1: \documentclass[12pt,preprint]{aastex}
2: %\documentclass{emulateapj}
3:
4: %\usepackage{showkeys}
5:
6: \newcommand{\etal }{{et al.} }
7: \newcommand{\msun}{\thinspace M_\odot}
8: \newcommand{\rsun}{\thinspace R_\odot}
9: \newcommand{\vect}[1]{\mbox{\boldmath$#1$}}
10: \newcommand{\dfrac}[2]{{\displaystyle \frac{#1}{#2}} }
11: \newcommand{\alfven}{Alfv\'{e}n}
12: \newcommand{\eqref}[1]{(\ref{#1})}
13:
14: \def\lesssim{\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$<$}}}}
15: \def\gtrsim{\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$>$}}}}
16:
17: \slugcomment{Accepted for Publication in ApJ}
18:
19: \shorttitle{Orbit in a Protoplanetary Disk}
20: \shortauthors{Muto and Inutsuka}
21:
22:
23: \begin{document}
24:
25:
26: \title{Orbital Evolution of a Particle
27: Interacting with a Single Planet in a Protoplanetary Disk}
28:
29:
30: \author{Takayuki Muto\altaffilmark{1} and Shu-ichiro Inutsuka}
31: \affil{Department of Physics, Kyoto University, \\
32: Kitashirakawa-oiwake-cho, Sakyo-ku, Kyoto, 606-8502, Japan}
33: \email{muto@tap.scphys.kyoto-u.ac.jp}
34:
35: \altaffiltext{1}{JSPS Research Fellow}
36:
37:
38: \begin{abstract}
39: We investigate the motion of a particle around a low mass planet
40: embedded in a non-turbulent gaseous disk. We take into account the
41: effect of the gas
42: structure that is modified by the gravitational
43: interaction between the planet. We derive an analytic formula that
44: describes the change of the semi-major axis of the particle due to the
45: encounter with the planet using
46: local approximation in distant encounter regime. Our final formula
47: includes the effects of steady, axisymmetric radial gas
48: flow, the global gas pressure gradient in the disk, planet gravity, and
49: the structure of the gas flow
50: modified by the planet's gravity. We compare
51: the analytic results with numerical calculations, and indicate
52: that our formula well describes the secular evolution of the dust
53: particles' semi-major axes well, especially
54: for small particles with large drag coefficient. We discuss the
55: conditions for dust gap opening around a low mass planet and radial
56: distribution of dust particles. Our formula may provide a
57: useful tool for calculating radial distribution of particles
58: in a disk around the planet.
59: \end{abstract}
60:
61: \keywords{planet and satellites: formation --- solar system: formation
62: --- celestial mechanics --- methods: analytical}
63:
64:
65:
66: \section{Introduction} \label{intro}
67:
68: The motion of small particles in a gas disk has been investigated in
69: detail since this is essential for planet formation theory.
70: In a number of work, an axisymmetric gas disk is assumed
71: and the effects of the structure of the gas disk on
72: particle orbit are investigated, neglecting the effects of planet's
73: gravitational force (Adachi et al. 1976, Weidenschilling 1977). Their
74: results show that particles with
75: stopping timescale of the order of Kepler timescale migrate towards the
76: central star rapidly, thereby imposing a serious barrier on the formation
77: of planetesimals.
78:
79: Although this poses a serious question on the formation of
80: planetesimals, there is another aspect of this fact in the formation of
81: gas giant core.
82: Gas giant core should reach the critical mass to accrete nearby
83: gas rapidly (Mizuno et al. 1978), and its formation timescale
84: seems to be very long (Pollack et al. 1993).
85: If the global pressure gradient exerted on the gas disk will
86: force the particles to migrate towards the central star, the particles
87: at outer disk may reach the orbit of an already formed core and may
88: accrete onto it. Kary et al. (1993) have shown that this process seems
89: to be possible but most of the particles may miss the planet since they
90: migrate very rapidly. Their discussion is based on the three-body
91: problem including the effect of axisymmetric gas disk.
92:
93: The assumption of axisymmetric gas disk is a simplification in order to
94: make the calculation tractable. The planet gravitationally interacts
95: with the gas disk and changes the structure of the disk around it,
96: producing the spiral density wave (see e.g., Goldreich and Tremaine
97: 1979). Only recently does appear the
98: investigation of particle motion around
99: the planet embedded in a gas disk, which fully takes into account
100: the structure around the planet (Paardekooper and Mellema 2004, 2006,
101: Paardekooper 2007, Fouchet et al. 2007, Lyra et al. 2008).
102: For instance, Paardekooper and Mellema (2004, 2006)
103: and Paardekooper (2007) numerically
104: calculated the particle motion around a high mass planet in the
105: presence of a gas disk and discussed the accretion of the particles onto
106: the planet and the formation of gap of particles of different size.
107: They concluded that minimum mass of the planet that opens up a
108: dust gap was $0.05M_J$, thereby indicating the possibility of detecting
109: a low mass planet by future observations by observing a dust gap.
110:
111: Many of such works involve numerical studies on particle motion.
112: In this paper, we present an analytic investigation of the motion of
113: particles around a low mass planet embedded in a gas disk, taking into
114: account of the structure around the planet.
115: This is the
116: full analytic solution of a particle motion around the planet in which
117: the structure of the gas disk is fully taken into account.
118: We consider a vertically
119: averaged disk and perform two-dimensional analysis for simplicity.
120: We use the results of
121: (quasi-)linear calculation of the gas structure around the planet and
122: calculate the distant encounter between a particle and a planet embedded
123: in a gaseous disk. We derive a formula which describes the secular
124: change of the semi-major axis of the particle caused by the encounter
125: with the planet using Hill's equations.
126: It is not possible to describe the resonant capture discussed
127: by Weidenschilling and Davis (1985) or Kary et al. (1993); however, we
128: take into account the structure of the gaseous disk that is modified by
129: the planet, which is not considered in most of the previous analytic
130: works. Our
131: paper is complementary with previous numerical studies in two respects.
132: Firstly, our analysis is focused on low mass planet, and secondly, our
133: treatment is analytic. It is challenging to follow numerically the
134: motion of dust particles embedded in a gas disk for long timescale
135: (e.g., $10^5$ Kepler time), solving gas and the motion of dust
136: simultaneously.
137: Therefore, analytic treatment may be necessary in order to understand
138: the secular motion of dust particles. Moreover, basic physical
139: processes in the problem become clear by analytic treatment.
140:
141: %%%%%%%Modified in Revision%%%%%%%%
142: The plan of this paper is as follows. In Section \ref{analytic}, we
143: show analytic treatment of the problem and derive the
144: formula for the evolution of semi-major axis, equation
145: \eqref{deltab_final}. This formula includes the effects of global
146: pressure gradient of gas disk, steady, axisymmetric mass accretion flow
147: of the gas, gravitational force from the planet, and
148: the gas velocity structure modified by the planet. We then compare our
149: results with numerical calculations in Sections \ref{numerical} and
150: \ref{global}, and the limitations of our analytic approach are
151: discussed.
152: In Section \ref{dustgap}, we discuss the conditions for dust gap
153: formation and long-term behavior of dust particles whose orbits are
154: close to an embedded planet. Section
155: \ref{summary} is for summary and future prospects.
156: %%%%%%%Modified in Revision end%%%%%%%%
157:
158: We expect that our results, equation \eqref{deltab_final}, can be used
159: as a tool for estimate radial motion of the particle
160: in a disk around the planet, since this is
161: written in an analytic form. It may be possible to calculate the radial
162: distribution of particles of various sizes using this formula and
163: derive, for example, the opacity of the disk. We recommend readers
164: who need only the final results of the particle motion just to refer
165: equation \eqref{deltab_final} directly and go to Section
166: \ref{discuss:dustdistribution} where we demonstrate the calculation of
167: radial dust distribution around a planet.
168:
169: %%%%%%%Added in Revision%%%%%%%
170: Before proceeding to the details of calculations, we note that there
171: are two relevant length scales in the problem at hand: pressure scale
172: height $H$ of the gas disk and Hill's radius of the planet, $r_{\rm H}$.
173: Pressure scale height is an important quantity for the
174: structure of gas modified by the gravitational force of the planet,
175: while Hill's radius is important for the orbit of particles.
176: Pressure scale height that is given by sound speed of the gas $c$
177: divided by Keplerian
178: angular velocity and Hill's radius of the planet are related by
179: \begin{equation}
180: \dfrac{r_{\rm H}}{H}
181: \sim \dfrac{r_{\rm p}}{H}
182: \left( \dfrac{M_{\rm p}}{M_{\rm c}} \right)^{1/3}
183: = \left( \dfrac{GM_{\rm p}}{Hc^2} \right)^{1/3},
184: \end{equation}
185: where $r_{\rm p}$ is the semi-major axis of the planet, $M_{\rm p}$ is
186: the mass of the planet, and $M_{\rm c}$ is the mass of the central star.
187: For low mass planets
188: considered in this paper, $r_{\rm H}/H$ is smaller than unity.
189: %%%%%%%Added in Revision end%%%%%%%
190:
191:
192: \section{Analytic Consideration} \label{analytic}
193:
194: In this section, we analytically investigate the particle motion under
195: the influence of gravitational force by the protoplanet and gas drag
196: using local approximation. We show that particles with friction time of
197: the order Kepler time will migrate towards the planet because of the gas
198: drag, although gravitational scattering tends to repel the particle from
199: the planet. This mechanism is different from
200: the inward migration of dust particles caused by the velocity difference
201: between the gas and dust [see e.g., Adachi et al. (1976) or
202: Weidenschilling (1977)]. We also derive the effects of the modification
203: of gas velocity due to the gravitational force of the planet.
204:
205: \subsection{Basic Equations}
206: \label{subsec:analytic_equation}
207:
208: We use Hill's approximation to investigate the encounter between the
209: particle and the planet. We consider the motion only in $(x,y)$ plane
210: where $x$ is in radial direction and $y$ is in the azimuthal direction.
211: We assume the planet is fixed at the origin $(x,y)=(0,0)$ in this
212: coordinate and neglect the mass of the particle. We write the velocity
213: of the gas $\vect{v}_{\rm g}$ by
214: \begin{equation}
215: \vect{v}_{\rm g} = -\dfrac{3}{2} \Omega_{\rm p} x \vect{e}_{y}
216: + \delta \vect{v}_{\rm g},
217: \end{equation}
218: where $\Omega_{\rm p}$ is the Keplerian angular velocity of the planet
219: and $\vect{e}_y$ is the unit vector in the $y$-direction.
220: The first term
221: represents the Keplerian motion around the central star and the second
222: term $\delta \vect{v}_{\rm g}$
223: represents the deviation from this motion, which,
224: for instance, may be caused by the global pressure gradient of the
225: disk.
226:
227: The equations of motion of the particle are
228: \begin{equation}
229: \ddot{x} - 2\Omega_{\rm p} \dot{y} = 3 \Omega_{\rm p}^2 x - \nu \dot{x}
230: + F_x
231: \label{EoM_x}
232: \end{equation}
233: \begin{equation}
234: \ddot{y} + 2\Omega_{\rm p} \dot{x} = - \nu \left( \dot{y} + \dfrac{3}{2}
235: \Omega_{\rm p} x \right) + F_y
236: \label{EoM_y}
237: \end{equation}
238: where $\nu$ stands for the drag coefficient.
239: %%%%%%Added in revision%%%%%%%%
240: The quantity $\nu$ is a reciprocal of the stopping time of a particle.
241: Note that
242: there are various notations for this parameter. In Weidenschilling
243: (1977), stopping time is denoted by $t_e$ so $\nu$ in this paper is
244: equal to $1/t_e$. In Adachi et al. (1976),
245: stopping time corresponds to their $A$ times gas density times relative
246: velocity.
247: %%%%%%Added in revision end%%%%%%%%
248: We write the sum of
249: the force exerted by the planet and the part of the friction force
250: caused by the non-Keplerian motion of the gas by $\vect{F}$,
251: \begin{equation}
252: \vect{F}(x,y) = \nabla \frac{GM_{\rm p}}{\sqrt{x^2 + y^2}}
253: + \nu \delta \vect{v}_{\rm g} (x,y) ,
254: \label{force_general}
255: \end{equation}
256: where $G$ is the gravitational constant and $M_{\rm p}$ denotes the mass
257: of the planet.
258:
259: We give a brief note on the relationship between the dust size and drag
260: coefficient. In this paper, we use the constant drag coefficient $\nu$
261: for mathematical convenience. The dust size $d$ and the drag
262: coefficient $\nu$ are related by $\nu \propto d^{-1}$ for Epstein law
263: and $\nu \propto d^{-2}$ for Stokes law. Simple relationship between
264: the dust size and drag coefficient may be given by
265: \begin{eqnarray}
266: \label{draglaw}
267: \dfrac{\nu}{\Omega_{\rm p}} =
268: \left\{
269: \begin{array}{cc}
270: 10^4 \left( \dfrac{\rho_0}{10^{-9} \mathrm{g/cm}^3} \right)
271: \left( \dfrac{d}{1\mathrm{cm}} \right)^{-1} &
272: \dfrac{d}{1\mathrm{cm}} < \dfrac{10^{-9} \mathrm{g/cm}^3}{\rho_0}
273: \\[15pt]
274: 10^4 \left( \dfrac{d}{1\mathrm{cm}} \right)^{-2} &
275: \dfrac{10^{-9} \mathrm{g/cm}^3}{\rho_0} < \dfrac{d}{1\mathrm{cm}}
276: \end{array}
277: \right.
278: \end{eqnarray}
279: where $\rho_0$ is the background density and $d$ is the size of the
280: particle. Figure \ref{fig:size_drag} shows the relationship between
281: dust size and drag coefficient given by this equation for
282: $\rho_0=10^{-8}\mathrm{gcm}^{-3}$, $10^{-10}\mathrm{gcm}^{-3}$, and
283: $10^{-12}\mathrm{gcm}^{-3}$.
284:
285: The assumption of constant drag coefficient is not true for large bodies
286: whose drag force is proportional to
287: the square of the velocity difference. However, we expect that the
288: results are qualitatively similar even for large bodies.
289:
290: We neglect the effects of gravitational potential caused by gas disk.
291: It is a good approximation for non-self-gravitating disk as shown in
292: Section \ref{subsec:db_spiral}.
293:
294: In the absence of the planet and the friction force, the particle
295: motion is the Keplerian motion around the central star. In our setup,
296: this is represented by four orbital elements $(b,h,k,y_0)$,
297: \begin{equation}
298: x(t) = b + r_{\rm p} h \cos (\Omega_{\rm p} t) + r_{\rm p} k \sin
299: (\Omega_{\rm p} t)
300: \label{orbit_x}
301: \end{equation}
302: \begin{equation}
303: y(t) = y_0 - \dfrac{3}{2} b \Omega_{\rm p} t - 2 r_{\rm p} h \sin
304: (\Omega_{\rm p} t) + 2 r_{\rm p} k \cos (\Omega_{\rm p} t) .
305: \label{orbit_y}
306: \end{equation}
307: Physically, $b$ is the difference of orbital semi-major axis between the
308: planet and the particle, $h$ and $k$ are quantities related to the
309: orbital eccentricity, and $y_0$ is related to the initial azimuthal
310: position of the particle.
311:
312: We calculate the evolution of the orbital
313: elements caused by the planet's gravitational force and the friction due
314: to the gas. The time evolution of the osculating elements is given by
315: \begin{equation}
316: \dot{b} = \nu r_{\rm p} \left[ h \cos (\Omega_{\rm p} t) + k \sin
317: (\Omega_{\rm p} t) \right] +
318: \dfrac{2}{\Omega_{\rm p}} F_y
319: \label{eq_b}
320: \end{equation}
321: \begin{equation}
322: \dot{h} = -\nu h
323: - \dfrac{1}{r_{\rm p} \Omega_{\rm p}}
324: \left[ F_x \sin (\Omega_{\rm p} t) + 2 F_y \cos (\Omega_{\rm p} t) \right]
325: \label{eq_h}
326: \end{equation}
327: \begin{equation}
328: \dot{k} = -\nu k
329: + \dfrac{1}{r_{\rm p} \Omega_{\rm p}}
330: \left[ F_x \cos (\Omega_{\rm p} t) - 2 F_y \sin (\Omega_{\rm p} t)
331: \right] .
332: \label{eq_k}
333: \end{equation}
334:
335: \subsection{Change of Semi-Major Axis by Distant Encounter: General
336: Treatment}
337: \label{subsec:db_formulation}
338:
339: We now solve equations \eqref{eq_b}-\eqref{eq_k}. We assume that the
340: particle is in circular orbit initially:
341: $h(t=-\infty)=k(t=-\infty)=0$. We derive the change of orbital
342: semi-major axis $\Delta b=b(t=\infty)-b(t=-\infty)$ by the encounter.
343:
344: The formal solutions
345: of equations \eqref{eq_h} and \eqref{eq_k} are given by
346: \begin{equation}
347: h(t) = - \dfrac{1}{r_{\rm p} \Omega_{\rm p}} \int_{0}^{\infty} du
348: e^{-\nu u} \left\{ F_x(t-u) \sin [ \Omega_{\rm p}(t-u) ] + 2F_y(t-u)
349: \cos [ \Omega_{\rm p} (t-u) ] \right\}
350: \label{sol_h}
351: \end{equation}
352: \begin{equation}
353: k(t) = \dfrac{1}{r_{\rm p} \Omega_{\rm p}} \int_{0}^{\infty} du
354: e^{-\nu u} \left\{ F_x(t-u) \cos [ \Omega_{\rm p}(t-u) ] - 2F_y(t-u)
355: \sin [ \Omega_{\rm p} (t-u) ] \right\} ,
356: \label{sol_k}
357: \end{equation}
358: where $F_x (t)$ denotes $F_x (x(t),y(t))$ where $(x(t),y(t))$ is
359: the location of the particle at time $t$ given by equations
360: \eqref{orbit_x} and \eqref{orbit_y} and so for $F_y$.
361: Substituting equations \eqref{sol_h} and \eqref{sol_k} into equation
362: \eqref{eq_b}, we find
363: \begin{eqnarray}
364: & \Delta b &\equiv \int_{-\infty}^{\infty} \dot{b}(t) dt \nonumber \\
365: && = \dfrac{\nu}{\Omega_{\rm p}}\int_{-\infty}^{\infty} dt
366: \int_{0}^{\infty} du
367: e^{-\nu u} \left\{ F_x(t-u) \sin(\Omega_{\rm p} u) - 2F_y(t-u) \cos
368: (\Omega_{\rm p} u) \right\} \nonumber \\
369: && \ \ \ + \dfrac{2}{\Omega_{\rm p}}
370: \int_{-\infty}^{\infty} dt F_y(t)
371: \end{eqnarray}
372: Changing the integration variable $t \to t-u$ in the first integral of
373: the second line and using the formula
374: \begin{equation}
375: \int_{0}^{\infty} du e^{-\nu u} \cos (\Omega_{\rm p} u) =
376: \dfrac{\nu}{\nu^2+\Omega_{\rm p}^2}
377: \end{equation}
378: and
379: \begin{equation}
380: \int_{0}^{\infty} du e^{-\nu u} \sin (\Omega_{\rm p} u) =
381: \dfrac{\Omega_{\rm p}}{\nu^2+\Omega_{\rm p}^2},
382: \end{equation}
383: we find
384: \begin{equation}
385: \Delta b = \dfrac{\nu}{\nu^2 + \Omega_{\rm p}^2}
386: \int_{-\infty}^{\infty} F_x(t)dt + 2 \dfrac{\Omega_{\rm p}}{\nu^2 +
387: \Omega_{\rm p}^2} \int_{-\infty}^{\infty} F_y(t) dt
388: \label{deltab_integral}
389: \end{equation}
390:
391: This equation formally describes the amount of change of semi-major
392: axis. We note that the integration with respect to $t$ must be
393: performed along the particle's trajectory.
394: The force $\vect{F}$ consists of two parts [see equation
395: \eqref{force_general}].
396: One is the gravitational force of the planet and the other is the
397: non-Keplerian motion of the gas velocity. We now see these
398: effects separately.
399:
400: \subsection{Planet Encounter}
401: \label{subsec:db_planet}
402:
403: First, we consider the change of the semi-major axis caused by the
404: planet. We assume that the trajectory of the particle is close to
405: circular orbit. We approximate $F_x$ by
406: \begin{equation}
407: F_x \sim - \mathrm{sgn}(b) \dfrac{GM_{\rm p}}{b^2}
408: \dfrac{1}{( 1+(9/4)\Omega_{\rm p}t^2 )^{3/2}} .
409: \end{equation}
410: With this approximation, we obtain
411: \begin{equation}
412: \int_{-\infty}^{\infty} F_x(t) dt = - \mathrm{sgn}(b) \dfrac{4GM_{\rm
413: p}}{3b^2}.
414: \label{fx_integral}
415: \end{equation}
416: For the integration of $F_y$, we can use the well-known result of
417: restricted three-body problem (see e.g., Goldreich \& Tremaine 1980,
418: H\'{e}non \& Petit 1986, or Hasegawa \& Nakazawa 1990). The result
419: is
420: \begin{equation}
421: \int_{-\infty}^{\infty} F_y(t) dt = \dfrac{64}{243} \dfrac{G^2 M_{\rm
422: p}^2}{b^5 \Omega_{\rm p}^3} \left[ K_1\left(\dfrac{2}{3}\right) + 2
423: K_0 \left(\dfrac{2}{3}\right) \right]^2,
424: \label{fy_integral}
425: \end{equation}
426: where $K_0$ and $K_1$ are modified Bessel function of zeroth and first
427: order. The derivation of this equation is outlined in Appendix
428: \ref{App:fy_int}.
429:
430: Substituting equations \eqref{fx_integral} and \eqref{fy_integral} into
431: equation \eqref{deltab_integral}, we obtain
432: \begin{equation}
433: \Delta b = -\mathrm{sgn}(b) 4 \dfrac{r_{\rm H}^3}{b^2} \dfrac{\nu
434: \Omega_{\rm p}}{\nu^2+\Omega_{\rm p}^2} + \alpha \dfrac{r_{\rm
435: H}^6}{b^5} \dfrac{\Omega_{\rm p}^2}{\nu^2 + \Omega_{\rm p}^2},
436: \label{deltab_planet}
437: \end{equation}
438: where $r_{\rm H}$ is the Hill's radius of the planet defined by
439: \begin{equation}
440: r_{\rm H} = \left( \dfrac{M_{\rm p}}{3 M_{\ast}} \right)^{1/3}
441: r_{\rm p}
442: \end{equation}
443: and $\alpha$ is a numerical factor
444: \begin{equation}
445: \alpha \equiv \dfrac{128}{27}
446: \left[ K_1\left(\dfrac{2}{3} \right) +
447: 2K_0 \left( \dfrac{2}{3} \right) \right]^2 = 30.094
448: \end{equation}
449:
450: The second term of equation \eqref{deltab_planet} represents the effect
451: of gravitational scattering. The particle and the planet tends to repel
452: each other by
453: mutual gravitational interaction, which is a well-known result of
454: restricted three-body problem. This effect remains when drag force
455: vanishes ($\nu \to 0$). When there is no drag force, the
456: gravitational interaction between the planet and the particle results in
457: the excitation of eccentricity and the difference in semi-major axes of
458: the two increases since Jacobi energy must be conserved in the absence
459: of any dissipative force.
460:
461: When friction force is large, gravitational
462: scattering is ineffective since the gas drag enforces the particle to
463: move with the fluid element. The first term of the equation
464: \eqref{deltab_planet} shows that the orbits of the
465: planet and the particle tend to attract each other and this is most
466: efficient when $\Omega_{\rm p} \sim \nu$. The intuitive
467: explanation of this effect is as follows. When the particle feels the
468: gravitational force of the planet, it is attracted towards the planet's
469: position at first. Then, when the gas drag is effective, the gas
470: enforces the particle to stay at the Kepler orbit at the location where
471: the particle is attracted, resulting in the attraction of the semi-major
472: axes of the particle and the planet. If the drag force is taken into
473: account, the semi-major axis difference between the planet and the
474: particle can shrink since Jacobi energy is not necessarily conserved.
475:
476: Since the attraction of the semi-major axes of the particle and
477: the planet in the presence of drag force represented by the first term
478: of equation \eqref{deltab_planet} is proportional
479: to $b^{-2}$ at large $b$,
480: this overwhelms the effect of scattering represented by the
481: second term, which is proportional to $b^{-5}$. Therefore, we conclude
482: that particles far away from the planet are attracted towards the planet
483: when the gas velocity equals Keplerian rotation velocity.
484:
485: In the vicinity of the planet, it is expected that gravitational
486: scattering is more effective since this effect increases as $b^{-5}$.
487: The value of $b$ where these two effects are comparable is given by
488: $\Delta b=0$, that is
489: \begin{equation}
490: |b| \sim 1.96 \left(\dfrac{\Omega_{\rm p}}{\nu} \right)^{1/3}
491: r_{\rm H}.
492: \label{b_eq_grav}
493: \end{equation}
494: This indicates that particles with $\nu>\Omega_{\rm p}$ may move
495: towards the planet even in the absence of the global
496: pressure gradient.
497:
498:
499: \subsection{Non-Keplerian Rotation of Gas Disk due to Pressure Gradient}
500: \label{subsec:db_pressure}
501:
502: In this subsection, we consider the change of the semi-major axis caused
503: by the effect of non-Keplerian rotation of the gas, which is due to the
504: presence of global pressure gradient. The result of this
505: section is already derived by previous studies such as Adachi et
506: al. (1976) or Weidenschilling (1977), but we briefly show the results in
507: this formulation for completeness.
508:
509: We parameterize the degree of non-Keplerian rotation by $\eta$ and write
510: the velocity difference $\delta \vect{v}_{\rm g}$ by
511: \begin{equation}
512: \delta \vect{v}_{\rm g} = \eta v_{\rm p} \vect{e}_{y} = \mathrm{const},
513: \label{parameter_nonkepler}
514: \end{equation}
515: where $v_{\rm p}$ is the rotation velocity of the planet around the
516: central star, $v_{\rm p} = r_{\rm p} \Omega_{\rm p}$. The parameter
517: $\eta$ is non-dimensional and its magnitude is the order of the square
518: of the disk aspect ratio. This is negative for the disk with negative
519: pressure gradient.
520:
521: Since $\delta \vect{v}_{\rm g}$ is constant, it is easy to integrate
522: \eqref{deltab_integral} to obtain
523: \begin{equation}
524: \Delta b = 2 \eta v_{\rm p} T \dfrac{\nu \Omega_{\rm p}}{\nu^2 +
525: \Omega_{\rm p}^2},
526: \label{deltab_nonkepler}
527: \end{equation}
528: where $T$ is the time taken for the particle to cross the box. As
529: is well-known, this effect is most efficient for the particle with
530: $\nu\sim\Omega_{\rm p}$.
531:
532:
533: \subsection{Steady Accretion Flow}
534:
535: In this subsection, we consider the change of the semi-major axis caused
536: by the steady accretion (or deccretion) flow. This may be modeled by
537: the constant axisymmetric radial velocity $\delta v_x$. We parameterize
538: the radial velocity by non-dimensional factor $\zeta$ as follows,
539: \begin{equation}
540: \delta \vect{v}_{\rm g} = \zeta v_{\rm p} \vect{e}_x = \mathrm{const}.
541: \label{parameter_accretion}
542: \end{equation}
543: The parameter $\zeta$ denotes the ratio between the radial flow velocity
544: and the Kepler velocity. If the dissipation of the gaseous nebula is
545: due to the gas accretion, the sign of $\zeta$ is negative and its
546: magnitude is the order of $10^{-6}$, which
547: is indicated from observation (see, e.g., Haisch et al. 2001).
548: Drag force due to the gas is given by $\nu \delta \vect{v}_{\rm g}$ and
549: therefore, using equation \eqref{deltab_integral}, we
550: find
551: \begin{equation}
552: \Delta b = \zeta v_{\rm p} T \dfrac{\nu^2}{\nu^2+\Omega_{\rm p}^2} .
553: \label{deltab_accretion}
554: \end{equation}
555: We note that the steady accretion flow effect is important for small
556: particles, which have large drag coefficient $\nu$, since such particles
557: move in the same way as the gas flow.
558:
559:
560: \subsection{Non-Keplerian Rotation of Gas due to the Presence of the
561: Planet}
562: \label{subsec:db_spiral}
563:
564: In this section, we consider the change of the orbital semi-major axis
565: of the particle caused by the deviation of the gas velocity from
566: Keplerian rotation velocity due to the presence of the planet.
567: In this section, $\delta \vect{v}_{\rm g}$ is caused by the
568: gravitational interaction between gas and the embedded planet, in
569: contrast to the previous sections. We show that only axisymmetric
570: structure of the disk causes the change of the semi-major axis of the
571: particle.
572:
573: In order to calculate the velocity perturbation caused by the planet, we
574: solve vertically averaged Euler equations using local shearing-sheet
575: approximation:
576: \begin{eqnarray}
577: \label{EoC_hydro}
578: & \dfrac{\partial \Sigma}{\partial t} + \nabla \cdot (\Sigma \,
579: \vect{v}) = 0,\\
580: \label{EoM_hydro}
581: & \dfrac{\partial \vect{v}}{\partial t} + (\vect{v} \cdot \nabla)
582: \vect{v} = 3\Omega_{\rm p}^2 x \vect{e}_x
583: - \dfrac{c^2}{\Sigma} \nabla \Sigma
584: - \nabla \psi_{\rm p} - 2 \Omega_{\rm p} (\vect{e}_z \times \vect{v}),
585: \end{eqnarray}
586: where $\Sigma$ denotes the surface density, $\vect{v}$ is the gas
587: velocity, $c$ is the sound speed, and $\Omega_{\rm p}$ is the Keplerian
588: angular velocity of the planet. We assume an isothermal disk, where $c$
589: is constant, for simplicity.
590: The gravitational potential of the
591: planet, $\psi_{\rm p}$ is given by
592: \begin{equation}
593: \psi_{\rm p} = -\dfrac{GM_{\rm p}}{\sqrt{x^2+y^2+\epsilon^2}},
594: \end{equation}
595: where $\epsilon$ is the softening parameter.
596: Taking the rotation of the equation of motion, we obtain the equation
597: for vorticity,
598: \begin{equation}
599: \left( \dfrac{\partial}{\partial t} + \vect{v}\cdot\nabla \right)
600: \left[ \dfrac{1}{\Sigma} \left\{ \dfrac{\partial v_y}{\partial x}
601: - \dfrac{\partial v_x}{\partial y}
602: + 2\Omega_{\rm p} \right\} \right] = 0 .
603: \label{EoV_hydro}
604: \end{equation}
605:
606: For small mass planets we are interested in this paper, it is possible
607: to calculate the velocity perturbation by linear analysis. However, as
608: shown later, it is necessary to calculate the flow up to the second
609: order in order to obtain the correct results for the particle motion.
610: We assume that background surface density $\Sigma_0$ is constant and
611: background gas is rotating with Kepler velocity
612: $\vect{v}_0=-(3/2)\Omega_{\rm p}x\vect{e}_y$. We neglect the effect of
613: the global pressure gradient or steady mass accretion since it is
614: calculated separately in the previous subsections.
615:
616: The first order perturbation of the flow is caused by the planet
617: gravity. We denote the first order perturbation of physical variables
618: by superscript $(1)$. The linearization of equations
619: \eqref{EoC_hydro} and \eqref{EoM_hydro} are
620: \begin{eqnarray}
621: \label{EoC_hydro_lin}
622: & \left(\dfrac{\partial}{\partial t} - \dfrac{3}{2}
623: \Omega_{\rm p} x \dfrac{\partial}{\partial y} \right) \dfrac{\delta
624: \Sigma^{(1)}}{\Sigma_0} + \dfrac{\partial}{\partial x} \delta v_x^{(1)}
625: + \dfrac{\partial}{\partial y} \delta v_y^{(1)} = 0 \\
626: \label{EoMx_hydro_lin}
627: & \left(\dfrac{\partial}{\partial t} - \dfrac{3}{2}
628: \Omega_{\rm p} x \dfrac{\partial}{\partial y} \right) \delta v_x^{(1)}
629: - 2 \Omega_{\rm p} \delta v_y^{(1)} = -c^2 \dfrac{\partial}{\partial x}
630: \dfrac{\delta \Sigma^{(1)}}{\Sigma_0}
631: - \dfrac{\partial}{\partial x} \psi_{\rm p} \\
632: \label{EoMy_hydro_lin}
633: & \left(\dfrac{\partial}{\partial t} - \dfrac{3}{2}
634: \Omega_{\rm p} x \dfrac{\partial}{\partial y} \right) \delta v_y^{(1)}
635: + \dfrac{1}{2} \Omega_{\rm p} \delta v_x^{(1)}
636: = -c^2 \dfrac{\partial}{\partial y} \dfrac{\delta
637: \Sigma^{(1)}}{\Sigma_0}
638: - \dfrac{\partial}{\partial y} \psi_{\rm p} .
639: \end{eqnarray}
640: Assuming that vorticity is not created by the formation of the planet,
641: we have
642: \begin{equation}
643: \dfrac{\partial}{\partial y} \delta v_x^{(1)} -
644: \dfrac{\partial}{\partial x}
645: \delta v_y^{(1)} + \dfrac{1}{2} \Omega_{\rm p} \dfrac{\delta
646: \Sigma^{(1)}}{\Sigma_0} =0 ,
647: \label{vorticity}
648: \end{equation}
649: see, e.g., Narayan et al. (1987).
650: Note that $\Sigma_0$ appears only in the form of
651: $\delta \Sigma^{(1)} / \Sigma_0$
652: and therefore, this value itself does not depend on the background
653: surface density.
654: The source term of these equations is given by planet potential,
655: $\psi_{\rm p}$ so the perturbation is proportional to the planet mass
656: $M_{\rm p}$. For homogeneous ($\psi_{\rm p}=0$) equations, only
657: dimensional parameters are $c$ and $\Omega_{\rm p}$. Taking
658: $\Omega_{\rm p}^{-1}$ as a unit of time and $H=c/\Omega_{\rm p}$ as a
659: unit of length, homogeneous equations become scale free.
660: Since the amplitude of the perturbation is proportional to the source
661: term, which is $\psi_{\rm p}$ in this set of equations, the
662: perturbation amplitude scales with $\psi_{\rm p}/c^2$, in other words,
663: \begin{equation}
664: \dfrac{\delta \Sigma^{(1)}}{\Sigma_0} \propto \dfrac{GM_{\rm p}}{H c^2}
665: \label{amp_order}
666: \end{equation}
667: as shown by Tanaka et al. (2002). Once we know the perturbation amplitude
668: of a specific value of $GM_{\rm p}/Hc^2$, we can easily obtain the actual
669: amplitude of density perturbation $\delta \Sigma$ for different
670: background density, protoplanet mass, sound speed and so on. Linear
671: calculation shows that the coefficient of proportionality of equation
672: \eqref{amp_order} is the order of unity.
673:
674: The first order perturbation becomes a source of second order
675: perturbations through non-linear terms of basic equations. We denote
676: the second order perturbation by superscript $(2)$. The set of second
677: order perturbation equations is
678: \begin{eqnarray}
679: \label{EoC_hydro_2nd}
680: & \left(\dfrac{\partial}{\partial t} - \dfrac{3}{2}
681: \Omega_{\rm p} x \dfrac{\partial}{\partial y} \right) \dfrac{\delta
682: \Sigma^{(2)}}{\Sigma_0} + \dfrac{\partial}{\partial x} \delta v_x^{(2)}
683: + \dfrac{\partial}{\partial y} \delta v_y^{(2)} \nonumber \\
684: & \ \ \ \ \ \ \ \ \ \
685: = - \left[ \dfrac{\partial}{\partial x} \left( \delta \Sigma^{(1)}
686: \delta v_x^{(1)} \right)
687: + \dfrac{\partial}{\partial y} \left( \delta \Sigma^{(1)} \delta
688: v_y^{(1)} \right)
689: \right] \\
690: \label{EoMx_hydro_2nd}
691: & \left(\dfrac{\partial}{\partial t} - \dfrac{3}{2}
692: \Omega_{\rm p} x \dfrac{\partial}{\partial y} \right) \delta v_x^{(2)}
693: - 2 \Omega_{\rm p} \delta v_y^{(2)} + c^2 \dfrac{\partial}{\partial x}
694: \dfrac{\delta \Sigma^{(2)}}{\Sigma_0} \nonumber \\
695: & \ \ \ \ \ \ \ \ \ \
696: = c^2 \dfrac{\delta \Sigma^{(1)}}{\Sigma_0} \dfrac{\partial}{\partial x}
697: \dfrac{\delta \Sigma^{(1)}}{\Sigma_0} - \delta v_x^{(1)}
698: \dfrac{\partial}{\partial x} \delta v_x^{(1)} - \delta v_y^{(1)}
699: \dfrac{\partial}{\partial y} \delta v_x^{(1)} \\
700: \label{EoMy_hydro_2nd}
701: & \left(\dfrac{\partial}{\partial t} - \dfrac{3}{2}
702: \Omega_{\rm p} x \dfrac{\partial}{\partial y} \right) \delta v_y^{(2)}
703: + \dfrac{1}{2} \Omega_{\rm p} \delta v_x^{(2)}
704: + c^2 \dfrac{\partial}{\partial y} \dfrac{\delta
705: \Sigma^{(2)}}{\Sigma_0} \nonumber \\
706: & \ \ \ \ \ \ \ \ \ \
707: = c^2 \dfrac{\delta \Sigma^{(1)}}{\Sigma_0} \dfrac{\partial}{\partial y}
708: \dfrac{\delta \Sigma^{(1)}}{\Sigma_0} - \delta v_x^{(1)}
709: \dfrac{\partial}{\partial x} \delta v_y^{(1)} - \delta v_y^{(1)}
710: \dfrac{\partial}{\partial y} \delta v_y^{(1)} .
711: \end{eqnarray}
712: The second-order vorticity equation is
713: \begin{equation}
714: \dfrac{\partial}{\partial y} \delta v_x^{(2)} -
715: \dfrac{\partial}{\partial x}
716: \delta v_y^{(2)} + \dfrac{1}{2} \Omega_{\rm p} \dfrac{\delta
717: \Sigma^{(2)}}{\Sigma_0} =
718: - \dfrac{\delta \Sigma^{(1)}}{\Sigma_0}
719: \left[ \dfrac{\partial}{\partial x} \delta v_y^{(1)} -
720: \dfrac{\partial}{\partial y} \delta v_x^{(1)}
721: \right] + \dfrac{1}{4} \left( \dfrac{\delta \Sigma^{(1)}}{\Sigma_0}
722: \right)^2 .
723: \label{vorticity_2nd}
724: \end{equation}
725: Since the first order perturbation scales with $GM_{\rm p}/Hc^2$ (see
726: equation \eqref{amp_order}), it is clear that the second order
727: perturbation scales with $(GM_{\rm p}/Hc^2)^2$.
728: We assume that the perturbation is stationary with respect to the planet
729: so in the corotating frame with the planet considered in this
730: subsection, $\partial/\partial t = 0$.
731:
732: %%%%%%%%%%%%%moved to later section in revision%%%%%%%%%
733: % In order to find the spiral
734: %pattern produced by the planet, we numerically integrate
735: %equations \eqref{EoC_hydro_lin}-\eqref{EoMy_hydro_lin} using the Fourier
736: %transform methods presented by Goodman and Rafikov (2001). This method
737: %is particularly useful in calculating the gravitational potential
738: %produced by the density fluctuation.
739: %We fix the spatial resolution
740: % (therefore, the maximum wave number in Fourier space)
741: %$(\Delta x/H, \Delta y/H)$ to be
742: %$(9.8 \times 10^{-3}, 7.8 \times 10^{-2} )$, which is enough to resolve
743: % the effective Lindblad resonance (see e.g., Artymowicz 1993). We fix
744: % the box size in $y$-direction such that $-80H<y<80H$ and impose a
745: % periodic boundary condition. For $x$-direction, the
746: % resolution in Fourier space in $k_x$ direction is varied from $10^4$ to
747: % $10^6$ according to $k_y$ in order
748: % to resolve very fast oscillation at $k_x \gg k_y$. We also use the
749: % linear window function given by Goodman and Rafikov (2001).
750: % We set softening parameter $\epsilon$ to be $0.01H$. This numerical
751: % data is used in integrating the particle orbits in section
752: % \ref{numerical}.
753: %%%%%%%%%%%%%moved to later section in revision end%%%%%%%%%
754:
755: We now see equations \eqref{EoC_hydro_lin}-\eqref{EoMy_hydro_lin}
756: in more detail to find an analytic expression of $\Delta b$. We first
757: give an order-of-magnitude discussion to show that the gravitational
758: force produced by the density fluctuation is small compared to the
759: gravitational force by the planet. We expect that the order of
760: magnitude of the gravitational force by the spiral density fluctuation
761: is
762: \begin{equation}
763: F_{\rm spiral} \sim \dfrac{G \delta \Sigma H^2}{H^2},
764: \end{equation}
765: since the characteristic length scale of the spiral is of the order
766: of the scale height. On the other hand, the gravitational force by the
767: planet exerted on the particle at the distance of the order of the scale
768: height is
769: \begin{equation}
770: F_{\rm planet} \sim \dfrac{GM_{\rm p}}{H^2}.
771: \end{equation}
772: Using equation \eqref{amp_order}, we expect that these two forces are
773: related by
774: \begin{equation}
775: \dfrac{F_{\rm spiral}}{F_{\rm planet}} \sim
776: \dfrac{G\Sigma_0}{c\Omega_{\rm p}} \equiv Q^{-1},
777: \end{equation}
778: where $Q$ is Toomre's $Q$ parameter of the disk. Since protoplanetary
779: disks in planet forming phase are
780: expected to be gravitationally stable in general, $Q \gg 1$.
781: Therefore, we conclude that the gravitational interaction between the
782: particle and the disk is negligible. We later confirm this
783: numerically. In this section, we consider only
784: the effect of the velocity fluctuation of the gas disk in the presence
785: of the planet.
786:
787: \subsubsection{First-Order Axisymmetric Mode}
788:
789: We first consider axisymmetric modes, or azimuthally averaged
790: quantity. For any perturbation quantities $\delta f$, we denote azimuthal
791: average by bars,
792: \begin{equation}
793: \overline{\delta f}(x) = \dfrac{1}{L_y} \int_{-L_y/2}^{L_y/2} \delta
794: f (x,y) dy .
795: \end{equation}
796: Using Euler equations \eqref{EoC_hydro_lin}-\eqref{EoMy_hydro_lin} and
797: vorticity equation \eqref{vorticity}, we can derive
798: \begin{eqnarray}
799: \label{dvx_asym}
800: & \overline{\delta v_x^{(1)}} = 0, \\
801: \label{dsigma_dvy}
802: & \dfrac{\overline{\delta \Sigma^{(1)}}}{\Sigma_0} =
803: \dfrac{2}{\Omega_{\rm p}}
804: \dfrac{d}{dx} \overline{\delta v_y^{(1)}}, \\
805: \label{dvy_asym_eqn}
806: & \dfrac{d^2}{dx^2} \overline{\delta v_y^{(1)}} - \dfrac{\Omega_{\rm
807: p}^2}{c^2} \overline{\delta v_y^{(1)}} = -\dfrac{\Omega_{\rm p}}{2c^2}
808: \dfrac{d\overline{\psi}}{dx}.
809: \end{eqnarray}
810: For $x\ll L_y$, azimuthal average of $\partial \psi/\partial x$ is given
811: by
812: \begin{equation}
813: \dfrac{d\overline{\psi}}{dx} \sim
814: \dfrac{1}{L_y}\int_{-\infty}^{\infty} \dfrac{GM_{\rm
815: p}x}{\left(x^2+y^2\right)^{3/2}} dy = \dfrac{2GM_{\rm p}}{L_y x},
816: \end{equation}
817: where the integration range is extended to infinity.
818: The appropriate boudary condition is such that the perturbation vanishes
819: for $|x|\to\infty$. Since homogeneous solution of equation
820: \eqref{dvy_asym_eqn} is $\exp[\pm x/H]$, Green's function of the
821: differential operater of equation \eqref{dvy_asym_eqn} is
822: \begin{equation}
823: G(x;x^{\prime}) = -\dfrac{H}{2}
824: \left[ e^{-(x-x^{\prime})/H} \theta(x-x^{\prime})
825: + e^{(x-x^{\prime})/H} \theta(x^{\prime}-x) \right] ,
826: \end{equation}
827: where $\theta(x)$ is step function.
828: Therefore, the solution
829: of equation \eqref{dvy_asym_eqn} is given by
830: \begin{equation}
831: \overline{\delta v_y^{(1)}} = \dfrac{H^2 \Omega_{\rm p}}{2L_y}
832: \dfrac{GM_{\rm p}}{Hc^2}
833: \left[ e^{-(x/H)} \mathrm{Ei}\left( \dfrac{x}{H} \right)
834: - e^{x/H} \mathrm{Ei}\left( -\dfrac{x}{H} \right) \right] .
835: \label{dvy_asym}
836: \end{equation}
837: Using equation \eqref{dsigma_dvy}, the surface density perturbation is
838: given by
839: \begin{equation}
840: \dfrac{\overline{\delta \Sigma^{(1)}}}{\Sigma_0} = - \dfrac{H}{L_y}
841: \dfrac{GM_{\rm p}}{Hc^2}
842: \left[ e^{-(x/H)} \mathrm{Ei}\left( \dfrac{x}{H} \right)
843: + e^{x/H} \mathrm{Ei}\left( -\dfrac{x}{H} \right) \right],
844: \label{dens_asym}
845: \end{equation}
846: where $\mathrm{Ei}(x)$ denotes exponential integral. Equations
847: \eqref{dvy_asym} and \eqref{dens_asym} are valid for locations
848: $x\ll L_y$. Figure \ref{fig:dvy_asym_profile} shows the profile of
849: $L_y\overline{\delta v_y^{(1)}}(x)/c$ normalized by $(GM_{\rm p}/Hc^2)$
850: and compares this with analytic expression \eqref{dvy_asym}. Difference
851: between numerical and analytic results is about $25 \%$ at $x\sim 5H$.
852: We have checked that this error significantly decreases when box size in
853: $x$-direction is smaller than that of $y$-axis.
854:
855:
856: \subsubsection{Second-Order Axisymmetric Mode and Path Line of the Flow}
857:
858: The second-order axisymmetric modes are obtained by equations
859: \eqref{EoMx_hydro_2nd}, \eqref{EoMy_hydro_2nd}, and
860: \eqref{vorticity_2nd}. We note that equation of continuity
861: \eqref{EoC_hydro_2nd} and the $y$-component of equation of motion
862: \eqref{EoMy_hydro_2nd} is not independent for axisymmetric modes.
863: We use only $\overline{\delta v_x^{(2)}}$ later. This is given by
864: \begin{equation}
865: \overline{\delta v_x^{(2)}} = - \dfrac{2}{\Omega_{\rm p}}
866: \overline{ \delta v_x^{(1)} \dfrac{\partial}{\partial x} \delta
867: v_y^{(1)}}.
868: \label{dvx_2nd}
869: \end{equation}
870:
871: We then consider the path line of the flow. This gives the motion of
872: the fluid element and therefore gives the motion of the particle with
873: $\nu\to\infty$. Since we are interested in the change of the orbital
874: semi-major axis of the particle, we consider the motion of the fluid
875: element in the $x$-direction. Path line is given by
876: \begin{eqnarray}
877: \label{pathline_x}
878: & \dfrac{dx}{dt} = \delta v_x (x(t),y(t)) \\
879: \label{pathline_y}
880: & \dfrac{dy}{dt} = v_c + \delta v_y(x(t),y(t))
881: \end{eqnarray}
882: where $v_c$ is the unperturbed velocity
883: \begin{equation}
884: v_c = -\dfrac{3}{2} \Omega_p x
885: \end{equation}
886: and $\delta v_x$ and $\delta v_y$ include both first and second order
887: perturbations; for instance,
888: $\delta v_x = \delta v_x^{(1)} + \delta v_x^{(2)}$.
889:
890: The path line of the unperturbed flow is given by
891: \begin{eqnarray}
892: & x_c(t) = b = \mathrm{const} \\
893: & y_c(t) = \dfrac{L_y}{2} -\dfrac{3}{2} b \Omega_{\rm p} t
894: \end{eqnarray}
895: where we assume that the fluid element is at $(b,L_y/2)$ at $t=0$. This
896: is simply a Keplerian circular motion. We
897: denote the time when the particle reaches the other end of the box in
898: the unperturbed flow by $t=T$, where
899: \begin{equation}
900: T = \dfrac{L_y}{(3/2) b \Omega_{\rm p}} .
901: \end{equation}
902:
903: We solve equations \eqref{pathline_x} and \eqref{pathline_y} using
904: perturbation methods and obtain the change of $x$-coordinate,
905: $\Delta x$, of the fluid element after it crosses the box
906: \begin{equation}
907: \Delta x \equiv x(T)-x(0) = \int_0^T dt \delta v_x(x(t), y(t)) .
908: \label{pathline_Dx_def}
909: \end{equation}
910: We divide the motion of the fluid element into the unperturbed motion
911: and perturbation
912: \begin{eqnarray}
913: & x(t) = x_c + \delta x(t) \\
914: & y(t) = y_c(t) + \delta y(t).
915: \end{eqnarray}
916: Expanding right hand side of equation
917: \eqref{pathline_Dx_def} upto the first order of $\delta x$ and
918: $\delta y$ and using equations \eqref{pathline_x} and
919: \eqref{pathline_y}, we obtain
920: \begin{eqnarray}
921: & \Delta x =& \int_0^T \delta v_x^{(2)} (x_c, y_c(t)) dt
922: + \int_0^T dt_1 \dfrac{\partial \delta v_x^{(1)}}{\partial x}
923: (x_c,y_c(t_1))
924: \int_0^{t_1} dt_2 \delta v_x^{(1)} (x_c,y_c(t)) \nonumber \\
925: && + \int_0^T dt_1 \dfrac{\partial \delta v_x^{(1)}}{\partial y}
926: (x_c,y_c(t_1))
927: \int_0^{t_1} dt_2 \delta v_y^{(1)} (x_c,y_c(t))
928: \label{Dx_calc}
929: \end{eqnarray}
930: where we have used $\overline{\delta v_x^{(1)}}=0$ and take the terms up
931: to the second order. Using the first order vorticity equation
932: \eqref{vorticity} and equation of continuity, it is possible to show
933: that the right hand side of equation \eqref{Dx_calc} vanishes, see
934: Appendix \ref{app:pathline}. Therefore, we conclude that fluid element
935: returns to the original radial position after crossing the spiral
936: density wave,
937: \begin{equation}
938: \Delta x = 0 .
939: \label{Dx_vanish}
940: \end{equation}
941: We note that second-order perturbation is essential to derive this
942: conclusion and therefore, it is necessary to consider the fluid motion
943: upto the second order in finding the correct motion of the particles in
944: a disk.
945:
946: The fact that the fluid element does not move in the radial direction
947: indicates that there is no radial mass flux up to second order
948: perturbation. It is possible to show
949: this directly by calculating $\Sigma v_x$,
950: \begin{eqnarray}
951: & \displaystyle{\int} dy \Sigma v_x
952: &= \int dy \Sigma_0 \delta v_x^{(2)} + \Sigma_0
953: \int dy \dfrac{\delta \Sigma^{(1)}}{\Sigma_0} \delta v_x^{(1)}
954: \nonumber \\
955: & &= -\dfrac{2}{\Omega_{\rm p}} \Sigma_0 \int dy \delta v_x^{(1)}
956: \dfrac{\partial}{\partial x} \delta v_y^{(1)}
957: + \Sigma_0 \int dy \dfrac{\delta \Sigma^{(1)}}{\Sigma_0} \delta
958: v_x^{(1)} \nonumber \\
959: & &= -\dfrac{1}{\Omega_{\rm p}}\Sigma_0 \int dy
960: \dfrac{\partial}{\partial y} \left( \delta v_x^{(1)} \right)^2
961: \nonumber \\
962: & &= 0,
963: \label{massflux_vanish}
964: \end{eqnarray}
965: where we have used equation \eqref{dvx_2nd} in the second line,
966: vorticity equation \eqref{vorticity} in the third line, and periodic
967: boundary condition in the last equality
968: %%%%%%%%footnote added in revision%%%%%%%%
969: \footnote{It is actually possible to show that, in the shearing-sheet
970: approximation, the radial mass flux
971: vanishes for all orders of perturbation in the same manner as presented
972: here.}.
973: %%%%%%%%footnote added in revision end%%%%%%%%
974: This is in contrast to the case of the sound wave propagating in a
975: static, homogeneous medium, which carries the linear momentum and
976: therefore mass flux is present (Landau and Lifshitz 1959). We
977: demonstrate this in Appendix \ref{app:soundwave}.
978:
979:
980: \subsubsection{$\Delta b$ Caused by Spiral Density Wave}
981:
982: We are now in the position to calculate $\Delta b$ due to the structure
983: of the gas disk modified by the planet's gravity. Using equation
984: \eqref{deltab_integral} and drag force is
985: given by $\nu \delta \vect{v}_{\rm g}$, we have
986: \begin{equation}
987: \Delta b = \dfrac{\nu^2}{\nu^2 + \Omega_{\rm p}^2}
988: \int_{-\infty}^{\infty} \delta v_x \left(x(t),y(t) \right) dt +
989: \dfrac{2\nu\Omega_{\rm p}}{\nu^2+\Omega_{\rm p}^2}
990: \int_{-\infty}^{\infty} \delta v_y \left( x(t),y(t) \right) dt.
991: \label{deltab_deltav_int}
992: \end{equation}
993: The integration must be performed along the path of the particle. The
994: first term of equation \eqref{deltab_deltav_int} dominates when
995: $\nu\to\infty$. In the perfect coupling limit, since the particle
996: traces the flow of the fluid element, this integral should vanish. We
997: expect that this integral always remains small and set the first term of
998: equation \eqref{deltab_deltav_int} to
999: be zero. We expect that this approximation is valid since the second
1000: term that is proportional to
1001: $\nu\Omega_{\rm p}/(\nu^2+\Omega_{\rm p}^2)$ dominates $\Delta b$ when
1002: $\nu$ is small. We later confirm this numerically.
1003:
1004: We then consider the second term of equation
1005: \eqref{deltab_deltav_int}. The leading order of this integration is
1006: obtained by approximating that the particle orbit is circular. Since
1007: $\overline{\delta v_y^{(1)}}\neq 0$, it is enough to consider this
1008: contribution. Therefore, we arrive at the following expression of the
1009: change of orbital semi-major axis caused by the spiral structure of the
1010: gas
1011: \begin{equation}
1012: \Delta b = \mathrm{sgn}(b) \dfrac{4}{3} \dfrac{1}{b\Omega_{\rm p}}
1013: \dfrac{\nu \Omega_{\rm p}}{\nu^2+\Omega_{\rm p}^2}
1014: L_y \overline{\delta v_y^{(1)}},
1015: \label{deltab_spiral}
1016: \end{equation}
1017: where $\overline{\delta v_y^{(1)}}$ is given by equation \eqref{dvy_asym}.
1018:
1019: We note that in deriving equation \eqref{deltab_spiral}, we have
1020: repeatedly used the assumption of no vorticity source, equation
1021: \eqref{vorticity}. In the presence of vorticity source,
1022: $\overline{\delta v_y^{(1)}}$ may be different and the mass flux may be
1023: present.
1024:
1025:
1026:
1027: \subsection{Analytic Expression of the Change of the Orbital Semi-Major
1028: Axis}
1029: \label{subsec:db_anal}
1030:
1031: Adding equations \eqref{deltab_planet}, \eqref{deltab_nonkepler},
1032: \eqref{deltab_accretion}, and \eqref{deltab_spiral}, we
1033: obtain the expression for the secular evolution of the orbital
1034: semi-major axis of the particle. We write in the
1035: form of the rate of the change by dividing $\Delta b$ by
1036: $T=L_y/(3/2)\Omega_{\rm p} |b|$, where $L_y$ is the box size of the
1037: $y$-direction of the coordinate system considered.
1038: Time $T$ may be interpreted as the time interval between successive
1039: conjunctions between the particle and the planet.
1040: The result is
1041: \begin{eqnarray}
1042: & \dfrac{\Delta b}{T} =& 2\eta v_{\rm p} \dfrac{\nu \Omega_{\rm p}}{\nu^2
1043: + \Omega_{\rm p}^2}
1044: + \zeta v_{\rm p} \dfrac{\nu^2}{\nu^2+\Omega_{\rm p}^2} \nonumber \\
1045: && - \mathrm{sgn}(b) \dfrac{4}{T} \dfrac{r_{\rm H}^3}{b^2} \dfrac{\nu
1046: \Omega_{\rm p}}{\nu^2 + \Omega_{\rm p}^2}
1047: + \dfrac{\alpha}{T} \dfrac{r_{\rm H}^6}{b^5}
1048: \dfrac{\Omega_{\rm p}^2}{\nu^2 + \Omega_{\rm p}^2} \nonumber \\
1049: && + \mathrm{sgn}(b) \dfrac{2}{T} \dfrac{r_{\rm H}^3}{bH}
1050: \left[ e^{-(b/H)} \mathrm{Ei}\left( \dfrac{b}{H} \right)
1051: - e^{b/H} \mathrm{Ei}\left( -\dfrac{b}{H} \right) \right]
1052: \dfrac{\nu \Omega_{\rm p}}{\nu^2+\Omega_{\rm p}^2},
1053: \label{deltab_final}
1054: \end{eqnarray}
1055: where $r_{\rm H}$ is Hill's radius of the planet, the definitions of
1056: $\eta$ and $\zeta$ are given by equations \eqref{parameter_nonkepler}
1057: and \eqref{parameter_accretion} respectively, and the value of $\alpha$
1058: is $30.094$.
1059:
1060: If the planet mass is sufficiently small, the leading contribution comes
1061: from the first term, which is due to the global pressure gradient
1062: exerted on the gas disk. Comparing the order of magnitude of the first
1063: and the third term of equation \eqref{deltab_final}, the third term
1064: becomes dominant for particles with
1065: \begin{equation}
1066: \dfrac{|b|}{H} \lesssim \left( \dfrac{r_{\rm H}}{H} \right)^{3/2},
1067: \end{equation}
1068: where we have assumed $L_y\sim r_{\rm p}$, $\eta\sim (H/r_{\rm p})^2$
1069: and all the numerical coefficients are neglected. For low mass planets
1070: which does not form a gap in the gas disk, Hill's radius is generally
1071: smaller than the scale height. Therefore, the motion of the particle is
1072: predominated by the effect of global pressure gradient. However, for
1073: large planets with mass of order Jupiter, the effect of the global
1074: pressure gradient is small compared to the gravitational interaction
1075: between the planet, and equation \eqref{deltab_final} indicates that if
1076: the profile of the gas disk is neglected, the particles may be attracted
1077: towards the planet. This seems to be consistent with what is found by
1078: Paardekooper (2007). He found that when axisymmetric gas disk profile
1079: is considered, plants with the mass of the order of Jupiter would
1080: accrete the
1081: particles with stopping time comparable to Kepler time, while accretion
1082: of such particles around planets with 0.1 times Jupiter mass would be
1083: inefficient (see figures 7 and 11 of Paardekooper 2007). When global
1084: pressure gradient of gas disk drives the
1085: particles to migrate towards the central star, a part of particles with
1086: $\nu\sim\Omega_{\rm p}$ may pass the orbit of the planet without
1087: accumulating onto it, since the particle migration rate
1088: is very fast and the perturbation of the orbital elements of the
1089: particles is not related to the planet location (Kary et al. 1993). If
1090: the third term of equation
1091: \eqref{deltab_final} dominates over the global pressure effect, on the
1092: other hand, particles will accumulate onto the planet since their
1093: orbital elements are strongly perturbed at the location of the planet.
1094:
1095: If the disk model with flat pressure profile is
1096: considered, the behavior of the particle around the planet is
1097: complicated. Figure \ref{fig:db_contour} shows the contour plot of equation
1098: \eqref{deltab_final} for the planet mass $GM_{\rm p}/Hc^2=10^{-2}$,
1099: flat pressure profile ($\eta=0$), and no mass accretion onto the central
1100: star ($\zeta=0$).
1101: For small particles ($\nu\to\infty$), there is no
1102: secular motion of
1103: particles since all the terms of equation \eqref{deltab_final} vanishes
1104: if $\zeta=0$. The motion of small particles is determined by the
1105: strength of accretion of the background flow.
1106: For large particles ($\nu\to 0$),
1107: on the other hand, the fourth term, which is the gravitational scattering
1108: by the planet, is dominant and the particles are always repeled away
1109: from the planet.
1110: For particles with intermediate size
1111: ($\nu\sim\Omega_{\rm p}$), all the terms must be considered.
1112: Particles located sufficiently far away from the planet are
1113: scattered away because of the contribution from the last term of
1114: equation \eqref{deltab_final}, which is due to the axisymmetric mode of
1115: the perturbed flow structure caused by the planet gravity. Particles at
1116: the intermediate distance
1117: ($b/H\sim 1$), on the other hand, are attracted towards the planet
1118: because of the third term. If the particles are close to the planet,
1119: they are again scattered away because of the contribution from
1120: gravitational scattering.
1121:
1122: The equilibrium distance between the
1123: third and fourth term is given by equation \eqref{b_eq_grav}, and this
1124: becomes close to the planet with increasing $\nu$, and it eventually
1125: becomes smaller than the Hill's radius, where we expect that the
1126: particle motion is largely dominated by the planet's gravity and our
1127: treatment of distant encounter breaks down. Therefore, we conclude that
1128: the particles with intermediate but still smaller drag coefficient, on
1129: one hand, will
1130: accumulate around the planet, but not fully accrete onto the planet.
1131: On the other hand, particles with larger drag coefficient will accrete
1132: onto the
1133: planet since gravitational scattering is ineffective in this case.
1134: Equating the equilibrium distance given by equation \eqref{b_eq_grav}
1135: to Hill's radius, the critical drag coefficient
1136: separating these two regimes is of order $10 \Omega_{\rm p}$.
1137:
1138: We note that since equation \eqref{deltab_final} is obtained under the
1139: assumption of distant encounter, it does not accurately describe the
1140: dynamics of
1141: particles located within the Hill's sphere of the planet.
1142: %%%%%%Added in revision%%%%%%%%%%%%%%%%%
1143: In Section \ref{numerical}, we perform a numerical calculation of Hill's
1144: equations and investigate the validity of assumptions in the
1145: derivation of equation \eqref{deltab_final}.
1146: %%%%%%Added in revision end%%%%%%%%%%%%%%%%%
1147: It needs to be checked whether particles will really
1148: accrete onto the planet when they approach towards the planet. For
1149: close encounter, the detailed structure of the gas around the planet
1150: must be considered. Inaba \& Ikoma (2003) have shown that the capture
1151: cross section of particles by the planet may be enhanced when the
1152: atmospheric structure is considered.
1153:
1154:
1155: \section{Numerical Calculation in Local Coordinate}
1156: \label{numerical}
1157:
1158: %\textbf{
1159: %Note: This section is completely rewritten in revision
1160: %}
1161:
1162: \subsection{Numerical Methods}
1163:
1164: To investigate the accuracy of the description by equation
1165: \eqref{deltab_final} for the evolution of
1166: orbital element of a particle whose semi-major axis is close to the
1167: planet embedded in a gaseous disk, we
1168: perform a series of numerical calculations and compare the results with
1169: equation \eqref{deltab_final}.
1170:
1171: In our calculation, we solve Hill's equations \eqref{EoM_x} and
1172: \eqref{EoM_y} using fifth-order Runge-Kutta method with variable time
1173: step. We are primarily interested in the effects of
1174: spiral density wave and the gravitational force by the planet centered
1175: at the coordinate system: we neglect the effect of the global pressure
1176: gradient and mass accretion, $\eta=\zeta=0$.
1177: %%%%%%%%Added in revision%%%%%%%%%%%%%%%%%%
1178: The velocity field of the gas is therefore
1179: \begin{equation}
1180: \delta \vect{v}_g = \delta \vect{v}^{(1)} + \delta \vect{v}^{(2)},
1181: \end{equation}
1182: where $\delta \vect{v}^{(1)}$ and $\delta \vect{v}^{(2)}$ are
1183: first and second order perturbations, respectively, described in section
1184: \ref{subsec:db_spiral}. We also
1185: note that for gravitational force exerted by the planet, we use
1186: \begin{equation}
1187: \nabla \dfrac{GM_{\rm p}}{\sqrt{x^2+y^2}}.
1188: \end{equation}
1189: We do not use the time-averaged force given by equations \eqref{app_fx}
1190: and \eqref{app_fy}.
1191: %%%%%%%%Added in revision end%%%%%%%%%%%%%%%%%%
1192:
1193: %%%%%%%%Added in revision (moved from previous section)%%%%
1194: The modification of gas velocity due to the gravitational force by the
1195: planet is obtained by second-order perturbation analysis.
1196: We numerically integrate
1197: equations \eqref{EoC_hydro_lin}-\eqref{EoMy_hydro_lin} for the first
1198: order solution and \eqref{EoC_hydro_2nd}-\eqref{EoMy_hydro_2nd} for
1199: the second order solutions.
1200: We make use of Fourier
1201: transform methods presented by Goodman and Rafikov (2001). This method
1202: is particularly useful in calculating the gravitational potential
1203: produced by the density fluctuation.
1204: We fix the spatial resolution
1205: (therefore, the maximum wave number in Fourier space)
1206: $(\Delta x/H, \Delta y/H)$ to be
1207: $(9.8 \times 10^{-3}, 7.8 \times 10^{-2} )$, which is enough to resolve
1208: the effective Lindblad resonance (see e.g., Artymowicz 1993). We fix
1209: the box size in the $y$-direction such that $-80H<y<80H$ and impose a
1210: periodic boundary condition. For the $x$-direction, the
1211: resolution in Fourier space in $k_x$ direction is varied from $10^4$ to
1212: $10^6$ according to $k_y$ in order
1213: to resolve very fast oscillation at $k_x \gg k_y$. We also use the
1214: linear window function given by Goodman and Rafikov (2001).
1215: We set softening parameter $\epsilon$ to be $0.01H$.
1216: %%%%%%%%Added in revision (moved from previous section end)%%%%
1217:
1218: The box size of our calculation in $y$-direction is
1219: $-80H<y<80H$, to be consistent with the calculation
1220: of the modification of the velocity field.
1221: In integrating Hill's equations, we impose
1222: a periodic boundary condition in the $y$-direction and integrate the
1223: equations of motion until $t=5000\Omega_{\rm p}^{-1}$.
1224: %%%%%%%%%Added in revision%%%%%%%%%%%%%%%%%
1225: By periodic boundary condition, we mean that when the particle has
1226: reached the boundary of the box in $y$-direction, $y=\pm 80H$, it is
1227: reintroduced into the other end of the box, $y=\mp 80H$, while the
1228: values of $x$, $v_x$, and $v_y$ are kept unchanged.
1229: In principle, Hill's equations describe the motion of the particle only
1230: in the vicinity of the planet, and hence there is no reason that the
1231: phase of the epicyclic motion is conserved when it is reintroduced.
1232: However, we
1233: expect that if the box size is the same as $2\pi r_{\rm p}$, which is
1234: the circumference of the disk at the orbit of the planet, this
1235: periodic boundary treatment may simulate qualitative properties of
1236: the orbit in the global calculations. The condition
1237: \begin{equation}
1238: L_y = 2\pi r_{\rm p}
1239: \end{equation}
1240: is realized, in the parameter of our calculations, if we consider a disk
1241: with the aspect ratio $H/r_{\rm p}=2\pi/160 \sim 0.04$.
1242: %%%%%%%%%Added in revision end%%%%%%%%%%%%%%%%%
1243:
1244: %%%%%%%%Added in revision%%%%%%%%%%%%%
1245: Once we know the data of $(x,\dot{x},y,\dot{y})$, we can calculate the
1246: osculating elements $(b,h,k)$ by, using equations \eqref{orbit_x} and
1247: \eqref{orbit_y},
1248: \begin{equation}
1249: b = 4x + \dfrac{2}{\Omega_{\rm p}} \dot{y},
1250: \label{b_instant}
1251: \end{equation}
1252: \begin{equation}
1253: h = - \dfrac{1}{r_{\rm p} \Omega_{\rm p}} \left[
1254: \dot{x}\sin (\Omega_{\rm p} t) + 2 \left(
1255: \dot{y}+\dfrac{3}{2}x\Omega_{\rm p} \cos (\Omega_{\rm p} t)
1256: \right)
1257: \right],
1258: \label{h_instant}
1259: \end{equation}
1260: \begin{equation}
1261: k = \dfrac{1}{r_{\rm p} \Omega_{\rm p}} \left[
1262: \dot{x}\cos (\Omega_{\rm p} t) - 2 \left(
1263: \dot{y}+\dfrac{3}{2}x\Omega_{\rm p} \sin (\Omega_{\rm p} t)
1264: \right)
1265: \right].
1266: \label{k_instant}
1267: \end{equation}
1268: In the following sections showing results of calculations, we refer to
1269: $(b,h,k)$ using these relations. These values are not averaged over the
1270: encounter, but are defined at any time of the orbit.
1271: %%%%%%%%Added in revision end%%%%%%%%%%%%%
1272:
1273: The initial condition
1274: of the calculation is that the particles are located at the edge of the
1275: box and moving in a circular
1276: motion. When a particle
1277: approaches to the planet and the mutual distance becomes smaller than
1278: half of the Hill's radius of the planet, we stop the calculation since
1279: we are primarily interested in the distant encounter. Such particles
1280: should be trapped by the planet's gravity and eventually accrete onto
1281: the planet. In order to obtain a smooth results in the runs with large
1282: drag coefficient, we make use of smoothing methods
1283: described in Appendix \ref{app:sphint} in calculating the perturbed gas
1284: velocity at the location of the particle from the data obtained by
1285: hydrodynamic equations.
1286: We fix the
1287: background density to be $10^{-10} \mathrm{g/cm}^3$, which is an order of
1288: magnitude smaller than the Minimum Mass Solar Nebula (Hayashi et
1289: al. 1985). Toomre's $Q$ parameter of the disk at $1 \mathrm{AU}$ with
1290: is
1291: \begin{equation}
1292: Q^{-1} \sim \dfrac{G\rho}{\Omega_{\rm p}^2} \sim 10^{-4}
1293: \end{equation}
1294: We assume the mass of the planet as
1295: $GM_{\rm p}/Hc^2=10^{-2}$ that corresponds to $0.2M_{\oplus}$ for
1296: $c=10^5\mathrm{cm/s}$ and $H=0.05\mathrm{AU}$.
1297: In the following sections, we show the results without the gravitational
1298: force by gas. We have checked that gas gravity does not affect the
1299: results in our parameter range. We also neglect the effect
1300: of global pressure gradient and steady mass accretion in our
1301: calculation, $\eta=\zeta=0$, in order to see the effect of spiral wave
1302: and planet's gravity.
1303:
1304:
1305:
1306: \subsection{Results}
1307: \label{subsec:numericalresults}
1308:
1309: \subsubsection{Properties of Orbital Evolution}
1310:
1311: %%%%%%%%%%Added in revision%%%%%%%%%%%%
1312: We first review the properties of orbital evolution of a particle with
1313: zero friction force, which is obtained by setting $\nu=0$. Figure
1314: \ref{fig:threebody_b1} shows the evolution of $b$ for a particle
1315: initially located at $b=H$. The values of $b$ are obtained according to
1316: equations \eqref{b_instant} when the particle is at the box boundary.
1317: Also indicated in Figure \ref{fig:threebody_b1} is the variation of
1318: $b$ for $0<t<250\Omega^{-1}$. The values of $b$ varies rapidly when the
1319: particle encounters with the planet, when $y\sim0$. We note that since
1320: the box size is $160H$, time taken for the particle to cross the box is
1321: $\sim 160/(3/2) \sim 100$ Kepler times.
1322:
1323: It is indicated that the semi-major axis difference $b$ first increases
1324: but then shows an oscillation with the period $\sim 1200$ Kepler times.
1325: This oscillation is caused by the excitation of eccentricity by the
1326: perturbation due to the planet. If the particle comes into the box with
1327: finite eccentricity, the distance between the particle and the planet at
1328: the closest approach is different depending on the phase of the
1329: epicyclic motion, and the semi-major axis of the particle can increase
1330: or decrease. The period and amplitude of this oscillation depend on the
1331: location of the particle. If the phase of the epicyclic motion among
1332: successive encounters between the
1333: planet and the particle matches, the period of the oscillation becomes
1334: longer and the amplitude becomes larger. The condition for strong
1335: amplification is given by
1336: \begin{equation}
1337: \Omega_{\rm p}T=2\pi l,
1338: \label{resonance_cond}
1339: \end{equation}
1340: where $l$ is an integer and $T$ is the time taken to cross the box in
1341: $y$-direction
1342: \begin{equation}
1343: T = \dfrac{L_y}{(3/2)b\Omega_{\rm p}}.
1344: \end{equation}
1345: Figure \ref{fig:threebody_res} compares the evolution of semi-major axis
1346: of the particle starting at $l=5$ position ($b=3.4H$) and at
1347: $b=3.8H$. It is clear that the particle at $l=5$ position experiences
1348: the long-term, strong oscillation of semi-major axis.
1349:
1350: Equation \eqref{resonance_cond} indicates that there is a resonance
1351: between
1352: the box crossing time and the epicyclic motion of the particle, and this
1353: appears as a result of our treatment of periodic boundary in Hill's
1354: coordinate.
1355: Thus, in general, we should not consider this resonance to be physical
1356: if we arbitrary choose the box size in the y-direction.
1357: However, if we take the box size $L_y$ as the circumference of
1358: the disk, $2\pi r_{\rm p}$, the crossing time $T$ corresponds
1359: to the period of synodic encounter. Therefore, it may be possible to
1360: interpret this resonance as $j/(j+1)$ resonances in global
1361: problems. In section \ref{global}, we explore the
1362: correspondence between the Hill's approximation and the full global
1363: problem.
1364: %%%%%%%%%%Added in revision end%%%%%%%%%%%%
1365:
1366: %%%%%%%%%Modified in revision%%%%%%%%%%%
1367: We now consider the problem with gas drag.
1368: Figure \ref{fig:bevol_nu} shows the evolution of $b$ for particles with
1369: various drag coefficients initially located at $b/H=1$ as a function of
1370: time. The values of $b$ when the particles reach at the edge of the box
1371: are plotted.
1372: Particles with drag coefficient larger than
1373: $\nu/\Omega_{\rm p}=10^{-2}$ shows a systematic change of semi-major
1374: axis while that with $\nu/\Omega_{\rm p}=10^{-4}$ shows a systematic
1375: change plus oscillation. In a simple restricted three-body problem,
1376: where drag coefficient $\nu$ equals zero, semi-major
1377: axis does not show
1378: a systematic change but only oscillation, as discussed in previous
1379: paragraphs (see also Figures \ref{fig:threebody_b1} and
1380: \ref{fig:threebody_res}). The gas drag
1381: causes a systematic change of Jacobi energy
1382: and thereby results in a systematic change of particle's angular
1383: momentum. The timescale during which the oscillation of semi-major
1384: axis damps is given by $\sim \nu^{-1}$.
1385: Since the oscillation of the semi-major axis comes from the excitation
1386: of eccentricity during the synodic encounter, this oscillation does not
1387: appear when the eccentricity is damped during successive encounters.
1388: If the eccentricity is damped during successive encounters, the
1389: assumption of initially circular orbit in deriving equation
1390: \eqref{deltab_final} is valid for every encounter. Therefore, although
1391: we have computed only one encounter, it is possible to use equation
1392: \eqref{deltab_final} to predict the
1393: long-term evolution of the orbital semi-major axis
1394: especially for small bodies with large drag coefficient.
1395: %%%%%%%%%Modified in revision end%%%%%%%%%%%
1396:
1397:
1398: \subsubsection{Comparison between Analytic and Numerical Calculations}
1399:
1400: %%%%%%%%%Added in revision%%%%%%%%%%%%
1401: We now show to what extent equation \eqref{deltab_final} describes
1402: the evolution of the orbit of a particle encountering with a planet
1403: embedded in a gaseous nebula. Since the particles show a systematic
1404: change of semi-major axis, we take the average radial velocity of the
1405: particle by calculating
1406: \begin{equation}
1407: \dfrac{b(t_{\rm fin}) - b(t_{\rm ini})}{t_{\rm calc}},
1408: \label{average_db_calc}
1409: \end{equation}
1410: where $t_{\rm ini}=0$ is the initial time,
1411: $t_{\rm fin}$ is the last time when the particle reaches
1412: the box boundary before the calculation is stopped at
1413: $t=5000\Omega^{-1}$, and $t_{\rm calc}=t_{\rm fin}-t_{\rm ini}$. The
1414: precise value of $t_{\rm fin}$ is obtained by numerical calculation, and
1415: it varies with the value of $b$.
1416: Semi-major axis of the particle at each time is calculated by equation
1417: \eqref{b_instant}.
1418: %%%%%%%%%Added in revision end%%%%%%%%%%%%
1419:
1420: Figures \ref{fig:bdecay_Hill_d10m}-\ref{fig:bdecay_Hill_d10cm} show
1421: the results of long-term evolution of semi-major axis of particles
1422: initially located at various position with
1423: $\nu/\Omega_{\rm p}=10^{-2}$, $1$, and $10^2$ respectively. Horizontal
1424: axes of Figures \ref{fig:bdecay_Hill_d10m}-\ref{fig:bdecay_Hill_d10cm}
1425: are the initial semi-major axis of the particle, $b(t_{\rm ini})$, and
1426: the vertical axes show the average rate of change of semi-major axes
1427: obtained by equation \eqref{average_db_calc}.
1428: We also
1429: plot our analytic formula given by \eqref{deltab_final}.
1430: It is clear
1431: that the analytic formula \eqref{deltab_final} describes the results of
1432: numerical calculations well, at least quantitatively, for all values
1433: of drag coefficient. Deviation from the analytic value for
1434: $\nu/\Omega_{\rm p}=1$ at $b/H \gtrsim 2$ comes from the deviation of
1435: the data of $\overline{\delta v_y^{(1)}}$ from the analytic value shown
1436: in Figure \ref{fig:dvy_asym_profile} while for small $b$, we expect that
1437: the numerical results deviate from the analytic calculation since
1438: eccentricity effect is not negligible.
1439: The match between numerical calculation and
1440: analytic formula is better for particles with large drag coefficient,
1441: since their eccentricities remain smaller. We think that the
1442: deviation from the analytic value for $\nu/\Omega_{\rm p}=10^2$ at
1443: $b/H\gtrsim 3$ is due to smoothing.
1444:
1445: One significant difference between analytic results and numerical
1446: calculation is the spike-like structures that appear in the calculations
1447: with $\nu/\Omega_{\rm p}=10^{-2}$.
1448: This structure is due to the resonance effect given by equation
1449: \eqref{resonance_cond}.
1450: We show in Figure
1451: \ref{fig:bdecay_Hill_d10m} the location of resonance with $l=4$, $5$,
1452: and $6$. It is clear that $\Delta b/T$ shows a spike at this location.
1453: We have not found such resonance structure for calculations with
1454: $\nu/\Omega_{\rm p} \gtrsim 1$. This is because eccentricity is
1455: sufficiently damped between successive encounters.
1456:
1457: %%%%%%%%Modified in revision%%%%%%%%%%
1458: Another interesting result is the motion of the particle with
1459: $\Omega_{\rm p}\sim \nu$.
1460: Figure \ref{fig:long_nu1}
1461: shows the evolution of the semi-major axes of the particles with
1462: $\nu/\Omega_{\rm p}=1$. It is clear that the particles migrate
1463: towards the planet and stagnate at $b\sim 0.35H$.
1464: Since we set $\eta=\zeta=0$ in our calculation, the orbital change of
1465: the particle is caused by the third, fourth, and fifth terms of equation
1466: \eqref{deltab_final}. The last term is negligible for small $b$
1467: compared to the third term.
1468: Equation \eqref{deltab_final} indicates that
1469: these particles accumulate near the planet's orbit where
1470: gravitational
1471: scattering represented by the fourth term of equation
1472: \eqref{deltab_final} balances with the attraction towards the planet
1473: represented by the third term, which is the result of combination of the
1474: planet's gravity and the gas drag, as
1475: discussed in Section \ref{subsec:db_planet}.
1476: The condition for the balance between these two effects is given by
1477: equation \eqref{b_eq_grav}. For chosen parameters in numerical
1478: calculations, $r_{\rm H}=0.15H$ so the equilibrium distance is given
1479: by $\sim 0.3 H$, which is close to the value obtained in numerical
1480: calculation $\sim 0.35H$.
1481: We consider that the small difference comes from the effect of
1482: eccentricity and close encounter.
1483: This effect seems analogous to ``the shepherding effect''.
1484: We note that this effect
1485: appears when the first term of equation
1486: \eqref{deltab_final}, $\Delta b$ by global
1487: pressure gradient, is small compared to the third term. This condition
1488: is expressed in terms of $\eta$:
1489: \begin{equation}
1490: |\eta| < \dfrac{r_{\rm p}}{b} \dfrac{r_{\rm H}^3}{r_{\rm p}^3}
1491: \sim \dfrac{r_{\rm p}}{b} \dfrac{M_{\rm p}}{M_{\ast}} .
1492: \end{equation}
1493: If $\eta$ is of the order of square of disk aspect ratio, as in
1494: standard parameters, this condition is rewritten as
1495: \begin{equation}
1496: \dfrac{b}{r_{\rm H}} \lesssim
1497: \left( \dfrac{M_{\rm p}}{M_{\ast}} \right)^{2/3}
1498: \left( \dfrac{r_{\rm p}}{H} \right)^2.
1499: \end{equation}
1500: For $M_{\rm p}/M_{\ast}=10^{-6}$ and $H/r_{\rm p}=0.05$, this gives
1501: $b/r_{\rm H}\lesssim 0.2$, which is too close to the planet and the
1502: assumption of distant encounter is violated. Kary et al. (1993) has
1503: shown that, in the presence of global pressure gradient, a particle with
1504: $\nu \sim \Omega_{\rm p}$ falls onto the
1505: central star so rapidly that it bypasses the planet without being
1506: trapped.
1507: %%%%%%%%Modified in revision end%%%%%%%%%%
1508:
1509: %%%%%%%%%section added in revision%%%%%%%%%
1510: \subsection{Limitation of Analytic Formula}
1511: \label{subsec:limitation_local}
1512:
1513: %\textbf{
1514: %Note: This section is added in revision
1515: %}
1516:
1517:
1518: Equation \eqref{deltab_final} is derived by assuming that the
1519: deflection angle of the particle by each encounter with the planet is
1520: small (distant encounter) and that the particle returns to a circular
1521: orbit before the next encounter.
1522:
1523: First, the assumption of initially circular orbit breaks down if we
1524: consider the particle with small friction. The oscillation of
1525: semi-major axis observed in figure \ref{fig:bevol_nu} for particles with
1526: $\nu=10^{-4}\Omega_{\rm p}$ indicates that the damping of eccentricity
1527: between successive encounters is insignificant for these particles and
1528: that the finite eccentricity effect should be taken into account.
1529: Quantitatively, the period of synodic encounter must be longer than the
1530: particle's stopping time. Therefore, our treatment is limited in the
1531: cases with
1532: \begin{equation}
1533: \dfrac{2\pi r_{\rm p}}{(3/2) b \Omega_{\rm p}} \gtrsim \dfrac{1}{\nu},
1534: \end{equation}
1535: where the left hand side is the synodic period and the right hand side
1536: is stopping time. Rearranging this condition, we have
1537: \begin{equation}
1538: \dfrac{b}{H} \lesssim \dfrac{4\pi}{3} \dfrac{r_{\rm p}}{H}
1539: \dfrac{\nu}{\Omega_{\rm p}}.
1540: \end{equation}
1541: Adopting standard parameters, we have
1542: \begin{equation}
1543: \dfrac{b}{H} \lesssim 80
1544: \left( \dfrac{0.05}{H/r_{\rm p}} \right)^{-1}
1545: \left( \dfrac{\nu/\Omega_{\rm p}}{1} \right) .
1546: \label{condition_e}
1547: \end{equation}
1548:
1549: We now consider the assumption of distant encounter. If the particle's
1550: semi-major axis is too close to the planet, the assumption of small
1551: deflection angle, or distant encounter, breaks down.
1552: In Figure \ref{fig:size_deflection}, we plot the amount of deflection
1553: of the particle's semi-major axis after the very first encounter with
1554: the planet. We quantify the amount of deflection by
1555: \begin{equation}
1556: \dfrac{|b(t_1) - b(t_2)|}{|b(t_1)|},
1557: \label{deflection_rate}
1558: \end{equation}
1559: where $t_1=t_{\rm ini}=0$ is the initial time, and
1560: $t_2 \sim L_y/(3/2)b\Omega_{\rm p}$ is the time when the particle
1561: reaches the box boundary after the first encounter. The precise value
1562: of $t_2$ is obtained by numerical calculations. Particles
1563: in black regions are trapped by the planet, that is, in our criterion,
1564: mutual distance between the planet and the particle becomes smaller than
1565: half of the Hill's radius, or in horseshoe orbit. Note the similarity
1566: of the shape of the contour of 1\% deflection with the contour shown in
1567: Figure \ref{fig:db_contour}.
1568:
1569: In three-body problem with zero drag force, it is known that the
1570: assumption of small encounter is violated for particles with
1571: $b\lesssim 2.5 r_{\rm H}$ if we consider the very first encounter. Our
1572: calculation is
1573: consistent with this for particles with small drag,
1574: $\nu \ll \Omega_{\rm p}$. However, for particles with large drag,
1575: $\nu\gg\Omega_{\rm p}$, it can be seen that this condition is relaxed
1576: and it is indicated that the assumption of small deflection angle can be
1577: used even down to $b \sim r_{\rm H}$. Figure \ref{fig:orbit_2hill}
1578: shows the orbits of particles initially located at $b=2r_{\rm H}$.
1579: Orbits of particles with $\nu/\Omega_{\rm p}=10^{-4}$ and $10^3$ are
1580: shown. Particles with very large drag coefficient can remain in the
1581: orbit close to the initial semi-major axis since gas drag prevents it
1582: from falling onto the planet.
1583:
1584: After multiple encounters, the condition of distant encounter is
1585: violated for particles with $b\lesssim2.5-2\sqrt{3}r_{\rm H}$ in
1586: three-body problem without gas drag. The contour of 1\% deflection in
1587: Figure \ref{fig:size_deflection} passes close to $2\sqrt{3}r_{\rm H}$ in
1588: the limit of small drag. Therefore, we infer that this line gives the
1589: limitation of our analytic treatment assuming small deflection. This
1590: assumption is better for particles in wider range of semi-major axes for
1591: larger drag.
1592:
1593: %%%%%%%%%section added in revision end%%%%%%%%%
1594:
1595:
1596: %%%%%section rearranged in revision%%%%%
1597: \section{Comparison between Local and Global Calculation}
1598: \label{global}
1599:
1600: %\textbf{
1601: %Note: This section is completely rewritten in revision
1602: %}
1603:
1604:
1605: %%%%%%%This section is completely rewritten in revision%%%%%%
1606: So far we have performed analytic and numerical calculations of the
1607: orbits of particles using local coordinate systems. It remains to be
1608: checked how our local treatment simulates the real orbit in global
1609: coordinate.
1610:
1611: We have performed a global calculation with the gas rotating at
1612: Keplerian rotation velocity where we have neglected
1613: the effect of the spiral pattern around the planet, global pressure
1614: gradient, and gas accretion towards the central star, $\eta=\zeta=0$.
1615: We solved the equations of
1616: motion in a frame rotating with the planet and integrate until
1617: the particle encounters with the planet for 100 times. We calculated
1618: the evolution of the semi-major
1619: axis of a particle at various initial semi-major axes and obtain
1620: $\Delta b/T$ over the calculation time.
1621: In the
1622: following, we show calculations in which the particle is located at the
1623: opposition point initially, but we have checked that the initial
1624: locations do not affect the results as long as we consider particles
1625: with $\nu/\Omega_{\rm p} \geq 10^{-2}$.
1626: Particles that reach the distance with the planet less than Hill's
1627: radius are removed. We set disk aspect ratio $H/r_{\rm p}=0.04$, which
1628: gives $2\pi r_{\rm p}=L_y$, where $L_y=160H$ is the box size we have
1629: used in calculations with Hill's equation. For planet mass, we assumed
1630: $M_{\rm p}/M_{\ast}=6\times 10^{-7}$, which corresponds to
1631: $GM_{\rm p}/Hc^2=10^{-2}$, the value we have adopted in calculations
1632: with Hill's equations. In order to compare the results of global
1633: calculations with Hill's equations, we also integrate Hill's equations
1634: \eqref{EoM_x} and \eqref{EoM_y} by setting $\delta \vect{v}_{\rm g}=0$.
1635:
1636: Figures \ref{fig:globalHill_d10m}-\ref{fig:globalHill_d10cm} compare the
1637: results of $\Delta b/T$ obtained by global and local calculations.
1638: Horizontal axes of Figures
1639: \ref{fig:globalHill_d10m}-\ref{fig:globalHill_d10cm} show the initial
1640: semi-major axes of the particle and the vertical axes show the average
1641: rate of change calculated by equation \eqref{average_db_calc}. In
1642: global calculations, $b(t_{\rm fin})$ is the semi-major axis difference
1643: between the particle and the particle calculated when
1644: the particle has reached the opposition point after 100 encounters.
1645:
1646: For all parameters, global and local calculations show a reasonably
1647: good agreement. Relative difference between global and local runs
1648: is most significant for particles with
1649: ${\nu}/{\Omega_{\rm p}}=1$. We infer that this is because in the case
1650: of ${\nu}/{\Omega_{\rm p}}=1$, the radial motion of the particle is the
1651: fastest, and the effect of curvature is most effective. In runs with
1652: ${\nu}/{\Omega}_{\rm p}=10^{2}$ and ${\nu}/{\Omega}_{\rm p}=10^{-2}$,
1653: radial motion of the particle is smaller by one or two orders of
1654: magnitude than calculations with ${\nu}/{\Omega}_{\rm p}=1$.
1655:
1656: In runs with $\nu/\Omega_{\rm p}=10^{-2}$, we
1657: observe a disagreement in resonance locations, since curvature effects
1658: are not taken into account. However, since the disagreement is
1659: less than 10\%, it may be still possible to say, for example, that $l=4$
1660: resonance in Hill's coordinate mimics $4:5$ resonance in global
1661: coordinate. Note that the relative strength of the resonance matches
1662: well with global runs.
1663:
1664: Global analysis of $j/(j+1)$ resonances including drag force is done by
1665: Greenberg (1978). It
1666: is shown that eccentricity and the angle between the longitude of
1667: conjunction and the longitude of pericenter of the particle converges to
1668: a fixed value with the timescale of $\nu^{-1}$, but depends on the
1669: particle's semi-major axis. The value of eccentricity is maximum at the
1670: resonance. We have observed similar behavior in local calculations.
1671: It is actually possible to do analytic calculation similar to Greenberg
1672: (1978) in Hill's system and derive orbital evolution including the
1673: effect of resonance. We do not show it here since it is necessary to
1674: do all the analyses presented in Section \ref{analytic} again including
1675: the effect of finite box size. We simply note here that this resonance
1676: effect does not appear in our formulation presented in Section
1677: \ref{analytic} since we have neglected the effect of finite length in
1678: the $y$-direction in deriving equation \eqref{deltab_final}.
1679:
1680: In runs with $\nu/\Omega_{\rm p}=1$, particles initially at
1681: $b\lesssim 1.4H$ shows the decrease of the average rate of semi-major
1682: axis decay. This is because the particle is trapped in the orbit in
1683: the vicinity of the planet as a result of shepherding effect discussed
1684: for calculations with Hill's
1685: equation in Section \ref{subsec:numericalresults}. Figure
1686: \ref{fig:long_nu1_global} shows the evolution of the
1687: semi-major axis of the particle. Osculating elements at the opposition
1688: points are plotted. Comparing with Figure \ref{fig:long_nu1}, it is
1689: clear that the shepherding effect discussed in local calculations also
1690: exists in global runs. We note that the amplitude of oscillation after
1691: particles are trapped is larger in global runs than local calculations,
1692: and the location of the shepherding orbit is slightly closer to the
1693: planet than the local calculations. Note in passing that in
1694: equation (4.11) of Adachi et al. (1976), it is indicated that for
1695: $\eta=0$, damping of eccentricity causes a decrease in semi-major axis.
1696: However, Adachi et al. (1976) do not take into account
1697: the effects of the planet's gravity. The shepherding presented here is
1698: entirely caused by the presence of the planet.
1699:
1700: By comparing global calculations with local runs, it is indicated that
1701: as long as the orbits close enough to the planet are considered, local
1702: calculations reproduce the results of global calculations well even if
1703: curvature effect is neglected. However, the effect of constant box size
1704: in the $y$-direction causes the small difference between the
1705: global and local runs. This is indicated by the mismatch between the
1706: locations of the resonances. It is also indicated that the local
1707: calculations reproduce the results of global calculations most
1708: efficiently for small particles,
1709: such as $\nu/\Omega_{\rm p} \lesssim 10^{-1} - 10^{-2}$.
1710: Local approach that is used in Section \ref{analytic} is valid when
1711: semi-major axis difference $b$ is much smaller than the scale of the
1712: semi-major axis of the planet:
1713: \begin{equation}
1714: \dfrac{b}{r_{\rm p}} \ll 1,
1715: \end{equation}
1716: which is equivalent to
1717: \begin{equation}
1718: \dfrac{b}{H} \ll \dfrac{r_{\rm p}}{H} .
1719: \label{condition_local}
1720: \end{equation}
1721: Note that in the usage of equation \eqref{deltab_final} there is another
1722: limitation of close encounter, which has already been addressed in
1723: Section \ref{subsec:limitation_local}.
1724: %%%%%%%This subsection is completely rewritten in revision%%%%%%
1725:
1726:
1727: %%%%%%%section rearranged in revision%%%%%%
1728: \section{Discussion: Dust Gap Opening}
1729: \label{dustgap}
1730:
1731: \subsection{Order-of-Magnitude Estimate of Dust Gap Opening Criterion}
1732:
1733: Paardekooper and Mellema (2004, 2006) showed that dust gap may be opened
1734: up even
1735: if the planet mass is too small to open up a gap in a gas disk using
1736: numerical calculation. We reconsider this result from the analytical
1737: point of view in this section.
1738:
1739: Since there is no long-term radial motion of fluid elements, it is
1740: essential in dust gap formation that gas and dust moves in a different
1741: velocity. Terms that are important in equation \eqref{deltab_final} are
1742: therefore those proportional to
1743: $\nu\Omega_{\rm p}/(\nu^2 + \Omega_{\rm p}^2)$ for small particles.
1744:
1745: We first consider the disk without global pressure gradient.
1746: In this case, the motion of small dust particles are mainly dominated by
1747: the third term of equation \eqref{deltab_final} when they are as close
1748: to the planet as $b\lesssim2H$ and by the last term when $b$ is larger.
1749: The third term is the attraction of particles by the planet's gravity
1750: and they migrate towards the planet. The last term on the other hand
1751: originates from the axisymmetric flow structure
1752: around the planet. This term causes dust particles migrate away from
1753: the planet. In either case, dust particles around the planet with
1754: $b \sim H$ forms a gap around the planet. Taking $b\sim H$, very rough
1755: order-of-magnitude estimate (neglecting numerical coefficients of order
1756: unity) gives the migration timescale of dust particles as
1757: \begin{equation}
1758: \tau_{\rm gap} \equiv b \dfrac{T}{\Delta b} \sim 10^{4-5} T_{\rm K}
1759: \left(\dfrac{10^{-5}}{M_{\rm p}/M_{\rm c}} \right)
1760: \left( \dfrac{H/r_{\rm p}}{0.05} \right)^2
1761: \left( \dfrac{(\nu^2+\Omega_{\rm p}^2)/\nu \Omega_{\rm p}}{10^2}
1762: \right)
1763: \end{equation}
1764: where $T_{\rm K}$ denotes the Kepler timescale and $M_{\rm c}$ is the
1765: mass of the central star. This is the timescale
1766: of dust particles changing its orbit by the order of scale height and
1767: gives the timescale of the formation of dust gap with width of the
1768: order of the scale height.
1769:
1770: If there is global pressure gradient, the first term of equation
1771: \eqref{deltab_final} may dominate over the
1772: terms considered above. Since global pressure gradient causes
1773: systematic inward or outward motion of particles, dust gap around the
1774: planet is not expected to form if this term dominates.
1775: The timescale of orbital change of dust
1776: particles by the order of scale height due to the first term of equation
1777: \eqref{deltab_final} is
1778: \begin{equation}
1779: \tau_{\eta} \sim 10^{3-4} T_{\rm K}
1780: \left( \dfrac{10^{-3}}{|\eta|} \right)
1781: \left( \dfrac{H/r_{\rm p}}{0.05} \right)
1782: \left( \dfrac{(\nu^2+\Omega_{\rm p}^2)/\nu \Omega_{\rm p}}{10^2}
1783: \right) .
1784: \end{equation}
1785: If the magnitude of $\eta$ is of order $(H/r_{\rm p})^2 \sim 10^{-3}$,
1786: which is a
1787: standard value, dust gap may not appear around the planet since
1788: $\tau_{\eta}<\tau_{\rm gap}$.
1789: However if $\eta$ is as small as $10^{-5}$, dust gap formation may be
1790: possible. The value of $\eta$ depends on disk model and can take values
1791: of wide range. From theoretical point of view, small value of $\eta$ is
1792: preferable in order for meter-size particles to survive in a Minimum
1793: Mass Solar Nebula. One mechanism of decreasing $\eta$ around ice line
1794: is suggested by Kretke and Lin (2007).
1795:
1796: Comparing the first and the third term of equation \eqref{deltab_final},
1797: we obtain the criterion of dust gap formation. Again, for $b\sim H$,
1798: and neglecting numerical coefficients of order unity, we have, for the
1799: gap formation criterion,
1800: \begin{equation}
1801: |\eta| \dfrac{M_{\rm c}}{M_{\rm p}} \dfrac{H}{r_{\rm p}} \lesssim 1 .
1802: \end{equation}
1803: We note that this criterion does not depend on drag coefficient of the
1804: particles. However, the implicit assumption is that gravitational
1805: scattering, which is the fourth term of equation \eqref{deltab_final},
1806: is ineffective, which is proportianl to
1807: $\Omega_{\rm p}^2/(\nu^2+\Omega_{\rm p}^2)$. Therefore, this condition
1808: applies to particles with $\nu\gtrsim \Omega_{\rm p}$.
1809: Taking $\eta \sim (H/r_{\rm p})^2$ and $H/r_{\rm p} \sim 0.05$, this
1810: gives $M_{\rm p}/M_{\rm c} \gtrsim 10^{-4}$, which explains the
1811: condition obtained by Paardekooper and Mellema (2004, 2006).
1812:
1813: Timescale constraint of dust gap formation may be obtained by comparing
1814: the above obtained timescale with the timescale of planet migration or
1815: gas dispersal timescale. Timescale of the type I planet migration of
1816: the radial distance comparable to the scale height is given by
1817: Tanaka et al. (2002) as
1818: \begin{equation}
1819: \tau_{\rm mig I} \sim \Omega_{\rm p} \dfrac{M_{\rm c}}{M_{\rm p}}
1820: \dfrac{M_{\rm c}}{\Sigma r_{\rm p}^2}
1821: \left( \dfrac{H}{r_{\rm p}} \right)^3 .
1822: \end{equation}
1823: Therefore, comparing this with $\tau_{\rm gap}$, we have
1824: \begin{equation}
1825: \dfrac{r_{\rm p}}{H}
1826: \dfrac{\Sigma r_{\rm p}^2}{M_{\rm c}}
1827: \dfrac{\nu^2+\Omega_{\rm p}^2}{\nu \Omega_{\rm p}} \lesssim 1 .
1828: \end{equation}
1829: Dust gap formation is efficient at late stage of planet formation when
1830: gas starts to dissipate and type I migration timescale becomes long.
1831: If $H/r_{\rm p}=0.05$ and $\Sigma r_{\rm p}^2/M_{\rm c}=10^{-3}$, this
1832: criterion gives $\nu/\Omega_{\rm p} \lesssim 50$.
1833: If there is no type I migration, on the other hand, this
1834: constraint is relaxed and dust gap formation constraint is given by the
1835: timescale of disk gas dispersal, $\tau_{\rm gas}$. If $\tau_{\rm gas}$
1836: is of the order of $10^6$ years, gap of particles with
1837: $\nu/\Omega_{\rm p} \lesssim 10^4$ may be opened up.
1838:
1839:
1840:
1841: \subsection{Model of Radial Distribution of Dust
1842: Particles}
1843: \label{discuss:dustdistribution}
1844:
1845: In order to investigate the qualitative behavior of dust particles
1846: around the planet in long time, we model the motion of particles in the
1847: disk by one-dimensional advection equation
1848: \begin{equation}
1849: \dfrac{\partial N(t,b)}{\partial t} + \dfrac{\partial}{\partial b}
1850: \left[ v_b (b) N(t,b) \right] = 0
1851: \label{dustadvection_eqn}
1852: \end{equation}
1853: where $v_b (b)$ is the radial velocity of the dust particles whose
1854: semi-major axis difference with the planet is $b$
1855: and $N(t,b)$ is the appropriately normalized number of the particles
1856: with semi-major axis difference $b$ at time $t$. We use equation
1857: \eqref{deltab_final} for $v_b$ and solve \eqref{dustadvection_eqn} for
1858: $10^6 \mathrm{yr}$ for particles with various drag coefficients. We
1859: have used second-order scheme with monotonicity
1860: condition (van Leer 1977). We fix $H$ and $c$ for simplicity in our
1861: computational domain and a planet with $GM_{\rm p}/Hc^2=10^{-1}$ is
1862: fixed at the origin.
1863: Planet mass corresponds to $2M_{\oplus}$ for
1864: $H=0.05\mathrm{AU}$ and $c=10^5\mathrm{cm/s}$.
1865: The computational domain is $r_{H}<b<10H$
1866: and dust particles are homogeneously distributed initially:
1867: $N_0\equiv N(t=0,x)=1$ in this region. We have assumed that there is no
1868: dust particle out of this domain. For $T$ that appears in equation
1869: \eqref{deltab_final}, we have assumed
1870: \begin{equation}
1871: T = 2\pi / |\Omega_K(r_p+b) - \Omega_K(r_p)|
1872: \end{equation}
1873: where $\Omega_K$ is Keplerian angular velocity and $r_p$ is the location
1874: of the planet.
1875:
1876: Figures \ref{fig:advection} shows the snapshots of the distribution of
1877: dust particles of various drag coefficients at
1878: $t\Omega_{\rm p}=5\times10^{3}$, $t\Omega_{\rm p}=10^{5}$, and
1879: $t\Omega_{\rm p}=10^{6}$ for $\eta=0$.
1880: The gap of particles with $\nu \sim 10 \Omega_{\rm p}$
1881: has been opened up after several $10^5$
1882: Kepler time. The gap of particles $\nu\lesssim \Omega_{\rm p}$
1883: may be the artefact
1884: of boundary condition, since the radial velocity of particles at the
1885: nearest edge of the planet is positive but we have assumed there is no
1886: dust particles outside the calculation domain. However, there is no
1887: boundary effect for small particles with large drag coefficient, since
1888: radial velocity of dust particle is negative at the inner edge and
1889: positive at the outer edge. We note that since it takes about $10^5$
1890: years for dust gap formation for the planet
1891: with $2M_{\oplus}$, this effect is small
1892: if the gas pressure gradient is significantly large, $\eta>0$.
1893: However,
1894: for larger mass planets, planet perturbation on the dust motion is large
1895: and dust gap formation is possible even in the presence of pressure
1896: gradient, as seen in the order-of-magnitude estimate provided in the
1897: previous section.
1898:
1899: Finally, we briefly discuss the observational implication of dust gap
1900: in the disk with zero pressure gradient.
1901: The width of the dust gap that forms around a low mass planet is
1902: several times the scale height.
1903: Therefore, it might be challenging to detect
1904: a low mass planet embedded in a gas disk at $1\mathrm{AU}$. However, if
1905: there is a planet at several tens of AU, the width of
1906: the gap can be of the order of several AU, which can be resolved in
1907: near future. The gap formation timescale
1908: in this case is of order $10^{5-6}$ years,
1909: which might be comparable to disk
1910: life time. Therefore, it may be possible to find a low mass planet by
1911: observing a dust gap in a gaseous nebula, even if the planet itself is
1912: too small to be observable.
1913:
1914:
1915: \section{Summary and Future Prospects} \label{summary}
1916:
1917: In this paper, we have investigated the motion of particles embedded in
1918: a gas disk in the presence of a small mass planet. This is the
1919: full analytic calculation of the particle motion including the gas disk
1920: and the planet considering the non-axisymmetric pattern of the gas around
1921: the planet produced by the gravitational interaction between the planet
1922: and the gas.
1923:
1924: Our main result is equation \eqref{deltab_final} which
1925: describes the change of the particle semi-major axis, which has been in
1926: circular motion initially, after one distant encounter with the planet.
1927: In addition to well-known particle migration towards the central star
1928: due to the pressure gradient (the first term) and the gravitational
1929: scattering by the planet (fourth term),
1930: we have derived (1) the effect of steady mass accretion (the second term),
1931: (2) attraction of particles towards the planet due to the planet's
1932: gravity (the third term), and (3) contribution of the
1933: structure of the gas disk produced by the planet's gravity (the fifth
1934: term).
1935: We find that only the axisymmetric structure of the gas
1936: contributes to the semi-major axis evolution of particles and the spiral
1937: structure is ineffective in the absence of vorticity source.
1938:
1939: All the terms considered, in the absence of the global
1940: pressure gradient and steady accretion flow towards the central star, we
1941: find that (see Figure \ref{fig:db_contour})
1942: (1) particles with
1943: $\nu \gtrsim \Omega_{\rm p}$ will migrate towards the
1944: planet if they are close to the planet because of the gravitational
1945: attraction by the planet, but they migrate away from the planet
1946: if they are far away, primarily due to the effect of the gas
1947: structure,
1948: (2) large particles with
1949: $\nu \ll \Omega_{\rm p}$ will be scattered away from the planet because
1950: of the gravitational scattering, and
1951: (3) there is a parameter range where particles are scattered away from
1952: the planet when they are close to the planet while attracted towards the
1953: planet if they are located at far away, thereby particles are
1954: accumulated at an equilibrium distance from the planet (i.e.,
1955: shepherding by gas drag and gravitational scattering).
1956: %%%%%%%Added in revision%%%%%%%%
1957: This shepherding
1958: mechanism acts for particles in both interior and exterior orbits as
1959: long as the radial pressure gradient of the gas is negligible.
1960: %%%%%%%Added in revision%%%%%%%%
1961:
1962: %%%%%%%modified in revision%%%%%%
1963: We have checked the validity of our formula by solving three-body
1964: problem with and without Hill's approximations. Our main assumptions in
1965: deriving equation
1966: \eqref{deltab_final} are (1) initially circular orbit, (2) distant
1967: encounter, and (3) local approximation.
1968: Validity of the first assumption is justified when the eccentricity of
1969: the particle is damped within the time taken for each synodic encounter
1970: that is given by equation \eqref{condition_e}. The valid region for the
1971: second
1972: assumption is shown by the contour of small deflection in Figure
1973: \ref{fig:size_deflection}. The condition for the third assumption is
1974: simply given by equation \eqref{condition_local}.
1975: Hill's approximation is good
1976: as long as we consider the orbits in the vicinity of the planet.
1977: In short, our formula is especially useful in predicting the motions of
1978: dust particles with stopping time $t_e \Omega_{\rm p} \lesssim 10^{2}$.
1979: For particles with smaller drag, our treatment predicts only
1980: quantitative behavior of non-resonant orbital evolution of particles
1981: whose semi-major axis is close to that of the planet.
1982: We also note that since we have made use of linear calculation for the
1983: flow, our treatment is
1984: restricted to low mass planets with the mass up to the orders of several
1985: tens of Earth mass.
1986: %%%%%%%modified in revision end%%%%%
1987:
1988: Using equation \eqref{deltab_final}, we have discussed the criterion for
1989: dust gap formation. The condition depends on the value of global
1990: pressure gradient, mass of the planet, and disk scale height. It is
1991: possible to derive the condition suggested by Paardekooper and Mellema
1992: (2004, 2006), and we have derived more general criteria.
1993: We have calculated the qualitative
1994: long-term behavior of dust particles around a planet with
1995: $2M_{\oplus}$.
1996: It is indicated that the gap of small dust particles with width of the
1997: order of several scale height may be opened up
1998: after several $10^5$ Kepler time. This long timescale has not been
1999: reached by full numerical simulations yet.
2000:
2001: We have simplified the description of the drag force by assuming the
2002: force proportional to relative velocity with the gas. For large particles,
2003: however, the drag force is proportional to the square of the relative
2004: velocity, which is not treated in our calculation. We have also made a
2005: simplification by considering two-dimensional problem. In considering
2006: the vertical motion of particles, it is necessary to do
2007: three-dimensional calculations. The effects of magnetic field,
2008: self gravity, and turbulence of the disk will be the subjects in this
2009: line of analytical study.
2010: It is also of interest to extend our formulation to include the
2011: finite box size and investigate how the resonance effect is described in
2012: Hill's system.
2013:
2014: \acknowledgments
2015: The authors thank T. Takeuchi, E. Kokubo, S. Ida, and T. Tanigawa for
2016: helpful discussions.
2017: %%%%%%%Added in Revision%%%%%%%%%
2018: We also thank the referee of this paper, Dr. Stuart Weidenschilling,
2019: for useful comments that improved this paper.
2020: %%%%%%%Added in Revision end%%%%%%%%%
2021: This work was supported by the Grant-in-Aid for the Global COE Program
2022: ``The Next Generation of Physics, Spun from Universality and Emergence''
2023: from the Ministry of Education, Culture, Sports, Science and Technology
2024: (MEXT) of Japan.
2025: The numerical calculations were carried out on Altix3700 BX2 at YITP in
2026: Kyoto University.
2027: T. M. is supported by Grants-in-Aid for JSPS Fellows (19$\cdot$2409)
2028: from MEXT of Japan.
2029: S. I. is
2030: supported by Grants-in-Aid (15740118, 16077202, and 18540238) from MEXT
2031: of Japan.
2032:
2033:
2034:
2035: \appendix
2036:
2037: \section{Derivation of Equation \eqref{fy_integral}}
2038: \label{App:fy_int}
2039:
2040: In this section, we briefly show the derivation of \eqref{fy_integral}.
2041: The method is the same as that of Goldreich and Tremaine (1980),
2042: H\'{e}non and Petit (1986), or Hasegawa \& Nakazawa (1990).
2043:
2044: We consider the restricted three-body problem. In other words, we
2045: consider the problem with $\nu=0$. In this case,
2046: Jacobi energy $E_J$ conserves throughout the particle orbit. Jacobi
2047: energy is given by
2048: \begin{equation}
2049: E_J = \dfrac{1}{2} r_{\rm p}^2 \Omega_{\rm p}^2 (h^2 + k^2) -
2050: \dfrac{3}{8} \Omega_{\rm p}^2 b^2 + \psi_{\rm p},
2051: \end{equation}
2052: where $\psi_{\rm p}$ is the gravitational potential of the planet.
2053: Therefore, orbital elements before and after the encounter with the
2054: planet are related by
2055: \begin{equation}
2056: - \dfrac{3}{8} \Omega_{\rm p}^2 b^2 = \dfrac{1}{2} r_{\rm p}^2
2057: \Omega_{\rm p}^2 (h(\infty)^2 + k(\infty)^2) - \dfrac{3}{8}
2058: \Omega_{\rm p}^2 (b+\Delta b)^2.
2059: \end{equation}
2060: Assuming $\Delta b \ll b$, we obtain
2061: \begin{equation}
2062: \Delta b \sim \dfrac{2 r_{\rm p}^2}{3b}
2063: \left( h(\infty)^2 + k(\infty)^2 \right).
2064: \end{equation}
2065: On the other hand, from equation \eqref{eq_b}, $\Delta b$ is given by
2066: \begin{equation}
2067: \Delta b = \dfrac{2}{\Omega_{\rm p}} \int_{-\infty}^{\infty} F_y(t) dt.
2068: \label{db_fy}
2069: \end{equation}
2070: Therefore, we obtain
2071: \begin{equation}
2072: \int_{-\infty}^{\infty} F_y(t) dt = \dfrac{r_{\rm p}^2 \Omega_{\rm
2073: p}}{3b} ( h(\infty)^2 + k(\infty)^2).
2074: \label{fyderive}
2075: \end{equation}
2076:
2077: From equation \eqref{sol_h} and \eqref{sol_k},
2078: the elements $h$ and $k$ after the encounter are given by
2079: \begin{equation}
2080: h(\infty) = - \dfrac{1}{r_{\rm p} \Omega_{\rm p}} \int_{-\infty}^{\infty} du
2081: \left\{ F_x(u) \sin [ \Omega_{\rm p}u ] + 2F_y(u)
2082: \cos [ \Omega_{\rm p} u ] \right\}
2083: \label{app_h}
2084: \end{equation}
2085: \begin{equation}
2086: k(\infty) = \dfrac{1}{r_{\rm p} \Omega_{\rm p}} \int_{-\infty}^{\infty} du
2087: \left\{ F_x(u) \cos [ \Omega_{\rm p}u ] - 2F_y(u)
2088: \sin [ \Omega_{\rm p} u ] \right\} .
2089: \label{app_k}
2090: \end{equation}
2091: Approximating the trajectory of the particle by circular orbit, we have
2092: \begin{equation}
2093: F_x(t) \sim - \mathrm{sgn}(b) \dfrac{GM_{\rm p}}{b^2}
2094: \dfrac{1}{\left( 1+(9/4)(\Omega_{\rm p}t)^2 \right)^{3/2}}
2095: \label{app_fx}
2096: \end{equation}
2097: \begin{equation}
2098: F_y(t) \sim \mathrm{sgn}(b) \dfrac{GM_{\rm p}}{b^2}
2099: \dfrac{(3/2)\Omega_{\rm p}t}{\left( 1+(9/4)(\Omega_{\rm p}t)^2
2100: \right)^{3/2}}.
2101: \label{app_fy}
2102: \end{equation}
2103: Integrating \eqref{app_h} and \eqref{app_k}, we have
2104: \begin{equation}
2105: h(\infty) = 0
2106: \end{equation}
2107: \begin{equation}
2108: k(\infty) = - \mathrm{sgn}(b) \dfrac{8}{9} \dfrac{1}{r_{\rm
2109: p}\Omega_{\rm p}^2} \dfrac{GM_{\rm p}}{b^2}
2110: \left[K_1\left(\dfrac{2}{3}\right) + 2 K_0 \left( \dfrac{2}{3} \right)
2111: \right].
2112: \end{equation}
2113: Substituting these results into equation \eqref{fyderive}, we finally
2114: obtain
2115: \begin{equation}
2116: \int_{-\infty}^{\infty} F_y(t) dt = \dfrac{64}{243} \dfrac{G^2 M_{\rm
2117: p}^2}{b^5 \Omega_{\rm p}^3} \left[ K_1\left( \dfrac{2}{3} \right) + 2
2118: K_0 \left( \dfrac{2}{3} \right)\right]^2
2119: \end{equation}
2120: We note that direct substitution of equation \eqref{app_fy} into
2121: equation \eqref{db_fy} results in
2122: zero, indicating that higher order of the expansion is essential.
2123:
2124:
2125: \section{Proof of Equation \eqref{Dx_vanish}}
2126: \label{app:pathline}
2127:
2128: In this section, we show that the right hand side of equation
2129: \eqref{Dx_calc} vanishes up to second order perturbation and prove
2130: equation \eqref{Dx_vanish}. The equation we prove is
2131: \begin{eqnarray}
2132: & -\dfrac{2}{\Omega_{\rm p}} \displaystyle{\int_0^T} dt
2133: \delta v_x^{(1)} \dfrac{\partial}{\partial x} \delta v_y^{(1)}
2134: + \displaystyle{\int_0^T} dt_1 \dfrac{\partial
2135: \delta v_x^{(1)}}{\partial x} (x_c,y_c(t_1))
2136: \displaystyle{\int_0^{t_1}} dt_2 \delta v_x^{(1)} (x_c,y_c(t_2))
2137: \nonumber \\
2138: & + \displaystyle{\int_0^T} dt_1 \dfrac{\partial
2139: \delta v_x^{(1)}}{\partial y} (x_c,y_c(t_1))
2140: \displaystyle{\int_0^{t_1}} dt_2 \delta v_y^{(1)} (x_c,y_c(t_2)) = 0
2141: \label{Dx_calc_app}
2142: \end{eqnarray}
2143: We Fourier transform the perturbation in the $y$-direction,
2144: \begin{equation}
2145: \delta f (x,y) = \overline{\delta f} +
2146: \sum_{k_y \neq 0} \tilde{\delta f}_{k_y}(x) e^{ik_y y}
2147: \end{equation}
2148: where $\delta f$ denotes any perturbation quantity,
2149: $\overline{\delta f}$ is the average over $y$, and $\tilde{f}_{k_y}$ is
2150: the Fourier component of non-axisymmetric modes. Since physical
2151: quantities are real,
2152: $\tilde{\delta f}_{k_y} = \tilde{\delta f}_{-k_y}^{\ast}$.
2153: In this section, we
2154: drop superscript $(1)$ since all the perturbed values refer to first
2155: order values.
2156:
2157: Since we integrate over the circular orbit, integration with respect to
2158: $t$ is converted to that with respect to $y$ by
2159: \begin{equation}
2160: y = -\dfrac{3}{2} x_c \Omega_{\rm p} t.
2161: \end{equation}
2162: It is possible to show that the right hand side of equation
2163: \eqref{Dx_calc_app} is equal to
2164: \begin{eqnarray}
2165: & -\dfrac{8L_y}{3\Omega_{\rm p}x_c} \sum_{k_y>0}
2166: \mathrm{Re} \left[ \tilde{d\delta v_x}_{k_y} \dfrac{\tilde{\delta
2167: v_y}_{k_y}^{\ast}}{dx} \right]
2168: - \dfrac{8L_y}{9\Omega_{\rm p}^2 x_c^2} \sum_{k_y>0}
2169: \dfrac{1}{k_y} \mathrm{Im} \left[\dfrac{d \tilde{\delta
2170: v_x}_{k_y}^{\ast}}{dx} \tilde{\delta
2171: v_x}_{k_y} \right] \nonumber \\
2172: & + \dfrac{8L_y}{9 \Omega_{\rm p}^2 x_c^2} \sum_{k_y>0}
2173: \left[ \tilde{\delta v_x}_{k_y}^{\ast} \tilde{\delta v_y}_{k_y}
2174: \right]
2175: + \dfrac{4L_y}{9\Omega_{\rm p}^2 x_c^2}
2176: \overline{\delta v_y} \delta v_x (x_c, L_y/2) .
2177: \end{eqnarray}
2178: The last term is zero as long as the box size is large enough. Using
2179: equation of continuity
2180: \begin{equation}
2181: -\dfrac{3}{2}ik_y x \Omega_{\rm p} \dfrac{\tilde{\delta
2182: \Sigma}_{k_y}}{\Sigma_0}
2183: + \dfrac{d}{dx}\tilde{\delta v_x}_{k_y} + ik_y \tilde{\delta v_y} =0
2184: \end{equation}
2185: and conservation of vorticity
2186: \begin{equation}
2187: \dfrac{d}{dx} \tilde{\delta v_y}_{k_y} - ik_y \tilde{\delta v_x}_{k_y}
2188: - \dfrac{1}{2} \Omega_{\rm p} \dfrac{\tilde{\delta
2189: \Sigma}_{k_y}}{\Sigma_0} = 0,
2190: \end{equation}
2191: we can show that the first three terms cancel. Hence,
2192: equation \eqref{Dx_vanish} is proved.
2193:
2194:
2195:
2196: \section{Mass Flux of Sound Wave Propagating in a Homogeneous,
2197: Static Medium}
2198: \label{app:soundwave}
2199:
2200: In this section, we demonstrate that the sound wave that propagates in a
2201: homogeneous, non-rotating medium carries mass flux, and therefore, show
2202: that the vanishing mass flux given by equation \eqref{massflux_vanish}
2203: is not a general conclusion but specific to the problem in
2204: consideration. We consider one-dimensional system which extends in
2205: $x>0$, and at $x=0$, there is a forcing that creates a sound wave, which
2206: is turned on at $t=0$. The full system of equations are the equation of
2207: continuity and Euler equation in one-dimensional system
2208: \begin{eqnarray}
2209: & \dfrac{\partial \rho}{\partial t} + \dfrac{\partial}{\partial x}(\rho
2210: x) = 0 \\
2211: & \dfrac{\partial v}{\partial t} + v\dfrac{\partial v}{\partial x} =
2212: - \dfrac{c^2}{\rho} \dfrac{\partial \rho}{\partial x} + S\cos(\Omega t)
2213: \delta_D(x)\theta(t) ,
2214: \end{eqnarray}
2215: where $\rho$ is density, $v$ is velocity, $c$ is sound speed, $S$ is the
2216: amplitude of the source, $\Omega$ is the frequency of the forcing,
2217: $\delta_D(x)$ is the Dirac's delta function, and $\theta(t)$ is step
2218: function. We consider background state with constant density $\rho_0$
2219: and vanishing velocity.
2220:
2221: The first order fluctuation caused by the forcing is given by
2222: \begin{eqnarray}
2223: & \dfrac{\partial}{\partial t} \dfrac{\delta \rho^{(1)}}{\rho_0} +
2224: \dfrac{\partial}{\partial x} \delta v^{(1)} = 0 \\
2225: & \dfrac{\partial}{\partial t} \delta v^{(1)} + c^2
2226: \dfrac{\partial}{\partial x} \dfrac{\delta \rho}{\rho_0} = S
2227: \cos(\Omega t) \delta(x) \theta(t)
2228: \end{eqnarray}
2229: and the second order perturbation is caused by the first order
2230: perturbation as follows
2231: \begin{eqnarray}
2232: & \dfrac{\partial}{\partial t} \dfrac{\delta \rho^{(1)}}{\rho_0} +
2233: \dfrac{\partial}{\partial x} \delta v^{(1)} =
2234: - \dfrac{\partial}{\partial x} \left[ \dfrac{\delta \rho^{(1)}}{\rho_0}
2235: \delta v^{(1)} \right] \\
2236: & \dfrac{\partial}{\partial t} \delta v^{(1)} + c^2
2237: \dfrac{\partial}{\partial x} \dfrac{\delta \rho}{\rho_0} =
2238: \dfrac{1}{2}\dfrac{\partial}{\partial x}
2239: \left[
2240: c^2 \left(\dfrac{\delta \rho^{(1)}}{\rho_0}\right)^2
2241: - \delta v^{(1)^2} \right].
2242: \end{eqnarray}
2243: Mass flux $F_M=\rho v$ is given by, up to the second order,
2244: \begin{equation}
2245: F_M = \rho_0 \delta v^{(1)} + \rho_0 \delta v^{(2)}
2246: + \delta \rho^{(1)} \delta v^{(1)}
2247: \end{equation}
2248:
2249: Since the sound wave propagates to $x\to\infty$ and there is no
2250: reflection, the appropriate solution for the first order is given by
2251: \begin{eqnarray}
2252: & \dfrac{\delta \rho^{(1)}}{\rho_0}
2253: = \dfrac{S}{c^2} \cos(\Omega t - kx), \\
2254: & \delta v^{(1)} = \dfrac{S}{c} \cos (\Omega t - kx)
2255: \end{eqnarray}
2256: where $k=\Omega/c$ and the boundary condition at $x=0$ is considered.
2257:
2258: Equation for second order velocity fluctuation is given by
2259: \begin{equation}
2260: \left[ -\dfrac{1}{c^2} \dfrac{\partial^2}{\partial t^2} +
2261: \dfrac{\partial^2}{\partial x^2} \right]
2262: \delta v^{(2)} =
2263: 2\dfrac{k^2 S^2}{c^3} \cos \left[ 2(\Omega t - kx) \right].
2264: \end{equation}
2265: Since all the perturbation must vanish at $t=0$, we obtain
2266: \begin{equation}
2267: \delta v^{(2)} = - \dfrac{S^2}{2c^3} \Omega t
2268: \sin \left[2(\Omega t - kx) \right] .
2269: \end{equation}
2270: Therefore, mass flux is given by
2271: \begin{equation}
2272: F_M = \dfrac{\rho_0 S^2}{2c^3}
2273: + \dfrac{\rho_0 S}{c} \cos \left[ \Omega t - kx \right]
2274: + \dfrac{\rho_0 S^2}{2c^3} \Omega t
2275: \sin \left[2(\Omega t -kx) \right]
2276: + \dfrac{\rho_0 S^2}{2c^3} \cos \left[ 2(\Omega t - kx) \right] .
2277: \end{equation}
2278: Therefore, there is a positive and finite mass
2279: flux $\rho_0 S^2/2c^3$ remains even after spatial and temporal averages
2280: are taken. It is also possible to show that
2281: a fluid element move to $x>0$ on average by calculating path line.
2282:
2283:
2284: \section{Interpolation Methods }
2285: \label{app:sphint}
2286:
2287: In the numerical calculation of large drag coefficient, stopping time of
2288: the particle $\nu^{-1}$ is very small that the particle remains in the
2289: same grid cell of the hydrodynamic calculation for several time steps.
2290: In this situation, it is necessary to have a smooth data of gas velocity
2291: even in sub-grid scale in order to obtain a smooth results. In this
2292: section, we describe how this can be realized.
2293:
2294: Our method of interpolation is to consider the value of one
2295: grid cell as a representative value of physical data around the grid
2296: cell with a certain width, and add the contribution from nearby grid
2297: cells to find the value of physical quantities at a location of
2298: interest. Let $f(x)$ be the value of a physical quantity of interest
2299: and the data of $f(x)$ is given at discrete set of grid cells at
2300: $x=x_i$. We denote the data of $f(x)$ at $x=x_i$ by $f_i=f(x_i)$ and
2301: $\Delta$ by grid size $\Delta=x_{i+1}-x_{i}$. Let
2302: the grid cell which is closest to the location of interest $x$ be
2303: $x_{i0}$. Our method of interpolation approximate the value of
2304: $f$ at $x$ by
2305: \begin{equation}
2306: f(x) \sim \dfrac{\sum_{i=i0-I}^{i=i0+I} f_i \exp[-(x_i/\eta
2307: \Delta)^2]}{\sum_{i=i0-I}^{i=i0+I}\exp[-(x_i/\eta \Delta)^2]}
2308: \end{equation}
2309: where $I$ is an integer and $\eta$ is a numerical factor. If $\eta$ is
2310: large, we have a more smoothed but more damped value. We find
2311: $I=20$ and $\eta=4$ gives a reasonably smooth results.
2312:
2313:
2314:
2315: \begin{thebibliography}{}
2316:
2317: \bibitem[Adachi et al. 1976]{AHN76} Adachi, I, Hayashi, C \& Nakazawa, K.
2318: 1976, Prog. Theo. Phys., 56, 1756
2319:
2320: \bibitem[Artymowicz 1993]{Art93} Artymowicz, P. 1993, \apj,
2321: 419, 155
2322:
2323: \bibitem[Fouchet et al. 2008]{F08} Fouchet, L., Maddison, S. T.,
2324: Gonzalez, J.-F., \& Murray, J.R., 2007, A\&A 474, 1037
2325:
2326: \bibitem[Goldreich and Tremaine 1979]{GT79} Goldreich, P. \& Tremaine, S.
2327: 1979, \apj, 233, 857
2328:
2329: \bibitem[Goldreich and Tremaine 1980]{GT80} Goldreich, P. \& Tremaine, S.
2330: 1980, \apj, 241, 425
2331:
2332: \bibitem[Goodman and Rafikov 2001]{GR01} Goodman, J \& Rafikov, R. R.
2333: 2001, \apj, 552, 793
2334:
2335: \bibitem[Greenberg 1978]{Gre78} Greenberg, R.
2336: 1978, Icarus, 33, 62
2337:
2338:
2339: \bibitem[Haisch et al. 2001]{HLL01} Haisch, K., E., Lada, E., A., \&
2340: Lada, C. J. 2001, \apj, 553, L153
2341:
2342: \bibitem[Hasegawa and Nalazawa 1990]{HN90} Hasegawa, M. \& Nakazawa, K.
2343: 1990, A\&A, 227, 619
2344:
2345: \bibitem[Hayashi et al. 1985]{HNN85} Hayashi, C., Nakazawa, K.,
2346: \& Nakagawa, Y. 1985, in Protostars and Planets II,
2347: ed. Black \& Matthews (Tucson: Univ. Arizona Press)
2348:
2349: \bibitem[H\'{e}non and Petit 1986]{HP86} H\'{e}non, M. \& Petit, J-M.
2350: 1986, Celes. Mech, 38, 67
2351:
2352: \bibitem[Inaba and Ikoma 2003]{II03} Inaba, S. \& Ikoma, M.
2353: 2003, A\&A, 410, 711
2354:
2355: \bibitem[Kary et al. 1993]{KLG93} Kary, D. M., Lissauer, J. J., \&
2356: Greenzweig, Y. 1993, Icarus, 106, 288
2357:
2358: \bibitem[Kretke and Lin 2007]{KL07} Kretke, A. K. \& Lin, D. N. C.
2359: 2007, \apj, 664, L55
2360:
2361:
2362: \bibitem[Landau and Lifshitz 1959]{LL59} Landau, L. D. \& Lifshitz
2363: E. M., 1959, Fluid Mechanics, Course of Theoretical
2364: Physics (Oxford: Pergamon Press)
2365:
2366: \bibitem[Lyra et al. 2008]{L08} Lyra, W, Johansen, A., Klahr, H., \&
2367: Piskunov, N. preprint (arXiv:0810.3192)
2368:
2369: \bibitem[van Leer 1977]{Leer77} van Leer, B.
2370: 1977, J. Comp. Phys., 23, 276
2371:
2372: \bibitem[Mizuno et al. 1978]{MNH78} Mizuno, H., Nakazawa, K., \&
2373: Hayashi, C. 1978, Prog. Theo. Phys, 60, 699
2374:
2375:
2376: \bibitem[Narayan et al. 1987]{NGG87} Narayan, R., Goldreich, P., \&
2377: Goodman, J. 1987, \mnras, 228, 1
2378:
2379: \bibitem[Paardekooper 2007]{P07} Paardekooper, S.-J.
2380: 2007, A\&A, 462, 355
2381:
2382: \bibitem[Paardekooper and Mellema 2004]{PM04} Paardekooper, S.-J. \&
2383: Mellema, G. 2004, A\&A, 425, L9
2384:
2385: \bibitem[Paardekooper and Mellema 2006]{PM06} Paardekooper, S.-J. \&
2386: Mellema, G. 2006, A\&A, 453, 1129
2387:
2388:
2389: \bibitem[Pollack et al. 1996]{PHBL96} Pollack, J. B., Hubickyj, O.,
2390: Bodenheimer, P., \& Lissauer, J. J.
2391: 1996, Icarus, 124, 62
2392:
2393:
2394: %\bibitem[Press et al. 1992]{NumRecipe} Press, W. H., Flannery, B. P.,
2395: % Teukolsky, S. A., \& Vetterling, W. T., 1992, Numerical Recipes
2396: % in Fortran 90: The Art of Scientific Computing ,
2397: % (Cambridge: Cambridge Univ. Press)
2398:
2399: \bibitem[Tanaka et al. 2002]{TTW02} Tanaka, H., Takeuchi, T. \& Ward,
2400: W. R. 2002, \apj, 565, 1257
2401:
2402:
2403: \bibitem[Weidenschilling 1977]{W77} Weidenschilling, S. J.
2404: 1977, \mnras, 180, 57
2405:
2406: \bibitem[Weidenschilling and Davis 1985]{WD85} Weidenschilling, S. J. \&
2407: Davis, D. R. 1985, Icarus, 62, 16
2408:
2409:
2410: \end{thebibliography}
2411:
2412:
2413: \clearpage
2414:
2415:
2416: %%%%%%%%%%Figure changed in revision%%%%%%%%%%
2417: \begin{figure}
2418: \plotone{f1.eps}
2419: \caption{An example of the relationship between the dust size and the
2420: drag coefficient
2421: $\nu$ used in our numerical calculations, given by equation
2422: \eqref{draglaw}. Reciprocal of stopping time, $\nu=t_e^{-1}$ is shown
2423: as a function of dust size in centimeters for gas densities
2424: with $\rho=10^{-8}\mathrm{g/cm}^3$, $\rho=10^{-10}\mathrm{g/cm}^3$, and
2425: $\rho=10^{-12}\mathrm{g/cm}^3$.}
2426: \label{fig:size_drag}
2427: \end{figure}
2428: %%%%%%%%%%Figure changed in revision end%%%%%%%%%%
2429:
2430:
2431: \begin{figure}
2432: \plotone{f2.eps}
2433: \caption{Comparison between the azimuthally averaged profile of
2434: $L_y \delta v_y/c$ obtained by the
2435: numerical calculation (solid line) and analytic expression
2436: \eqref{dvy_asym} (dotted line).
2437: The results of the numerical calculation is normalized by
2438: $GM_{\rm p}/Hc^2$.}
2439: \label{fig:dvy_asym_profile}
2440: \end{figure}
2441:
2442:
2443: \begin{figure}
2444: \plotone{f3.eps}
2445: \caption{Contour plot of $\Delta b/T$ given by equation
2446: \eqref{deltab_final} normalized by $H\Omega_{\rm p}$ for a planet with
2447: $GM_{\rm p}/Hc^2=10^{-2}$. The horizontal axis shows the distance from
2448: the planet and the vertical axis shows the drag coefficient of the
2449: particle. Global pressure gradient and steady mass accretion is
2450: neglected ($\eta=\zeta=0$). We note that our analytic approach is
2451: limited by assumptions of local approximation, initially circular orbit
2452: of the particle, and distant encounter. This figure has only
2453: qualitative meanings for particles with
2454: $\nu/\Omega_{\rm p}\lesssim10^{-2}$, with initial semi-major axis
2455: difference $b \lesssim 3r_{\rm H}$ for
2456: $\nu/\Omega_{\rm p} \lesssim 1-10$, or with
2457: $b \lesssim r_{\rm H}$ for $\nu/\Omega_{\rm p} \gtrsim 10$. Detailed
2458: discussions in the limitation of analytic calculations are given in
2459: Sections \ref{subsec:limitation_local} and \ref{global}, and are
2460: summarized in Section \ref{summary}.}
2461: \label{fig:db_contour}
2462: \end{figure}
2463:
2464: %%%%%%%figure added in revision%%%%%%%%
2465: \begin{figure}
2466: \plotone{f4.eps}
2467: \caption{Evolution of $b$ for a particle with zero drag coefficient
2468: initially located at $b=H$. Symbols show the values of $b$
2469: obtained at the box boundary. The
2470: value of $b$ during the encounter is also indicated by dashed line upto
2471: $t=250\Omega_{\rm p}^{-1}$.}
2472: \label{fig:threebody_b1}
2473: \end{figure}
2474: %%%%%%%figure added in revision end%%%%%%%%
2475:
2476: %%%%%%%figure added in revision%%%%%%%%
2477: \begin{figure}
2478: \plotone{f5.eps}
2479: \caption{Evolution of $b$ for a particle with zero drag coefficient
2480: initially located at $b_{\rm ini}=3.4H$ (plus) that corresponds
2481: to $l=5$ case in equation \eqref{resonance_cond} and
2482: $b_{\rm ini}=3.8H$ (cross) that does not satisfy equation
2483: \eqref{resonance_cond}. Vertical axis indicates the value of
2484: $(b(t)-b_{\rm ini})/H$. The value of $b(t)$ is obtained when the
2485: particle reaches the box boundary. Note that vertical axis of this
2486: figure is in logarithmic scale and the values smaller than $10^{-9}$
2487: are not plotted. Horizontal axis shows time.}
2488: \label{fig:threebody_res}
2489: \end{figure}
2490: %%%%%%%figure added in revision end%%%%%%%%
2491:
2492: \begin{figure}
2493: \plotone{f6.eps}
2494: \caption{Evolution of $b$ for particles initially located at $x=H$ with
2495: different drag
2496: coefficient. Planet mass corresponds to $GM_{\rm p}/Hc^2=10^{-2}$.
2497: Particles with small drag coefficients shows oscillation,
2498: while particles with large drag coefficient shows a systematic decrease
2499: or increase of semi-major axis.}
2500: \label{fig:bevol_nu}
2501: \end{figure}
2502:
2503:
2504:
2505: \begin{figure}
2506: \plotone{f7.eps}
2507: \caption{Comparison between the local numerical calculation (solid
2508: line) and analytical formula (dashed line) of $\Delta b / T$ for
2509: particles with $\nu/\Omega_{\rm p}=10^{-2}$.
2510: The value of $\Delta b/T$ is the average of a number of encounters and
2511: calculated using equation \eqref{average_db_calc}.
2512: The gravitational force by the planet and the
2513: spiral pattern of the gas around the planet are both considered.}
2514: \label{fig:bdecay_Hill_d10m}
2515: \end{figure}
2516:
2517: \begin{figure}
2518: \plotone{f8.eps}
2519: \caption{Same as Figure \ref{fig:bdecay_Hill_d10m} but for
2520: particles with $\nu/\Omega_{\rm p}=1$.}
2521: \label{fig:bdecay_Hill_d1m}
2522: \end{figure}
2523:
2524: \begin{figure}
2525: \plotone{f9.eps}
2526: \caption{Same as Figure \ref{fig:bdecay_Hill_d10m} but for
2527: particles with $\nu/\Omega_{\rm p}=10^2$.}
2528: \label{fig:bdecay_Hill_d10cm}
2529: \end{figure}
2530:
2531:
2532: \begin{figure}
2533: \plotone{f10.eps}
2534: \caption{Long-term evolution of the semi-major axis of the particles
2535: with $\nu=\Omega_{\rm p}$. The mass of the planet corresponds to
2536: $GM_{\rm p}/Hc^2=10^{-2}$. The calculations are done for the disk model
2537: with zero pressure gradient and zero accretion flow, $\eta=\zeta=0$.}
2538: \label{fig:long_nu1}
2539: \end{figure}
2540:
2541: %%%%%%%figures added in revision%%%%%%%
2542: \begin{figure}
2543: \plotone{f11.eps}
2544: \caption{Deflection of the particle's semi-major axis after the first
2545: encounter with the planet. The amount of deflection is quantified by
2546: equation \eqref{deflection_rate}. Horizontal axis denotes
2547: the particle's initial semi-major axis and vertical axis denotes the
2548: particle's drag, or equivalently, size. Initial semi-major axes are
2549: shown in terms of scale height (bottom axis) and Hill's radius (top
2550: axis). Gray scale shows the amount of deflection, and the contours of
2551: 10\% (solid line) and 1\% (dashed line) deflection are shown.
2552: Particles in black regions are trapped by the planet (that is, mutual
2553: distance between the particle and the planet becomes smaller than half
2554: of Hill's radius) or in horseshoe orbit.}
2555: \label{fig:size_deflection}
2556: \end{figure}
2557: %%%%%%%figures added in revision end%%%%%%%
2558:
2559: %%%%%%%figures added in revision%%%%%%%
2560: \begin{figure}
2561: \plotone{f12.eps}
2562: \caption{Examples of the orbit of the particle when they encounter with
2563: the planet. Particles with $\nu/\Omega_{\rm p}=10^{-4}$ (solid line)
2564: and $\nu / \Omega_{\rm p}=10^3$ (dashed line) are shown. Both
2565: particles are initially located at $b=2r_{\rm H}$. Only the region
2566: $(1r_{\rm H}<x<2r_{\rm H},-50r_{\rm H}<y<50r_{\rm H})$ is shown. The
2567: particle with small drag is strongly perturbed by the planet and
2568: eventually captured within the Hill's radius, while that with large
2569: drag can escape owing to the drag by the background gas.}
2570: \label{fig:orbit_2hill}
2571: \end{figure}
2572: %%%%%%%figures added in revision end%%%%%%%
2573:
2574:
2575: %%%%%figures added in revision%%%%%%
2576: \begin{figure}
2577: \plotone{f13.eps}
2578: \caption{Comparison between the global numerical calculation (solid
2579: line) and local calculation (dashed line) of
2580: $\Delta b / T$ for
2581: particles with $\nu/\Omega_{\rm p}=10^{-2}$. The average of the change
2582: of the semi-major axis of the particle after 100
2583: encounters with the planet is shown.
2584: Horizontal axis shows the initial location of the particle normalized
2585: by disk scale height, $b/H=(r-r_{\rm p})/H$.
2586: We use disk aspect ratio $H/r_{\rm p}=0.04$ and the mass ratio between
2587: the planet and the central star is $M_{\rm p}/M_{\ast}=6\times10^{-7}$.
2588: The disk aspect ratio is chosen in such a way that the value of $L_y$
2589: we have adopted in the integration of Hill's equation corresponds to
2590: $2\pi r_{\rm p}$, and the planet mass is chosen in such a way that
2591: $GM_{\rm p}/Hc^2=10^{-2}$, which is also the value we have used in
2592: local calculations. Keplerian rotation of the gas is assumed so
2593: $\eta=\zeta=0$ and there is no modification of the gas structure by the
2594: planet's gravity. Note that global calculations and local calculations
2595: are in good agreement except for the location of the resonances. The
2596: position of $4:5$, $5:6$, and $6:7$ resonances are indicated by arrows.
2597: }
2598: \label{fig:globalHill_d10m}
2599: \end{figure}
2600: %%%%%figures added in revision end%%%%%%
2601:
2602: %%%%%figures added in revision%%%%%%
2603: \begin{figure}
2604: \plotone{f14.eps}
2605: \caption{Same as Figure \ref{fig:globalHill_d10m} but for
2606: particles with $\nu/\Omega_{\rm p}=1$. Particles initially located at
2607: $b/H < 1.4$ has been trapped in the vicinity of the resonance so the
2608: average rate of orbital change over 100 encounters is small. See also
2609: Figures \ref{fig:long_nu1} and \ref{fig:long_nu1_global} for the time
2610: evolution of semi-major axis difference $b$
2611: for particles with $b/H \lesssim 1.4$.}
2612: \label{fig:globalHill_d1m}
2613: \end{figure}
2614: %%%%%figures added in revision end%%%%%%
2615:
2616: %%%%%figures added in revision%%%%%%
2617: \begin{figure}
2618: \plotone{f15.eps}
2619: \caption{Same as Figure \ref{fig:globalHill_d10m} but for
2620: particles with $\nu/\Omega_{\rm p}=10^{2}$.}
2621: \label{fig:globalHill_d10cm}
2622: \end{figure}
2623: %%%%%figures added in revision end%%%%%%
2624:
2625: %%%%%figures added in revision%%%%%%
2626: \begin{figure}
2627: \plotone{f16.eps}
2628: \caption{Evolution of semi-major axis of particles with
2629: $\nu/\Omega_{\rm p}=1$ in global run with $\eta=\zeta=0$ and the
2630: gas is assumed to rotate at Kepler velocity. Osculating elements of
2631: the particle at the opposition point is plotted. Particles in both
2632: interior and exterior orbits migrate towards the planet and trapped in
2633: the orbit of $\sim 0.3H$. This figure corresponds to Figure
2634: \ref{fig:long_nu1} that shows the results of local calculations.}
2635: \label{fig:long_nu1_global}
2636: \end{figure}
2637: %%%%%figures added in revision end%%%%%%
2638:
2639: \begin{figure}
2640: \plotone{f17.eps}
2641: \epsscale{0.95}
2642: \caption{Time evolution of the distribution of dust particles around
2643: $2M_{\oplus}$ embedded
2644: in a solar nebula at $t\Omega_{\rm p}=5\times 10^3$ (top left),
2645: $t\Omega_{\rm p}=10^5$ (top right), and
2646: $t\Omega_{\rm p}=10^6$ (bottom left) with various $\nu$.
2647: The horizontal axis denotes the semi-major axis difference between
2648: the planet and the particle. Gray scale indicates the fraction of
2649: the remaining particles. The
2650: pressure gradient factor $\eta$ is set to be zero.}
2651: \label{fig:advection}
2652: \end{figure}
2653:
2654:
2655: \end{document}
2656: