0804.0975/ms.tex
1: \documentclass[12pt,preprint]{aastex}
2: %%\usepackage{graphicx}
3: %% manuscript produces a one-column, double-spaced document:
4: 
5: %\documentclass[manuscript]{aastex}
6: 
7: %% preprint2 produces a double-column, single-spaced document:
8: 
9: % \documentclass[preprint2]{aastex}
10: 
11: %% If you want to create your own macros, you can do so
12: %% using \newcommand. Your macros should appear before
13: %% the \begin{document} command.
14: %%
15: %% If you are submitting to a journal that translates manuscripts
16: %% into SGML, you need to follow certain guidelines when preparing
17: %% your macros. See the AASTeX v5.x Author Guide
18: %% for information.
19: \def\dfrac{\frac}
20: \newcommand{\vdag}{(v)^\dagger}
21: \newcommand{\var}{\Omega}
22: \newcommand{\tha}{\theta}
23: \newcommand{\alp}{\alpha}
24: \newcommand{\beq}{\begin{equation}}
25: \newcommand{\eeq}{\end{equation}}
26: \newcommand{\bea}{\begin{array}}
27: \newcommand{\eea}{\end{array}}
28: \newcommand{\dw}{\Delta \Omega}
29: 
30: %% You can insert a short comment on the title page using the command below.
31: 
32: %\slugcomment{Not to appear in Nonlearned J., 45.}
33: 
34: %% If you wish, you may supply running head information, although
35: %% this information may be modified by the editorial offices.
36: %% The left head contains a list of authors,
37: %% usually a maximum of three (otherwise use et al.).  The right
38: %% head is a modified title of up to roughly 44 characters.  Running heads
39: %% will not print in the manuscript style.
40: 
41: \shorttitle{Hot Jupiter} \shortauthors{Zhou et
42: al.}
43: 
44: %% This is the end of the preamble.  Indicate the beginning of the
45: %% paper itself with \begin{document}.
46: 
47: \begin{document}
48: 
49: 
50: \title{Interaction of Close-in Planets with the Magnetosphere of their
51: Host Stars I: Diffusion, Ohmic Dissipation of Time Dependent Field, 
52: Planetary Inflation, and Mass Loss}
53: 
54: 
55: \author{Randy O. Laine\altaffilmark{1},  
56: Douglas N.C. Lin\altaffilmark{2,3}, \&
57: Shawfeng Dong\altaffilmark{2}}
58: \altaffiltext{1}{Ecole Normale Superieure, Paris, France, randy.laine@ens.fr}
59: \altaffiltext{2}{UCO/Lick Observatory, University of California, Santa Cruz, CA 95064, USA lin@ucolick.org, dong@ucolick.org}
60: \altaffiltext{3}{Kavli Institute of Astronomy \& Astrophysics, Peking University, Beijing,  China}
61: 
62: 
63: \begin{abstract} 
64: The unanticipated discovery of the first close-in planet around 51 Peg
65: has rekindled the notion that shortly after their formation outside
66: the snow line, some planets may have migrated to the proximity of
67: their host stars because of their tidal interaction with
68: their nascent disks. After a decade of discoveries, nearly 20 \% of
69: the 200 known planets have similar short periods. If these 
70: planets indeed migrated to their present-day location, their survival
71: would require a halting mechanism in the proximity of their host
72: stars. Here we consider the possibility that a magnetic coupling
73: between young stars and planets could quench the planet's orbital
74: evolution. Most T Tauri stars have magnetic fields of several thousand
75: gausses on their surface which can clear out a cavity in the
76: innermost regions of their circumstellar disks and impose
77: magnetic induction on the nearby young planets. After a brief
78: discussion of the complexity of the full problem, we focus our
79: discussion on evaluating the permeation and ohmic
80: dissipation of the time dependent component of the stellar magnetic
81: field in the planet's interior. Adopting a model first introduced by
82: Campbell for interacting binary stars, we determine the
83: modulation of the planetary response to the tilted magnetic field of a
84: non-synchronously spinning star. We first compute
85: the conductivity in the young planets, which indicates
86: that the stellar field can penetrate well into the planet's envelope in
87: a synodic period.  For various orbital configurations, we show that the
88: energy dissipation rate inside the planet is sufficient to induce
89: short-period planets to inflate. This process
90: results in mass loss via Roche lobe overflow and in the halting of the
91: planet's orbital migration.
92: \end{abstract}
93: 
94: \keywords{Planetary systems: formation, planetary disks: protoplanetary disks, stars: magnetic field, MHD, accretion disks, stars: individual (Peg 51b)}.
95: 
96: \section{Introduction}
97: Perhaps the most surprising finding in the search for extrasolar
98: planets is the discovery of short-period ($P < 1$ week) Jupiter-mass
99: ($M_J$) companions around Solar-type main sequence stars (Mayor \&
100: Queloz 1995, Marcy et al. 2000).  Among the inventory of $>$200
101: presently-known extrasolar planets, 20\% have $P = 1-7$ days. Nearly
102: 20 short-period planets have measured radii ($R_p$) that are comparable to or
103: larger than that of Jupiter ($R_J$). While these information may be
104: biazed because of observational selection effects, these planets are
105: most probably gas giants.
106: 
107: According to the conventional sequential-accretion scenario (Pollack
108: {\it et al.} 1996), the most likely birth place for gas giant planets
109: is just outside the snow line where volatile heavy elements can
110: condense and coagulate into large planet building blocks (Ida \& Lin
111: 2004). In protostellar disks with comparable surface density
112: ($\Sigma$), metallicity ([Fe/H]), and temperature ($T$) distributions
113: as those of minimum mass nebula model (Hayashi {\it et al.} 1985),
114: protoplanets with $M_p \sim M_J$ induce the formation of a gap near
115: their orbit as a consequence of their tidal torque on the nascent
116: disks (Goldreich \& Tremaine 1978, 1980, Lin \& Papaloizou 1980,
117: 1986a, 1993).  In relatively massive and fast evolving disks, the
118: outward transfer of angular momentum due to the disks' intrinsic
119: turbulence can lead to an inward mass flux ($\dot M_d$) and the
120: migration of the gas giant planets (Lin \& Papaloizou 1986b). This
121: process is commonly referred to as type II migration (Ward 1997).
122: 
123: This migration scenario was resurrected to account for the origin of
124: the first known short-period extra solar planet (Lin {\it et al.}
125: 1996).  Although type-II migration provides a natural avenue for
126: relocating some gas giants, a mechanism is needed to retain these
127: planets close to their host stars. Moreover, many stars are born with
128: rapid rotation (Stassun 2001). When young planets venture close to
129: their host stars, angular momentum would be transferred from the
130: stellar spin to the planet's orbit if the stellar spin frequency
131: $\omega_\ast$ is still larger than the planet's orbital frequency
132: ($\Omega_k$).  The rate of the star-to-planet angular momentum
133: transfer intensifies rapidly and may exceed that from the planet to
134: the disk.
135: 
136: Two basic physical effects were suggested as potential migration barriers.
137: The first one is tidal interaction between the host star and the planet. The
138: gravitational perturbation of the star and close-in planet leads to
139: responses in both the star and planet. For tidal frequencies smaller than twice the
140: spin frequency, inertial waves are excited in the convective envelope
141: of the host star and are dissipated there by turbulent viscosity
142: (Ogilvie \& Lin 2007).  But, the tide excited by a close-in gas giant
143: planet in a star, with a structure similar to that of the present Sun,
144: marginally fails to achieve nonlinearity so that their survival is
145: ensured. Nevertheless, during the formation epoch of solar-type stars,
146: conditions at the center of the star evolve, so that nonlinearity may
147: set in at a critical age, resulting in a relatively intense
148: star-planet tidal interaction.
149: 
150: The second effect suggested is based on the magnetic interaction 
151: between the host star and the planet. 
152: Young T Tauri stars also have radii ($R_\ast$) 2-3 times that of the
153: present-day Sun ($R_\odot$) and several thousand gausses fields
154: ($B_\ast$) on their surface (Johns-Krull 2007). The stellar magnetosphere
155: threads across the inner regions of the disk and clears a cavity out
156: to a critical radius ($R_c$) which is determined by both the magnitude
157: $B_\ast$ and $\dot M_d$ (Konigl 1991). The subsequent complex interplay
158: between accretion and outflow leads to angular momentum exchange which
159: induces $\omega_\ast$ to evolve toward $\Omega_k$ at $R_c$ (Shu 1994).
160: When the planet's orbital semi major axis ($a$) reduces well inside
161: $R_c$, its Lindblad resonances relocate inside the star's
162: magnetospheric cavity. In principle, the planet's migration would
163: stall due to its diminishing tidal torque on the disk.
164: 
165: However, if the star's magnetospheric interaction with the disk can
166: lead to $\omega_\ast = \Omega_k (R_c)$, the planet inside the
167: magnetospheric cavity would have $\Omega_k > \omega_\ast$.  In this
168: limit, the star-planet tidal interaction would induce a transfer of
169: angular momentum from the planet to the star. In addition, the
170: differential motion between the planet and the stellar spinning
171: magnetosphere induces an electromagnetic field with a potential to
172: generate a large current analogeous to the interaction between the Jovian
173: magnetosphere with its satellite Io (Goldreich \& Lynden-Bell 1969).
174: The associated Lorentz force drives an orbital evolution toward a
175: synchroneous state, in which case, angular momentum would be
176: transferred from the planets with $\Omega_k > \omega_\ast$ to their
177: host stars, and the planets would continue their orbital decay.
178: 
179: In order to determine the necessary condition for the retention of
180: close-in young planets, we examine, in this paper, their interaction
181: with the magnetosphere of their host T Tauri stars.  
182: In \S2, we briefly recapitulate the essential concepts and validity of previous
183: investigations on some related topics and give an overview of the key 
184: phenomena that will be discussed in this and later papers. In \S3, we
185: adopt an existing model in order to examine the interaction between a
186: planet and the magnetosphere of its host star. In \S4, we compute
187: precisely the planet's magnetic diffusivity for a specific set of
188: parameters, as well as the corresponding ohmic dissipation rate within
189: that planet. In \S5, we suggest that the ohmic dissipation can
190: generate sufficient heat to inflat the planet.  In \S6, we construct an idealized 
191: self-consistent model in which the polytropic and isothermal equations
192: of state are utilized.  These equations represent the expected outcome
193: of radiation transfer within the fully convective interior and the
194: isothermal surface of a close-in planet which is exposed to the
195: intense radiation from its host star.  This internal model allows us
196: to compute the magnetic diffusivity.  With these tools, we discuss in
197: \S7 the structural adjustment of the planet in response to this
198: heating source and we compute the ohmic dissipation and mass loss rate
199: for different set of parameters. Finally in \S8, we summarize our
200: results and discuss their implications.
201: 
202: \section{Planetary and astrophysical analogue}
203: 
204: Two previous analyses are directly relevant to the present study: 1) the
205: interaction of Io with the magnetosphere of Jupiter and 2) the spin-orbit
206: synchronization in binary stars containing a magnetized white dwarf
207: and its main sequence or white dwarf or planetary companion. 
208: 
209: \subsection{Unipolar Induction in Io}
210: Io orbits around Jupiter inside its magnetosphere once every 1.7 days
211: which is considerably longer than Jupiter's 10 hours spin period. This
212: relative motion imposes a periodic variation in Jupiter's decametric
213: emission (Duncan 1966).  A class of models that accounts for the
214: origin of this emission was developped based on the assumption that Io
215: has a sufficiently high conductivity. In Io's rest frame, the
216: electric field vanishes and the steady component of Jupiter's magnetic
217: field permeates in Io's interior over time. When a steady state is established, 
218: the tube of constant magnetic flux is firmly frozen into Io 
219: (Piddington and Drake 1968) due to its high conductivity. 
220: The flux tube carried by Io moves through the surrounding field lines (which
221: corotate with Jupiter) and slips through Jupiter's less conductive ionosopheric
222: surface. Plasma in Jupiter's ionosphere flows around the tube and
223: introduces a potential difference across it (Goldreich \& Lynden-Bell
224: 1969). The associated electric field drives a current which 
225: travels down one half of the flux tube from Io and is sent back to Io
226: along the other half. Within the flux tube connecting Io and Jupiter's
227: ionosphere and those across it on Io, the electric field vanishes as a
228: consequence of high conductivity. Thus, this DC circuit is closed by
229: Io as a unipolar inductor.
230: 
231: The magnitude of the electric current is primarily determined by the Pederson
232: conductivity at the foot of the flux tube {\it i.e.} on the ionosphere
233: of Jupiter. Finite conductivity also determines the magnitude of the
234: drag against the slipage of the flux tube through Jupiter.  This drag
235: results in energy dissipation on the surface of Jupiter and in a
236: torque on the orbital motion of Io, driving the system towards a state
237: of synchronization.  This configuration is justified by the assumption
238: that a constant flux tube is firmly anchored on and dragged along by
239: Io which requires the conductivity in Io to be much higher than that
240: in Jupiter's ionosphere. A 10$^o$ inclination between Jupiter's
241: magnetic dipole and rotation axis does introduce a periodic variation
242: (over a synodic period) in the field felt by Io. The permeation and
243: dissipation of this time-dependent AC field may be negligible in the
244: limit of high conductivity in Io.
245: 
246: The validity of the key assumption for the unipolar induction model
247: ({\it i.e.} conductivity on Io is larger than that in Jupiter's
248: ionosphere) has also been challenged by Dermott (1970).  A modest
249: resistence in Io would distort the field which may lead to field
250: slippage through Io.  In this case, the passage of Io through the
251: magnetosphere of Jupiter would lead to the generation of Alfven waves
252: along the flux tube (Drell {\it et al.}  1965, Neubauer 1980).  But,
253: due to the field displacement, the waves, partially reflected at the
254: foot of the flux tube on Jupiter's surface, may not be able to return
255: to Io, in which case the DC circuit would be broken and the motion of
256: Io would be decoupled from the that of the flux tube.  Nevertheless,
257: the Alfven waves are dissipated inside both Io and Jupiter, leading to
258: a torque which must depend on their penetration depth.
259: 
260: An alternative class of scenarios has been proposed based on the
261: assumption that the magnetosphere is everywhere anchored on Jupiter
262: and the flux tube moves freely through Io (Gurnett 1972).  This model
263: requires the conductivity in Jupiter's ionosphere to be larger than
264: that in Io.  It assumes that the presence of Io creates a plasma
265: sheath with an electric field to cancel the induced EMF associated
266: with the motion of Io relative to Jupiter's magnetosphere (Shawhan
267: 1976).  The simplifying approximations in the development of this
268: theory have been challenged by Piddington (1977) who questioned both
269: the validity of the sheath-creation mechanism and the self-consistency
270: of the internal and external field configurations, subjected to the
271: electric currents in and around Io.
272: 
273: On the observational side, UV emissions from Io's footprint on
274: Jupiter has been observed. But, it extends
275: well beyond the intersection between Io's flux tube and Jupiter's
276: ionosphere and the emssion downstream is protracted (Clarke {\it et
277: al.} 1996). These observations do not agree with the simple
278: interpretation of either the unipolar induction or the plasma sheath
279: scenarios.
280: 
281: \subsection{Magnetic Coupling in Interacting Binary Stars}
282: There are many close binary-star systems with a white dwarf as their
283: primary component.  These systems also contain main sequence stars and
284: other white dwarfs as secondary components in compact and circular
285: orbits around each other.  In some cases, mass is transferred from the
286: secondary to the primary.  In other cases, gravitational radiation may
287: play an important role in determining the evolution of these systems.
288: 
289: A sub class of such interacting binary stars, AM Her systems, is
290: composed of a magnetized white-dwarf primary and a lower-mass main
291: sequence star as its secondary in a fully synchroneous orbit despite
292: the ongoing mass transfer between them (Warner 1995).  This orbital
293: configuration is very similar to that of the Jupiter-Io system despite
294: the enormous difference between the mass ratio in the two cases. The
295: motivation for studying the impact of magnetic coupling between these
296: stellar components is to assess whether this synchronous state can be
297: achieved through the ohmic dissipation of the white dwarf's field in
298: the main sequence star's surface (Joss {\it et al.} 1979).  Toward
299: this goal, Campbell (1983, hereafter C83) adopted a novel approach by
300: considering the penetration and dissipation of a periodically variable
301: field, associated with an asynchronously spinning primary. 
302: 
303: Campbell's approach is fundamentally different from that of the
304: unipolar induction model.  In this analysis, Campbell focussed on the
305: flow in the envelope of the secondary and neglected the possibility of
306: current flowing through the flux tube between the secondary star/satellite
307: and the surface of the primary star/planet. This vacuum-surrounding
308: approximation is justifiable since conductivity on the primary is
309: likely to be much larger than that on the secondary and 
310: the stationary component of the field is
311: frozen in the white-dwarf primary but not in the main-sequence
312: secondary. Campbell analyzed the time dependent response of
313: the secondary, including the modification of the field by the induced
314: (AC) current in it (Campbell 2005), to the periodic modulation of the
315: field.  In contrast, the unipolar induction model depends on the
316: explicit assumption that the field is anchored on the secondary and
317: its distortion near the secondary must be small so that a complete
318: current loop can be established between the primary and the
319: secondary. Campbell determined the periodic diffusion of the field and
320: the ohmic dissipation of the induced AC current in the companion
321: whereas that of the induced DC current is assumed to occur in the
322: primary in the unipolar induction model.
323: 
324: In recent applications of the unipolar induction model in the context
325: of interaction between white dwarf binary stars, the current induced
326: by the unperturbed field has been computed but the induced field
327: generated by the current was neglected (Wu {\it et al.}  2002,
328: Dall'Osso {\it et al.}  2006). A tidal torque is computed at the
329: footprint of the flux tube, which is attached to the secondary white
330: dwarf, on the surface of the magnetized primary white dwarf.  A
331: totally self-consistent solution of this diffult and complex problem
332: remains outstanding.  In addition to the uncertain anchorage location
333: of the field, it is not clear whether the resulting misalignment of
334: the total (original plus induced) field and current may be
335: sufficiently large to break the circuit, in which case, Campbell's
336: model may be more appropriate.
337: 
338: \subsection{Mathematical approximations made by the two previous models}
339: In this subsection, we summarize the physical description of the two
340: models just presented (the unipolar inductor versus the periodic
341: diffusion) by explicating the mathematical approximations made in each
342: one of them.  The complete MHD induction equation can be expressed as
343: \begin{equation}
344: \frac{\partial \textbf{B}}{\partial t}=
345: \nabla \wedge (\upsilon\wedge 
346: \textbf{B})
347: -\nabla \wedge (\eta \nabla \wedge\textbf{B})
348: \label{complete mhd induction equation}
349: \end{equation} 
350: where the magnetic diffusivity $\eta=1/\mu_{0}\sigma$ and the
351: electrical conductivity $\sigma$ functions of position and $\mu_{0}$
352: is the permeability.  If $\eta$ is constant, the second term on the
353: right hand side becomes $-\nabla \wedge (\eta \nabla
354: \wedge\textbf{B})= \eta\ \nabla^{2} \textbf{B}$ which reduces to a
355: common expression of diffusion.  The two models (unipolar induction
356: versus periodic diffusion) consider two complementary approximations
357: of equation \ref{complete mhd induction equation}.  In the problem
358: where Io is treated as an unipolar inductor, its conductivity is
359: explicitly assumed to be large so that the second term on the right
360: hand (the diffusion term) is negligible compared to the first ({\it
361: i.e.} the induction term).  In this configuration, one can show that
362: the field lines of the steady component of the magnetic field are
363: moving with Io and appear to be "frozen" on Io (see appendix A).
364: Alternatively, in the model considered by Campbell, it is the first
365: term on the right hand side that is being neglected. Moreover, only
366: the diffusion of the time dependent component of the field is being
367: considered.  This approximation is valid if the two interacting bodies
368: are almost in corotation ({\it i.e.} the relative speed $\upsilon$
369: that appears in equation (\ref{complete mhd induction equation}) is
370: small), or if the conductivity in the secondary is small.
371: 
372: 
373: \subsection{Overview of the phenomena that will be discussed}
374: \label{subsection overview of our model}
375: The process under investigation in this paper is analogeous to both
376: the Jupiter-Io and the interacting binary star problems.  In fact, the
377: unipolar induction model has already been applied to study the orbital
378: evolution of terrestrial planets or the cores of gas giants around
379: white dwarfs (Li {\it et al.} 1998). There are even follow-up
380: determination of the radio flux densities from potential
381: white-dwarf/planet systems (Willes \& Wu 2005). In this analysis,
382: although the dissipation of the induced current due to the finite
383: conductivities in the white dwarf was considered, the feedback
384: modification of the field and the dissipation within the planet have
385: been neglected (Li {\it et al.} 1998).  As discussed above, it is not
386: clear whether a DC circuit can be closed to promote the unipolar
387: induction mechanism.
388: 
389: In light of these uncertainties, we consider both classes of models
390: for the interaction of close-in planets with their magnetized host
391: stars. In this paper, we focus our discussion on the
392: mechanism described by Campbell, 
393: and apply it to a hot-jupiter revolving around its star. We
394: will return to the unipolar induction problem in a later paper.
395: 
396: When young planets first arrive at the vicinity of their host stars,
397: they are unlikely to be in a totally synchronized state.  The stellar
398: magnetic field felt by the planet may be dominated by the periodic
399: modulation associated with the synodic (between the stellar spin and
400: the planet's orbit) motion. In addition, the temperature in the
401: planet's surface is expected to be $\sim 10^3$ K and the conductivity
402: there may be moderate.  In response to the modulation of the field,
403: the interior of the planet continually adjusts to the magnetization
404: effects so that the flux tube cannot be effectively frozen in the
405: planet.  All of these boundary condition suggest that at least over
406: some regions of the planet (especially on the night side where the
407: photo ionization due to the stellar flux is negligible), the
408: modulation of the field may lead to an induced current inside the
409: planet which does not contribute to the close circuit of a unipolar
410: inductor.
411: 
412: Following the geometry introduced by Campbell (C83), we consider a
413: close-in gas giant planet, with a finite conductivity, interacting
414: with a time-dependent magnetic field generated by the star. An induced
415: current is generated inside the planet, which is associated with an
416: ohmic dissipation rate. Our main contributions to the model used by
417: Campbell are: 1) the relevant diffusivity inside the gas giant
418: planets, 2) the effects of the ohmic dissipation on the planet's
419: internal structure, and 3) the resulting orbital evolution of the
420: planet. (Items 2 and 3 have negligible consequence in the interacting
421: binary star problem considered by Campbell). Since we are only
422: considering the dissipation in the planet's interior, the associated
423: torque applied on its orbit should be regarded as a lower limit.
424: 
425: In our scenario, we postulate that at sufficiently close proximity to
426: the host star, the stellar magnetic field is sufficiently intense that
427: the ohmic dissipation of the periodically diffused field inside the
428: planet is adequate to heat and to inflat the planet until it overflows
429: its Roche lobe. The hemisphere of the planet facing its host star is also exposed to
430: the intense flux of UV radiation during the stellar infancy. It is
431: possible for the planet to develop a substantial ionosphere regardless
432: the state of synchronization between the planet's orbit and spin (the
433: time scale for establishing local ionization equilibrium is much
434: faster than the planet's spin and orbital periods).  
435: 
436: We will separately study these two phenomena (angular momentum transfert due 
437: to mass loss and presence of a ionosphere) in the follow-up papers of this serie. 
438: We will show that the angular momentum transfer associated with the mass transfer can halt the
439: orbital evolution of the planet. We will also present an analysis on the conductivities in the planet's day-side
440: ionosphere and on the host stars surface. This will lead to an
441: analysis on the condition for the unipolar induction to effectively
442: operate and apply a significant slow-down torque on the planet's
443: orbit.
444: 
445: \section{Magnetic induction} 
446: \label{chapter magnetic induction}
447: 
448: In this section, we are going to derive the governing equations that
449: we use to compute the ohmic dissipation rate. Various equations are
450: presented here for the purpose of introducing the algorithm of the
451: numerical models to be presented in subsequent chapters.  Although we
452: follow closely the approach made in C83, for brevity, we do not
453: repeatedly cite this reference. But, wherever similarities occur,
454: referal of Campbell's earlier work is implicitly implied. Also, throughout the paper, we use SI units. 
455: 
456: \begin{figure}[htbp]
457: \begin{center}
458: \epsscale{0.5}
459: \plotone{f1}  
460: \end{center}
461: \caption{The geometry of the system. The star is on the left, at the center of the 
462: set of axies ($x_{0}$,$y_{0}$,$z_{0}$), and the planet is on the right, at the center of 
463: the set of axies (x,y,z).}
464: \end{figure}
465: 
466: 
467: We consider a protoplanetary system with a gas giant planet revolving
468: around a T Tauri star with an angular frequency $\Omega_p$. Well
469: beyond the planet's semi major axis, there is also a protoplanetary
470: disk. The host star has a dipolar moment $\textbf{m}$
471: tilted with an angle $\alpha$ with respect to its spinning axis (see figure 1). 
472: The angular frequency of the stellar spin is $\omega_\ast$.  The orbital
473: axis of the disk and the planet are parallel to the star's spinning
474: axis. The following analysis is applied to a frame of reference
475: centered on the star and rotating with the planet. 
476: 
477: In this frame, the planet is a fixed object (the planet's spin is
478: neglected) in a periodic magnetic field with a frequency $\omega =
479: \omega_\ast -\Omega_p$.  From Ohm's law 
480: $\textbf{J}=\sigma \textbf{E}$ 
481: and Maxwell's equations $\dfrac{\partial \textbf{B}}{\partial t} = 
482: -\nabla \wedge \textbf{E}$ and 
483: $\nabla \wedge \textbf{B} = 
484: \mu_{0} \textbf{J}$, 
485: the equation on the magnetic field becomes:
486: \begin{equation}
487: \dfrac{\partial\textbf{B}}{\partial t}
488: =- \nabla\wedge(\eta \nabla \wedge\textbf{B}) .
489: \label{B time dependent}
490: \end{equation}
491: 
492: It follows that in the mechanism considered by Campbell 
493: (as well as in this paper), it is the time dependent stellar 
494: magnetic field, diffusing inside the 
495: secondary (for Campbell) or the hot Jupiter (in our paper), 
496: as well as the planet's induced magnetic field, 
497: that generate the current inside the planet, following the equation 
498: $\nabla \wedge \textbf{B} = 
499: \mu_{0}\textbf{J}$.
500: The relative speed between the planet and the stellar magnetic 
501: field thus intervenes not through 
502: $\textbf{E} = 
503: -\upsilon \wedge \textbf{B}$ 
504: but through the time dependence in the stellar magnetic field that diffuses in the planet. 
505: 
506: Following C83, we only consider the poloidal component $\phi$ of the magnetic field: 
507: \begin{equation}
508: \textbf{B}\ =\  
509: \nabla\wedge(\nabla\wedge(\phi\ \mathbf{e_{r}}))
510: \label{definition poloidal scalar}\\
511: \end{equation}
512: where $\phi$ is a function of r, $\theta$, $\varphi$. and can be expanded 
513: in terms of the sperical harmonics
514: $Y_{l}^{m}(\theta,\varphi)$ (equation \ref{decomposition of
515: Phi}). Moreover the variation in time of the magnetic field felt by
516: the planet is periodic. In the limit where the field penetrates
517: quickly in the planet compared to the time scale on which the field
518: changes (so that the planet can respond ``adiabatically''), we can
519: account for the time dependence of $\phi$ by mutiplying its spatial
520: part by $e^{i\omega t}$:
521: \begin{eqnarray}
522: \phi(\textbf{r},t)
523: =\mu_{0}\left[\sum_{l,m}\ C_{l}^{m}\
524: G_{l}(r)\ Y_{l}^{m}(\theta,\varphi)\right]e^{i\omega t}
525: \label{decomposition of Phi} \ \ \ \ \ \ \ (l \geqslant 0\ and\
526: -l\leqslant m\leqslant l)
527: \end{eqnarray}
528: where $C_{l}^{m}$ are constant coefficients and $G_{l}(r)$ is a
529: function of $r$ to be determined. We then replace \textbf{B} in 
530: the left hand of (2) by its expression in (3). After integration, we obtain 
531: \begin{equation}
532: \nabla\wedge\mathbf{B}\
533: =-\dfrac{i\omega}{\eta}\nabla\wedge
534: \left(\phi\ \mathbf{e_{r}}\right).
535: \label{rhs=lhs}
536: \end{equation}
537: We then replace \textbf{B} in the left hand side of this equation using (\ref{definition poloidal scalar}), and 
538: develop both sides of the equation. After identification, we obtain: 
539: \begin{equation}
540: \dfrac{d^{2}G_{l}}{dr^{2}}(r)-\left(\dfrac{l(l+1)}{r^{2}}+
541: \dfrac{i\omega}{\eta}\right)G_{l}(r)=0 \ \ \ \ \ \ \ \ \ \ \ 
542: {\rm inside} \ {\rm the} \ {\rm planet.} \label{first equation on G}
543: \end{equation}
544: This equation holds inside and outside the planet (same as eqs 16 and
545: 18 in C83). However, outside the planet, the conductivity is assumed
546: to be very low and, therefore, the magnetic diffusivity is extremely
547: high compared to the diffusivity inside the planet. In the limit where
548: the magnetic diffusivity outside tends to infinity (equivalent to a
549: vacuum surrounding), the equation (\ref{first equation on G}) becomes:
550: \begin{equation}
551: \dfrac{d^{2}G_{l}}{dr^{2}}(r)-\left(\dfrac{l(l+1)}{r^{2}}\right)G_{l}(r)
552: =0 \ \ \ \ \ \ \ \ \ 
553: \ \ \ \ \ \ \ \ \ \ {\rm outside} \ {\rm the} \ {\rm planet.} \label{second equation on G}
554: \end{equation}
555: 
556: We consider the radial part of the poloidal scalar outside the planet. Following C83,  
557: we introduce $\phi_{star}$ the radial part of the poloidal scalar outside the planet 
558: due to the star's magnetic field, and $\phi_{planet}$ the radial part of the poloidal scalar 
559: outside the planet due to the planet (cf C83 eqs 21-22): 
560: \begin{eqnarray}
561: \phi_{star}=\dfrac{\mu_{0}m {\rm \sin}\alpha}{8\pi d^{3}}r^{2} 
562: \left(2{\rm cos}\varphi\ {\rm \sin}\omega t+\ {\rm \sin}\varphi\ 
563: {\rm cos}\omega t\right)P_{1}^{1} 
564: + \dfrac{\mu_{0}m {\rm \sin}\alpha}{8\pi\ d^{4}}r^{3}\left[P_{2}^{0}
565: {\rm \sin}\omega t \right.   \nonumber \\
566: -\left. \left(\dfrac{1}{2} {\rm cos}2\varphi\ {\rm \sin}\omega t\ 
567: + \dfrac{1}{3}{\rm \sin} 2\varphi\ {\rm cos}\omega t\right)P_{2}^{2}\right] 
568: \label{phi_star outside}
569: \end{eqnarray}
570: 
571: \begin{eqnarray}
572: \phi_{planet}=\mu_{0}P_{1}^{1}\left[\dfrac{{\rm cos}\varphi}{r} 
573: (\alpha_{1}\ {\rm \sin}\omega t\ +\alpha_{2}\ {\rm cos}\omega t)
574: +\dfrac{ {\rm \sin}\varphi} {r}(\alpha_{3}\ {\rm \sin}\omega t\ 
575: +\alpha_{4}\ {\rm cos}\omega t)\right]\ 
576: \nonumber \\
577: +\ \dfrac{\mu_{0}P_{2}^{0}}{r^{2}}(\beta_{1}\ {\rm \sin}
578: \omega t\ +\ \beta_{2}\ {\rm cos}\omega t) 
579: +\mu_{0}P_{2}^{2}\left[\dfrac{ {\rm cos}2\varphi}
580: {r^{2}}(\gamma_{1}\ {\rm \sin}\omega
581: t\ +\ \gamma_{2}\ {\rm cos}\omega t)
582:  +\dfrac{{\rm \sin}2\varphi}{r^{2}}(\gamma_{3} {\rm \sin}\omega t\ 
583: +\ \gamma_{4}\ {\rm cos}\omega t)\right]
584: \label{phi_planet outside}
585: \end{eqnarray}
586: where $P_{1}^{1}=- {\rm \sin}\theta,\ P_{2}^{0}=\frac{1}{2}(3 {\rm cos}^{2}
587: \theta-1)$, and $P_{2}^{2}=3 {\rm \sin}^{2}\theta$ are the associated 
588: Legendre polynomials (our convention for $P_{1}^{1}$ has an opposite 
589: sign as that adopted by Campbell). In addition, $\phi_{planet}$ has the 
590: same time and angular dependence as $\phi_{star}$ because the field 
591: inside the planet is induced by the stellar's magnetic field. 
592: 
593: The sum $\phi_{outside}=\phi_{star}+\phi_{planet}$ is the total
594: poloidal scalar outside the planet, and
595: $\phi_{outside}$ (given by (\ref{phi_star outside}) and
596: (\ref{phi_planet outside})) is equal to $\phi_{inside}$ (given by
597: (\ref{decomposition of Phi})) at the surface of the planet
598: ($r=R_{p}$).
599: 
600: 
601: \subsection{Poloidal scalar inside the planet}
602: \label{subsection poloidal scalar inside}
603: In order to determine the poloidal scalar inside the planet, we first
604: numerically calculate the values of $G_{l}(r)$ (the radial part of
605: $\phi$, cf. equation \ref{decomposition of Phi}) and $G_{l}'(r)$
606: inside the planet by solving equation (\ref{first equation on G}). We
607: then calculate the coefficients $C_{l}^{m}$, which appears in the
608: decomposition of~$\phi$. They are determined by the boundary
609: conditions which connect the interior and exterior solutions.  In the
610: rest of this section (\S3), we assume that the conductivity profile is
611: known, and we describe the procedure used to compute the ohmic
612: dissipation rate inside the planet.  In the following sections, we
613: apply the method described in \S3 to compute the ohmic dissipation
614: rate inside the planet.
615: 
616: In the following sections, we compute the magnetic dissipation and
617: numerically obtain the value of the ohmic dissipation rate.
618: 
619: \subsubsection{Computation of G(r)}
620: If the diffusivity $\eta(r)$ is known, we can solve (\ref{first
621: equation on G}) numerically, for $l=1$ and $l=2$, with a two-point
622: boundary solver using the Newton-Raphson-Kantorovich method, and the
623: equations and boundary conditions are given below:
624: \begin{eqnarray}
625: \left\{
626: \begin{array}{l}
627: Y_{1}(r)=\mathcal{R}e(G(r))\\
628: Y_{2}(r)=\mathcal{I}m(G(r))
629: \end{array}
630: \right.
631: \left\{
632: \begin{array}{l}
633: {\rm Equations:} \\
634: Y'_{1}(r)=Y_{3}(r) \\
635: Y'_{2}(r)=Y_{4}(r) \\
636: Y'_{3}(r)=-\dfrac{\omega}{\eta(r)}Y_{2}(r)+\dfrac{l(l+1)}{r^{2}}Y_{1}(r) \\
637: Y'_{4}(r)=\dfrac{\omega}{\eta(r)}Y_{1}(r)+\dfrac{l(l+1)}{r^{2}}Y_{2}(r)
638: \end{array}
639: \right.
640: \left\{
641: \begin{array}{l}
642: {\rm Boundary} \ {\rm conditions} \ {\rm at} \ r=R_{p}\ \& \ 0 \\
643: G'_{l}(R_{p})+\dfrac{l}{R_{p}}G_{l}(R_{p})-(2l+1)R_{p}^{l}=0\\
644: G'_{l}(r\simeq 0)-\dfrac{l+1}{r}G_{l}(r\simeq 0)=0 \label{detailed equations}
645: \end{array}
646: \right.
647: \end{eqnarray}
648: 
649: \subsubsection{Computation of the C(l,m)}
650: The complex coefficients $C_{l}^{m}=\mu_{l}^{m}+i\nu_{l}^{m}$ have
651: real and imaginary parts $\mu_{l}^{m}=Re(C_{l}^{m})$ and
652: $\nu_{l}^{m}=Im(C_{l}^{m})$. We equate the real part of the
653: decomposition of the poloidal scalar inside the planet given in
654: (\ref{definition poloidal scalar}) at $r=R_{p}$ (radius of the planet)
655: with the expression of $\phi_{outside}=\phi_{star}+\phi_{planet}$
656: given in (\ref{phi_star outside}) and (\ref{phi_planet outside}) at
657: $r=R_{p}$.
658: 
659: Moreover, using the fact that ($P_{1}^{1}$,$P_{2}^{0},P_{2}^{2}$) and
660: then (${\rm cos} \omega t {\rm cos} \varphi$, ${\rm cos} \omega t {\rm \sin}
661: \varphi$, ${\rm \sin} \omega t {\rm cos}\varphi$, ${\rm \sin}\omega t
662: {\rm \sin}\varphi$) (${\rm cos} \omega t$,${\rm \sin}\omega t$),
663: (${\rm cos}\omega t {\rm cos} 2\varphi$, ${\rm cos} \omega t {\rm \sin}
664: 2\varphi$,${\rm \sin} \omega t {\rm cos} 2\varphi$, ${\rm \sin} \omega t
665: {\rm \sin} 2\varphi$) are a set of bases, we get a
666: set of linear equations which can be solved for
667: ($\mu_{1}^{1}$, $\mu_{1}^{-1}$, $\nu_{1}^{1}$, $\nu_{1}^{-1}$,
668: $\alpha_{1}$, $\alpha_{2}$,$\alpha_{3}$,$\alpha_{4}$), ($\mu_{2}^{0}$,
669: $\nu_{2}^{0}$, $\beta_{1}$,$\beta_{2}$), and ($\mu_{2}^{2}$,
670: $\mu_{2}^{-2}$, $\nu_{2}^{2}$, $\nu_{2}^{-2}$, $\gamma_{1}$,
671: $\gamma_{2}$, $\gamma_{3}$, $\gamma_{4}$) (the linear systems verified
672: by these unknowns are given in appendix B).
673: 
674: \subsection{Computation of the ohmic energy dissipation rate}
675: \label{computation of the ohmic dissipation rate}
676: The potential generates an electric field $\mathbf{E}$ which induces a
677: volumic current $\mathbf{J}$ inside the planet.  The associated ohmic
678: dissipation inside the planet is
679: $\mathcal{P}_{volumic}=Re(\mathbf{J})\ Re(\textbf{E})$.  Using
680: $\textbf{E}=\dfrac{1}{\sigma} \mathbf{J}$ and
681: $\mathbf{J}=\frac{1}{\mu_{0}} \nabla\wedge\textbf{B}$, we can write:
682: \begin{equation}
683: \mathcal{P}=\int_V \dfrac{1}{\sigma(r)} \left(\mathcal{R}e(\mathbf{J})
684: \right)^{2}dV=\int_V \dfrac{1}{\sigma(r)} \left[\mathcal{R}e
685: \left(\dfrac{\nabla\wedge\textbf{B}}{\mu_{0}}\right)\right]^{2}.  dV
686: \label{W provisoire}
687: \end{equation}
688: Moreover, using equation (\ref{rhs=lhs}), we can write
689: \begin{eqnarray}
690: \mathcal{P}=\dfrac{\omega^{2}}{\mu_{0}}\int
691: \dfrac{1}{\eta(r)}\left[\dfrac{1}{ {\rm \sin} \theta}\left(\dfrac{\partial
692: Im(\phi)}{\partial \varphi}(r, \theta,\varphi)\right)^{2} + {\rm \sin}
693: \theta\left(\dfrac{\partial Im(\phi)}{\partial \theta}(r, 
694: \theta\varphi)\right)^{2}\right]dr\ d\theta\
695: d\varphi. 
696: \label{dependence of P}
697: \end{eqnarray}
698: We use eq. (\ref{decomposition of Phi}) to express the real and
699: imaginary parts of $\Phi$.  After integrating over $\theta$ and
700: $\varphi$, we are left with:
701: \begin{equation}
702: \mathcal{P}= \int_r 
703: < \mathcal{P}_{volumic} > r^2 dr
704: \label{energy_integrand_expression}
705: \end{equation}
706: where the angle-integrated volumic power and  
707: \begin{footnotesize}
708: \begin{eqnarray}
709: < \mathcal{P}_{volumic} > =
710: \dfrac{\mu_{0} \omega^{2}}{\eta r^2 }\left\{ {\rm cos}^{2}\omega t
711: \left[\left(A_{11}^{2}+A_{12}^{2}\right)
712: +3\left(A_{17}^{2}+A_{18}^{2}\right)
713: +\dfrac{3}{\pi} A_{15}^{2}\right]
714: \right. \nonumber \\ \left.
715: +{\rm \sin}^{2}\omega t\left[\left(A_{13}^{2}
716: +A_{14}^{2}\right)+3\left(A_{19}^{2}
717: +A_{20}^{2}\right)+\dfrac{3}{\pi}A_{16}^{2}\right] \right. 
718: \nonumber \\ \left.
719: + {\sin \omega t} {\cos \omega t}
720: \left[\left(A_{12}A_{14}+
721: A_{11}A_{13}\right)+3
722: \left(A_{17}A_{19}+A_{18}A_{20}\right)+\dfrac{3}{\pi}
723: A_{15}A_{16}\right]\right\} \nonumber \label{energy integral}
724: \end{eqnarray}
725: \end{footnotesize}
726: where the expressions for $A_{ij}$ are given in Appendix C. 
727: 
728: \section{Conductivity profile and ohmic dissipation rate}
729: \label{shawfeng's conductivity}
730: The general setting of the problem and the basic equations have been
731: laid down. We have seen that once a conductivity profile is chosen,
732: one can solve (\ref{detailed equations}) and determine $G_{l}(r)$
733: inside the planet (the radial part of the poloidal scalar inside the
734: planet). Then, one can compute the $C(l,m)$, and finally obtain the
735: ohmic dissipation rate $\mathcal{P}$ inside the planet.
736: 
737: \subsection{Computation of $\mathcal{P}$ for one specific set of parameters} 
738: We compute the conductivity inside the planet with two parallel
739: approaches. In \S\S 5-7, we develop an idealized self-consistent internal
740: structure model to determine the response of the planet to the ohmic
741: dissipation of the induced current in it. But, in this section, we
742: first introduce a realistic, but non self consistent, model with the
743: following set of parameters:
744: \begin{flushleft}
745: \textbf{Planet's mass and radius:} $0.63 M_J$ 
746: \& $R_{p}=1.4\ R_{J}=10^{8}\ m$, \\
747: \textbf{Semi-major axis:} $a=0.04\ AU=6\times 10^{9}\ m$. \\
748: \end{flushleft}
749: These and other (such as mass and luminosity) stellar parameters are
750: appropriate for the short-period planet around HD209458 (Bodenheimer
751: {\it et al.} 2001).  We compute the internal conductivity due to the
752: ionization of the alkaline metals (see Appendix D for details).
753: Although the planet is heated on the day side, thermal circulation can
754: redistribute the heat and reduce the temperature gradient between
755: the two side of the planet (Burkert {\it et al.}  2005, Dobbs-Dixon \&
756: Lin 2007).  We adopt a spherically symmetric approximation for the
757: surface temperature of the planet to be 1,360K. Here, we neglect the
758: modification in the internal structure due to the ohmic dissipation
759: which will be considered with self-consistent models in the next
760: sections.  In Paper IV, we will also consider the conductivity on the
761: planet's upper atmosphere due to photoionization which only occurs on
762: the dayside of the planet.
763: 
764: Using this conductivity profile, we can approximate the magnetic
765: diffusivity $\eta(r)=1/{\mu_{0}\sigma(r)}$ by:
766: \begin{equation}
767: \eta(r) \simeq 10^{3} {\rm exp}
768: \left[25\left(\dfrac{r}{R_{p}}\right)^{2}\right]
769: \label{expression of eta}
770: \end{equation}
771: where the effects of the photoionization have been neglected in this
772: paper.
773: 
774: To apply the procedure described in \S3, we also need to specify: 
775: \begin{flushleft}
776: \textbf{Relative angular velocity:} $\omega=10^{-5}\ s^{-1}$, \\
777: \textbf{Star's magnetic dipole:} $m=4\times 10^{34}\ A\ m^{2}$, \\
778: \textbf{Value of the tilt of the magnetic dipole:} ${\rm \sin}(\alpha)=1$. \\
779: \end{flushleft}
780: We then obtain the following $\mathcal{P}$ (also see figure 2 for the
781: graphs of $G_{l}(r)$)
782: \begin{equation}
783: \mathcal{P}(t)=2.26\times 10^{21}cos^{2}\omega t+2.1\times 
784: 10^{21}\sin^{2}\omega t+1.3\times
785: 10^{21} {\sin \omega t} {\cos \omega t} \label{W}
786: \end{equation}
787: We then take the average in time over one synodic period and obtain
788: $\mathcal{P}\approx 2.18 \times 10^{21}\ Js^{-1}$ \\
789: 
790: \begin{figure}[htbp]
791: \begin{center}
792: \epsscale{1.3}
793: \plottwo{f2a}{f2b}  
794: \label{Graph of G}
795: \end{center}
796: \caption{$Re(G_{l=1})(r)$, $Im(G_{l=1})(r)$, $Re(G_{l=2})(r)$,
797: $Im(G_{l=2})(r)$, and their first derivatives, for $R_{p}=10^{8}m=1.4\
798: R_{J}$, $a=0.04\ AU$, and $\eta(r) \simeq 10^{3} {\rm exp}
799: \left[25\left(\dfrac{r}{R_{p}}\right)^{2}\right]$ The shape of $G_{l}$
800: for $l=1$ and $l=2$ are very close but the amplitudes for $l=2$ are
801: about $10^{8}=R_{p}$ higher than for $l=1$.  Indeed, the major
802: difference between $l=1$ and $l=2$ is found in the equations
803: describing the boundary conduitions (see equation \ref{detailed
804: equations} where a factor $10^{8}$ between $l=1$ and $l=2$ comes from
805: the term $R_{p}^{l}$). In addition, we found $|C(l,m)|$ for $l=1$ is
806: about $10^{10}$ times larger than for $l=2$. Therefore $|G_{l}
807: C(l,m)|$ for $l=1$ is much larger than for $l=2$, which allows us to
808: keep only the terms corresponding to $l=1$ and $l=2$ in the
809: decomposition of $\phi$ on spherical harmonics.}
810: \end{figure}
811: 
812: The conductivity profile we have obtained here is sensitive to the
813: planetary structure model.  At the epoch of planet formation, the gas
814: accretion and planetesimal bombardment history are stochastic (Zhou \&
815: Lin 2007).  The opacity in the accretion envelope of proto gas giant
816: planets may also be subjected to variations due to dust coagulation
817: (Iaroslavitz {\it et al.} 2007).  The thermal evolution of these planets
818: can be highly diverse.  There may, therefore, be a dispersion in the
819: magnitude of $\eta$.
820: 
821: \subsection{Comments on the skin depth and the dependence of the ohmic 
822: dissipation on the conductivity and on the sign of $\omega$}
823: \label{sec:skin}
824: Once the conductivity profile within the planet is determined, we are
825: able to compute the energy dissipation rate inside the planet of the
826: current induced by the star's magnetic field.  In light of the
827: possible uncertainties in the magnitude of $\eta$, we compute the
828: ohmic dissipation rate for different $\eta$ by artificially modifying
829: the above determined $\eta$ with a multiplicative factor. The
830: resulting magnitude of the time-averaged value of $\mathcal{P}$ is
831: listed below (table \ref{artificially modified eta}).
832: 
833: 
834: \begin{table}[htbp]
835: \begin{center}
836: \begin{tabular}{|l|c|}
837: \hline
838: $\eta(r)=10^{-3}\ {\rm exp} (25\ (\dfrac{r}{R_{p}})^{2})$ & $\mathcal{P}=1.26\times 10^{21}\ Js^{-1}$  \\
839: \hline
840: $\eta(r)=10^{0} {\rm exp} (25\ (\dfrac{r}{R_{p}})^{2})$ & $\mathcal{P}=2.7\times 10^{21}\ Js^{-1}$ \\
841: \hline
842: $\eta(r)=10^{3}\ {\rm exp} (25\ (\dfrac{r}{R_{p}})^{2})$ & $\mathcal{P}=2.18\times 10^{21}\ Js^{-1}$ \\
843: \hline
844: $\eta(r)=10^{5}\ {\rm exp} (25\ (\dfrac{r}{R_{p}})^{2})$ & $\mathcal{P}=1.71\times 10^{21}\ Js^{-1}$ \\
845: \hline
846: $\eta(r)=10^{7}\ {\rm exp} (25\ (\dfrac{r}{R_{p}})^{2})$ & $\mathcal{P}=1.12\times 10^{21}\ Js^{-1}$ \\
847: \hline
848: $\eta(r)=10^{9}\ {\rm exp} (25\ (\dfrac{r}{R_{p}})^{2})$ & $\mathcal{P}=2.33\times 10^{20}\ Js^{-1}$ \\
849: \hline
850: $\eta(r)=10^{10}\ {\rm exp} (25\ (\dfrac{r}{R_{p}})^{2})$ & $\mathcal{P}=2.5\times 10^{19}\ Js^{-1}$ \\
851: \hline
852: $\eta(r)=10^{12}\ {\rm exp} (25\ (\dfrac{r}{R_{p}})^{2})$ & $\mathcal{P}=2.5\times 10^{17}\ Js^{-1}$  \\
853: \hline
854: \end{tabular}
855: \end{center}
856: \caption{Table giving $\mathcal{P}$ as a function of $\eta$, where the
857: value of $\eta$ is artificially modified from the value
858: computed in section \S4}
859: \label{artificially modified eta}
860: \end{table}
861: 
862: 
863: These results indicate that the energy dissipation rate is insensitive
864: to a change in the amplitude of the conductivity by several orders of
865: magnitude (this conclusion is in agreement with a conjecture that
866: Campbell made (C83)). A high conductivity increases the energy
867: dissipation in a given volume, but it also tends to prevent the
868: magnetic field from penetrating inside the planet. On the other hand,
869: a lower conductivity corresponds to less dissipation per unit of
870: volume, but it also allows the field to penetrate deeper inside the
871: planet (and therefore increasing the volume where energy can be
872: dissipated).
873: 
874: The skin depth (for reasonable values of $\eta(r)$) is of order
875: $\delta=\sqrt{\dfrac{\eta}{\omega}}$.  For $\eta(r)=10^{3}{\rm
876: exp}\left[25\left( \dfrac{x}{R_{p}} \right)^{2} \right]$ and
877: $\omega=10^{-5}s^{-1}$, we have $\delta(r_{pn}) \approx 4\times
878: 10^{7}m$ (we define $r_{pn}$ to be the radius of penetration, or the
879: radius to which the magnetic field can diffuse inside the
880: planet). This estimate is consistent with the numerical values of
881: $G_{l}(r)$ inside the planet (see figure 2) in which
882: we find that $G_{1}(r)$ for $r< r_{pn} \simeq 6.5\times 10^{7}m$ is
883: negligibly small compared to its value elsewhere.
884: 
885: These considerations suggest that the total rate of energy dissipation
886: is well determined though the location where it occurs is less well
887: established due to the uncertainties in $\eta$. Moreover, with our
888: definition $\omega=\omega_{\star}-\Omega_{p}$, $\omega$ is positive
889: outside corrotation and negative inside corrotation. However the ohmic
890: dissipation rate inside the planet $\mathcal{P}$ only depends on the
891: absolute value of $\omega$.
892: 
893: \subsection{Energy source and direct influence on the planet's orbit.}
894: 
895: The induced current $\mathbf{J}$ deduced in
896: the previous section is due to the diffusion of a time dependent
897: magnetic field.  This time dependence comes from the relative motion
898: of the planet's orbit and the stellar magnetosphere. Thus, the ohmic
899: dissipation must be supplied by the orbital kinetic energy of the
900: planet and the rotational energy of the star.  Our stated goal in the
901: introduction is to consider whether the migration of some planets may
902: be halted by their magnetic coupling with their rapidly spinning
903: magnetized host stars.  In the case where $\omega_\ast > \Omega_p$,
904: the rotational energy of the star is transferred to the total orbital
905: energy of the planet and provides a supply for the ohmic
906: dissipation. The torque T associated with the ohmic dissipation  
907: is linked with the ohmic dissipation rate $\mathcal{P}$ and the relative angular velocity $\omega$ 
908: according to the following equation (cf. C83, eq. 55),
909: \begin{equation}
910: \mathcal{P}=-\omega |T|.     
911: \label{torque}
912: \end{equation}
913: 
914: Since the transfer of angular momentum requires the
915: torque associated with the ohmic dissipation, a similar fraction of
916: energy is being transfered to the planet's orbit and supplied to the
917: ohmic dissipation.  For this purpose, we qualitatively compare the
918: magnitude of $\mathcal{P}(t)$ with the rate of energy change needed
919: stall the migration of a protoplanet. A detailed computation on the
920: orbital evolution of the planet will be presented in Paper III.
921: 
922: For illustration purpose, we first consider the power associated with
923: the migration ($P_{mig}$) of a planet with a 0.63 Jupiter mass and a
924: 1.4 Jupiter radius toward a sun-like star.  At any semi major axis
925: $a$, the total energy of the Keplerian orbit is $|E| = G M_p M_s/2a$.
926: If its orbit decays on a characteristic planet-disk interaction time
927: scale ($\tau_{\rm mig}$) of about 3 million years, the torque needed
928: to halt the planet's migration would correspond to a power $P_{mig}$
929: such that
930: \begin{center}
931: $P_{mig} = |\dot E| \simeq \dfrac{\mathcal{G}
932: M_{p} M_{\star}}{2\ a\ \tau} \simeq 7.4\ \times 10^{22} Js^{-1} $.
933: \label{migration kinetic power}
934: \end{center}
935: Since this power is more than an order of magnitude larger than the
936: time average value of $\mathcal{P}$ (see Table 1), it seems,
937: therefore, not possible for the magnetic coupling to directly stall
938: the planet's migration at a 0.04 AU Keplerian orbit within a few
939: millions years, even in the limit of a positive $\omega$.  
940: 
941: However in the model we have considered here, $\mathcal{P}\propto B^2
942: \propto m^2 a^{-6}$.  The power needed to drive the planet to
943: migration with a specified speed is proportional to $a^{-2}$ (these
944: scalings are confirmed by numerical calculations that neglect any
945: changes in the relative frequencies $\omega$ and the planetary
946: internal structure). It means that there is a semi-major axis
947: $a_{stop} (\sim 0.01$AU) at which $\mathcal{P}$ and $P_{mig}$ are
948: comparable.  This distance is comparable to the radius of a typical T
949: Tauri stars.  Note that the requirement for $\omega_\ast > \Omega_p$
950: also implies that the planet must be outside the corotation
951: radius. This condition is satisfied only in a disk with a low gas
952: accretion rate around a rapidly spinning and weakly magnetized
953: stars. In paper II, we will consider such a model for the newly
954: discovered planet around TW Hyd (Setiawan {\it et al.} 2008).  Under
955: these circumstances, the planet-star magnetic interaction may also
956: be overwhelmed by their tidal interaction.  
957: 
958: \section{Planetary inflation and mass loss}
959: In this section, we propose that ohmic dissipation in the planet's
960: interior can indirectly halt its migration.  The main physical
961: mechanisms involve the heating of the planet's interior, its inflation
962: and mass loss through Roche lobe overflow, and angular momentum
963: transfer from the transferred material to the orbit of the planet.
964: 
965: Up to now, we have computed the planet's conductivity for one
966: particular set of parameters ($M_{p}$, $a$, etc.), and the corresponding ohmic energy
967: dissipation inside the planet due to the star's magnetic field.
968: Although, this dissipation rate for most close-in planets is generally
969: too small to directly provide the power needed to halt their migration
970: over the time scale of a few Myr, it can modify their internal
971: structure.
972: 
973: The ohmic dissipation is likely to increase the temperature, the ionization
974: fraction, and the conductivity around the region where most of the
975: dissipation occurs.  In principle, the extra energy source would
976: reduce the skin depth. However, the envelope of the young planet is
977: likely to be fully convective, similar to the low-mass main sequence
978: secondary in interacting binaries.  Campbell (C83) suggested that the
979: dominant diffusivity may be due to turbulence (Cowling 1981).  In
980: \S\ref{sec:skin}, we have already indicated that even though the skin
981: depth may be affected by the magnitude of the diffusivity, the total
982: energy dissipation rate in the planet's interior is not sensitively
983: determined by the profile of $\eta$.
984: 
985: Nevertheless, the heat released by the dissipation is comparable to
986: that associated with the Kelvin-Helmholtz contraction during the early
987: stage of the planet's evolution (Bodenheimer, Lin, Mardling 2001 (BLM)).  In
988: the proximity of its host star, this extra energy source may cause a
989: planet to inflat beyond its Hill's radius and lose mass (Gu {\it et
990: al.} 2004).  
991: 
992: In the following sections (\S\S5-7), we adopt an idealized and
993: self-consistent model of the planet's internal structure.  This
994: approach allows us to compute the conductivity of the planet for
995: different sets of parameters. Considering the low dependence 
996: of the total ohmic dissipation $\mathcal{P}$ on $\eta$, an idealized but versatile
997: prescription is adequate for the computation of $\mathcal{P}$ and the
998: mass loss rate ($\dot M$) for different values of the important
999: parameters (the planet's mass and radius, the star's mass, luminosity,
1000: and dipolar magnetic field strength, the tilt between the magnetic 
1001: dipole and the stellar spinning axis, and the relative orbital period).  In
1002: \S5, we show how the mass loss rate $\stackrel{\centerdot}{M}$ is
1003: related to the ohmic dissipation $\mathcal{P}$.  In \S6, we describe the
1004: model we used for the planet's interior and in \S7 we calculate $\mathcal{P}$
1005: and $\stackrel{\centerdot}{M}$ for different sets of parameters.
1006: 
1007: \subsection{A qualitative description}
1008: The planet receives energy, at its surface, from the star's radiation
1009: and, in the interior, from the ohmic dissipation. The surface heating
1010: diffuses inwards until an isothermal structure is established in the
1011: planet's outer envelope. But, well below the surface region, the heat
1012: flux is generated by the planet's Kelvin-Helmholtz contraction and
1013: ohmic dissipation and transported by convection.  In the limit that
1014: convection is efficient, the envelope attains a constant entropy
1015: profile. For computational simplicity, we adopt an isothermal model
1016: near the surface of the planet and a polytrope model for its deep
1017: interior. 
1018: 
1019: There are two regions of interest. Very close to the host star, the
1020: ohmic dissipation rate is larger than that ($L_p= L_\ast (R_p/2a)^2$)
1021: due to the stellar irradiation ($L_\ast$) received by the planet. In
1022: this limit, the planet would rapidly expand beyond its Roche lobe and
1023: become tidally disrupted. In accordance with the results of the
1024: previous section, (in which the effect of $\mathcal{P}$ on the
1025: internal structure of the planet has been neglected), $\mathcal{P}
1026: \propto a^{-6}$ and $L_p \propto a^{-2}$.  Thus, the stellar heating
1027: dominates at larger semi major axis. In this section, we consider the
1028: effect of planet's inflation due to the ohmic dissipation and show
1029: that $\mathcal{P}$ also increases with the planetary radius $R_p$ at
1030: nearly the same rate as $L_p (\propto R_p^2)$. Thus, during the
1031: thermal expansion of the planet, the ratio of $L_p/\mathcal{P}$ does
1032: not change.  In the region where $L_p > \mathcal{P}$, the effective
1033: temperature at the planet's surface, with or without the contribution
1034: from the ohmic dissipation remains to be the equilibrium value $T_p$.
1035: But, the planet's radius for thermal equilibrium increases with
1036: $\mathcal{P}$ which adds to the energy generation in the planet's
1037: interior (BLM). If the new equilibrium $R_p$ is larger than the
1038: planet's Roche radius, $R_H$, mass would be lost gradually through
1039: Roche overflow.
1040: 
1041: \subsection{Mass loss rate} 
1042: \label{mass loss rate section}
1043: We now derive the equations that will allow us to calculate the mass
1044: loss rate $\dot M$ and angular momentum transfer rate as functions of
1045: $\mathcal{P}$. We are in the second region where the ohmic dissipation is less
1046: than the radiation flux from the star ($\mathcal{P} \leqslant  L_{p}$), and we set the Bond albedo to zero. 
1047: We therefore assume that the
1048: equilibrium temperature at the surface of the planet is fixed by the
1049: radiation from the star $T_{p}^{4}=\frac{L_{star}}{16\pi\ \sigma_{r}\
1050: a^{2}}$ ($L_{star}$ the total luminosity of the star, $\sigma_{r}~=~
1051: 5.67\times 10^{-8}~Js^{-1}m^{-2}T^{-4}$), and that the ohmic dissipation
1052: provides the additional energy to inflate the planet.
1053: 
1054: An irradiated short-period planet establishes an isothermal surface
1055: layer. The hot interior continues to transport heat to this region and
1056: then radiates to infinity with a luminosity $L_i$ despite the surface
1057: heating. Note that
1058: \begin{equation}
1059: L_i < <  L_p = 4 \pi \sigma_{r} T_p^4 R_p^2
1060: \end{equation}
1061: so that the modification to $T_p$ is negligible.  The magnitude of
1062: $L_i$ is a function of $R_p$, $M_p$, $T_p$, and the existence of the
1063: core.  We have previous computed an equilibrium model for the
1064: parameters for several short-period planets (BLM). 
1065: In the range [$10^{-8} L_\odot 10^{-5} L_\odot$],
1066: the numerical results of BLM can be approximated by 
1067: \begin{equation}
1068: {\rm log} {R_p \over R_\odot}  = A (M_p)
1069: + B(M_p) {\rm log} {L_i \over L_\odot} 
1070: + C(M_p) \left( {\rm log} {L_i \over L_\odot} \right)^2.
1071: \label{eq:req}
1072: \end{equation}
1073: For HD209458b (the $0.63 M_J$
1074: model we presented in the previous section), (A, B, C)= (3.11, 1.01,
1075: 0.0642).  BLM also determined the value of these coefficients for more
1076: massive planets around a solar type star (they are modified by the
1077: stellar irradiation so that they are also function of $M_\star$).  The
1078: planet's radius $R_p$ would contract unless there is an adequate energy
1079: source to replenish its loss of internal energy. If the ohmic
1080: dissipation can provide such a source, $R_p=R_e$ and $L_i = \mathcal{P}$ in a
1081: thermal equilibrium.
1082: 
1083: At a=0.04 AU, the Roche radius of the planet is 
1084: \begin{equation}
1085: R_H=a\left(\dfrac{M_{p}}{3M_{\star}}\right)^{\frac{1}{3}}.
1086: \label{definition hill's radius}
1087: \end{equation}
1088: From equation(\ref{eq:req}), we find that the equilibrium $R_e \sim
1089: R_H$ if $L_i \sim 10^{-5} L_\odot$ which is approximately the value of
1090: $\mathcal{P} (\sim 10^{21}$ J s$^{-1})$ we have determined for HD209458b.  During
1091: the planet and star's infancy, this planet would inflate to fill its
1092: Roche lobe when it has migrated to this location.
1093: 
1094: Outside $\sim 0.04$ AU, $\mathcal{P}$ decreases rapidly with
1095: $a$. Consequently, $R_p$ reduces to the value which is essentially not
1096: modified by the ohmic heating. For the calculation presented in the
1097: previous section, we neglected the inflation of the planet. In the
1098: next section, we construct a self consistent model taking into account
1099: of the modification of the dissipation rate due to the internal
1100: structural changes.  For $a=0.04$ AU, the intense ohmic dissipation
1101: rate (with $L_i = \mathcal{P} \simeq 2 \times 10^{21}$J s$^{-1}$)
1102: modifies the planet's internal structure and inflates its radius to
1103: $R_e \sim 0.5 R_\odot$.  The inflation is more severe at $a < 0.04$ AU
1104: because $\mathcal{P}$ is a rapidly decreasing function of $a$.  If
1105: $R_e >R_H$ at this location, the planet would overflow its Roche lobe
1106: and loss mass.  For the rest of this paper, we assume that the we are
1107: in the case where the planet fills its Roche lobe, \textit{i.e.}
1108: $R_{p}=R_{H}$
1109: 
1110: Two remarks are appropriate here. Firstly, in order for the Roche lobe
1111: overflow to provide angular momentum, the actual shape of the Roche
1112: lobe should be taken into account. However, for computational
1113: simplicity, we adopt in this paper a spherically symmetric
1114: approximation (refer to Gu, Bodenheimer, Lin 2003 for a detailed study
1115: (GBL)). Secondly, we have only considered the contribution of ohmic
1116: dissipation $\mathcal{P}$ to the planetary inflation. The tidal
1117: (gravitational) interaction between the star and the planet can also
1118: significantly enhance the planet's inflation in some cases. More
1119: precisely, this tidal interaction can be strong for small semi-major
1120: axis (the tidal effect varies as $a^{-13/2}$) and for large radius
1121: (thus, the more inflated the planet, the stronger this effect
1122: becomes).
1123: 
1124: \subsection{The governing equations}
1125: Mass loss process is initiated when $R_{e} \geq R_{H}$. 
1126: In this limit, a contiuous flow would be
1127: established in which the inflation of the envelope's drives a steady
1128: supply of gas to the Roche lobe region.  Well inside the Roche lobe,
1129: the gravitational potential is primarily determined by the mass of the
1130: planet $M_{p}$: $\phi_{g}=-\frac{GM_{p}}{r}$, but near $R_H$, we need
1131: to take into account of both the planet and the star.  In a frame which 
1132: corotates with the planet, the gravitational potential U(r) 
1133: \begin{equation}
1134: U(r)=\dfrac{-GM_{\star}}{a}\left[\left(1-\dfrac{M_{p}}{M_{\star}}\right)\ 
1135: \left(\dfrac{a}{a-r}+\dfrac{(a-r)^{2}}{2a^{2}}\right)+\dfrac{M_{p}}
1136: {M_{\star}}\left(\dfrac{a}{r}+\dfrac{r^{2}}{2a^{2}}\right)\right]
1137: \label{expression of U}
1138: \end{equation}
1139: where $r$ is the distance to the center of the planet. The value of
1140: $R_H$ is determined from ${dU}/{dr}(R_H)=0$.
1141: 
1142: In principle, this potential introduces a complex multi-dimensional
1143: flow pattern, especially near the Roche lobe.  But the expansion of
1144: the envelope originates deep in the envelope where the ohmic
1145: dissipation occurs.  In this region, spherical symmetry is adequate.
1146: Near the Roche lobe, we adopt the results obtained by GBL. 
1147: For computational convenience, we neglect the planet's spin.
1148: 
1149: We consider a low-velocity quasi-hydrostatic expansion of the envelope.
1150: Under this gravitational potential in eq. (\ref{expression of U}), the
1151: radial component of the hydrodynamics momentum equation for a volume
1152: of gas is reduced to:
1153: \begin{equation}
1154: \upsilon\dfrac{d\upsilon}{dr}(r)+\dfrac{1}{\rho(r)}\dfrac{dP}{dr}(r)
1155: =-\dfrac{1}{\rho(r)}\dfrac{dU}{dr}(r)
1156: \label{momentum 1}
1157: \end{equation}
1158: where $v$ is the radial velocity.  The radial component of the
1159: equation of mass conservation: $\nabla (\rho (r) \upsilon(r))=0$ gives
1160: $r^{2}\rho (r)\upsilon(r)=Constant$, and then the mass loss rate
1161: $\stackrel{\centerdot}{M}$ is constant:
1162: \begin{equation}
1163: \stackrel{\centerdot}{M}=4\pi\ r^{2}\rho (r)\upsilon(r)=Constant
1164: \label{definition mass loss rate}
1165: \end{equation}
1166: or equivalently,
1167: $\dfrac{1}{\rho}\dfrac{d\rho}{dr}=-\dfrac{1}{r^{2}\upsilon}
1168: \dfrac{dr^{2}\upsilon}{dr}$. Then using $\dfrac{dP}{dr} =\dfrac{dP}
1169: {d\rho}\dfrac{d\rho}{dr} =\dfrac{d\rho}{dr}c_{s}^{2}$ ($c_{s}^{2}(r)$
1170: is the sound speed), the momentum equation becomes:
1171: \begin{eqnarray}
1172: \left(1-\dfrac{c_{s}^{2}}{\upsilon^{2}}\right)\dfrac{d}{dr}
1173: \left(\dfrac{\upsilon^{2}}{2}\right)&=&-\dfrac{1}{\rho}
1174: \dfrac{dU}{dr}\left(1-\dfrac{2c_{s}^{2}\rho}{r}
1175: \dfrac{1}{\frac{dU}{dr}}\right) .
1176: \label{momentum equation}
1177: \end{eqnarray}
1178: At a (sonic) radius $r_{2}$ near the inner Langragian point, the flow
1179: velocity becomes comparable to the sound speed (GBL) {\i.e.}
1180: \begin{equation}
1181: \upsilon(r_{2})=c_{s}(r_{2}).
1182: \end{equation}
1183: where the magnitude of $r_{2}$ is the largest solution of
1184: $(r^{2}-rR_{H}+\frac{2c_{s}^{2}a^{3}}{LGM_{\star}})=0$, with
1185: \begin{equation} 
1186: L=\left(1-\dfrac{M_{p}}{M_{\star}}\right)\ \left(\dfrac{2a^{3}}{(a-R_{H})^{3}}+1\right) 
1187: + \dfrac{M_{p}}{M_{\star}}\left(2\left(\dfrac{a}{R_{H}}\right)^{3}+1\right)\simeq 9 \label{l}.
1188: \end{equation}
1189: 
1190: The expansion rate is determined by the rate of ohmic energy
1191: dissipation within the planet. In a steady state, the energy equation
1192: reduces to
1193: \begin{equation}
1194: \dfrac{1}{r^{2}}\dfrac{d}{dr}\left[r^{2}\rho(r)\upsilon(r)\left(
1195: \dfrac{\upsilon^{2}}{2}+h(r)+\phi_{g}(r)\right)\right]=\mathcal{P}_{vol}
1196: \label{energy equation}
1197: \end{equation}
1198: where $\mathcal{P}_{vol}$ is the volumic ohmic energy dissipation, and
1199: $\phi_{g}$ the gravitational potential (of the planet only or of both
1200: the planet and the star, depending on the location).  In this
1201: approximation, we assume that the distribution of enthalpy $h$ is
1202: determined by both efficient convective transport (in term of an
1203: abiabat) and radiative diffusion inside the planet.
1204: 
1205: Using equation (\ref{definition mass loss rate}), we replace 
1206: $r^{2}\rho(r)\upsilon(r)$ by $\dfrac{\stackrel{\centerdot}{M}}{4\pi}$ 
1207: in eq. \ref{energy equation}. We then can then integrate
1208: equation (\ref{energy equation}) between the radius $r_{pn}$
1209: (the radius at which the field can no longer penetrate into the
1210: planet) and $r_{2}$ so that
1211: \begin{equation}
1212: \stackrel{\centerdot}{M}\left(\dfrac{\upsilon^{2}}{2}(r_{2})-
1213: \dfrac{\upsilon^{2}}{2}(r_{pn})+h(r_{2})-h(r_{pn})
1214: +\phi_{g}(r_{2})-\phi_{g}(r_{pn})\right)=\int
1215: 4\pi r^{2}\mathcal{P}_{vol}(r)dr=\mathcal{P}.
1216: \label{eq:eofw}
1217: \end{equation}
1218: Within an order of magnitude, $\bigtriangleup(h)\approx-\dfrac{1}{3}
1219: \bigtriangleup(\phi_{g})$ and $\bigtriangleup(\dfrac{1}{2}v^{2})
1220: \approx\dfrac{1}{10} \bigtriangleup(\phi_{g})$ (this comes from calculating the order of magnitude of these 3 terms using an 
1221: order of magnitude for the temperature, for the sound speed, and for $r_{pn}$). In addition, the
1222: integrated energy equation is the result of an approximation as the
1223: total ohmic dissipation rate $\mathcal{P}$ should be the integral of
1224: $\mathcal{P}_{vol}$ between $r_{pn}$, and $R_{p}$ ($=R_{H}$ because we
1225: assumed that the planet fills its Roche lobe).  However, this
1226: approximation is reasonable since $h(R_{H}) =h(r_{2})$ (the surface
1227: region is approximately isothermal), $\upsilon^{2}(r_{2})\approx
1228: \upsilon^{2}(R_{H})$, and $\phi_{g}(r_{2})\approx \phi_{g}(R_{H}$
1229: because $r_{2}$ is very close to $R_{H}$).
1230: 
1231: We now can calculate $\rho(r_{2})$ and $P(r_{2})$. Equation
1232: (\ref{definition mass loss rate}) for $r=r_{2}$ with
1233: $\upsilon(r_{2})=c_{s}(r_{2})=10^{4}\sqrt{\dfrac{T}{10^{4}}}$ meters
1234: gives:
1235: \begin{eqnarray}
1236: \left\{
1237: \begin{array}{l}
1238: \rho(r_{2})=\dfrac{\stackrel{\centerdot}{M}}{4\pi r_{2}^{2}c_{s}(r_{2})}\\
1239: P(r_{2})=\alpha\ \rho(r_{2}) T(r_{2})\\
1240: \alpha=\dfrac{\mathcal{N}_{a}k_{B}}{\mu\mathcal{M}_{H}}\
1241: \label{values of P and volumic mass at r2}
1242: \end{array}
1243: \right.
1244: \label{P(r2)}
1245: \end{eqnarray}
1246: where $\mathcal{N}_{a}$ is the Avogadro constant, $k_{B}$ the
1247: Boltzmann constant, $\mathcal{M}_{H}$ is the hydrogen molar mass, 
1248: $\mu$ a coefficient which depend on the ionisation
1249: rate ($\mu=1$ for hydrogen atoms, and
1250: $\mu=0.5$ for fully ionized hydrogen gas, and we usually choose $\mu$ close to unity).
1251: 
1252: \section{Isothermal and polytropic model} 
1253: \label{isothermal and polytropic model}
1254: 
1255: In \S4, the permeation and dissipation of the time-dependent external
1256: field is analyzed by neglecting any resulting changes in the planet's
1257: interior.  In \S5, we show that the resulting ohmic dissipation can
1258: substantially modify the temperature and density distribution within
1259: the planet.  Increases in the ionization rate modify the skin depth
1260: and relocate the region of maximum ohmic dissipation. However, the
1261: expansion of the planet's envelope does not affect the rate of ohmic
1262: dissipation.  In this section, we present a set of approximately
1263: self-consistent calculations to analyze the feedback effect of ohmic
1264: dissipation on the planet's internal structure.
1265: 
1266: \subsection{A Roche-lobe filling structural Model}
1267: In principle, the structure of the planet should be solved numerically
1268: with the standard planetary structure equations (BLM).  However, a
1269: semi analytic model based on simplifying assumption may provide
1270: insight on the inter-dependent relation between various physical
1271: parameters.  Based on the BLM's numerical models, we approximate the
1272: internal structure of the planet with an idealized model in which the
1273: outer region is isothermal (due to the stellar irradiation) and the
1274: inner region is polytropic (due to an efficient mix of entropy by
1275: thermal convection).  In the computation of $\eta$, we only take into
1276: account the ionization of the hydrogen because the internal
1277: temperature distribution is mostly determined by heat transfer rather
1278: than heat dissipation and the rate of $\mathcal{P}$ is a relatively
1279: insensitive function of $\eta$. The advantage of this approximation is
1280: that its application for the self consistent analysis is relatively
1281: straightforward.
1282: 
1283: \begin{description}
1284: \item[The isothermal region:] it extends from the surface to a
1285: transition radius $r_{=}$ which is to be determined self consistently
1286: in \S \ref{subsection transition between the two models}.  In this
1287: region, the equation of state and the equation describing the
1288: hydrostatic equilibrium are:
1289: \begin{eqnarray}
1290: T(r)&=& {\rm Constant}, 
1291: \label{isothermal T} 
1292: \\ P(r)&=&\alpha\ \rho (r)\ T(r), \\ 
1293: \label{equation of state}
1294: \dfrac{dP}{dr}(r)&=&-\dfrac{GM_{int}(r)}{r^{2}}\rho (r), \
1295: \label{hydrostatic equilibrium}
1296: \end{eqnarray}
1297: where $\alpha=\dfrac{\mathcal{N}_{a}k_{B}}{\mu\mathcal{M}_{H}}$ and
1298: $M_{int}(r)$ is the planet's mass inside a sphere of radius r centered
1299: on the planet's center. For all practical purpose, $\rho$ is
1300: sufficiently low in the isothermal region that we can approximate
1301: $M_{int}(r) \simeq M_{planet}$ (one can verify, a posteriori, that the
1302: neglected mass is less than a few percent of the total mass). To
1303: calculate P(r), we integrate (\ref{hydrostatic equilibrium}) using
1304: (\ref{isothermal T}) and (\ref{equation of state}). We then can
1305: calculate $\rho(r)$ using (\ref{equation of state}):
1306: \begin{eqnarray}
1307: \left\{
1308: \begin{array}{l}
1309: P(r)=C\ {\rm exp} \left(\dfrac{GM_{planet}}{r}\ \dfrac{1}{\alpha
1310: T}\right) \\ 
1311: \label{P isothermal model} 
1312: \rho(r)=\dfrac{1}{\alpha T}\ C\ {\rm exp}\left(\dfrac{GM_{planet}}{r}\
1313: \dfrac{1}{\alpha T}\right)
1314: \label{volumic mass isothermal model}
1315: \end{array}
1316: \right.
1317: \label{P in isothermal region}
1318: \end{eqnarray}
1319: where $C$ is an integration constant, which value is obtained 
1320: by injecting $r_{2}$ in the previous equations. 
1321: 
1322: \item[Polytrope region:] it extends from the center of the planet to
1323: $r_{=}$.  In this region, we use the following equations:  
1324: \begin{eqnarray}
1325: P(r)&=&K\rho^{\gamma}(r), \label{polytrope equation of state} \\
1326: \dfrac{d\phi_{g}}{dr}(r)&=&\dfrac{G\ M_{int}(r)}{r^{2}}, \label{definition gravitational potential}\\
1327: \dfrac{dP}{dr}(r)&=&-\dfrac{GM_{int}(r)}{r^{2}}\rho (r) \label{polytrope hydrostatic equilibrium}\\
1328: \bigtriangleup\phi_{g}(r)&=&4\pi\ G\ \rho(r) \label{poisson equation}
1329: \end{eqnarray}
1330: where $\phi_{g}$ is the gravitational potential, and $\bigtriangleup$ 
1331: is the laplacian (in the Poisson equation). 
1332: Using equations (\ref{polytrope equation of state}) and (\ref{definition
1333: gravitational potential}), equation (\ref{polytrope hydrostatic equilibrium})
1334: becomes: $K\gamma \rho^{\gamma
1335: -1}(r)\dfrac{d\rho}{dr}(r)=-\rho(r) \dfrac{d\phi_{g}}{dr}$\\ And
1336: after integration:\ \ \ \ \ \ $\phi_{g}(r)=Constant\ -\
1337: \dfrac{K\gamma}{\gamma-1}\rho^{\gamma-1}(r)$\\ We then replace
1338: $\phi_{g}$ in the poisson equation (\ref{poisson equation}):
1339: \begin{equation}
1340: \bigtriangleup \rho^{\gamma-1}(r)=-\frac{\gamma -1}{K\ \gamma}\ 
1341: 4\pi G\rho(r)
1342: \end{equation}
1343: For the condition appropriate in the interior of planets, the equation
1344: of state is reasonably approximated by a $\gamma=2$ polytrope (de
1345: Pater \& Lissauer 2001). In spherical coordinates, the previous
1346: equation becomes:
1347: \begin{equation}
1348: \dfrac{1}{r^{2}}\dfrac{d}{dr}\left(r^{2}\dfrac{d}{dr}\rho(r)\right)
1349: =-\dfrac{2\pi G}{K}\rho(r) .
1350: \label{final poisson equation}
1351: \end{equation}
1352: This equation has an analytical solution (Ogilvie \& Lin
1353: 2004), and we can calculate $\rho(r)$, P(r) and T(r) in the region
1354: described by the polytropic equation of state:
1355: \begin{eqnarray}
1356: \left\{
1357: \begin{array}{l}
1358: \rho(r)=\rho_{0}\dfrac{\sin kr}{kr} \\ \label{volumic mass
1359: polytropic model}
1360: P(r)=K\rho^{2}(r)=K\rho_{0}^{2}\left(\dfrac{\sin kr}{kr}\right)^{2} \\
1361: \label{P polytropic model} 
1362: T(r)=\dfrac{P}{\alpha \rho}(r)=\dfrac{1}{\alpha} K\rho_{0}\dfrac{\sin kr}{kr}
1363: \\ \label{T polytropic model} k=\sqrt{\dfrac{2\pi G}{K}}.
1364: \label{relation k and K}
1365: \end{array}
1366: \right.
1367: \end{eqnarray}
1368: \end{description}
1369: 
1370: \subsection{Transition between the two models}
1371: \label{subsection transition between the two models}
1372: In principle, the transition between the two regions is determined by
1373: the onset of convection.  In the construction of hydrostatic
1374: equilibrium structure models (to be presented in Paper II), we will
1375: indeed use that condition to determine its photospheric radius.
1376: Qualitatively, we expect the transition radius which separates the two
1377: regions, $r_=$ to be larger than $r_{pn}$, because only in the
1378: region interior to $r_=$ do we expect the temperature, ionization
1379: fraction, and conductivity to be sufficiently large to halt the
1380: penetration of the field.  In a hydrostatic equilibrium, the actual
1381: value of $r_=$ is determined by the ratio of the ohmic dissipation
1382: rate in the convective region to the sum of the ohmic dissipation rate
1383: in the entire planet's interior and the stellar irradiative flux on
1384: the planet's surface.  A set of fully self-consistent solution
1385: requires the matching of the ohmic dissipation rate to be expected
1386: from the planetary structure and that which determines its density and
1387: temperature distribution (see paper II).
1388: 
1389: In the present context, we are considering the situation in which the
1390: planet's radius is constrained by its Roche lobe and the density and
1391: temperature of the outer boundary is determined by 
1392: equation(\ref{P(r2)}). In this configuration, heat is also transported
1393: by advection which modifies the location of $r_=$.  Moreover, the
1394: density ratio between the planet's center and the outer boundary is
1395: much larger than the temperature ratio.  Therefore, the polytropic
1396: region cannot fill the entire interior region.  Since the pressure
1397: scale height on the planet's surface is much smaller than its radius,
1398: the isothermal region also cannot occur in the entire planet's interior
1399: while containing all of its mass. Instead, the planet's interior
1400: adjusts to attain a balance between the requirement of mass loading
1401: and constraints set by hydrostatic equilibrium for appropriate
1402: equations of state.  
1403:  
1404: In order to construct such an equilibrium model, we now determine
1405: $\rho_{0}$ and k at $r_{=}$ where the transition between the two
1406: regions occur. There are three equations that constrain these parameters:
1407: $T_{isothermal}(r_{=}) =T_{polytropic}(r_{=})$, $P_{isothermal}(r_{=})
1408: =P_{polytropic}(r_{=})$, and the total mass is constant. The first two
1409: conditions also imply
1410: $\rho_{isothermal}(r_{=})=\rho_{polytropic}(r_{=})$.  Therefore, we
1411: solve the following equations for $\rho_{0}$, k, and $r_{=}$:
1412: \begin{eqnarray}
1413: \left\{
1414: \begin{array}{l}
1415: k^{2}=\dfrac{2\pi G\ C}{(\alpha T)^{2}} {\rm exp}\left(
1416: \dfrac{GM_{\rm planet}}{\alpha T\ r}\right)\\
1417: \rho_{0}=\dfrac{\alpha T}{2\pi G}\ k^{2}\
1418: \dfrac{kr_{=}}{\sin kr_{=}} \\ \int_0^a 4\pi
1419: r^{2}\rho_{\rm polytropic}(r)dr\ +\ \int_a^R 4\pi
1420: r^{2}\rho_{\rm isothermal}(r)dr=M_{p}.
1421: \end{array}
1422: \right.
1423: \label{3 unknowns}
1424: \end{eqnarray}
1425: 
1426: By assuming an isothermal structure in the outer envelope,
1427: we have neglected an outward heat flux.  This approximation is
1428: only adequate if the dissipation rate is above that which is
1429: need to inflat $R_{p}$ to the planet's Roche radius.  If this
1430: condition is not satisfied, the planet's radius would attain
1431: equilibrium values for which the surface cooling is balanced by the
1432: Ohmic dissipation and stellar irradiation. We will construct, in Paper II,
1433: the equivalent of equation 21 (for a 0.63 $M_{\odot}$ planet) which takes into
1434: account the effect of ohmic dissipation in the planetary interior.
1435: 
1436: Whereas the temperature on the planet's surface is determined by the
1437: stellar irradiation, the density at $r_2=R_H$ is determined by the
1438: magnitude of $\dot M$ (through equation \ref{P(r2)}) which in term is
1439: determined by the rate of ohmic energy dissipation $\mathcal{P}$ (see \S
1440: \ref{sec:selfcon}).
1441: 
1442: For very large values of $\mathcal{P}$, a set of fully self-consistent solutions
1443: also modifies the temperature at the disk surface as well as the
1444: thermal content of the outflowing gas. However, provided $\mathcal{P}$ is small
1445: compared with the stellar irradiative flux, a transition for
1446: convective stability occurs near $r_=$.  
1447: 
1448: \subsection{Calculation of the magnetic diffusivity}
1449: With these internal structure specified, we consider Saha's equation
1450: for the hydrogen atoms which gives the ionization
1451: fraction $x$ (Kippenhahn \& Weigertal, pages 107-111):
1452: \begin{equation}
1453: \dfrac{x^{2}}{1-x^{2}}=K_{H}=\dfrac{1}{P(r)}\dfrac{(2\pi m_{e})^{
1454: \frac{3}{2}}}{h^{3}}(kT)^{\frac{5}{2}}exp\left(-\frac{E}{kT}\right) \ \ \ \ \ \ \ 
1455:  \label{kh}
1456: \end{equation}
1457: where the ionization energy of hydrogen is $E=13.6eV$. We also
1458: neglect here the radiation pressure as we write $P_{gas}(r)=P(r)$.
1459: 
1460: If the ionization fraction $x$ is small, $x^{2}\approx K_{H}$ (this is typically the case 
1461: in the region where the ohmic dissipation occurs).  We use
1462: $\sigma=\dfrac{N_{e}e^{2}}{m_{e}\nu_{e}}$, with 
1463: $\nu_{e}=N_{n}10^{-19}(\dfrac{128kT}{9\pi m_{e}})^{\dfrac{1}{2}}$.  
1464: 
1465: The electric conductivity we would obtain 
1466: does not take into account higher ionization states or the ionization 
1467: of elements other than hydrogen atoms. We then use for the following 
1468: calculations an electric conductivity that is 10 times higher than 
1469: that we would obtain with the Saha equation (eq. (\ref{kh})) for the hydrogen atom only. 
1470: We saw in table 1 that the ohmic dissipation rate was quite insensitive to the 
1471: magnetic diffusivity $\eta(r)=(\mu_{0}\sigma(r))$, and we verified that this is also the case with 
1472: the internal model we used for the planet in sections 5 to 8  
1473: (for example, in this model, a uniform change in the magnetic diffusivity by a factor 10 changes 
1474: $\mathcal{P}$ and $\stackrel{\centerdot}{M}$ by less than 20\%, and  
1475: a uniform change in the magnetic diffusivity by a factor 100 changes 
1476: $\mathcal{P}$ and $\stackrel{\centerdot}{M}$ by less than 40\%).
1477: 
1478: We then obtain the following expression for the magnetic diffusivity inside
1479: the planet: 
1480: \begin{equation}
1481: \eta(r)=1.28\times 10^{-2}\frac{\sqrt{P(r)}}{T^{\frac{3}{4}}(r)}exp\left(\frac{78909}{T}\right). \label{eta}
1482: \end{equation}
1483: where T(r) and P(r) are the temperature and pressure of
1484: the isothermal or polytropic region, depending on the radius r.
1485: 
1486: \section{Ohmic dissipation rate and the mass loss rate for 
1487: different sets of parameters} 
1488: \label{sec:selfcon}
1489: With the above idealized prescription for the planet's internal
1490: structure, we now calcultate self-consistently the ohmic dissipation
1491: rate $\mathcal{P}$ inside the planet as well as the mass loss rate
1492: $\stackrel{\centerdot}{M}$.
1493: 
1494: \subsection{Parameters involved in the calculation}
1495: The model parameters involved in the calculation of the ohmic
1496: dissipation rate inside the planet are: 1) the planet's mass $M_{p}$,
1497: 2) semi-major axis $a$, 3) the relative angular velocity $\omega$ (the
1498: angular velocity of the field seen in a frame centered on the star and
1499: rotating with the planet), 4) the strength of the star's magnetic
1500: dipole moment $m$, and 5) the angle $\alpha$ between the spin axis of
1501: the star and the star's magnetic dipole.
1502: 
1503: We use the isothermal and polytropic prescription described in the
1504: previous section to model the planet's internal structure and
1505: calculate the conductivity profile inside the planet. To do so, we
1506: also need to specify 6) the mass of the star $M_{\star}$, and 7) the
1507: star's luminosity $\L_{\star}$.
1508: 
1509: \subsection{Methodology}
1510: The construction of a self-consistent model requires a loop of
1511: retroaction involving the determination of the internal structure of
1512: the planet and that of the ohmic energy dissipation.  For a specified
1513: internal structure of the planet, one can compute (following \S3) the
1514: conductivity profile and then the total ohmic dissipation rate
1515: $\mathcal{P}$. However, this energy dissipated inside the planet
1516: corresponds to an input of heat, which triggers an adjustment in the
1517: planet's internal parameters. Due to the efficient convection inside
1518: the planet, we assume that the adjustment of the internal parameters
1519: to this external heating is quick.  We consider that the
1520: characteristic time scale for the planet to evolve from one
1521: equilibrium state to another is small compared to the variation time
1522: scale of the seven parameters mentioned in the previous
1523: paragraph. Therefore, we do not need to follow the planet's dynamical
1524: evolution at all times. Instead, we can take a series of "snap shots"
1525: of the planet in its equilibrium state for different set of
1526: parameters.
1527: 
1528: Because of this feedback loop between the ohmic dissipation rate and
1529: the planet's internal parameters, we use an iterative method. For any
1530: chosen set of parameters, we start from an estimate for the ohmic
1531: dissipation rate $\mathcal{P}_{0}$ and internal structure $T_{0}(r)$,
1532: $P_{0}(r)$, and $\rho_{0}(r)$ corresponding to a magnetic diffusivity
1533: profile $\eta_{0}(r)$ (In our parametric analyses, we typically make
1534: small incremental changes in the model parameters from those for which
1535: we have already obtained equilibrium values).  We then compute the new
1536: internal structure $T_{1}(r)$, $P_{1}(r)$, and $\rho_{1}(r)$
1537: associated with $\mathcal{P}_{0}$. This enables us to compute the
1538: corresponding magnetic diffusivity $\eta_{1}(r)$. Finally, we use
1539: $\eta_{1}(r)$ to calculate the corresponding ohmic dissipation rate
1540: $\mathcal{P}_{1}$ and mass loss rate $\stackrel{\centerdot}{M_{1}}$.
1541: This process is iterated until convergence of $\mathcal{P}$,
1542: $\stackrel{\centerdot}{M}$, and of the internal parameters. Moreover,
1543: for some specific set of parameters $M_{p}$, $a$, $\omega$, $m$,
1544: $\sin(\alpha)$, $M_{\star}$, $L_{\star}$), we have started the
1545: iterative process from two different initial states in order to verify
1546: that they both converge to the same solution. Therefore, the iterative
1547: process does converge to a unique solution.
1548: 
1549: We consider the following fiducial model in which the mass of the planet and semi-major axis corresponds to 
1550: HD 209458 b, and in which the other parameters are reasonable ones for the type of systems considered. 
1551: An estimate of the order of magnitude for the strength of the magnetic dipole can be found in Johns-Krull 2007. 
1552: \begin{flushleft}
1553: \textbf{Mass of the planet:} $M_{p}=0.63\ M_{J}=1.26\times 10^{27}\
1554: kg$, \\ \textbf{Semi-major axis:} $a=0.04\ AU=6\times 10^{9}\ m$, \\
1555: \textbf{Relative angular velocity:} $\omega=10^{-5}\ s^{-1}$, \\
1556: \textbf{Star's magnetic dipole:} $m=4\times 10^{34}\ A\ m^{2}$, \\
1557: \textbf{Value of the tilt of the magnetic dipole:} $\sin(\alpha)=1$,
1558: \\ \textbf{Mass of the star:} $M_{\star}=M_{\odot}=2\times 10^{30}\
1559: kg$, \\ \textbf{Luminosity of the star:} $\L_{\star}=1.5\
1560: L_{\odot}=5.7\times 10^{26}$ W. \\
1561: \end{flushleft}
1562: 
1563: \subsection{Computation of $\mathcal{P}$ and $\stackrel{\centerdot}{M}$, 
1564: plots, and mathematical relations}
1565: 
1566: We present seven groups of plots (figures 3, 4, 5, 6, 7, 8, 9), one
1567: group for each parameter mentioned just above. For each group, we vary
1568: one parameter (x-axis), while keeping the others at the reference
1569: values mentioned above. On the y-axis, we plotted the ohmic
1570: dissipation rate $\mathcal{P}$, mass loss rate
1571: $\stackrel{\centerdot}{M}$, and characteristic time scale $\tau_M =
1572: \dfrac{M}{\stackrel{\centerdot}{M}}$. Note that the magnitude of 
1573: $\tau_M \sim$ for a Jupiter mass planet is about 1Myr. In addition, 
1574: the mass loss rate determined here is many orders of magnitude larger than
1575: that due to photo evaporation.  Only with such large mass loss rate, can we
1576: compensate for the angular momentum transfer due to the planet-disk
1577: and planet-star tidal interaction.
1578: 
1579: We emphasize once again that in the construction of these models, we
1580: assume that there is adequate energy dissipation to inflate the planet
1581: with $R_e > R_H$. In later papers that use the roche-filling model, we verify
1582: that $\mathcal{P} (R_H) > L_i (R_{H})$ before these results are
1583: applied.  If this condition is not satisfied, the planet would not
1584: fill its Roche lobe and not lose mass.
1585: 
1586: 
1587: \begin{figure}[hbt]
1588: \begin{center}
1589: \epsscale{0.65}
1590: \plotone{f3}  
1591: \end{center}
1592: \label{Varying mass of planet}
1593: \caption{Ohmic dissipation rate, mass loss rate, and time scale for
1594: different planetary masses}
1595: \end{figure} 
1596: 
1597: \begin{figure}[hbt]
1598: \begin{center}
1599: \epsscale{0.65}
1600: \plotone{f4}  
1601: \label{Varying semi-major axis of the planet}
1602: \end{center}
1603: \caption{Ohmic dissipation rate, mass loss rate, and 
1604: time scale for different semi-major axies}
1605: \end{figure} 
1606: 
1607: \begin{figure}[hbt]
1608: \begin{center}
1609: \epsscale{0.65}
1610: \plotone{f5}  
1611: \label{Varying the relative angular velocity}
1612: \end{center}
1613: \caption{Ohmic dissipation rate, mass loss rate, and 
1614: time scale for different relative angular velocities}
1615: \end{figure} 
1616: 
1617: \begin{figure}[hbt]
1618: \begin{center}
1619: \epsscale{0.65}
1620: \plotone{f6}  
1621: \label{Varying the star magnetic moment}
1622: \end{center}
1623: \caption{Ohmic dissipation rate, mass loss rate, 
1624: and time scale for different stellar magnetic moment}
1625: \end{figure}
1626: 
1627: \begin{figure}[hbt]
1628: \begin{center}
1629: \epsscale{0.65}
1630: \plotone{f7}  
1631: \end{center}
1632: \caption{Ohmic dissipation rate, mass loss rate, 
1633: and time scale for different tilt of the stellar magnetic dipole 
1634: with regard to the stellar spin axis}
1635: \label{Varying sin}
1636: \end{figure}
1637: 
1638: \begin{figure}[hbt]
1639: \begin{center}
1640: \epsscale{0.65}
1641: \plotone{f8}  
1642: \label{Varying the stellar mass}
1643: \end{center}
1644: \caption{Ohmic dissipation rate, mass loss rate, 
1645: and time scale for different stellar masses}
1646: \end{figure}
1647: 
1648: \begin{figure}[hbt]
1649: \begin{center}
1650: \epsscale{0.65}
1651: \plotone{f9}  
1652: \label{Varying the stellar total flux}
1653: \end{center}
1654: \caption{Ohmic dissipation rate, mass loss rate, and time scale 
1655: for different stellar total flux}
1656: \end{figure}
1657: 
1658: 
1659: From each group of plots, we obtain $\mathcal{P}$ and
1660: $\stackrel{\centerdot}{M}$ as a function of the parameter that is
1661: being varied (all the others are kept constant at the value of the
1662: fiducial model given above).  These functions are given in the table
1663: below (table \ref{functions}).
1664: 
1665: 
1666: \begin{table}[htbp]
1667: \begin{center}
1668: \begin{tabular}{|c|c|}
1669: \hline
1670: Ohmic dissipation rate $\mathcal{P}$ and mass loss rate $\stackrel{\centerdot}{M}$ & Varying Parameter   \\
1671: \hline
1672: $\mathcal{P}_{1} = 3.3\times 10^{21}\ \left(\frac{M_{p}}{0.63\ M_{J}}\right)^{2.16}$ W&
1673: $0.25\ M_{J} \leq M_{p} \leq 1.7\ M_{J}$  \\
1674: $ \stackrel{\centerdot}{M}_{1} = 1.2\times 10^{13}\ \left(\frac{M_{p}}{0.63\ M_{J}}\right)^{2.4} {\rm kg \ s}^{-1}$  & \\
1675: \hline
1676:  $\mathcal{P}_{2} = 3.3\times 10^{21}\ \left(\frac{a}{0.04\ AU}\right)^{-4}$  W&
1677: $0.015\ AU \leq a \leq\ 0.08\ AU$ \\
1678: $ \stackrel{\centerdot}{M}_{2} = 1.2\times 10^{13}\ \left(\frac{a}{0.4\ AU}\right)^{-3.8} {\rm kg \ s}^{-1}$ & \\
1679: \hline
1680: $\mathcal{P}_{3} = 3.5\times 10^{21}\ \left(\frac{|\omega|}{10^{-5}}\right)-1.5\times 10^{20}$ W &
1681: $7\times 10^{-6}\ s^{-1} \leq |\omega| \leq\ 7.3\times 10^{-5}\ s^{-1}$ \\
1682: $ \stackrel{\centerdot}{M}_{3} = 1.4\times 10^{13}\ \left(\frac{|\omega|}{10^{-5}}\right)-1.3\times 10^{12} {\rm kg \ s}^{-1}$ & \\
1683: \hline
1684: $\mathcal{P}_{4} = 3.3\times 10^{21}\ \left(\frac{m}{4\times 10^{34}}\right)^{2.18}$ W & 
1685: $6\times 10^{33}\ Am^{2} \leq m \leq\ 4\times 10^{34}\ Am^{2}$ \\
1686: $ \stackrel{\centerdot}{M}_{4} = 1.2\times 10^{13}\ \left(\frac{m}{4\times 10^{34}}\right)^{2.3} {\rm kg \ s}^{-1}$ \\
1687: \hline
1688: $\mathcal{P}_{5}=3.3\times 10^{21} \sin^{2.17}(\alpha)$ W & 
1689: $0.3 \leq \sin(\alpha) \leq 1$ \\
1690: $\stackrel{\centerdot}{M}_{5} = 1.2\times 10^{13}\sin^{2.28}(\alpha) {\rm kg \ s}^{-1} $&  \\
1691: \hline
1692: $\mathcal{P}_{6} = 3.3\times 10^{21}\ \left(\frac{M_{\star}}{M_{\odot}}\right)^{-0.53}$ W & 
1693: $0.5\ M_{\odot} \leq M_{\star} \leq\ 1.5\ M_{\odot}$ \\
1694: $ \stackrel{\centerdot}{M}_{6} = 1.2\times 10^{13}\ \left(\frac{M_{\star}}{M_{\odot}}\right)^{-0.5} {\rm kg \ s}^{-1}$& \\
1695: \hline
1696: $\mathcal{P}_{7} = 3.3\times 10^{21}\ \left(\frac{L_{\star}}{1.5\ L{\odot}}\right)^{-0.5} $W & 
1697: $0.5\ L_{\odot} \leq L_{\star} \leq\ 2.6\ L_{\odot}$ \\
1698: $ \stackrel{\centerdot}{M}_{7} = 1.2\times 10^{13}\ \left(\frac{L_{\star}}{1.5\ L_{\odot}}\right)^{-0.8} {\rm kg \ s}^{-1}$& \\
1699: $\mathcal{P}_{7} = 7.5\times 10^{19}\ \left(\frac{L_{\star}}{1.5\ L_{\odot}}\right)^{5.9} $W  & 
1700: $2.6\ L_{\odot} \leq L_{\star} \leq\ 5\ L_{\odot}$ \\
1701: $ \stackrel{\centerdot}{M}_{7} = 1.25\times 10^{11}\ \left(\frac{L_{\star}}{1.5\ L_{\odot}}\right)^{5.8} {\rm kg \ s}^{-1}$ & \\
1702: \hline
1703: \end{tabular}
1704: \end{center}
1705: \caption{Table giving $\mathcal{P}$ and $\stackrel{\centerdot}{M}$ as  
1706: a function of the parameter that is being varied.}
1707: \label{functions}
1708: \end{table}
1709: 
1710: 
1711: \subsection{Model parameter dependence}
1712: \begin{enumerate}
1713: \item \textbf{Mass of the planet $\mathbf{M_{p}}$.}  The total ohmic
1714: dissipation is a volumic integral over the entire region where
1715: dissipation occurs.  Therefore, one might expect $\mathcal{P}$ to be
1716: proportional to the volume.  Since the planet fills its Roche lobe,
1717: the volume is detemined by the mass (cf. eq. \ref{definition hill's
1718: radius}).  However, $M_{p}$ also gives a constraint on the volumic
1719: mass at the center of the planet (cf. equation \ref{3 unknowns}),
1720: which makes $\mathcal{P}$ mostly proportional to $M^{2}$. There is
1721: also a minor correction due to $\mathcal{P}$'s weak dependence on
1722: $\eta$ which depends on $M_{p}$ through the calculation of the
1723: internal parameters $T$, $P$, $\rho$ (eqs. \ref{P isothermal model}
1724: and \ref{P polytropic model}).
1725: 
1726: \item \textbf{Semi major axis $\mathbf{a}$.}  In the model we adopted in \S4,
1727: or more generally, in a model that would not take into account the
1728: planet's internal adjustment to the ohmic dissipation (especially in a
1729: model in which the radius of the planet is independent of the
1730: semi-major axis), the ohmic dissipation rate would be related to the
1731: semi-major axis according to the following law: $\mathcal{P} \propto
1732: B^{2} \propto a^{-6}$.  In the self-consistent model we adopted here
1733: and in the limit where the planet fills its Roche lobe
1734: ($R_{p}=R_{H}$), the radius of the planet varies with the semi-major
1735: axis. For example, when a planet moves closer to its host star, its
1736: Roche radius decreases linearly with the radius
1737: (cf. eq. \ref{definition hill's radius}) and, therefore, $\mathcal{P}$
1738: increases less quickly than if the planet kept the same radius.
1739: Again, $\mathcal{P}$ also has a weak dependence on $\eta$ which is
1740: depends on the semi-major axis through the planet's surface
1741: temperature and through the dependence of $r_{2}$ on $a$.  From these
1742: arguments, we thus expect $\mathcal{P}$ the exponent in
1743: $\mathcal{P}_{2}$ to be greater than -6 and less than -3.
1744: 
1745: \item \textbf{Relative angular velocity $\mathbf{\omega}$.}  In equation
1746: (\ref{dependence of P)}, the multiplicative constant in front
1747: of the volumic integral comes comes from the induction by the time
1748: dependent stellar field (cf. eq. (\ref{rhs=lhs}) and eq. (\ref{W
1749: provisoire})) and gives to $\mathcal{P}$ a dependence on
1750: $\omega^{2}$. However, in addition, $\omega$ intervenes inside the
1751: volumic integral in eq. \ref{dependence of P} through the
1752: dependence of $G_{l}(r)$ on the $\delta^{-1}$, with
1753: $\delta=\sqrt{\frac{\eta}{\omega}}$, skin depth. Therefore, the
1754: volumic integral is proportional to $\omega^{-1}$ and 
1755: $\mathcal{P}$ is proportional to $\omega$.
1756: 
1757: 
1758: \item \textbf{Magnetic dipole $\mathbf{m}$ and tilt of the magnetic dipole
1759: $\alpha$.}  Without any adjustment of the planet's interior to the
1760: ohmic dissipation, we would expect $\mathcal{P}_{4}$ and
1761: $\mathcal{P}_{5}$ to vary respectively in $m^{2}$ and
1762: $\sin^{2}(\alpha)$. The fact that both exponents that have been
1763: computed numerically are slightly larger than 2 means, in the
1764: self-consistent model we adopted here, that the adjustment of the
1765: planet's interior tends to have a small positive retroaction on the
1766: amount of enegy that is deposited inside the planet through ohmic
1767: dissipation.
1768: 
1769: \item \textbf{Mass of the star $\mathbf{M_{\star}}$.}  The mass of the star
1770: intervenes in the computation of the Roche radius ($R_{H} \propto
1771: M_{\star}^{-1/3}$).  $\mathcal{P}$ being a volumic integral, one
1772: would, therefore, expect it to vary proportionally to
1773: $M_{\star}^{-1}$.  However, $M_{\star}$ also intervenes in the
1774: computation of $r_{2}$ ($r_{2}$ is the sonic point, or the largest
1775: solution of $(r^{2}-rR_{H}+\frac{2c_{s}^{2}a^{3}}{LGM_{\star}})=0$,
1776: see. eq \ref{l}), which brings a correction to the dependence of
1777: $\mathcal{P}$ on $M_{s}$. Indeed, $r_{2}$ is used to compute
1778: $P(r_{2})$ and $\rho(r_{2})$ (cf. eq. \ref{P(r2)}) which are the
1779: boundary conditions we adopted to calculate the pressure and volumic
1780: mass in the isothermal region (cf. eq. \ref{P in isothermal
1781: region}). As a side note, this dependence of $\mathcal{P}$ on $r_{2}$
1782: could also affect the dependence of $\mathcal{P}$ on $a$ and
1783: $M_{\star}$.)
1784: 
1785: 
1786: \item \textbf{Stellar total Luminosity $\mathbf{L_{\star}}$.}  From figure 9,
1787: one can see that $\L_{\star}=10^{27}\ W$ correponds to a minimum for
1788: $\mathcal{P}$ and that $\mathcal{P}$ varies slowly for $\L_{\star}\leq
1789: 10^{27}\ W$ and much faster for $\L_{\star} \geq 10^{27}\ W$.  In the
1790: model we adopted here, the stellar total luminosity fixes the planet's
1791: equilibrium surface temperature, which is also the temperature of the
1792: isothermal region (for $\L_{\star}=10^{27}\ W$, $T_{p}\simeq 1767 K.$).
1793: It in turns determines the temperature profile inside the planet (the
1794: surface temperature is used as a boundary condition) and influences
1795: the internal structure and magnetic diffusivity profile inside the
1796: planet $\eta(r)$.  $T_{p}$ varies proportionally to $L_{\star}^{1/4}$
1797: (for constant semi-major axis) and, therefore, $\eta$ is roughly
1798: proportional to $exp(\dfrac{78909}{L_{\star}^{1/4}})$. \\
1799: 
1800: We ploted the integrand of the ohmic dissipation ($< \mathcal{P}_{volumic} > r^{2}$ 
1801: in eq. \ref{energy_integrand_expression} ) as well as the
1802: magnetic diffusivity (see figures 10 and 11) for
1803: $L_{\star}=2\times 10^{26}$ W, $7\times 10^{26}$ W, $10^{27}$ W,
1804: $1.2\times 10^{27}$ W, $1.5\times 10^{27}$ W, 
1805: and $2\times 10^{27}$ W respectively.  One can notice two parts
1806: corresponding to the isothermal and the polytropic region.  In the
1807: isothermal region, the energy integrand decreases slowly from the
1808: surface to the center of the planet.  The transition between the two
1809: regions corresponds to a sharp increase in conductivity and,
1810: therefore, a quick increase in the ohmic dissipation. This accounts
1811: for the sharp increase in the integrand (around $1.5\times 10^8$
1812: meters), and most of the remaining magnetic energy is dissipated in
1813: this region.  Moreover, an increase in the stellar luminosity results
1814: in a decrease in the magnetic diffusivity as well as a deeper
1815: penetration and a deeper transition between the isothermal and the
1816: polytropic region.\\
1817: 
1818: When one increases $L_{\star}$ starting from low values
1819: ($L_{\star}=2\times 10^{26}$W), the energy integrand also
1820: increases in the isothermal region. Nevertheless, the extent and amplitude of the sharp
1821: increase at the transition between the isothermal and the polytropic
1822: region are also reduced.  Therefore, these two effects compensate each
1823: other, and for low stellar luminosity (e.g. $L_{\star}$ between
1824: $2\times 10^{26}$W and $10^{27}$W, the total ohmic dissipation in the
1825: planet increases slowly with the stellar luminosity). \\
1826: 
1827: On the other hand, for higher values of the stellar luminosity, the
1828: penetration depth as well as the transition depth saturates (when the coupling term in eq. (\ref{first equation on G})
1829: $\dfrac{\omega}{\eta(r)}$ between the real and imaginary parts of $G_{l}(r)$ 
1830: becomes comparable to the other term $(\dfrac{l(l+1)}{r^{2}})$).  One can
1831: see that an increase in the stellar luminosity results only in changes
1832: in the energy integrand that would increase the total ohmic
1833: dissipation. This result in a much sharper increase of the ohmic
1834: dissipation with the stellar luminosity.
1835: 
1836: 
1837: \begin{figure}[hbt]
1838: \begin{center}
1839: \epsscale{0.75}
1840: \plotone{f10}  
1841: \label{energy_integrand_plot}
1842: \end{center}
1843: \caption{Integrand of the ohmic dissipation (log scale) for different values of the stellar luminosity}
1844: \end{figure}
1845: 
1846: \begin{figure}[hbt]
1847: \begin{center}
1848: \epsscale{0.75}
1849: \plotone{f11}  
1850: \label{plot_of_eta}
1851: \end{center}
1852: \caption{Magnetic diffusivity (log scale) for different values of the stellar luminosity}
1853: \end{figure}
1854: 
1855: 
1856: \item \textbf{Attempts of generalized function.}
1857: We consider the generalized expression of the ohmic dissipation rate 
1858: and mass loss rate, in the case 
1859: where the variables are separable: 
1860: \begin{eqnarray}
1861: \mathcal{P}_{gen}=3.3\times 10^{21} W
1862: \left(\frac{M_{p}}{0.63\ M_{J}}\right)^{2.16} 
1863: \left(\frac{a}{0.04\ AU}\right)^{-4}
1864: \left[
1865: \left(\frac{|\omega|}{10^{-5} {\rm s}^{-1}}
1866: -0.045\right)\right]\nonumber \\
1867: \left(\frac{m}{4\times 10^{34} {\rm A \ m}^2 }\right)^{2.18}
1868: \left(\sin\ \alpha \right)^{2.17}
1869: \left(\frac{M_{\star}}{M_{\odot}}\right)^{-0.53}
1870: \left(\frac{L_{\star}}{1.5\ L{\odot}}\right)^{-0.5} 
1871: \label{generalized P}
1872: \end{eqnarray}
1873: 
1874: \begin{eqnarray}
1875: \mathcal{\stackrel{\centerdot}{M}}_{gen}=1.2\times 10^{13} {\rm kg \
1876: s}^{-1} \left(\frac{M_{p}}{0.63\ M_{J}}\right)^{2.4}
1877: \left(\frac{a}{0.04\ AU}\right)^{-3.8}
1878: \left[\left(\frac{|\omega|}{10^{-5} {\rm s}^{-1}} -0.1
1879: \right)\right]\nonumber \\ \left(\frac{m}{4\times 10^{34} {\rm A \
1880: m}^2 }\right)^{2.3} \left(\sin\ \alpha \right)^{2.28}
1881: \left(\frac{M_{\star}}{M_{\odot}}\right)^{-0.5}
1882: \left(\frac{L_{\star}}{1.5\ L{\odot}}\right)^{-0.8}. 
1883: \label{generalized Mdot}
1884: \end{eqnarray}
1885: 
1886: The previous formulas have been written for the first interval of 
1887: $\mathcal{P}_{7}$ in table 2, but one can write the corresponding formulas for 
1888: the second interval of by using the corresponding expression of $\mathcal{P}_{7}$. 
1889: We compute, using the iterative procedure described above (\S 7.2),
1890: the ohmic dissipation $\mathcal{P}$ and mass loss rate
1891: $\mathcal{\stackrel{\centerdot}{M}}$ for sets of parameters in which
1892: more than 2 parameters are different from the fiducial parameters.  We
1893: then compare these values with the value of $\mathcal{P}_{gen}$ and
1894: $\mathcal{\stackrel{\centerdot}{M}}_{gen}$ for these sets of
1895: parameters. This test enables us to determine if the hypothesis of
1896: separation of variable is accurate or not.  The first set of
1897: parameters that we consider is $M_{p}=1.5 M_{J}$ and $ a=0.03 AU$, all
1898: the other parameters being kept equal to the fiducial parameters. We
1899: get: $\mathcal{P}=4.2\times 10^{22}$W and
1900: $\mathcal{\stackrel{\centerdot}{M}}=1.5\times 10^{14} {\rm kg \
1901: s}^{-1}$.  However, using the formula of the ohmic dissipation rate
1902: and mass loss rate with the approximation of separation of variables,
1903: we get: $\mathcal{P}_{gen}=7\times 10^{22}$W and
1904: $\mathcal{\stackrel{\centerdot}{M}}=2.9\times 10^{14}{\rm kg \ s}^{-1}$.
1905: 
1906: We now consider a set of parameters in which all parameters are taken
1907: differnet from the fiducial value: $M_{p}=1.5M_{j},\ a=0.03AU,\
1908: \omega=2.9\times 10^{-5}s^{-1},\ m=2\times 10^{34}\ Am^{2},\ M_{\star}=1.5
1909: M_{\odot}, \L_{\star}=0.8 \L_{\odot}$ (this value of $\omega$
1910: corresponds, for example to a system in which the planet is at
1911: keplerian augular velocity and the star has a period of 4 days).  We
1912: get $\mathcal{P}=1.7\times 10^{22}$W and
1913: $\mathcal{\stackrel{\centerdot}{M}}=7\times 10^{13} {\rm kg \
1914: s}^{-1}$, and using the formula of the ohmic dissipation rate and mass
1915: loss rate with the approximation of separation of variables, we get
1916: $\mathcal{P}_{gen}=3.5\times 10^{22}$W and
1917: $\mathcal{\stackrel{\centerdot}{M}}=1.6\times 10^{14} {\rm kg \
1918: s}^{-1}$. \\
1919: 
1920: From these two tests, we deduce that the approximation of separation
1921: of variables gives a reasonable order of magnitude, but nevertheless,
1922: does not seem to be accurate. This results means that the exponents in
1923: the functions given in table 2 can themselves be a function of the
1924: parameters, and the generalized functions that describe the ohmic
1925: dissipation and mass loss rate as a function of all parameters can be
1926: fairly complicated.  Nevertheless, even if a generalized formula is
1927: still unknown, for a given set of parameters, one could still compute
1928: the mass loss rate and ohmic dissipation using the procedure described
1929: in \S\S 3-7.
1930: \end{enumerate}
1931: 
1932: \subsection{Mass loss and migration stalls}
1933: From equation (\ref{generalized Mdot}), we find
1934: \begin{eqnarray}
1935: \tau_m= \frac{M_p}{\mathcal{\stackrel{\centerdot}{M}}_{gen}}
1936: \simeq 3 {\rm Myr} \left(\frac{0.63\ M_{J}}{M_{p}}\right)^{2.4}
1937: \left(\frac{a}{0.04\ AU}\right)^{3.8}
1938: \left[\left(\frac{|\omega|}{10^{-5} {\rm s}^{-1}} -0.1 \right)\right]^{-1}
1939: \left(\frac{m}{4\times 10^{34} {\rm A m}^2}\right)^{-2.3} \nonumber \\
1940: \left(\sin\ \alpha \right)^{-2.28}   
1941: \left(\frac{M_{\star}}{M_{\odot}}\right)^{0.5}
1942: \left(\frac{L_{\star}}{1.5\ L{\odot}}\right)^{0.8} \label{taum}
1943: \end{eqnarray}
1944: 
1945: The same mass loss provides angular momentum to the planet.  We
1946: neglect any variations in excentricity and assume that all the mass is
1947: accreted into the star.  Using (GBL eq. 96), we link the mass loss
1948: rate to a rate of change of semi-major axis,
1949: \begin{equation}
1950: \frac{\stackrel{\centerdot}{a}}{a}=-2
1951: \frac{\stackrel{\centerdot}{M}} {Mp} 
1952: \label{rate of semi-major axis change}
1953: \end{equation}
1954: Thus, $\tau_a \equiv \vert {\dot a} \vert /a = - 2 \tau_m$.
1955: 
1956: These relations indicate that, within $\sim 0.04$ AU, the Ohmic
1957: dissipation within the planet may indeed generate sufficient energy to
1958: inflat their radius beyond their Roche lobe.  The resulting mass
1959: transfer not only reduces the planets' mass but also stall their
1960: orbital migration.  The impact of this process on the mass-period 
1961: distribution of gas giants will be discussed in paper III.
1962: 
1963: \section{Summary}
1964: In this paper, we applied a model described by Campbell (in the
1965: context of interacting binary stars) to the situation of a planet in a
1966: protoplanetary disk interacting with the stellar periodic magnetic
1967: field.  In \S 3, we showed that with a well determined electrical
1968: conductivity profile inside the planet as well as the characteristic
1969: parameters of the system (such as the stellar magnetic field strength
1970: and angular velocity spin, planet radius and semi-major axis), one can
1971: compute the total ohmic dissipation rate $\mathcal{P}(t)$ inside the
1972: planet and its average value over one synodic period. This dissipation
1973: rate gives a good estimate of the strength of the Lorrentz torque
1974: exerted on the planet due to the interaction between the stellar
1975: magnetic field and the induced current inside the planet.  When the
1976: planet is outside corrotation, this torque will provide angular
1977: momentum to the planet from the star and slow down the planet's
1978: migration.  In \S4, we computed $\mathcal{P}$ for one specific set of
1979: parameters ($R_{p}=0.63\ R_{J}$, $a=0.04$ AU), and also showed that
1980: the conductivity profile (all the other parameters being kept
1981: constant) had some influence on the location of maximum dissipation,
1982: but fairly little influence on the total dissipate rate $\mathcal{P}$.
1983: We noted that this value of $\mathcal{P}$ seemed too low to directly
1984: provide an adequate rate of angular momentum transfert to the planet
1985: to stop its migration toward the host star. However, this energy input
1986: can inflat the planet's enveloppe and trigger mass loss
1987: $\stackrel{\centerdot}{M}$ through Roche lobe overflow.  The mass that
1988: overflows toward the central star provides angular momentum to the
1989: planet (GBL). In order to estimate this mass loss rate, we first
1990: linked the ohmic dissipation rate to the mass loss rate (\S5).  Then
1991: we used an isotermal-polytropic model to describe the adjustment of
1992: the planet's interior to the heat deposited through ohmic dissipation
1993: (\S6).  Finally, we computed $\mathcal{P}$ and
1994: $\stackrel{\centerdot}{M}$ at equilibrium for several set of
1995: parameters (\S7). A detailed calculation on the orbital evolution of
1996: the planet due to this process will be presented in paper III.
1997: 
1998: 
1999: \acknowledgments
2000: 
2001: We thank Drs. P. Garaud, J.E. Pringle and F. Rasio for constructive
2002: discussions. This work is supported by NASA (NAGS5-11779, NNG04G-191G,
2003: NNG06-GH45G, NNX07AL13G, HST-AR-11267), JPL (1270927), and NSF(AST-0507424). 
2004: 
2005: 
2006: \appendix
2007: \section{Perfect conductor moving relative to a magnetic field}
2008: Let's consider the flux $\Phi$ of the magnetic field $\textbf{B}$ accross a surface $S(t)$ that 
2009: changes with time or moves in space. One can show that 
2010: \begin{eqnarray}
2011: \frac{d\Phi}{dt}\stackrel{def}{=}\frac{1}{dt}\left[\int_a \textbf{B}(\textbf{r},t+dt)\, d\mathbf{S} 
2012: -\int_b \mathbf{B}(\textbf{r},t)\, d\mathbf{S} \right]
2013: =\int_a \left[\frac{\partial\textbf{B}}{\partial t}-\nabla\wedge(\upsilon\wedge
2014: \textbf{B})\right]\, d\mathbf{S}
2015: \end{eqnarray}
2016: with $a=S(t+dt)$, and $b=S(t)$. Therefore, using the mhd induction equation 
2017: \begin{equation}
2018: \frac{\partial\textbf{B}}{\partial t}=
2019: \nabla\wedge(\upsilon\wedge
2020: \textbf{B})
2021: -\nabla\wedge(\dfrac{1}{\mu_{0}\sigma}\nabla\wedge\textbf{B}),
2022: \end{equation} 
2023: we get: 
2024: \begin{equation}
2025: \frac{d\Phi}{dt}=-\frac{1}{\sigma}\int_c \mathbf{J}(\textbf{r},t)\, d\mathbf{l}
2026: \end{equation} 
2027: which tends to zero when the electric conductivity is large (the previous integral is 
2028: a closed integral along a close curve). Therefore, the magnetic field's flux will be constant if $\sigma$ 
2029: is large enough that the second term in the right hand side 
2030: is negligible. In such a case, the field lines will move with the body and appear to be 'frozen.'
2031: 
2032: \section{Set of linear equations}
2033: \label{set of linear equations}
2034: We give below the linear set of equation that we solved for  
2035: ($\mu_{1}^{1}$, $\mu_{1}^{-1}$, $\nu_{1}^{1}$,
2036: $\nu_{1}^{-1}$, $\alpha_{1}$, $\alpha_{2}$,$\alpha_{3}$,$\alpha_{4}$),
2037: ($\mu_{2}^{0}$, $\nu_{2}^{0}$, $\beta_{1}$,$\beta_{2}$), and
2038: ($\mu_{2}^{2}$, $\mu_{2}^{-2}$, $\nu_{2}^{2}$, $\nu_{2}^{-2}$,
2039: $\gamma_{1}$, $\gamma_{2}$, $\gamma_{3}$, $\gamma_{4}$). 
2040: The values of $G_{l}(r)$ considered are for $r=R_{p}$ the radius of the planet. 
2041: 
2042: \begin{eqnarray}
2043: \left\{
2044: \begin{array}{l}
2045: (\mu_{1}^{1}-\mu_{1}^{-1})Re(G_{1})+(-\nu_{1}^{1}+\nu_{1}^{-1})Im(G_{1}) -\alpha_{2}\frac{1}{R_{p}}\sqrt{\frac{8\pi}{3}}=0\\
2046: (-\nu_{1}^{1}-\nu_{1}^{-1})Re(G_{1})+(-\mu_{1}^{1}-\mu_{1}^{-1})Im(G_{1}) -\alpha_{4}\frac{1}{R_{p}}\sqrt{\frac{8\pi}{3}}=\frac{m\ \sin\alpha}{8\pi\ d^{3}}R_{p}^{2}\sqrt{\frac{8\pi}{3}}\\
2047: (-\nu_{1}^{1}+\nu_{1}^{-1})Re(G_{1})+(-\mu_{1}^{1}+\mu_{1}^{-1})Im(G_{1}) -\alpha_{1}\frac{1}{R_{p}}\sqrt{\frac{8\pi}{3}}=2\frac{m\ \sin\alpha}{8\pi\ d^{3}}R_{p}^{2}\sqrt{\frac{8\pi}{3}}\\
2048: (-\mu_{1}^{1}-\mu_{1}^{-1})Re(G_{1})+(\nu_{1}^{1}+\nu_{1}^{-1})Im(G_{1}) -\alpha_{3}\frac{1}{R_{p}}\sqrt{\frac{8\pi}{3}}=0 \\
2049: (\mu_{1}^{1}-\mu_{1}^{-1})Re(\stackrel{\centerdot}{G}_{1})+(-\nu_{1}^{1}+\nu_{1}^{-1})Im(\stackrel{\centerdot}{G}_{1})+ \alpha_{2}\frac{1}{R^{2}_{p}}\sqrt{\frac{8\pi}{3}}=\frac{m\ \sin\alpha}{8\pi\ d^{3}}2R_{p}\sqrt{\frac{8\pi}{3}}\\
2050: (-\nu_{1}^{1}-\nu_{1}^{-1})Re(\stackrel{\centerdot}{G}_{1})+(-\mu_{1}^{1}-\mu_{1}^{-1})Im(\stackrel{\centerdot}{G}_{1})+ \alpha_{4}\frac{1}{R^{2}_{p}}\sqrt{\frac{8\pi}{3}}=2\frac{m\ \sin\alpha}{8\pi\ d^{3}}2R_{p}\sqrt{\frac{8\pi}{3}}\\
2051: (-\nu_{1}^{1}+\nu_{1}^{-1})Re(\stackrel{\centerdot}{G}_{1})+(-\mu_{1}^{1}+\mu_{1}^{-1})Im(\stackrel{\centerdot}{G}_{1})+ \alpha_{1}\frac{1}{R^{2}_{p}}\sqrt{\frac{8\pi}{3}}=0 \\
2052: (-\mu_{1}^{1}-\mu_{1}^{-1})Re(\stackrel{\centerdot}{G}_{1})+(\nu_{1}^{1}+\nu_{1}^{-1})Im(\stackrel{\centerdot}{G}_{1})+ \alpha_{3}\frac{1}{R^{2}_{p}}\sqrt{\frac{8\pi}{3}}=0\\
2053: \end{array}
2054: \right.
2055: \end{eqnarray}
2056: \begin{eqnarray}
2057: \left\{
2058: \begin{array}{l}
2059: (\mu_{2}^{2}+\mu_{2}^{-2})Re(G_{2}(R_{p}))+(-\nu_{2}^{2}-\nu_{2}^{-2})Im(G_{2}(R_{p})) -\gamma_{2}\frac{1}{R_{p}^{2}}12\sqrt{\frac{2\pi}{15}}=0\\
2060: (-\nu_{2}^{2}+\nu_{2}^{-2})Re(G_{2}(R_{p}))+(-\mu_{2}^{2}+\mu_{2}^{-2})Im(G_{2}(R_{p})) -\gamma_{4}\frac{1}{R_{p}^{2}}12\sqrt{\frac{2\pi}{15}}=-\frac{1}{3}\dfrac{m\ \sin\alpha}{8\pi\ d^{4}}R_{p}^{3}12\sqrt{\dfrac{2\pi}{15}}\\
2061: (-\nu_{2}^{2}-\nu_{2}^{-2})Re(G_{2}(R_{p}))+(-\mu_{2}^{2}-\mu_{2}^{-2})Im(G_{2}(R_{p})) -\gamma_{1}\frac{1}{R_{p}^{2}}12\sqrt{\frac{2\pi}{15}}=-\frac{1}{2}\dfrac{m\ \sin\alpha}{8\pi\ d^{4}}R_{p}^{3}12\sqrt{\dfrac{2\pi}{15}}\\
2062: (-\mu_{2}^{2}+\mu_{2}^{-2})Re(G_{2}(R_{p}))+(\nu_{2}^{2}-\nu_{2}^{-2})Im(G_{2}(R_{p})) -\gamma_{3}\frac{1}{R_{p}^{2}}12\sqrt{\frac{2\pi}{15}}=0\\
2063: (\mu_{2}^{2}+\mu_{2}^{-2})Re(\stackrel{\centerdot}{G}_{2}(R_{p}))+(-\nu_{2}^{2}-\nu_{2}^{-2})Im(\stackrel{\centerdot}{G}_{2}(R_{p})) +2\gamma_{2}\frac{1}{R_{p}^{3}}12\sqrt{\frac{2\pi}{15}}=0\\
2064: (-\nu_{2}^{2}+\nu_{2}^{-2})Re(\stackrel{\centerdot}{G}_{2}(R_{p}))+(-\mu_{2}^{2}+\mu_{2}^{-2})Im(\stackrel{\centerdot}{G}_{2}(R_{p})) +2\gamma_{4}\frac{1}{R_{p}^{3}}12\sqrt{\frac{2\pi}{15}}=-\dfrac{m\ \sin\alpha}{8\pi\ d^{4}}R_{p}^{2}12\sqrt{\dfrac{2\pi}{15}}\\
2065: (-\nu_{2}^{2}-\nu_{2}^{-2})Re(\stackrel{\centerdot}{G}_{2}(R_{p}))+(-\mu_{2}^{2}-\mu_{2}^{-2})Im(\stackrel{\centerdot}{G}_{2}(R_{p})) +2\gamma_{1}\frac{1}{R_{p}^{3}}12\sqrt{\frac{2\pi}{15}}=-\frac{3}{2}\dfrac{m\ \sin\alpha}{8\pi\ d^{4}}R_{p}^{2}12\sqrt{\dfrac{2\pi}{15}}\\
2066: (-\mu_{2}^{2}+\mu_{2}^{-2})Re(\stackrel{\centerdot}{G}_{2}(R_{p}))+(\nu_{2}^{2}-\nu_{2}^{-2})Im(\stackrel{\centerdot}{G}_{2}(R_{p})) +2\gamma_{3}\frac{1}{R_{p}^{3}}12\sqrt{\frac{2\pi}{15}}=0
2067: \end{array}
2068: \right.
2069: \end{eqnarray}
2070: \begin{eqnarray}
2071: \left\{
2072: \begin{array}{l}
2073: \mu_{2}^{0}Re(G_{2}(R_{p}))-\nu_{2}^{0}Im(G_{2}(R_{p})) -\beta_{2}\frac{1}{R_{P}^{2}}\sqrt{\frac{4\pi}{5}}=0\\
2074: -\nu_{2}^{0}Re(G_{2}(R_{p}))-\mu_{2}^{0}Im(G_{2}(R_{p})) -\beta_{1}\frac{1}{R_{P}^{2}}\sqrt{\frac{4\pi}{5}}=\sqrt{\dfrac{4\pi}{5}}\dfrac{m\ \sin\alpha}{8\pi\ d^{4}}R_{p}^{3} \\
2075: \mu_{2}^{0}Re(\stackrel{\centerdot}{G}_{2}(R_{p}))-\nu_{2}^{0}Im(\stackrel{\centerdot}{G}_{2}(R_{p})) +\beta_{2}\frac{2}{R_{P}^{3}}\sqrt{\frac{4\pi}{5}}=0\\
2076: -\nu_{2}^{0}Re(\stackrel{\centerdot}{G}_{2}(R_{p}))-\mu_{2}^{0}Im(\stackrel{\centerdot}{G}_{2}(R_{p})) +\beta_{1}\frac{2}{R_{P}^{3}}\sqrt{\frac{4\pi}{5}}=\sqrt{\dfrac{4\pi}{5}}\dfrac{m\ \sin\alpha}{8\pi\ d^{4}}3R_{p}^{2}
2077: \end{array}
2078: \right.
2079: \end{eqnarray}
2080: 
2081: \section{Coefficients intervening in the expression of the ohmic dissipation rate}
2082: \begin{eqnarray}
2083: \left\{
2084: \begin{array}{l}
2085: A_{11}(r)\stackrel{def}{=}(\nu_{1}^{1}-\nu_{1}^{-1})Re(G_{1}(r))+(\mu_{1}^{1}-\mu_{1}^{-1})Im(G_{1}(r)) \\
2086: A_{12}(r)\stackrel{def}{=}(\mu_{1}^{1}+\mu_{1}^{-1})Re(G_{1}(r))-(\nu_{1}^{1}+\nu_{1}^{-1})Im(G_{1}(r)) \\
2087: A_{13}(r)\stackrel{def}{=}(\mu_{1}^{1}-\mu_{1}^{-1})Re(G_{1}(r))-(\nu_{1}^{1}-\nu_{1}^{-1})Im(G_{1}(r)) \\
2088: A_{14}(r)\stackrel{def}{=}-(\nu_{1}^{1}+\nu_{1}^{-1})Re(G_{1}(r))-(\mu_{1}^{1}+\mu_{1}^{-1})Im(G_{1}(r)) \\
2089: A_{15}(r)\stackrel{def}{=}\nu_{2}^{0}Re(G_{2}(r))+\mu_{2}^{0}Im(G_{2}(r)) \\
2090: A_{16}(r)\stackrel{def}{=}\mu_{2}^{0}Re(G_{2}(r))-\nu_{2}^{0}Im(G_{2}(r)) \\
2091: A_{17}(r)\stackrel{def}{=}(\nu_{2}^{2}+\nu_{2}^{-2})Re(G_{2}(r))+(\mu_{2}^{2}+\mu_{2}^{-2})Im(G_{2}(r)) \\
2092: A_{18}(r)\stackrel{def}{=}(\mu_{2}^{2}-\mu_{2}^{-2})Re(G_{2}(r))+(-\nu_{2}^{2}+\nu_{2}^{-2})Im(G_{2}(r)) \\
2093: A_{19}(r)\stackrel{def}{=}(\mu_{2}^{2}+\mu_{2}^{-2})Re(G_{2}(r))-(\nu_{2}^{2}+\nu_{2}^{-2})Im(G_{2}(r)) \\
2094: A_{20}(r)\stackrel{def}{=}(-\nu_{2}^{2}+\nu_{2}^{-2})Re(G_{2}(r))-(\mu_{2}^{2}-\mu_{2}^{-2})Im(G_{2}(r)) 
2095: \end{array}
2096: \right.
2097: \end{eqnarray}
2098: 
2099: \section{Conductivity and resistance of short-period extrasolar planets}
2100: 
2101: We now calculate the conductivity and resistance (the reciprocal of
2102: conductivity) of short-period extrasolar planets.  The conductivity of
2103: the planet is determined by the density of charged particles.  
2104: We consider separately the contributiond from the collisional 
2105: ionization within the planet's interior and from the photoionization
2106: near its surface.  
2107: 
2108: \subsection{Ionization of the planet's interior}
2109: 
2110: The cores and the inner envelopes of Jovian planets are mostly
2111: ionized, they are shielded by cool, mostly neutral gaseous envelopes,
2112: where the ionization is dominated by elements with low ionization
2113: potentials, such as sodium and potassium.
2114: 
2115: The planetary model used in our calculation of ionization fraction is
2116: model A3 presented by Bodenheimer et al. (2001). It is a spherically
2117: symmetric model for the short-period planet around HD 209458. The
2118: following parameters are assumed: the planetary mass is 0.63 Jupiter
2119: masses ($M_J$); the equlibrium temperature at the surface of the
2120: planet due to irradiation from the star is $T_s = 1360$~K; there is a
2121: solid core with a density $\rho_c = 5.5\ 10^{3} \ \rm{kg \ m^{-3}}$ and a mass
2122: 0.139 $M_J$ ( $ = 44 M_{\oplus} = 0.22 M_p $ ) in the center. An
2123: energy source, uniformly distributed through the gaseous part of the
2124: planet, with an energy input rate $\dot{E}_d = 8.5 \times 10^{19} \
2125: \rm{J \ s^{-1}}$, is also imposed to take into account the effect
2126: of tidal dissipation of energy. Those model parameters result in an
2127: asymptotic radius of $1.41 R_J$ at $t = 4.5$~Gyr, which is consistent
2128: with the photometric occultation observations of HD 209458 (Henry et
2129: al. 2000b; Charbonneau et al. 2000)
2130: 
2131: The cores and the inner envelope of Jovian planets are mostly ionized
2132: by the pressure ionization effect, where the Saha equation breaks
2133: down.  We have to resort to various equations of state (EOS) tables,
2134: including the equation of state tables for hydrogen and helium by
2135: Saumon et al. (1995).  These tables are calculated using the free
2136: energy minimization methods, with a careful study of the nonideal
2137: interactions.  They cover temperatures in the range $2.10 < \log
2138: T({\rm K}) < 7.06$ and pressure in the range $5 < \log P ({\rm N \
2139: m^{-2}}) < 20$.  The calculations based on which these tables were
2140: constructed also include the treatments of partial dissociation and
2141: ionization caused by both pressure and temperature effects. Given the
2142: internal structure data of $\rho$, $T$ and $P$ for model A3, we use
2143: those EOS tables to calculate electron number density distribution for
2144: the inner part of the planet.  In this approximation, we bear in mind
2145: that the ionization fraction calculated from the free-energy
2146: minimization are of limited accuracy.  Moreover, in the interpolation
2147: regions of both the H and He equation of state, the data have very
2148: little physical basis but are reasonably well behaved by construction.
2149: 
2150: In the envelope, hydrogen and helium are mostly neutral, and free
2151: electrons are exclusively provided by thermal ionization of elements
2152: with low ionization potentials. Among them, potassium and sodium have
2153: highest concentrations with a relative abundance $\log (N_K / N_H)
2154: \simeq -6.88$ and $\log (N_{Na} / N_H) \simeq -5.67$ for the solar
2155: composition.  The lowest ionization potentials for these two elements
2156: are $\chi_{K} = 4.339 $~eV and $\chi_{Na} = 5.138 $~eV. We identify
2157: these two elements to be the sources of most of the free electrons in
2158: the planetary envelope.
2159: 
2160: We solve the Saha equations for ionization of Na \& K jointly
2161: (cf. Allen 1955)
2162: \begin{equation}
2163: \frac{N^1_{Na}}{N^0_{Na}}\ 10^{6}\times N_e = - \chi_{Na} \Theta - {\frac{3}{2}} \log
2164: \Theta + 20.9388
2165: \end{equation}
2166: and
2167: \begin{equation}
2168: \frac{N^1_{K}}{N^0_{K}}\ 10^{6}\times N_e = - \chi_{K} \Theta - {\frac{3}{2}} \log
2169: \Theta + 20.9388
2170: \end{equation}
2171: where $\Theta = 5040 K / T$, $N^1_{Na}$ and $N^0_{Na}$ are,
2172: respectively, the singly ionized and neutral number density of sodium (SI units),
2173: $N^1_{K}$ and $N^0_{K}$ are, respectively, the singly ionized and
2174: neutral number density of potassium, and $N_e$ is the electron
2175: density.
2176: 
2177: 
2178: 
2179: 
2180: \subsection{Conductivity and resistivity}
2181: 
2182: Using the electron number density profile, we then calculate the
2183: conductivity using the formulas given by Fejer (1965). Three kinds of
2184: conductivity are of interests here. The conductivity $\sigma_0$, which
2185: determines the current parallel to the magnetic lines of force, and
2186: which would exist for all direction in the absence of the magnetic
2187: field, is given by
2188: \begin{equation}
2189: \sigma_0 = \frac{N_e e }{B} \left(\frac{\omega_{i}}{\nu_{i}} 
2190: -\frac{\omega_{e}}{\nu_{e}} \right)
2191: \cong \frac{N_e e^2}{m_e \nu_{e}} 
2192: \label{eq:sigma0}
2193: \end{equation}
2194: where $e$ is the electron charge, $\omega_{e} = - \frac{e B}{m_{e}}$
2195: and $\omega_{i} = \frac{e B}{m_{i}}$ are the gyrofrequencies of
2196: electron and ion, respectively, while $\nu_{e}$ and $\nu_{i}$ are the
2197: collisional frequencies associated with the momentum transfer of
2198: electrons and ions. 
2199: 
2200: The Pedersen conductivity, which determines current parallel to the 
2201: electric field, is given by  
2202: \begin{equation}
2203: \sigma_p = \frac{N_e e }{B} \left( \frac{\nu_{i} \omega_{i}}{\nu_{i}^{2}+
2204: \omega_{i}^{2}}  - \frac{\nu_{e} \omega_{e}}{\nu_{e}^{2}+
2205: \omega_{e}^{2}} \right)
2206: \cong \frac{\sigma_0 }{1 + (\omega_e / \nu_e)^2}.
2207: \label{eq:sigmap}
2208: \end{equation}
2209: The Hall conductivity, which determines the current perpendicular to both the
2210: electric and magnetic fields, is given by
2211: \begin{equation}
2212: \sigma_h = \frac{N_e e }{B} \left( \frac{\omega_{e}^{2}}{\nu_{e}^{2}+
2213: \omega_{e}^{2}} - \frac{\omega_{i}^{2}}{\nu_{i}^{2}+
2214: \omega_{i}^{2}} \right) 
2215: \cong \frac{\sigma_0 (\omega_e / \nu_e) }{1 + (\omega_e / \nu_e)^2} .
2216: \label{eq:sigmah}
2217: \end{equation}
2218: 
2219: 
2220: In the limit of low ionization fraction, $\nu_{e}$ is closely related
2221: to the mean collisional frequencies of the electrons with molecules of
2222: the neutral gas such that (cf. Draine et al. 1983)
2223: \begin{equation}
2224: \nu_{e}^{(1)} = N_n <\sigma v>_{e-n}
2225: \simeq N_n  \times 10^{-19}
2226:        \left( \frac{128 k T}{9 \pi m_e} \right)^{1/2} .
2227: \label{eq:nue1}
2228: \end{equation}
2229: where $N_n$ is the number density of neutral particles, $<\sigma
2230: v>_{e-n}$ is the average product of collisional cross section and the
2231: relative speed between electrons and neutral particles.  In the other
2232: limit of a completely ionized gas, we take into account both
2233: electron-ion and electron-electron encounters.  The collisional
2234: frequency of the electrons is given by (cf. Sturrock 1994)
2235: \begin{equation}
2236: \nu_{e}^{(2)} \simeq 10^{8.0} N_e T^{-3/2}
2237: \label{eq:nue2}
2238: \end{equation}
2239: In the intermediate range, we use $\nu_{e} = \max (\nu_{e}^{(1)}, 
2240: \nu_{e}^{(2)})$. 
2241: 
2242: The resistance of the planet cannot be specified exactly because an
2243: unknown shape factor is involved. Following Dermott (1970), we use the
2244: following expression for the resistance:
2245: \begin{equation}
2246: R_p = \frac{f_R}{r_p^2} \int_0^{r_p} \frac{dr}{\sigma_p (r)} ,
2247: \label{eq:resis}
2248: \end{equation}
2249: where $f_R$ is a order-of-unity parameter for the geometry.  In the
2250: aligned geometry we consider here, the planet's resistance comes from
2251: Pedersen resistivity (conductivity). The core of the planet is assumed
2252: to be a perfect conductor. From Eq.~\ref{eq:resis}, we can see that it
2253: is the cold, mostly neutral envelope that gives rise to most of the
2254: resistance. 
2255: 
2256: \begin{thebibliography}{}
2257: \bibitem[Allen 1955]{2321} Allen, C. W. 1955, 
2258: {\it Astrophysical quantities} (London, University of London Press)
2259: \bibitem[Bodenheimer, Lin, and Mardling 2001]{2323} Bodenheimer, P., Lin, D. N. C., and Mardling, R. A. 2001, ApJ, 548, 466
2260: \bibitem[Burkert, et al. 2005]{2324}  Burkert, A., Lin,D. N. C.; Bodenheimer, P. H., Jones, C. A., Yorke, H. W. 2005, ApJ, 618, 512
2261: \bibitem[Campbell, 1983]{2325} Campbell, C. G., 1983, MNRAS, 205, 1031
2262: \bibitem[Campbell 2005]{2326} Campbell, C. G. 2005, MNRAS, 359, 835
2263: \bibitem[ch2000]{2327} Charbonneau, D., Brown, T. M., Latham, D. W., Mayor, M. 2000, ApJ, 529, 45
2264: \bibitem[Clarke, et al. 1998]{2328}  Clarke, J. T. {\it et al} 1996, Sci, .274, 404
2265: \bibitem[Cowling 1981]{2329} Cowling, T. G. 1981, ARA\& A, 19, 115
2266: \bibitem[Dall'Osso, et al. 2005]{2330} Dall'Osso, S.,  Israel, G. L., Stella, L. 2006, A\& A, 447, 785
2267: \bibitem[Dermott 1970]{2331} Dermott, S. F. 1970, MNRAS, 149, 35
2268: \bibitem[Dobbs-Dixon, and Lin 2007]{2332} Dobbs Dixon, I. and Lin, D. N. C. 2008, ApJ, 673, 513 
2269: \bibitem[Draine 1983]{2333} Draine, B. T., Roberge, W. G., Dalgarno, A. 1983, ApJ, 264, 485
2270: \bibitem[Drell, et al. 1965]{2334} Drell, S. D., et al. 1965 JGR, 70, 3131
2271: \bibitem[Duncan 1966]{2335} Duncan, R. A. 1966, P\& SS, 14, 173
2272: \bibitem[Fejer 1965]{2336} Fejer, J. A. 1965, JGR, 70, 4972
2273: \bibitem[Goldreich, and Lynden-Bell 1969]{2337} Goldreich, P., and Lynden-Bell, D., 1969, ApJ, 156, 59
2274: \bibitem[Goldreich, and Tremaine 1978]{2338} Goldreich, P., and Tremaine, S. 1978, ApJ, 222, 850
2275: \bibitem[Goldreich, and Tremaine 1980]{gl80} Goldreich, P., and Tremaine, S. 1980, ApJ, 241, 425
2276: \bibitem[Gu 2003]{2340} Gu, P-G., Bodenheimer, P., Lin, D. N. C. 2003, ApJ, 588, 509
2277: \bibitem[Gu 2004]{2341} Gu, P-G., Bodenheimer, P., Lin, D. N. C. 2004, ApJ, 608, 1076
2278: \bibitem[Gurnett 1972]{2342} Gurnett, D. A. 1972, ApJ, 175, 525
2279: \bibitem[Hayashi, et al. 1985]{2343} Hayashi, C., Nakazawa, K.,  and Nakagawa, Y. 1985
2280: {\it Protostars and Planets II}, (Tucson: University of Arizona Press), 1100
2281: \bibitem[henry2000]{2345} Henry, G. W., Marcy, G. W., Butler, R. P., Vogt, S. S. 2000, ApJ, 529, 41
2282: \bibitem[Iaroslavitz, et al. 2007]{2366}  Iaroslavitz, E., Podolak, M. 2007, Icar, 187, 600 
2283: \bibitem[Ida, and Lin 2004]{2346} Ida, S., and Lin, D. N. C. 2004, ApJ, 604, 388
2284: \bibitem[Johns-Krull  2007]{2347} Johns-Krull, C. M. 2007, ApJ, 664, 975
2285: \bibitem[Joss, Rappaport, and Katz. 1979]{2348} 	Joss, P. C., Rappaport, S. A., and Katz, J. I. 1979, ApJ, 230, 176
2286: \bibitem[Stellar struct]{2349} Kippenhahn, R., Weigert, A. 1994, 
2287: {\it Stellar Structure and Evolution} (Germany: Springer-Verlag)
2288: \bibitem[konigl 1991]{2351} Konigl, A. 1991, ApJ, 370, 39
2289: \bibitem[Li, et al. 1998]{2352} Li, J., Ferrario, L., Wickramashinghe, D. 1998, ApJ, 503, 151
2290: \bibitem[Lin, and Papaloizou 1980]{lp80} Lin, D. N. C., and Papaloizou, J. C. B. 1980, MNRAS, 191, 37
2291: \bibitem[Lin, and Papaloizou 1986a]{lp86} Lin, D. N. C., and Papaloizou, J. C. B. 1986, ApJ, 307, 395
2292: \bibitem[Lin, and Papaloizou 1986b]{2355} Lin, D. N. C., and Papaloizou, J. C. B. 1986, ApJ, 309, 846
2293: \bibitem[Lin, and Papaloizou 1993]{lp93} Lin, D. N. C., and Papaloizou, J. C. B. 1993 
2294: {\it Protostars and Planets III}, (Tucson: University of Arizona Press), 749
2295: \bibitem[Lin, et al. 1996]{2358} Lin, D. N. C., Bodenheimer, P., Richardson, D. C. 1996, Nature, 380, 606
2296: \bibitem[Marcy, et al. 2000]{m00} Marcy, G. W., Cochran, and W. D., Mayor, M . 2000 
2297: {\it Protostars and Planets IV}, eds. V. Mannings, A.P. Boss, S.S Russell, (Tucson: University of Arizona Press), 1285
2298: \bibitem[Mayor, and Queloz 1995]{mq95} Mayor, M., and Queloz, D. 1995, Nature, 378, 355
2299: \bibitem[Neubauer 1980]{2362} Neubauer, F. M. 1980, JGR, 85, 1171
2300: \bibitem[Ogilvie and Lin 2004]{2300} Ogilvie, G. I., and Lin, D.N.C. 2004, ApJ, 610, 477
2301: \bibitem[Ogilvie, and Lin 2007]{2363} Ogilvie, G. I., and Lin, D. N. C. 2007, ApJ, 661, 1180
2302: \bibitem[de Pater]{2302} de Pater, I., and  Lissauer, J. 2001, Planetary Sciences, Cambridge University
2303: \bibitem[Piddington, and Drake]{2364} Piddington, J. H., and Drake, J. F., 1968, Nature, 217, 935
2304: \bibitem[Piddington 1977]{2365} Piddington, J. H. 1977, Moon, 17, 373
2305: \bibitem[Pollack, et al. 1996]{pol96} Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., Podolak, M., Greenzweig, Y. 1996, Icarus, 124, 62
2306: \bibitem[Saumon 1995]{2368} Saumon, D., Chabrier, G, Van Horn, H. M. 1995, ApJS, 99, 713
2307: \bibitem[Setiawan et al 2008]{2307} Setiawan, J., Henning, Th., Launhardt, R., Muller, A.; Weise, P., Kurster,
2308: M. 2008, Nature, 451, 38
2309: \bibitem[Shawhan 1976]{2369} Shawhan, S. D. 1976, JGR, 81, 3373
2310: \bibitem[Shu 1994]{2370} Shu, F. Najita, J., Ostriker, E., Wilkin, F., Ruden, S., Lizano, S. 1994, ApJ, 429, 781
2311: \bibitem[Stassun 2001]{2371} Stassum, K. G., Mathieu, R. D., Vrba, F. J., Mazeh, T., Henden, A. 2001, AJ, 121, 1003 
2312: \bibitem[Sturrock1994]{2372} Sturrock, P. A. 1994
2313: {\it Plasma Physics}, (Cambridge: Cambridge Universtiy Press)
2314: \bibitem[Ward 1997]{2374}  Ward, W. R. 1997, ApJ, 482, 211
2315: \bibitem[Warner 1995]{2375} Warner, B. 1995, 
2316: {\it Cataclysmic Variable Stars}, (Cambridge: Cambridge University Press)
2317: \bibitem[Willes, and Wu 2005]{2377} Willes, A. J., and Wu, K. 2005, A\& A, 432, 1091
2318: \bibitem[Wu, et al. 2002]{2378}  Wu, K., Cropper, M., Ramsay, G., Sekiguchi, K. 2002, MNRAS, 331, 221
2319: \bibitem[Zhou, and Lin 2007]{2379}   Zhou, J. L. and Lin, D. N. C. 2007, ApJ, 666, 447
2320: \end{thebibliography}
2321: \end{document}
2322: 
2323: