140: assumption requires the criterion of $v_p^2 \gg G (m_b+m_p)/b$. If
141: $b$ is small, the perturber may collide with a member of the binary.
142: In this case the assumption that the path of the perturber is
143: unaffected by the gravity of the binary is true under the condition
144: that $v_p$ is much greater than the escape velocity of that member of
145: the binary. The escape velocity from body 1 is defined $v^2_{{\rm
146: esc,1}}=2 G m_1/ R_1$, where $R_1$ is the radius of body 1.
147:
148: \begin{figure}[t!]
149: \center
150: \includegraphics[angle=-90]{b-1-3.ps}
151: \caption{An illustration of the notation we use to denote the geometry of each
152: perturbation. The dotted line is the almost circular orbit of the binary viewed
153: at an angle. The dashed line is the path of the perturber, given by
154: $\vec r_p(t)= \vec b + \vec v_p t$.}
155: \label{figCartoon}
156: \end{figure}
157:
158: We are assuming that the timescale of the interaction is much shorter than
159: the orbital timescale, such that the perturbation instantaneously changes the
160: velocities of the binary objects.
161: The impulse provided to a specific member of the binary is found by
162: integrating the acceleration caused by the perturber over its path:
163:
164: \begin{equation}
165: \label{eqDeltaV1}
166: \Delta \vec v_j = \int_{-\infty}^{\infty}
167: \frac{G m_p(\vec b_j + \vec v_p t)}{|\vec b_j+ \vec v_p t|^3}
168: dt = 2 \frac{G m_p}{v_p}
169: \frac{\hat b_j}{b_j},
170: \end{equation}
171:
172: \noindent
173: where the index $j$ specifies whether the impulse $\Delta \vec v_j$
174: and impact parameter $\vec b_j$ are
175: with respect to either the primary ($j=1$) or the secondary
176: ($j=2$). For the primary, $\vec b_1 = \vec b$ as we have defined it
177: above. For encounters with the secondary, $\vec b_2$ is related to
178: $\vec b$ by enforcing that it is also perpendicular to $\vec v_p$.
179: Thus we find $\vec b_2 = \vec b - \vec r_b + \hat v_p ( \vec r_b \cdot
180: \hat v_p)$.
181:
182:
183: We
184: consider the effects of such impulses on the full Laplace-Runge-Lenz
185: vector, $\vec e = (\vec v_b \times \vec H)/G m_b - \hat r$, where
186: $\vec H=\vec r_b \times \vec v_b$, the angular momentum per unit mass
187: of the binary. The vector $\vec e$ has a magnitude equal to the
188: eccentricity of the orbit, and points from body 1 towards
189: the periapse. It responds to a small impulse $\Delta \vec v$ according to
190: the formula
191:
192: \begin{equation}
193: \label{eqEdot}
194: \Delta \vec e = \frac{1}{G m_b} \left[ 2 \vec r_b (\Delta \vec v \cdot \vec v_b)
195: - \vec v_b (\Delta \vec v \cdot \vec r_b)
196: - \Delta \vec v (\vec r_b \cdot \vec v_b)
197: \right],
198: \end{equation}
199:
200: \noindent
201: keeping terms up to linear order in $\Delta \vec v$.
202: Since we have assumed the binary has very small eccentricity, the
203: third term in equation \ref{eqEdot} is negligible compared to the
204: other two.
205:
206: The orbital plane of the binary is defined by the angular momentum
207: vector $\vec H$, and evolves according to $\Delta \vec H = \vec
208: r_b \times \Delta \vec v$. The impulses affect the
209: direction of the angular momentum vector, and therefore alter the
210: orientation of the orbital plane of the binary. We use the
211: two-dimensional vector $\vec i$ to denote the components of $\hat H$
212: in the plane defined by the initial angular momentum. This vector,
213: $\vec i$, has a magnitude equal to $\sin i$, the sine of the
214: inclination of the binary with respect to the initial orbital plane,
215: and points from body 1 towards the longitude of the ascending node.
216:
217: The change in relative velocity given by a general gravitational scattering
218: is given by $\Delta \vec v = \Delta \vec v_2 - \Delta \vec v_1$. The
219: resulting change in the eccentricity vector is
220:
221: \begin{equation}
222: \label{eqDeltaegeneral}
223: \Delta \vec e = 2 \frac{m_p}{m_b} \frac{v_b}{v_p}
224: \left[ 2 \hat r_b \left( \frac{\hat b_2 \cdot \hat v_b}{b_2/r_b}
225: - \frac{\hat b \cdot \hat v_b}{b/r_b} \right)
226: - \hat v_b \left( \frac{\hat b_2 \cdot \hat r_b}{b_2/r_b} -
227: \frac{\hat b \cdot \hat r_b}{b/r_b} \right) \right].
228: \end{equation}
229:
230: \noindent
231: %We simplify this expression with the abbreviation
232: %$|\Delta \vec e| = A_e (m_p/m_b)(v_b/v_p)$.
233: %Here $A_e$ contains all of the
234: %dependence of the perturbation on the relative orientation of the binary
235: %and perturber as well as the dependence on impact parameter $b$.
236: The change in $\vec i$ is
237:
238: \begin{equation}
239: \label{eqDeltaigeneral}
240: \Delta \vec i = - 2 \frac{m_p}{m_b} \frac{v_b}{v_p}
241: \left[ \hat v_b \left(\frac{\hat b_2 \cdot \hat n}{b_2/r_b}
242: - \frac{\hat b \cdot \hat n}{b/r_b} \right) \right],
243: \end{equation}
244:
245: \noindent
246: where $\hat n$ is the unit normal vector to the binary's orbital
247: plane.
248: %We abbreviate this expression as $|\Delta \vec i|=A_i (m_p/m_b)(v_b/v_p)$.
249: For both the farthest perturbers and the closest, the dependence of
250: equations \ref{eqDeltaegeneral} and \ref{eqDeltaigeneral} on the impact parameter
251: can be simplified. We discuss these limits in the following sections.
252:
253: \subsection{Close Encounters}
254: \label{secVClose}
255:
256: Interactions with impact parameters greater than the radius of the
257: primary or secondary but much less than the semi-major axis of the
258: binary belong to what we call the ``close-encounter regime.'' By
259: definition the encounters in this regime of impact parameter are much
260: closer to one member of the binary than the other. As a result the
261: relative impulse experienced is dominated by the single impulse
262: delivered to that body, $|\Delta \vec v| \approx |\Delta \vec v_j|$.
263: The changes in $\vec e$ and $\vec i$ are then given not by the
264: difference of the impulses on each body, as in equations
265: \ref{eqDeltaegeneral} and \ref{eqDeltaigeneral}, but by the effects of
266: only the largest impulse. For the change in eccentricity we find,
267:
268: \begin{equation}
269: \label{eqEClose}
270: \Delta \vec e = 2 \frac{m_p}{m_b} \frac{v_b}{v_p} \frac{r_b}{b}
271: \left[ 2 \hat r_b (\hat b_j \cdot \hat v_b) - \hat v_b (\hat b_j \cdot \hat r_b)
272: \right],
273: \end{equation}
274:
275: \noindent
276: and for the inclination,
277:
278: \begin{equation}
279: \label{eqIClose}
280: \Delta \vec i = - 2 \frac{m_p}{m_b} \frac{v_b}{v_p} \frac{r_b}{b}
281: \left[ \hat v_b (\hat b_j \cdot \hat n) \right].
282: \end{equation}
283: %\noindent
284: %We abbreviate the angular information of close encounters with the
285: %quantity $B_e$ such that $|\Delta \vec e|=B_e (m_p/m_b)(v_b/v_p)(r_b/b)$,
286: %leaving the scaling with $b$ explicit.
287: %We use angled brackets
288: %to denote the average over the possible angular configurations of the
289: %encounter:
290: %$\langle B_e \rangle = \int B_e d \Omega/ 8 \pi^2 = E(-3) \approx 1.2$,
291: %where $E(-3)$ is the complete Elliptic integral and $d \Omega$ denotes
292: %the differential of the angular phase space. Inclination?
293:
294:
295: \subsection{Distant Encounters}
296:
297: For interactions where $b \gg r_b$, the impulse delivered to each
298: member of the binary is almost the same. In this limit only the tidal
299: difference in impulse affects the eccentricity of the binary. The
300: perturbation delivered to the lowest order in $r_b/b$ is
301:
302: \begin{equation}
303: \label{eqDeltae}
304: \Delta \vec e = 2 \frac{m_p}{m_b} \frac{v_b}{v_p} \left(\frac{r_b}{b}\right)^2
305: \left[ \hat r_b \left( 4 (\hat r_b \cdot \hat b)( \hat v_b \cdot \hat b)
306: + 2 (\hat r_b \cdot \hat v_p)( \hat v_b \cdot \hat v_p) \right)
307: + \hat v_b \left(1 - 2 (\hat r_b \cdot \hat b)^2 - (\hat r_b \cdot \hat v_p)^2
308: \right) \right].
309: \end{equation}
310:
311: \noindent
312: %To shorten this expression, we write $|\Delta \vec e|=C_e
313: %(m_p/m_b)(v_b/v_p)(r_b/b)^2$.
314: \citet{P92} derives the special case of
315: this formula for interactions that take place entirely in the plane of
316: the binary. This formula is also equivalent to equation A24 of
317: \citet{HR96}.
318:
319: The change in $\vec i$ due to distant encounters is given by:
320:
321: \begin{equation}
322: \label{eqDeltah}
323: \Delta \vec i = \frac{m_p}{m_b} \frac{v_b}{v_p} \left(\frac{r_b}{b}\right)^2
324: \hat v_b \left[ 4(\hat r_b \cdot \hat b)(\hat b \cdot \hat n)
325: + 2 (\hat r_b \cdot \hat v_p)(\hat v_p \cdot \hat n) \right].
326: \end{equation}
327:
328:
329: \subsection{Collisions}
330: \label{secCollisions}
331:
332: Physical collisions between perturbers and body 1 or 2 cause the orbit
333: to evolve impulsively. We define collisions to be any encounters
334: where the impact parameter is smaller than the radius of the primary
335: or secondary: $b<r_1$ or $b_2<r_2$. In this case the impulse is given
336: by the conservation of linear momentum of the encounter: $\Delta \vec
337: v = \chi (m_p/m_j) \vec v_p$, where $m_j$ is the mass of the binary
338: member involved in the collision ($j=1$ or $2$). The coefficient
339: $\chi$ accounts for the final momentum of the perturber. For an
340: inelastic collision with $m_p \ll m_j$, $\chi=1$. If the perturber is
341: perfectly reflected, $\chi=2$. The momentum loss from an impact
342: crater can enhance this factor above 2 depending on the properties of
343: the colliding bodies \citep{MNZ94}. For simplicity we
344: assume that the mass of each binary
345: member remains unchanged after each collision.
346:
347: The collisional impulse changes the eccentricity according to equation
348: \ref{eqEdot} and the orbital plane according to the change in angular
349: momentum $\vec r_b \times \Delta \vec v$.
350:
351: \section{Boltzmann Equation}
352: \label{secBE}
353:
354: The evolution of the eccentricity and inclination (relative to the
355: initial orbital plane) is given by the sum of the perturbations the
356: binary receives as it travels through a swarm of perturbers. From the
357: average properties of the perturbing population, we can calculate a
358: distribution function that describes the evolution of the orbit in a
359: statistical sense.
360:
361: \subsection{Eccentricity}
362: \label{secBEecc}
363:
364:
365: Since the perturbation in eccentricity is a two-dimensional vector,
366: each component is added to the components of the existing eccentricity
367: vector separately. As the binary experiences many perturbations, its
368: eccentricity vector travels throughout this two-dimensional space. We
369: write a distribution function $f(\vec e,t)$ that describes the
370: probability that the binary will have an eccentricity
371: in a small region $d^2 \vec e$. Assuming isotropic perturbations, there
372: is no preferred longitude of periapse for the binary. It follows that
373: $f(\vec e,t)=f(e,t)$ and the likelihood of finding the
374: eccentricity in a small range $de$ around $e$ is $2 \pi e f(e, t) de$.
375:
376: We define ${\cal R}(e')$ to be the frequency at which the binary
377: experiences perturbations of magnitudes between $e'$ and $e'+de'$.
378: The frequency of perturbations with magnitudes on the order of
379: $|\Delta \vec e|=e'$ is given
380: schematically by $e' {\cal R}(e') \sim n v b^2$, where
381: $n$ is the number density of the perturbers, $v$ is the speed at
382: which the binary encounters those perturbers, and $b$ is the
383: distance at which the binary encounters perturbers that cause a
384: perturbation of strength $e'$. We make this calculation precise with
385: the following integral:
386:
387: \begin{equation}
388: \label{eqRofe}
389: {\cal R}(e')= \int \delta(| \Delta \vec e(\vec v_p,\vec b,m_p)| - e')
390: {\cal F}(\vec v_p,m_p) v_p \delta(\vec b \cdot \hat v_p) d^3 \vec b d^3 \vec v_p dm_p,
391: \end{equation}
392:
393: \noindent
394: where ${\cal F}(\vec v_p,m_p)$ is the phase space density per unit
395: mass of the perturbers. The integral of ${\cal F}(\vec v_p,m_p)$ over
396: $d^3 \vec v_p dm_p$ is the number density of the perturbers. We
397: assume this density is uniform in the spatial dimensions and isotropic
398: in velocity. It is normalized such that the total mass density of
399: perturbers is given by $\rho= \int m_p {\cal F}(\vec v_p,m_p) d^3 \vec
400: v_p dm_p$. The factor of $v_p$ in the integrand of equation
401: \ref{eqRofe} represents the velocity at which the binary encounters
402: perturbers. The second delta function in equation \ref{eqRofe}
403: converts the volume element $d^3 \vec b$ to an element of
404: cross-sectional area. The first delta function, $\delta(| \Delta \vec
405: e(\vec v_p,\vec b,m_p)| - e')$, restricts the integral to include only
406: the combinations of $\vec b$, $\vec v_p$, and $m_p$ that cause a
407: $|\Delta \vec e|=e'$.
408:
409: The evolution of the distribution function as a result of these perturbations
410: is given by a Boltzmann equation that links the rate of change of $f(e,t)$
411: to the interaction frequency. We write this equation as:
412:
413: \begin{equation}
414: \label{eqFofe}
415: \frac{\partial f(e,t)}{\partial t}
416: = \int p(\vec e') \left[ f(|\vec e'+\vec e|) - f(e)\right] d^2\vec e'
417: \end{equation}
418:
419: \noindent
420: The function $p(\vec e')$ describes the frequency per unit of
421: eccentricity space ($d^2 \vec e'$) at which a binary with
422: eccentricity $\vec e$ is perturbed to the value $\vec e+\vec e'$.
423: Since there is no preferred direction for the encounters, this
424: function is axisymmetric, $p(\vec e') = p(e')$. It is related to
425: ${\cal R}(e')$ by integrating over the angular
426: direction of the phase space, ${\cal R}(e') = \int p(e') e' d \omega = 2 \pi e' p(e')$.
427:
428: We first derive $p(e')$ for a simple scenario: a population of
429: perturbers each with mass $m_p$ and velocity $v_p$. To clarify this
430: derivation, we present a qualitative treatment. The eccentricity
431: excited by such a perturber with an impact parameter of order $b
432: \gg r_b$ is about $e' \sim (m_p/m_b)(v_b/v_p)(r_b/b)^2$ (Section
433: \ref{secSingleEnc}). Since the frequency of encounters with impact
434: parameters $b$ is proportional to $b^2$, and the size of the
435: perturbation $e' \propto b^{-2}$, the frequency at which the binary is
436: perturbed by an amount of order $e'$ is therefore a power law:
437: $e'^2 p(e') \propto e'^{-1}$.
438: This power law is valid from very low
439: $e'$, caused by the farthest possible impulsive encounter, to $e' \sim
440: (m_p/m_b)(v_b/v_p)$, the rare encounters with $b \sim r_b$. We take
441: into account the very rare occurrence of a physical collision, which
442: excite eccentricities of order $e' \sim (m_p/m_j)(v_p/v_b)$, in section
443: \ref{secCollisions}.
444:
445: Evaluating equation \ref{eqRofe} using $\Delta \vec e (\vec v_p,\vec b,m_p)$
446: given by equation \ref{eqDeltae} provides the exact form of
447: $p(e')$ for this scenario.
448: We find:
449:
450: \begin{equation}
451: \label{eqPofE}
452: p(e') = %\frac{148.9}{32 \pi^3} G n_{\rm 3d} m_p T_{\rm orb} \frac{1}{e^3},
453: \frac{\langle C_e \rangle}{4 \pi} G \rho T_{\rm orb} \frac{1}{e'^3},
454: \end{equation}
455:
456: \noindent
457: where $T_{\rm orb}$ is the orbital period of the binary, and $\langle
458: C_e \rangle = 1.89$ is the average value of the angular terms of
459: equation \ref{eqDeltae} (see Appendix).
460: We note that the frequency of perturbations depends not on
461: $m_p$, but only on the total mass density of perturbers. It is also
462: independent of $v_p$, as the lowered effectiveness of the faster
463: perturbations is directly canceled by their higher frequency.
464: These properties are typical of distant encounters with
465: binaries, as evident in earlier work on binary dynamics \citep{BHT85}.
466:
467: We can generalize equation \ref{eqFofe} by including a term to
468: account for dissipation of the binary's eccentricity: $\partial f(\vec
469: e,t)/\partial t = {\rm -div}(f(\vec e,t) {\dot \vec e})$. We restrict
470: our attention to mechanisms that reduce $\vec e$ at a timescale that
471: is independent of $\vec e$, ${\dot \vec e} = - \vec e / \tau_d$. The
472: tidal dissipation of eccentricity obeys this form and is our main
473: motivation for including such terms.
474:
475: Since $p(e')$ is a power-law, we can look for self-similar solutions
476: to the time-dependent integro-differential Boltzmann equation,
477: equation \ref{eqFofe}. The frequency of perturbations $p(e')$ does
478: not depend on any special eccentricity, so the distribution function
479: should depend only on the time $t$. We separate the
480: distribution function into three parts: the time-dependent
481: normalization, $F(t)$, the time-independent shape of the function,
482: $g(x)$, and the time-dependent eccentricity scale, $e_c(t)$. These
483: quantities obey the relation $f(e,t)=F(t)g(e/e_c(t))$. We choose the
484: normalization of $g(x)$ such that $\int g(x) d^2 x=1$. We further
485: choose that $f(e,t)$ be normalized to 1 for all times; this constrains
486: the normalization function to be $F(t)=1/e_c(t)^2$.
487:
488: Substituting $f(e,t)=e_c(t)^{-2}g(e/e_c(t))$ into equation \ref{eqFofe},
489: we find two equations. The first specifies the time-independent
490: shape of the distribution as a function of the dimensionless
491: parameter $x \equiv e/e_c(t)$:
492:
493: \begin{equation}
494: \label{eqFofg}
495: 2 g(x)
496: +x \frac{d g(x)}{d x}
497: + \frac{1}{2 \pi} \int\int
498: \frac{g(x_n)-g(x)}{|\vec x_n - \vec x|^3} d^2\vec {x}_n = 0,
499: \end{equation}
500:
501: \noindent
502: The solution to this equation has been presented in several earlier
503: works, where we investigate the eccentricity distribution of the
504: oligarchs in a protoplanetary disk \citep{CS06,CSS07}:
505:
506: \begin{equation}
507: \label{eqGofx}
508: g(x)=\frac{1}{2\pi} (1+ x^2)^{-3/2}.
509: \end{equation}
510:
511: \noindent
512: This function is the two-dimensional Cauchy distribution.
513: The median and mode of this distribution are
514: $x_{\rm med}=\sqrt{3}$ and $x_{\rm mode}=1/\sqrt{2}$. The mean of
515: this distribution is formally divergent; assuming there is a maximum
516: value of $x$, $x_u \gg 1$, then $x_{\rm mean}\approx 2.3 \log_{10} (0.74 x_u)$.
517:
518: The eccentricity scale $e_c(t)$ is set by an ordinary differential
519: equation,
520:
521: \begin{equation}
522: \label{eqDiffeoft}
523: \dot e_c(t) = - e_c(t)/\tau_d + \langle C_e \rangle G \rho T_{\rm orb}/2
524: %+ \sum_{i=1,2} \chi D_e G \frac{\rho}{\Lambda} T_{\rm orb}
525: %\left(\frac{ A v_p}{v_{\rm esc}}\right)^2 \frac{r_i}{r_b}.
526: \end{equation}
527: %+ G \rho T_{\rm orb} \left( \frac{C_e}{2}
528: %+ \sum_{i=1,2} \frac{D_e}{\Lambda} \chi
529: %\left(\frac{ A v_p}{v_{\rm esc}}\right)^2 \frac{r_i}{r_b}
530:
531: \noindent
532: We note that $\tau_d$ and the terms on the right
533: hand side of equation \ref{eqDiffeoft} do not need to be constant in time;
534: evolution of the binary ($T_{\rm orb}(t)$), the perturbing swarm
535: ($\rho(t)$), or the damping mechanism ($\tau_d(t)$)
536: can be treated by including the time-dependence of these quantities.
537:
538: We offer a reminder that $e_c(t)$ is the characteristic value of the
539: entire distribution of eccentricity that the binary may attain. The
540: probability is highest that the binary will have an eccentricity near
541: the mode of the distribution, which is smaller than $e_c(t)$ by a
542: factor of 0.7. The distribution is somewhat wide, and the confidence
543: levels around the median value are large. The 66 percent confidence
544: interval of $x$ is $0.67-5.8$, and the 95 percent interval is
545: $0.23-40.0$.
546:
547: Equations \ref{eqGofx} and \ref{eqDiffeoft} present a new picture of
548: the stochastic evolution of the binary's eccentricity. Often the
549: evolution of a random variable is characterized by Brownian motion,
550: in which the distribution of the random variable is
551: set by the long term accumulation of many small perturbations. The
552: typical value of such a variable grows as the square-root of time
553: (written $\sqrt{\langle x^2 \rangle} \propto t^{1/2}$), and the
554: probability of finding the system very far away from the typical value
555: is exponentially low. The eccentricity of the binary evolves
556: differently. The probability of finding the binary with an
557: eccentricity larger than $e_c(t)$ only diminishes as a power law
558: (equation \ref{eqGofx}). Physically, this reflects the probability
559: that the binary received a single large perturbation to that state.
560: The characteristic eccentricity, $\sim e_c(t)$ corresponds to the size
561: of the perturbation that occurs with a frequency of about $1/t$. The
562: linear growth of $e_c(t)$ demonstrated by equation
563: \ref{eqDiffeoft} reveals that the eccentricity of the binary does not
564: reflect the accumulation of many small perturbations, but the single
565: largest perturbation occurring in its history. This kind of random
566: walk is called a ``L\'{e}vy flight'' \citep{SZF95}.
567:
568: \subsection{Inclination}
569: \label{secBEinc}
570:
571: The same analysis applies to the changes in angular momentum of the
572: binary. Since $|\Delta \vec i| \sim |\Delta \vec e|$, it follows that
573: $p(i') \sim p(e')$. The evolution of inclination differs only in the
574: coefficients that depend on the geometrical configuration of the
575: encounter. The calculation of the coefficients is described in the
576: Appendix. The self-similar distribution shape is a function of the
577: dimensionless variable $i/i_c(t)$, where $i_c(t)$ is the
578: time-dependent characteristic inclination. The following equation
579: describes the evolution of $i_c(t)$:
580:
581: \begin{equation}
582: \label{eqDiffioft}
583: \dot i_c(t) = - i_c(t)/\tau_{d,i}+ \langle C_i \rangle G \rho T_{\rm orb}/2
584: %+ \sum_{i=1,2} \chi D_i G \frac{\rho}{\Lambda} T_{\rm orb}
585: %\left(\frac{ A v_p}{v_{\rm esc}}\right)^2 \frac{r_i}{r_b},
586: \end{equation}
587:
588: \noindent
589: where we have used $\tau_{d,i}$ to distinguish the timescale at which
590: the inclination of the binary is damped, and $ \langle C_i \rangle =
591: 0.75$, the average of the angular terms in equation \ref{eqDeltah}.
592: The inclination is always measured relative to the orbital plane at
593: $t=0$. The distribution given by equation \ref{eqGofx} then describes
594: the probability of the binary being inclined by $i=x~ i_c(t)$ relative
595: to its original orbital plane.
596:
597: \section{A Spectrum of Colliding Perturbers}
598: \label{secMassSpec}
599:
600: For many physical applications we must consider a range of perturbing
601: masses and velocities and the effects of collisions onto the binary.
602: In the single mass case discussed in section
603: \ref{secBEecc}, the interaction frequency $p(e')$ is set by the
604: likelihood that the binary encounters a perturber at the impact parameter that
605: causes such a change of $e'$. For perturbers that have different
606: masses, the chance of experiencing a perturbation of magnitude $e'$
607: depends on the combined likelihood that the perturber has the required
608: impact parameter and the required mass to excite such a change.
609:
610: To extend our analysis we set up several pieces of notation.
611: Assuming that the mass and velocity distributions are independent, we
612: consider ${\cal F}(m_p,v_p) = {\cal F}_v(v_p) {\cal F}_m(m_p)$. We
613: restrict our analysis to velocity distributions with a characteristic
614: value, $v_0$, such as a Gaussian distribution. We consider systems
615: with differential mass spectra characterized by a power law: ${\cal
616: F}_m(m_p) \propto m_p^{-\gamma}$, valid from a minimum mass $m_{\rm
617: min}$ to a maximum $m_{\rm max}$. These functions are consistent
618: with conditions in the Kuiper belt, where a power law mass spectrum
619: and roughly Gaussian velocity spectrum are observed \citep{LJ02}.
620: We define the differential mass spectrum by
621:
622: \begin{equation}
623: \label{eqMassSpec}
624: {\cal F}_m(m_p)=(n_0 (\gamma-1)/m_0) (m_0/m_p)^{\gamma},
625: \end{equation}
626:
627: \noindent
628: where $n_0$ is the number density of bodies larger than mass $m_0$.
629: In the literature the differential size spectrum of Kuiper belt
630: objects is characterized as a power law in
631: radius with index $q$; this is related to our index by $\gamma =
632: (q+2)/3$. In this section we discuss the $p(e')$ and $p(i')$ that
633: result from several values of $\gamma$.
634:
635: \subsection{$\gamma<2$}
636: \label{secShallow}
637:
638: The total mass density of perturbers for $\gamma < 2$ is dominated by
639: the perturbers with the largest mass, $m_{\rm max}$. While
640: perturbations of size $e'$ are excited by all of the perturbers, the
641: most likely perturber to cause a perturbation of this strength is the
642: largest mass perturber. The dynamics of the binary are then the same as described
643: in section \ref{secBEecc} with $m_p=m_{\rm max}$. The power law of
644: $p(e') \propto e'^{-3}$, based on distant encounters, is valid up to
645: the eccentricity excited by a perturber of mass $m_{\rm max}$
646: interacting at a $b \sim r_b$, or for $e' \ll (m_{\rm
647: max}/m_b)(v_b/v_p)$ (equation \ref{eqDeltae}). It is necessary only
648: to know the total mass density $\rho$ of the perturbing swarm in order
649: to calculate the excitation frequency in this scenario, given by
650: equation \ref{eqPofE}.
651:
652: \subsection{$\gamma=2$}
653: \label{secEq2}
654:
655: The power law $\gamma=2$ describes a special mass distribution where
656: the frequency of encountering the few large perturbers at large impact
657: parameters is the same as encountering the more abundant smaller perturbers at
658: smaller impact parameters. Thus each logarithmic interval in impact
659: parameter contributes the same amount to the frequency of perturbations by
660: $e'$, $p(e')$.
661: The upper limit of impact parameters that can contribute to
662: excitations of a given $e'$, however, is given by the maximum mass
663: perturber. The total range of contributing impact parameters then
664: diminishes as $e'$ approaches the eccentricity caused by the largest
665: perturber interacting with $b \sim r_b$, $e'_{\rm max}\equiv (m_{\rm
666: max}/m_b)(v_b/v_0)$. Mathematically this behavior is determined by
667: the integral of equation \ref{eqRofe}, which yields an excitation
668: frequency of:
669:
670: \begin{equation}
671: \label{eqRdistanteq2}
672: p(e') = \frac{G n_0 m_0 T_{\rm orb}}{e'^3}
673: \frac{\log \left( 2.1 (e'_{\rm max}/e') \right)\langle C_e \rangle}
674: {4 \pi},
675: \end{equation}
676:
677: \noindent
678: for $e' \ll e'_{\rm max}$. The
679: equivalent formula for the inclination excitations is:
680:
681: \begin{equation}
682: \label{eqRIdistanteq2}
683: p(i') =
684: \frac{G n_0 m_0
685: T_{\rm orb}}{i'^3} \frac{\log \left((e'_{\rm max}/i') \right)\langle C_i \rangle}{4 \pi}.
686: \end{equation}
687:
688: \noindent
689: For the smallest $e'$ and $i'$, the entire range of perturbing masses
690: contributes to the interaction frequency. This occurs for excitations of the order
691: $(m_{\rm min}/m_b)(v_b/v_0)$, below which the perturbation frequency is given by
692: equation \ref{eqPofE}.
693:
694:
695: \subsection{$2<\gamma<3$}
696: \label{secInt}
697:
698: The mass density of the perturbers when $2<\gamma<3$ is
699: dominated by perturbers of the smallest mass, $m_{\rm min}$. Distant
700: encounters by perturbers with this mass produce very small
701: perturbations; for very low $e'$ then, $p(e') \propto e'^{-3}$, given
702: by the simple model of section \ref{secBEecc}. The upper limit of
703: $e'$ caused by these perturbers interacting with impact parameters $b
704: \sim r_b$ is $e' \sim (m_{\rm min}/m_b)(v_b/v_p)$.
705:
706: Perturbers with
707: $m_{\rm min}$ cause eccentricity changes larger than this via close
708: encounters, but these encounters are less frequent than
709: interactions with perturbers of a higher mass and an impact parameter
710: of order $r_b$. Perturbations with a strength $(m_{\rm
711: min}/m_b)(v_b/v_p) \gg e' \gg (m_{\rm max}/m_b)(v_b/v_p)$ are most
712: often excited by perturbers with impact parameters of $\sim r_b$ and
713: masses $m \sim e' (v_p/v_b) m_b$. In other words the frequency of
714: perturbations is directly proportional to the slope and normalization
715: of the mass spectrum.
716:
717: In this case, the functions $p(e')$ and $p(i')$ cannot be determined
718: using the simplifications to equation \ref{eqDeltaegeneral} afforded
719: by very small or very large impact parameters. In general, the
720: perturbation frequency for a mass spectrum of $2<\gamma<3$ follows the
721: power law $p(e') \propto e'^{-(\gamma+1)}$. As an example we present
722: the perturbation frequency for $\gamma=25/12$. This corresponds to
723: $q=4.25$, the best fit to observations of the Kuiper belt size
724: distribution presented by \citet{FKH08}. We calculate from equation
725: \ref{eqRofe},
726:
727: \begin{equation}
728: \label{eqPintermediate}
729: p(e')=2.6 \frac{G n_0 m_0 T_{\rm orb}}{e'^{37/12}}
730: \left( \frac{m_0}{m_b} \frac{v_b}{v_0} \right)^{1/12}.
731: \end{equation}
732:
733: \noindent
734: It is simple to understand the relationship between equations
735: \ref{eqPofE} and \ref{eqPintermediate} with the following argument.
736: A perturbation of size $e'$ that occurs via an interaction at a distance
737: $r_b$ requires a perturber of mass about $m' \sim e' (v_0/v_b) m_b$.
738: If we interpret the total density in equation \ref{eqPofE} as only
739: the density in bodies around $m'$, then $\rho' \sim m' {\cal F}_m(m') \sim
740: n_o (m_0/m')^{\gamma-1}$, and we recover the scaling of equation
741: \ref{eqPintermediate}.
742:
743: The integral over
744: $b$ and the angular variables of equation \ref{eqDeltaigeneral} yield
745: a different coefficient for the perturbations to inclination:
746:
747: \begin{equation}
748: \label{eqPofIintermediate}
749: p(i')= \frac{G n_0 m_0 T_{\rm orb}}{i'^{37/12}}
750: \left( \frac{m_0}{m_b} \frac{v_b}{v_0} \right)^{1/12}.
751: \end{equation}
752:
753: \noindent
754: We relegate to the appendix the details of the integrals that
755: produce the coefficients of equations \ref{eqPintermediate} and
756: \ref{eqPofIintermediate}.
757:
758: %\subsection{$\gamma>3$}
759:
760: %For all $\gamma >3$, the distribution is given by a Gaussian. This is
761: %because the accumulation of the smaller excitations overwhelms the
762: %rare largest excitations. The eccentricity and relative inclination
763: %then random walk diffusively, causing $e_c(t) \propto t^{1/2}$.
764:
765: \subsection{Collisional Perturbations}
766: \label{secColRate}
767:
768: The integral of equation \ref{eqRofe} over impact parameters from 0 to
769: $r_j$ produces the frequency of perturbations to the binary by
770: collisions on member $j$. Since the size of the impulse from a
771: collision does not depend on the impact parameter, it is the mass of the
772: perturber that dictates the size of the eccentricity perturbation.
773: Accordingly, the frequency of perturbations as a function of $e'$
774: reflects the frequency of collisions as a function of $m_p$. The frequency
775: of collisional perturbations does not depend on $m_{\rm max}$ or
776: $m_{\rm min}$ regardless of the slope. However, the limits of the mass
777: distribution specify the lowest and highest perturbations achievable
778: via collisions: $\chi (m_{\rm min}/m_j)(v_0/v_b) \leq e' \leq \chi (m_{\rm max}/m_j)(v_0/v_b)$.
779: In this range of $e'$, for any value of $\gamma$, the
780: perturbation frequency due to collisions is
781:
782: \begin{equation}
783: \label{eqPcol}
784: p(e') = \frac{G n_{0} m_b T_{\rm orb} }{e'^{\gamma+1}}
785: \left (\chi \frac{m_0}{m_j} \right)^{\gamma-1}
786: \left(\frac{v_0}{v_b}\right)^{\gamma} \left(\frac{r_j}{r_b}\right)^2
787: V_{\gamma} \frac{ (\gamma-1) \langle D_e^{\gamma-1}\rangle}{2 \pi},
788: \end{equation}
789:
790: \noindent
791: where $\langle D_e^{\gamma-1} \rangle$ is the average of the angular
792: dependence of $\Delta \vec e$ from collisions to the power of
793: $\gamma-1$, and $V_\gamma \equiv v_0^{-\gamma} \int v_p^{\gamma+2}
794: {\cal F}_v(v_p) dv_p$. If ${\cal F}_v(v_p)$ is proportional to a
795: delta function, $\delta(v_p-v_0)$, then $V_{\gamma}=1$ for all
796: $\gamma$. If the velocity spectrum were Gaussian, such that ${\cal
797: F}_v(v_p) \propto \exp (-(v_p/v_0)^2)$, then $V_\gamma=2
798: \Gamma((3+\gamma)/2)/\sqrt{\pi}$.
799: The frequency of perturbations to the relative inclination by
800: collisions is the same as equation \ref{eqPcol}, replacing the
801: integrated coefficient $\langle D_e^{\gamma-1} \rangle$ with the
802: appropriate calculation made from the coefficients of $|\Delta \vec
803: i|$.
804:
805: Although we use $r_j$ to represent either member of the binary, it is
806: clear from equation \ref{eqPcol} that the collisions onto the smallest
807: body have the largest effect on the orbit. The ratio of the
808: perturbation frequency through collisions, $p(e')_{\rm collisions}$
809: (equation \ref{eqPcol}) to the frequency of gravitational scatterings,
810: $p(e')_{\rm gravity}$ (equation \ref{eqPintermediate}), is, for mass
811: distributions of $2 < \gamma <3$,
812:
813: \begin{equation}
814: \label{eqRatio}
815: \frac{p(e')_{\rm collisions}}{p(e')_{\rm gravity}}
816: = 0.03 \left( \frac{r_j}{r_b} \right)^2
817: \left[\chi \frac{m_b}{m_j} \left( \frac{v_0}{v_b} \right)^2 \right]^{\gamma-1},
818: \end{equation}
819:
820: \noindent
821: where we have evaluated the coefficients for $\gamma=25/12$. The choice
822: of $\gamma$ does not change these coefficients dramatically.
823:
824: %The total frequency of perturbations to the binary is the
825: %sum of the gravitational scattering and the collisional perturbation frequencies.
826: %The distribution set by both follows equation
827: %\ref{eqFofe} with the sum of both $p(e)$ in the scattering integrand.
828: %For $2<\gamma<3$ the dependence on eccentricity is cancelled as seen above
829: %in equation \ref{eqRatio}. For shallower mass distributions, the
830: %gravitational scattering rate, which goes as $e^{-3}$, is
831: %overwhelmed by the effects of collisions, whose eccentricity
832: %dependence follows the mass distribution, $e^{-(\gamma+1)}$.
833:
834: %The contribution of collisions with this mass spectrum to the
835: %excitation rate, $p_{\rm col}(e)$, relative to that of gravitational
836: %interactions, $p_{\rm grav}(e)$, is given by the ratio
837:
838: %\begin{equation}
839: %\label{eqGravVsCol}
840: %\frac{p_{\rm col}(e)}{p_{\rm grav}(e)}=\frac{2}{\Lambda}
841: %\frac{\langle D_e \rangle}{\langle C_e \rangle} \chi
842: %\left( \frac{ A v_p}{v_{\rm esc}} \right)^2 \frac{r_i}{r_b}
843: %\end{equation}
844:
845: %For the Pluto
846: %Charon binary collisions contribute about 10 percent of the
847: %excitation rate of distant encounters (see section \ref{secPluto}).
848: %For the outer satellites that are at least 10 times smaller but have
849: %orbits larger by a factor of several, the collisions increase the
850: %perturbation rate by about 25 percent.
851:
852: \subsection{Eccentricity Distributions}
853:
854: %The distribution given by equation \ref{eqGofx} and the equation to
855: %calculate $e_c(t)$ were derived in the context of $p(e') \propto
856: %e'^{-3}$. The fact that $p(e')$ is a power law allows the distribution
857: %function $f(e,t)$ to be self-similar; its shape then follows the same
858: %power-law at high $e$, and its characteristic value $e_c(t)$ evolves
859: %linearly with time. It is these two properties that identify the
860: %evolution of the eccentricity as a L\'{e}vy flight: a power-law
861: %distribution function, and an average traversal of phase space
862: %that is faster than $t^{1/2}$ \citep{SZF95}.
863:
864: The distribution given by equations \ref{eqGofx} and \ref{eqDiffeoft}
865: were derived in the context of $p(e') \propto e'^{-3}$. As long as
866: $p(e')$ follows a power law with $e'$, we can write a self-similar
867: distribution function $f(e,t)$. We write a generic function,
868: $p(e')=P_0 e'^{-(1+\eta)}$, to account for the different slopes caused
869: by different mass distributions (for $3 > \gamma > 2$, $\eta =
870: \gamma$; for $\gamma < 2$, $\eta=2$). The derivation of the
871: distribution function proceeds analogously as in section
872: \ref{secBEecc}. Equation \ref{eqFofe} becomes two equations: a
873: dimensionless integro-differential equation that specifies the shape,
874: and an ordinary differential equation to specify the evolution of the
875: eccentricity scale $e_c(t)$. The general version of equation
876: \ref{eqDiffeoft} is:
877:
878: \begin{equation}
879: \label{eqDiffeoftgeneral}
880: \dot e_c(t) = - e_c(t)/\tau_d + 2 \pi P_0 / e_c(t)^{\eta-2}.
881: \end{equation}
882:
883: \noindent
884: In the limit of no eccentricity dissipation ($\tau_d \rightarrow
885: \infty$), equation \ref{eqDiffeoftgeneral} shows that $e_c(t) \propto
886: t^{1/(\eta-1)}$. For all of the $p(e')$ discussed in section \ref{secMassSpec},
887: the growth of $e_c(t)$ is always faster than $t^{1/2}$.
888:
889: The shape of the distribution function is determined through a Fourier
890: transform of the general version of equation \ref{eqFofg}. For slopes
891: of $1<\eta<3$, $g(x)=\int \cos (\vec k \cdot \vec x) \exp(-|\vec
892: k|^{\eta-1})d^2 \vec k $ \citep{S99,CSS07}. While there is only a
893: closed form solution for $\eta=2$, given by equation \ref{eqGofx}, all
894: of these functions are flat at low $x$ and fall off like
895: $x^{-(\eta+1)}$. In fact, it is easy to show from equation
896: \ref{eqFofe} that the high $e$ tail is given by
897:
898: \begin{equation}
899: \label{eqHighTail}
900: f(e \gg e_c(t)) = p(e) t / (\gamma-1),
901: \end{equation}
902:
903: \noindent
904: when $t \ll \tau_d$. For equilibrium distributions where $\dot
905: e_c(t)=0$, $t$ is replaced with
906: $\tau_d$, the timescale for the dissipation.
907:
908: When $p(e') \propto e'^{-4}$ or steeper, the accumulation of the
909: smallest perturbations over time is more effective at raising the
910: eccentricity of the binary than single large perturbations. In this
911: case, the evolution of the eccentricity follows standard Brownian
912: motion, where the distribution function is a Gaussian, and $e_c(t)
913: \propto t^{1/2}$.
914:
915: \section{Kuiper Belt Binaries}
916: \label{secKBB}
917:
918: In this section we compute $e_c(t)$ and $i_c(t)$ for several Kuiper
919: belt binaries. The ``binary'' of section \ref{secSingleEnc} now refers to
920: a bound pair of Kuiper belt objects, and the ``perturbers'' are all of the
921: other members of the Kuiper belt.
922:
923: For the highest mass KBOs, the size spectrum is well determined to be
924: a power law with an index slightly greater than $q=4$. The lowest
925: mass bodies, of about 30 km in radius, are less frequent than
926: predicted by a single power law, however the parameters of a more
927: general model are still under investigation
928: \citep{TB01,LJ02,PS05,FKH08,FH08}. For this section we use the best
929: fit of a single power law model to the high mass part of the spectrum
930: provided by \citet{FKH08}, who find $q=4.25$ and a number density of 1
931: body per square degree brighter than magnitude 23.4. We assume an
932: average distance of 40 AU to the Kuiper belt and a depth of 20 AU to
933: find a volumetric number density $n_0=3 \times 10^{-41} ~{\rm
934: cm^{-3}}$. To convert the magnitudes of the objects to physical
935: sizes, we assume a constant geometric albedo of 0.04, a constant
936: physical density of 1 g ${\rm cm^{-3}}$, and take the R-band apparent
937: magnitude of the Sun to be -27.6. We find that the magnitude 23.4
938: corresponds to a mass $m_0=1.75\times 10^{21}~{\rm g}$, equivalent to
939: a radius of 75 km. Most of the objects found between 30-50 AU are
940: inclined by about 5-15 degrees relative to the plane of the solar
941: system, and have heliocentric eccentricities of 0.1-0.2.
942:
943: \subsection{Perturbations by a Disk}
944: \label{secDisk}
945:
946: Our analysis so far has treated the perturbing bodies as unbound
947: objects moving relative to the binary with a constant velocity. When
948: the perturbers are part of a disk orbiting the central star, the
949: orbital elements of the disk set the parameters of the perturbation
950: frequencies we calculate in section \ref{secBE}.
951:
952: The relative velocity between KBOs, when they interact, is set by the
953: size of their eccentricities and inclinations, $v_p \sim e_{H} a
954: \Omega_H$, where the subscript ``H'' denotes a heliocentric orbital
955: quantity. We assume a constant perturbing velocity with $v_p = 1~{\rm
956: km/s}$, which corresponds to the typical heliocentric eccentricities
957: and inclinations of KBOs. We assume that these encounters occur
958: isotropically in the frame of a binary, however this is not accurate.
959: A more detailed calculation of the angular distribution of relative
960: velocities will only affect the coefficients of the perturbations.
961: The disk does not specify a special direction for the perturbation vector
962: $\Delta \vec e$, so the perturbing frequency and the distribution
963: function retain their axisymmetry. The influence of the central
964: star on the binary and the perturbers adds
965: another constraint to our assumption of impulsive encounters: the
966: timescale for an interaction must be shorter than the orbital period
967: around the star: $b/v_p \ll 1/\Omega_H$, or equivalently, $b \ll e_H
968: a$. This guarantees that the relative velocity is constant
969: during the interaction.
970:
971: If the orbit of the binary is much different than the typical KBO
972: orbit, there are several modifications to perturbation frequencies
973: experienced by the binary. One modification is due to the finite
974: height of the disk of perturbers. This height is set by their
975: inclinations around the central star; for the Kuiper belt we refer to
976: the average inclination as $\langle i \rangle_{KB}$. A binary with
977: heliocentric inclination $i_{\rm CoM} \ll \langle i \rangle_{KB}$
978: never travels above or below the perturbing disk height and therefore
979: experiences the maximal frequency of perturbations. If $i_{\rm CoM} \gg \langle
980: i \rangle_{KB}$, the binary spends most of its orbit outside of the
981: perturbing swarm. The frequency of perturbations to such a binary is
982: reduced by the fraction of the time the binary leaves the disk,
983: proportional to $\langle i \rangle_{KB}/i_{\rm CoM}$. The
984: eccentricity of the binary in the disk reduces the effective density
985: of perturbers in a similar manner if the epicycle of the binary
986: carries it outside of the region populated by perturbers.
987:
988: If the heliocentric eccentricity or inclination of the binary is much
989: greater than the typical values for the Kuiper belt, the relative
990: velocity between the binary and a perturber is primarily due to
991: the non-circular heliocentric motion of the binary.
992: Gravitational interactions depend weakly on $v_0$ so their frequency
993: does not change much in this case. Perturbations by collisions,
994: however, become more important if $v_0$ is increased due to this
995: effect (equation \ref{eqRatio}).
996:
997: \subsection{Pluto et. al.}
998: \label{secPluto}
999:
1000: Pluto is the second largest known Kuiper belt object, with a radius of
1001: about 1100 km. It has a semi-major axis of 39.5 AU and its orbit is
1002: inclined relative to the ecliptic by $17^{\circ}$. Its largest
1003: satellite, Charon, contains about one tenth of the total mass of the
1004: system. Recent observations have revealed two smaller satellites, Nix
1005: and Hydra \citep{W06}. These satellites have small eccentricities and
1006: are roughly co-planar with Charon. Numerical simulations of
1007: collisions between similarly sized objects by \citet{C05} produce
1008: binaries with orbits similar to Pluto and Charon. The circularity and
1009: co-planarity of Nix and Hydra lend additional weight to a collisional
1010: origin of the system.
1011:
1012: The triple system of Pluto and its moons is a valuable test case for
1013: the dynamics we have presented. For an isolated binary it is
1014: impossible to know the initial orbital plane. The relative
1015: inclinations of the moons of Pluto can be measured directly assuming
1016: their formation was co-planar. Furthermore, the perturbing swarm for
1017: all three Pluto-moon pairs is the same. A major issue in comparing
1018: our analytic calculations to the observations is that the large mass
1019: ratio of Charon to Pluto causes significant non-Keplerian effects in
1020: the orbits of the outer satellites. We first re-examine the published
1021: observational model of their orbits to separate the relevant motion
1022: of the outer satellites from the forced motion due to Charon. We then
1023: compare the resulting eccentricity with our predicted values.
1024:
1025: \subsubsection{Orbital Model of Tholen et al}
1026:
1027: A model of the observations of the Pluto system
1028: has been presented by \citet{TBGE07}, who fit the parameters of a
1029: four-body numerical integration such that the simulation agrees with
1030: the observations. Such work is necessary, as it has been
1031: shown that the observations cannot be consistently modeled by
1032: three non-interacting two-body orbits \citep{W06}.
1033:
1034: The model of \citet{TBGE07} presents a full set of osculating elements
1035: describing the orbits of Charon, Nix, and Hydra. The orbit of Charon
1036: is virtually unaffected by Nix and Hydra; \citet{TBGE07} measure the
1037: eccentricity of Charon to be $3.48 \pm 0.04 \times 10^{-3}$, and the
1038: period of its orbit is $6.387$ days. Since
1039: the combined potential of Charon and Pluto is significantly
1040: non-Keplerian, the elements of Nix and Hydra vary significantly during
1041: their orbits. \citet{TBGE07} average the osculating semi-major axis
1042: to find an orbital period for these satellites of 25.49 days and 38.73
1043: days for Nix and Hydra respectively. The osculating eccentricities of
1044: Nix and Hydra both oscillate between zero and about 0.2; for each
1045: satellite oscillations at the frequencies of its own orbit and
1046: that of Charon are visible (their figure 4). The orbital planes of the
1047: satellites relative to Charon's are tilted by 0.15 degrees for Nix and
1048: 0.18 degrees for Hydra. Each plane precesses relative to the plane of Charon,
1049: however the angle of the offset remains constant.
1050:
1051: \subsubsection{A Different Interpretation}
1052: \label{secInterpret}
1053:
1054: \begin{figure}[t!]
1055: \center
1056: \includegraphics[angle=-90,width=0.8\columnwidth]{binaries-fig3.ps}
1057: \caption{The distance of Nix (lower panel) and Hydra (upper panel) from the Pluto-Charon barycenter, in units of Pluto radii, as a function of time, in
1058: an integration of the parameters found by \citet{TBGE07}. Nix and Hydra are treated
1059: as massless test particles. The origin of the time coordinate is arbitrary.}
1060: \label{figRNixHydra}
1061: \end{figure}
1062:
1063:
1064: %The osculating Keplerian elements of Nix and Hydra are misleading.
1065: For two body motion, the Keplerian elements are constant and indicate
1066: the shape of the orbit in space. Osculating elements that describe
1067: motion in significantly non-Keplerian potentials, such as the combined
1068: potential of Pluto and Charon, may vary on timescales shorter than the
1069: orbital period of the satellite. When this is true, relating the
1070: osculating elements to the shape of the orbit can be misleading.
1071: The average value of the osculating eccentricity of Nix is
1072: 0.015 in the model of \citet{TBGE07}, however the motion of Nix
1073: relative to Pluto never resembles an ellipse with such an
1074: eccentricity.
1075:
1076: We re-examine the model provided by \citet{TBGE07} by reproducing the
1077: numerical integration based on the Pluto-centric positions and
1078: velocities of Charon, Nix, and Hydra published in their table 1. We
1079: set the masses of Nix and Hydra to zero to eliminate their secular
1080: interactions with each other. Instead of examining the
1081: osculating elements, we adopt the approach of \citet{LP06}
1082: and characterize the orbits of Nix and Hydra based
1083: on their position as a function of time from the Pluto-Charon
1084: barycenter, plotted in figure \ref{figRNixHydra}. The units of distance
1085: are Pluto radii, defined as $R_P = 1147~{\rm km}$.
1086:
1087: Although short oscillations on the timescale of Charon are visible in
1088: the top panel of figure \ref{figRNixHydra}, they are very small
1089: compared to the oscillations that occur on the timescale of Hydra's
1090: orbital period. To parametrize Hydra's orbit we fit the function
1091: $r_0(1+e \cos (\kappa_1 t + \omega_1))$ to the first 200 days of the
1092: numerical model. Because for a non-Keplerian potential the radial
1093: epicyclic frequency differs from the orbital frequency, we calculate
1094: the average angular frequency by fitting a straight line to the
1095: angular position of Hydra as a function of time, $f(t) = \Omega_1
1096: t+\lambda_0$. The results are written in table \ref{tabFitResults}.
1097: We interpret $e_1$ as the orbital degree of freedom in the combined
1098: potential of Pluto and Charon that is analogous to the eccentricity of
1099: a two-body orbit.
1100:
1101: \begin{deluxetable}{lcccccccc}
1102: \tablecaption{Best fit values to the epicyclic models of the radial motion
1103: of Nix and Hydra.}
1104: \tablehead{
1105: \colhead{} & \colhead{$r_0/R_P$} & \colhead{$e_1$}
1106: & \colhead{$2 \pi/\kappa_1$}& \colhead{$e_2$}
1107: & \colhead{$2 \pi/\kappa_2$}& \colhead{$e_3$}
1108: & \colhead{$2 \pi/\kappa_3$}& \colhead{$2 \pi/\Omega_1$} \\
1109: \colhead{} & \colhead{} & \colhead{$\times 10^{-3}$} & \colhead{(days)}
1110: & \colhead{$\times 10^{-3}$} & \colhead{(days)}
1111: & \colhead{$\times 10^{-3}$} & \colhead{(days)} & \colhead{(days)}}
1112: \startdata
1113: Nix & 46.805(5) & 2.96(3) & 25.22(2) &
1114: 1.25(3) & 8.599(8) & 1.38(3) & 4.298(1) & 24.8505(5) \\
1115: Hydra & 62.237(1) & 5.595(2) & 38.535(15) & & & & & 38.20(1) \\
1116: \enddata
1117: \tablecomments{The motion of Nix is fit with three epicyclic
1118: terms, while the motion of Hydra is only fit with one. The parenthesis
1119: indicate the 95 \% confidence level of the fit around the last digits.}
1120: \label{tabFitResults}
1121: \end{deluxetable}
1122:
1123: The motion of Nix (bottom panel of figure \ref{figRNixHydra}) appears
1124: more irregular than that of Hydra. We find the position of Nix to be
1125: well-described by a model of three epicycles with different
1126: frequencies: $r(t)=r_0(1+\sum_{k=1,2,3} e_k \cos (\kappa_k t +
1127: \omega_k))$. The best fit values are printed in table 1. We
1128: distinguish the cause of each epicycle by its period. The combined
1129: potential of Pluto and Charon oscillates with frequency of
1130: $\Omega_{\rm Charon}-\Omega_{\rm Nix}$; motion being forced by this
1131: potential should occur on integer multiples of this frequency.
1132: Using the numbers in table \ref{tabFitResults}, we see that
1133: $2 \pi/(\Omega_{\rm Nix}+\kappa_2) = 2\pi/(\Omega_{\rm Nix}+\kappa_3/2)=
1134: 6.39$ days. The second and third epicycles in our fit correspond
1135: to motion at the first and second harmonic of Nix's relative
1136: orbital frequency. We therefore interpret the first term, with a size of
1137: $e_1=3 \times 10^{-3}$ and a period close to Nix's orbital period,
1138: as analogous to the two-body eccentricity.
1139:
1140: We perform another integration of the best fit initial conditions from
1141: \citet{TBGE07} to investigate the secular effects between Nix and
1142: Hydra. We use the best fit masses from \citet{TBGE07} for the two
1143: outer satellites. Since the motion of Hydra is dominated by a single
1144: epicyclic frequency, the variation in the size of its epicycle is
1145: apparent on the timescale of several years. To determine the effect
1146: of secular variations on Nix, we fit the same three-component
1147: epicyclic model to five orbits at $t \sim 5$ years. In the best-fit model to these
1148: later orbits, the only difference compared to the model of table
1149: \ref{tabFitResults} is in $e_1$, the epicycle with a frequency close
1150: to Hydra's orbital frequency. This is further confirmation that the
1151: degree of freedom represented by $e_1$ is analogous to the two-body
1152: eccentricity.
1153: %The difference between the epicyclic motion we find for
1154: %Nix and Hydra, both of order $10^{-3}$, and the osculating
1155: %eccentricities that vary between 0 and 0.2 offer a strong warning
1156: %against the physical interpretation of osculating orbital elements.
1157:
1158: \subsubsection{Theoretical Distribution}
1159:
1160: To compute the distribution of eccentricities and inclinations
1161: expected of Pluto's moons, we solve equation \ref{eqDiffeoftgeneral}
1162: for each of the moons, given the interaction frequencies specified by
1163: equations \ref{eqPintermediate} and \ref{eqPofIintermediate}. The
1164: only remaining parameters to evaluate are the damping timescales for
1165: the eccentricity and inclinations of each satellite. We use the standard
1166: formula for the damping of eccentricity due to the tidal
1167: force of the primary acting on a secondary that is in synchronous
1168: rotation \citep{YP81,MD99}:
1169:
1170: \begin{equation}
1171: \label{eqTidalDamping}
1172: \tau_{d,2} = \frac{4}{63} Q_2 (1+{\tilde \mu_{2}}) \frac{m_2}{m_1}
1173: \left( \frac{r_b}{r_2} \right)^5 \frac{1}{\Omega},
1174: \end{equation}
1175:
1176: \noindent
1177: where $Q_2$ is the dissipation function of the secondary, and
1178: ${\tilde \mu_2} = 19 \mu r_2 / (2 \rho G m_2)$
1179: is its effective rigidity, a ratio between the
1180: material strength of the secondary and its self-gravity.
1181: The damping rate of
1182: eccentricity due to tides of the primary acting on the secondary,
1183: $\tau_{d,1}$, if the primary is also rotating synchronously with the
1184: orbit of the satellite, is given by equation \ref{eqTidalDamping} with
1185: the quantities specific to the primary switched with those of the
1186: secondary and vice versa.
1187:
1188: Pluto and Charon are known to be in a double-synchronous state of
1189: rotation, where the spin period of each body is equal to the 6.4 day
1190: orbital period. In many binaries, only the spin of the secondary is
1191: synchronous with the orbital period. Tides on the primary then raise
1192: the eccentricity. Double-synchronous systems, however, experience
1193: damping due to both the tides on the secondary and those on the
1194: primary. Assuming a water-ice composition for Pluto
1195: ($\mu=4~\times 10^{10}~{\rm dynes~cm^{-2}}$), we calculate the
1196: eccentricity damping timescale due to tides raised by Charon,
1197: $\tau_{d,1}$ from equation \ref{eqTidalDamping} to be 5.1 Myrs. The
1198: shortest damping timescale due to tides from Pluto acting on Charon,
1199: $\tau_{d,2}$ is found by assuming Charon is also made of water-ice; we
1200: find in this case a timescale of 8.2 Myr. The longest timescale
1201: assumes a rocky composition ($\mu=6.5~\times 10^{11}~{\rm dynes~cm^{-2}}$);
1202: we find this corresponds to 133 Myr. The
1203: overall damping of the system is given by the sum of the damping
1204: rates. The short damping timescale of tides on Pluto prevents Charon
1205: from contributing significantly to the combined effect of both tides,
1206: reducing the importance of its composition. The longest eccentricity
1207: damping timescale that results from both tides is 4.9 Myr. The
1208: inclinations of the outer satellites relative to the Pluto-Charon
1209: plane are also damped by tidal dissipation. For a circular synchronous
1210: orbit the timescale for inclination damping is longer than the
1211: timescale for eccentricity damping by a factor of $\sim i^{-2}$. We
1212: ignore the damping of inclinations in equation \ref{eqDiffioft} for
1213: all three satellites.
1214:
1215: As discussed in \citet{TBGE07} and section \ref{secInterpret}, secular
1216: interactions between the satellites are visible in the long term
1217: calculations of their orbits. For the best-fit values of the masses
1218: of Nix and Hydra, their eccentricities are modulated on the order of
1219: $10 \%$ over timescales of years; we neglect these fluctuations for
1220: this work. It is more important in this model to determine whether
1221: secular evolution can cause the eccentricity of Nix or Hydra dissipate
1222: via Charon's orbit.
1223:
1224: We use linear secular theory to describe the coupled evolution of the
1225: eccentricity and longitude of periapse of each satellite \citep{MD99}.
1226: We find that the undamped secular evolution agrees qualitatively with
1227: the numerical orbit determinations. We add a term to the differential
1228: equations describing Charon's eccentricity that reduces it at a
1229: constant timescale ($\dot e_{\rm Charon} = - e_{\rm Charon}/\tau_d$).
1230: The frequencies of the oscillations of the eigenmodes of the solution
1231: are practically unchanged by this term, however each eigenmode gains a
1232: dissipative factor. Quantitatively, only one eigenmode is damped on
1233: timescales shorter than than 4.5 Gyr. By integrating the damped
1234: secular equations with different initial periapses, we determined that
1235: the secular interactions do not cause substantial damping of Nix and
1236: Hydra.
1237:
1238: Equation \ref{eqRatio} gives the
1239: frequency of perturbations due to collisions of perturbers onto each
1240: moon relative to the frequency of perturbations caused by
1241: gravitational scattering, equation \ref{eqPintermediate}. For Charon,
1242: the collisional perturbations increase $p(e)$ by only 2 percent.
1243: Since Nix and Hydra are smaller, perturbations by collisions have a
1244: greater relative effect; however it is only a 20 percent contribution
1245: to the total perturbation frequency for Nix and 15 percent for Hydra.
1246: We solve equation \ref{eqDiffeoftgeneral} to find $e_c(t)$ and
1247: $i_c(t)$ for each of Pluto's moons.
1248:
1249: For Charon we find $e_c=2.6 \times 10^{-6}$, and $i_c=0.029^{\circ}$.
1250: This value of $e_c$ corresponds to the most likely perturbation during
1251: a damping timescale of 4.9 Myr, and is much smaller than the
1252: observed value of $3.5 \times 10^{-3}$ \citep{TBGE07}. Using equation
1253: \ref{eqHighTail}, we calculate that given this value of $e_c$, the
1254: probability of Charon's eccentricity being as high as its observed
1255: value is 0.2 percent.
1256:
1257: For Nix we calculate $e_c(4.5 ~{\rm Gyr})=4.8\times 10^{-3}$ and
1258: $i_c(4.5 ~{\rm Gyr})=0.1^{\circ}$, and for Hydra, $7.1 \times 10^{-3}$
1259: and $0.15^{\circ}$ respectively. The distributions specified by these
1260: values are quite consistent with the free eccentricity we determine in
1261: table \ref{tabFitResults}.
1262:
1263: \subsection{Other Interesting KBOs}
1264: \label{secOthers}
1265:
1266: Two other Kuiper belt objects have satellites on low eccentricity
1267: orbits: \el, and Eris. Along with Pluto these are three of the four
1268: most massive KBOs known, all with radii of about 1000 km.
1269: \el has two known satellites. The largest has a 50 day orbit and a
1270: measured orbital eccentricity of $0.050 \pm 0.003$ \citep{BBR05}. An
1271: additional smaller satellite orbits \el~ with a period of about 35
1272: days \citep{BVB06}. The orbital parameters of the inner satellite are
1273: unconstrained, however the relative inclination between the two is
1274: about $40^{\circ}$. The masses of the satellites are negligible compared
1275: to the mass of \el. The heliocentric inclination of the system is
1276: $28^{\circ}$.
1277:
1278: \citet{BBR05} argue that if the tidal response of \el~ and its large
1279: satellite are fluid-like, tidal interactions should damp their
1280: eccentricity on a timescale of about 300 Myr. With these parameters
1281: we use equation \ref{eqDiffeoftgeneral}
1282: to calculate an equilibrium $e_c=4.3 \times 10^{-4}$. The
1283: distribution with this eccentricity scale predicts an observed
1284: eccentricity of 0.05 at a probability of three percent. However, for
1285: smaller bodies, internal elastic forces dominate the tidal
1286: deformation of their shape; it is more reasonable to assume that the
1287: tidal response of the satellite is characterized by its material
1288: strength. Then, the tides raised on the primary have the greatest effect
1289: and the eccentricity of the system grows on the same timescale as the
1290: growth of the semi-major axis. Forced eccentricity growth and an
1291: evolving orbital period can be incorporated into equation
1292: \ref{eqDiffeoftgeneral}. However, these corrections are only an order
1293: unity correction since the growth timescale, by definition, is
1294: comparable to the age of the system. Assuming $T_{\rm orb}$ is fixed
1295: and ignoring the eccentricity growth, we calculate $e_c(4.5~{\rm
1296: Gyr})=0.0052$. The 95 percent confidence interval around this $e_c$ is
1297: 0.001-0.2; the observed eccentricity of \el is within this range.
1298:
1299: The dwarf planet Eris is orbited by the satellite Dysnomia.
1300: Observations have shown an upper limit to their eccentricity of 0.013
1301: \citep{BVB06}. The system has a 15 day orbital period, and orbits the
1302: sun at a semi-major axis of 67.7 AU with an eccentricity of 0.44 and a
1303: heliocentric inclination of $44^{\circ}$. In addition to the
1304: reduction in effective perturbing density caused by the inclination,
1305: the high eccentricity reduces the effective perturber density by an
1306: additional factor of 0.09. The semi-major axis of the binary is
1307: consistent with 4.5 Gyr of tidal evolution away from an initially very
1308: close orbit; if the tidal response of the secondary that of a
1309: strength-less fluid, then its eccentricity is damped on a timescale of
1310: 50 Myr. These parameters yield an $e_c=2.2\times 10^{-6}$. However,
1311: if the material strength of the secondary is stronger than its own
1312: self-gravity, then the tides raised on the primary cause the
1313: eccentricity of the satellite to grow. In this case the relevant
1314: timescale is the age of the system, and we find that $e_c(4.5~{\rm
1315: Gyr})=1.0\times 10^{-4}$. Both values are below the observed upper
1316: limit.
1317:
1318: In addition to the high mass ratio and low eccentricity Kuiper belt
1319: binaries, there are other known binaries of almost equal mass on
1320: moderately eccentric orbits. The binary $1998~{\rm WW}_{31}$ is an
1321: example of such an object: both members have a radius of about 50 km,
1322: an orbital period of 574 days, and a mutual eccentricity is 0.817
1323: \citep{VPG02}. Even though our analysis is derived in the
1324: low eccentricity limit, we can use equation \ref{eqDiffeoftgeneral}
1325: to estimate approximately the eccentricity expected from impulsive
1326: encounters; we find $e_c(4.5~{\rm Gyr}) = 0.31$. This moderate
1327: characteristic eccentricity is consistent with the high observed
1328: value. Other binaries with orbital periods on the order of a year
1329: will have acquired large eccentricities through their interactions
1330: with the other Kuiper belt objects.
1331:
1332: \section{Other Binary Systems}
1333: \label{secApp}
1334:
1335: Our analysis holds for any two-body orbit perturbed isotropically in
1336: the impulsive limit. As binary orbits are prevalent in astrophysics,
1337: we briefly discuss several other examples.
1338:
1339: The asteroid belt harbors many binaries with well determined
1340: eccentricities. The mass spectrum of the asteroid belt, however, is
1341: much shallower than that of the Kuiper belt: the largest asteroid,
1342: Ceres, contains a third of the total mass of all asteroids. A binary
1343: asteroid is then perturbed mostly by the largest objects that it
1344: encounters. To calculate $p(e')$ accurately, it is necessary to model
1345: the neighborhood of that binary. The asteroid belt is also
1346: collisionally active so its binaries may not be coeval with the whole
1347: solar system. We postpone a detailed analysis of the binary asteroid
1348: population for a future work.
1349:
1350: A well-measured class of binaries outside the solar system are
1351: millisecond pulsars with white dwarf companions. The tidal damping
1352: between the pulsar and its companion in the phase before the companion
1353: becomes a white dwarf is very short, indicating that during this phase
1354: the eccentricity of the binary should be smaller than the observed
1355: values of around $10^{-4}-10^{-5}$ \citep{S04}. To explain the
1356: observations, \citet{P92} presents the following model. As the
1357: companion star becomes a white dwarf, random fluctuations in the
1358: atmosphere of the star cause irregular motion in the orbit of the
1359: binary. These motions are reflected by a small eccentricity that
1360: remains since the tidal interactions between the white dwarf and the
1361: neutron star cannot damp the system. The model of \citet{P92} produces
1362: eccentricities for these systems that match the observations well.
1363:
1364: These binaries are perturbed by encounters with other stars in the
1365: galaxy; we can calculate the contribution to their eccentricities by
1366: the distant stellar interactions. The perturbation of these systems
1367: by other stars falls into the simple regime of only distant
1368: interactions described in section \ref{secBEecc}. A typical
1369: volumetric mass density for field stars is $0.1 {\rm M_{\odot}~
1370: pc^{-3}}$ \citep{HF00}. Given this density, we calculate the
1371: characteristic eccentricity of these systems to be
1372:
1373: \begin{equation}
1374: \label{eqEcpulsars}
1375: e_c(t)= 1.2 \times 10^{-9} \left( \frac{T_{\rm orb}}{1~{\rm day}}\right)
1376: \left( \frac{t}{1~{\rm Gyr}} \right)
1377: \left( \frac{\rho}{0.1 M_{\odot}~{\rm pc}^{-3}}\right).
1378: \end{equation}
1379:
1380: \noindent
1381: Typical orbital periods are between 1 and 10 days, and the ages of
1382: these systems are on the order of Gyrs. We find then that $e_c(t)$ is
1383: several orders of magnitude lower than the observed eccentricities.
1384: \citet{P92} also concludes that the perturbations from other stars
1385: cannot be responsible for the eccentricities of the binary pulsars.
1386: Since we have calculated the distribution, however, we can estimate
1387: more accurately the likelihood of achieving these eccentricities
1388: by only distant stellar perturbations: less than 0.1 percent.
1389:
1390: Globular clusters can have densities many orders of magnitudes higher than
1391: the average galactic density, such that distant perturbations to the
1392: binaries may be important. However, in a cluster the
1393: interactions between a binary and a star are not
1394: typically in the impulsive interaction regime. Instead the orbits of
1395: the perturbers are affected by the gravity of the binary, and the
1396: interactions occur over several orbital periods. Analytic work on the
1397: eccentricity perturbations in this regime has been performed by
1398: \citet{RH95} and \citet{HR96}.
1399:
1400: The characteristic eccentricity caused by distant stellar passages on
1401: the orbits of extra-solar planets is also given by equation
1402: \ref{eqEcpulsars}. These eccentricities are too low to be
1403: reflected in the current sample of known extra-solar planets. As with
1404: the pulsar binaries, the distant interactions may play a role in
1405: setting the eccentricity distribution of long period planets found in a dense
1406: stellar cluster. For most extra-solar planets however, planet-disk
1407: interactions \citep{GS03} or planet-planet scatterings \citep{RF96}
1408: are probably the source of their eccentricity.
1409:
1410: \section{Conclusions}
1411: \label{secConclusions}
1412:
1413: We have calculated the effects of impulsive perturbations and
1414: collisions on a nearly circular Keplerian orbit. If the swarm of
1415: perturbers encounter the binary isotropically in space, we can write a
1416: distribution function that describes the probability density for the
1417: binary to have a given eccentricity or inclination relative to its
1418: initial plane. The growth rate of the binary's likeliest eccentricity
1419: and inclination depends on the mass spectrum of the perturbers. For
1420: shallow mass distributions ($q<4$) it is the distant encounters that
1421: set the binary's eccentricity and only the total mass density of perturbers
1422: is important to the evolution of the binary. For steeper mass distributions of
1423: $q=4-7$, it is the interactions at about the semi-major axis of the binary
1424: that dominate the frequency of perturbations. Only the normalization and
1425: slope of the
1426: mass spectrum set the distribution of eccentricities in this regime.
1427:
1428: The assumptions of this model are valid in the Kuiper belt. Our
1429: calculations match the observations of Nix and Hydra very well. For
1430: Eris and \el, the observations lie within the 95 percent confidence
1431: intervals of the distributions we calculate, assuming the tidal
1432: response of the secondaries is dominated by material strength. For
1433: Charon our theory is consistent with the numerical simulations of
1434: \citet{SBL03}, predicting an eccentricity about 3 order of magnitudes
1435: smaller than observed. However, our analysis alleviates their need
1436: for numerical simulations as well as predicts the entire distribution
1437: of the eccentricity. The distributions measured by \citet{SBL03} are
1438: not all correct as their model includes only impact parameters out to
1439: twice the semi-major axis. In their simulations where $q=3.5$ and 4.0
1440: this excludes the impacts that are most relevant over an eccentricity
1441: damping timescale. Our results show that for $q=3.5$ the interactions
1442: that dominate Charon's eccentricity are Pluto-sized perturbers
1443: interacting at about 200 times the semi-major axis!
1444:
1445: %For Eris, we predict an eccentricity of the order $10^{-4}$, several
1446: %orders of magnitudes below current observational limits. The main
1447: %satellite of the binary \el~ has an eccentricity much higher than
1448: %predicted by this model, indicating that another process has
1449: %interfered with its orbit. The presence of an additional satellite
1450: %that is highly inclined may be evidence that its violent past
1451: %continues to dictate the dynamics of its satellites.
1452:
1453: Even without eccentricity dissipation through tides, perturbations
1454: from other Kuiper belt objects are too weak to excite eccentricities
1455: of order 1 or inclination changes of order a radian for binaries that
1456: have orbital periods of a few days or weeks. It is not likely that
1457: the orbital planes of the close binaries have been affected
1458: significantly by other Kuiper belt objects given our current
1459: understanding of the history of the Kuiper belt. It falls on theories
1460: of binary formation to explain the distribution of orbital
1461: inclinations relative to the ecliptic for close binaries. Since
1462: $e_c(t)$ grows faster for binaries with large orbital periods, it is
1463: plausible that the smaller wide binaries ($1998 {\rm WW}_{33}$ for
1464: example) have been brought to large eccentricities and inclinations by
1465: interacting with the rest of the Kuiper belt.
1466:
1467: When many binaries share the same perturbing swarm, such as in the
1468: Kuiper belt, we can use the eccentricities of all the binaries to
1469: probe the properties of the entire system. For example, if the mass
1470: spectrum is steeper than $q=4$, the distribution of
1471: eccentricity is directly related to the slope and
1472: normalization of the mass spectrum.
1473: Conversely, the observed eccentricity can be used to place limits on
1474: the damping timescale of a binary and therefore the rigidity of those
1475: bodies. The small sample of Kuiper belt binaries with well measured
1476: eccentricities limits the current effectiveness of such a calculation.
1477: However, the Pan-STARRS project plans to detect around 20000 more
1478: members of the Kuiper belt \citep{KAB02}; from these the number of
1479: orbit-determined Kuiper belt binaries will surely increase.
1480:
1481: The distribution we describe with equation \ref{eqGofx} is a special
1482: case of a L\'{e}vy distribution \citep{S99}. This class
1483: of functions arise in the generalization of the central limit theorem
1484: to variables distributed with an infinite second moment.
1485: Alternatively, these functions can be characterized by the properties
1486: of the L\'{e}vy flight they describe. For the eccentricity of the
1487: binaries discussed in this work, the frequency of a step is inversely
1488: proportional to a power of its size that depends on the
1489: mass spectrum of perturbers. It follows that the largest single step
1490: dominates the growth from accumulated smaller steps, causing, in the
1491: absence of damping, the typical eccentricity to grow faster than in a
1492: normal diffusive random walk. The slope of the distribution of
1493: excitations dictates the shape of the distribution. This explains the
1494: coincidence of the distribution we derive in this work being exactly
1495: that of the distribution of eccentricity of protoplanets in a
1496: shear-dominated planetesimal disk, where the probability of changing the
1497: eccentricity of a protoplanet is inversely proportional to the size of
1498: that change \citep{CS06,CSS07}.
1499:
1500: The authors thank Dmitri Uzdensky and Scott Tremaine for valuable
1501: discussions. R.S. is a Packard Fellow and an Alfred P. Sloan Fellow.
1502: This research was partially supported by the ERC.
1503:
1504: \appendix
1505:
1506: To calculate the excitation rates presented in sections \ref{secBE}
1507: and \ref{secMassSpec}, it is necessary to integrate over all possible
1508: configurations of angles $\vec b$ and $\vec v_p$ relative to $\vec
1509: r_b$ and $\vec v_b$. In this appendix we clarify the relation between
1510: the coefficients and equations \ref{eqDeltaegeneral} through
1511: \ref{eqDeltah}.
1512:
1513: We choose spherical polar
1514: coordinates for $\vec b$ and $\vec v_p$ to integrate
1515: equation \ref{eqRofe}. This requires a
1516: polar and azimuthal angle for $\vec b$, $\theta_b$ and $\phi_b$, and a
1517: polar and azimuthal angle for $\vec v_p$, $\theta_v$ and $\phi_v$.
1518: By defining $\theta_v$ relative to $\vec b$, the requirement that
1519: $\vec b$ and $\vec v_p$ be perpendicular fixes $\theta_v = \pi/2$.
1520:
1521: The magnitude of the perturbation only depends on the vectors
1522: $\vec b$ and $\vec v_p$ relative to $\hat r_b$ and $\hat v_b$, so we use
1523: these vectors and their cross product, $\hat n$ to describe the
1524: components of $\hat b$: $\hat b = b_r \hat r_b + b_v \hat v_b + b_n
1525: \hat n $. The components are related to $\theta_b$ and $\phi_b$ in
1526: the typical way: $b_r=\cos \phi_b \sin \theta_b$, $b_v=\sin \phi_b
1527: \sin \theta_b$, and $b_n= \cos \theta_b$.
1528: We define the components
1529: of $\vec v_p$ relative to the same unit vectors.
1530: The angle $\phi_v$ describes the direction of $\vec v_p$ in
1531: the plane given by $\hat b$; the components of $\vec v_p$ follow
1532: from a rotation of this plane to align with $\hat n$.
1533: We find the relations:
1534: %\begin{eqnarray}
1535: %\label{eqVconversion}
1536: %v_r & =& \cos \theta_b \cos \phi_v - ( \sin^2 \theta_b \sin \phi_b ) ( \cos \phi_b \sin \phi_v - \sin \phi_b \cos \phi_v) / (1+\cos \theta_b), \nonumber \\
1537: %v_v & =& \cos \theta_b \sin \phi_v + (\sin^2 \theta_b \cos \phi_b) (\cos \phi_b \sin \phi_v - \sin \phi_b \cos \phi_v) / (1+\cos \theta_b),\\
1538: %v_n& =& -\sin \theta_b (\cos \phi_b \cos \phi_v + \sin \phi_b \sin \phi_v). \nonumber
1539: %\end{eqnarray}
1540:
1541: \begin{eqnarray}
1542: \label{eqVconversion}
1543: v_r & =& b_n\cos \phi_v
1544: - b_v ( b_r \sin \phi_v - b_v \cos \phi_v) / (1+b_n), \nonumber \\
1545: v_v & =& b_n \sin \phi_v + b_r (b_r \sin \phi_v - b_v \cos \phi_v)
1546: / (1+ b_n),\\
1547: v_n& =& - b_r \cos \phi_v - b_v \sin \phi_v. \nonumber
1548: \end{eqnarray}
1549:
1550: The coefficient from equations \ref{eqPofE} and \ref{eqRdistanteq2},
1551: $\langle C_e \rangle$, is defined to be the integral of
1552: $|\Delta \vec e|/(8 \pi^2 (m_p/m_b)(v_b/v_p)(r_b/b)^2)$ as given by
1553: equation \ref{eqDeltae}:
1554:
1555: \begin{equation}
1556: \label{eqAvgCe}
1557: \langle C_e \rangle =
1558: \frac{1}{4 \pi^2} \int_0^{2 \pi} \int_0^{2 \pi} \int_{0}^{\pi}
1559: \left[ (4 b_r b_v +2 v_r v_v)^2+(1-v_r^2-2 b_r)^2
1560: \right]^{1/2} \sin \theta_b d\theta_b d\phi_b d\phi_v =1.89
1561: \end{equation}
1562:
1563: We similarly define $\langle C_i \rangle$ from equation \ref{eqDeltah}:
1564:
1565: \begin{equation}
1566: \label{eqAvgCi}
1567: \langle C_i \rangle =
1568: \frac{1}{4 \pi^2} \int_0^{2 \pi} \int_0^{2 \pi} \int_{0}^{\pi}
1569: | 2 b_r b_n+ v_r v_n | \sin \theta_b
1570: d\theta_b d\phi_b d\phi_v = 0.75.
1571: \end{equation}
1572:
1573: To calculate the coefficients used in the collisional excitation rate,
1574: equation \ref{eqPcol}, we use the $|\Delta \vec e|$ discussed in
1575: section \ref{secCollisions}.
1576:
1577: \begin{equation}
1578: \label{eqAvgDe}
1579: \langle D_e^{\gamma-1} \rangle =
1580: \frac{1}{4 \pi^2} \int_0^{2 \pi} \int_0^{2 \pi} \int_{0}^{\pi}
1581: (4 v_v^2+v_r^2)^{(\gamma-1)/2}
1582: \sin \theta_b d\theta_b d\phi_b d\phi_v
1583: \end{equation}
1584:
1585: \noindent
1586: For $\gamma=2$, the integral has a closed form solution of
1587: $\langle D_e \rangle = E(-3)$, the complete Elliptic integral.
1588: For the inclination,
1589:
1590: \begin{equation}
1591: \label{eqAvgDi}
1592: \langle D_i^{\gamma-1} \rangle
1593: = \frac{1}{4 \pi^2} \int_0^{2 \pi} \int_0^{2 \pi} \int_{0}^{\pi}
1594: |(v_z)^{\gamma-1}|
1595: \sin \theta_b d\theta_b d\phi_b d\phi_v = \frac{1}{\gamma}
1596: \end{equation}
1597:
1598: The coefficients for the excitation rates in the regime of
1599: $2<\gamma<3$ are more complicated as the dependence on
1600: $b/r_b$ cannot be factored out of the coefficient. In addition to
1601: integrating over all angles, we must integrate over impact parameter.
1602: For any $\gamma$, equation \ref{eqPintermediate} is written:
1603:
1604: \begin{equation}
1605: \label{eqPIntGeneral}
1606: p(e) = \frac{G n_0 m_0 T_{\rm orb}}{e^{\gamma+1}}
1607: \left( \frac{m_0}{m_b} \frac{v_b}{v_0} \right)^{\gamma-2}
1608: \frac{\gamma-1}{2 \pi} V_{2-\gamma} \langle A_e^{\gamma-1} \rangle,
1609: \end{equation}
1610:
1611: \noindent
1612: where $V_{\gamma-2}$ is discussed in section \ref{secColRate}; for a
1613: Gaussian distribution of perturber velocities, $V_{\gamma-2}=2
1614: \Gamma((1+\gamma)/2)$. The term $\langle A_e^{\gamma-1} \rangle$
1615: again contains the angular information. Excitations for $2 < \gamma <
1616: 3$ are most important at $b \sim r_b$ so we can not assume that $\vec
1617: b_2 \approx \vec b$. We introduce explicit notation for the the
1618: components of the unit vector $\hat b_2 = b_{2,r} \hat r_b + b_{2,v}
1619: \hat v_b + b_{2,n} \hat n$. Then the angular average coefficient is:
1620:
1621: \begin{equation}
1622: \label{eqAvgAe}
1623: \langle A_e^{\gamma-1} \rangle =
1624: \frac{1}{8 \pi^2}
1625: \int_0^{2 \pi} \int_0^{2 \pi} \int_{0}^{\pi} \int_{0}^{\infty}
1626: \left[
1627: 16 \left(\frac{b_{2,v}}{x_2}-\frac{b_v}{x_1} \right)^2
1628: + 4 \left(\frac{b_{2,r}}{x_2}-\frac{b_{r}}{x_1}\right)^2
1629: \right]^{(\gamma-1)/2} x_1 \sin \theta_b d x_1 d \theta_b d \phi_b d \phi_v,
1630: \end{equation}
1631:
1632: \noindent
1633: with $x_1=b/r_b$ and $x_2=b_2/r_b$. The magnitude and components
1634: of $\vec b_2$ are related to $\vec b$ and $\vec v_p$ as described
1635: in section \ref{secSingleEnc}:
1636: $\vec b_2 = \vec b - \vec r_b + \hat v_p (\vec r_b \cdot \hat v_p)$.
1637: For $\gamma=25/12$ as discussed in \ref{secInt},
1638: $\langle A_e^{13/12} \rangle \approx 15$. For other $\gamma$ between
1639: 2 and 3, this factor is of the same order, 10-15.
1640:
1641: \bibliographystyle{apj}
1642: \bibliography{ms}
1643:
1644: \end{document}
1645: