1: %\documentclass[preprint]{aastex}
2: %\documentclass[preprint2]{aastex}
3: %\documentclass[12pt,preprint]{aastex}
4: \documentclass{emulateapj}
5: \usepackage{graphicx}
6: \begin{document}
7:
8: \def\ba{\begin{eqnarray}}
9: \def\ea{\end{eqnarray}}
10: \def\etal{et al.\ \rm}
11:
12: \title{Cooling of young stars growing by disk accretion.}
13:
14: \author{Roman R. Rafikov\altaffilmark{1,2}}
15: \altaffiltext{1}{CITA, McLennan Physics Labs, 60 St. George St.,
16: University of Toronto, Toronto ON M5S 3H8 Canada; rrr@cita.utoronto.ca}
17: \altaffiltext{2}{Canada Research Chair}
18:
19: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
20:
21: \begin{abstract}
22: In the initial formation stages young stars must
23: acquire a significant fraction of their mass by accretion from a
24: circumstellar disk that forms in the center of a collapsing protostellar
25: cloud. Throughout this period mass accretion rates through the disk
26: can reach $10^{-6}-10^{-5}$ M$_\odot$ yr$^{-1}$ leading to
27: substantial energy release in the vicinity of stellar surface.
28: We study the impact of irradiation of the stellar surface produced
29: by the hot inner disk on properties of accreting fully
30: convective low-mass stars, and also look at objects such as young
31: brown dwarfs and giant planets. At high accretion rates irradiation
32: raises the surface temperature of the equatorial region
33: above the photospheric temperature $T_0$ that a star would
34: have in the absence of accretion. The high-latitude (polar)
35: parts of the stellar surface, where disk irradiation is weak,
36: preserve their temperature at the level of $T_0$. In strongly
37: irradiated regions an almost isothermal outer radiative zone
38: forms on top of the fully convective interior, leading
39: to the suppression of the local internal cooling flux derived from
40: stellar contraction (similar suppression occurs in
41: irradiated ``hot Jupiters''). Properties of this radiative
42: zone likely determine the amount of thermal energy that gets advected
43: into the convective interior of the star.
44: Total intrinsic luminosity integrated over the whole stellar
45: surface is reduced compared to the non-accreting case, by
46: up to a factor of several in some systems (young brown dwarfs,
47: stars in quasar disks, forming giants planets),
48: potentially leading to
49: the retardation of stellar contraction.
50: Stars and brown dwarfs irradiated by their disks tend to lose energy
51: predominantly through their cool polar regions while young giant
52: planets accreting through the disk cool through their whole surface.
53: \end{abstract}
54: \keywords{planets and satellites: formation ---
55: solar system: formation}
56:
57:
58: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
59: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
60:
61: \section{Introduction.}
62: \label{sect:intro}
63:
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65:
66:
67: Our understanding of advanced stages of star formation (T Tauri
68: and later phases) has been significantly
69: improved with the advent of infrared,
70: submillimeter, and high-resolution optical observatories
71: such as HST and Spitzer. At the same time a great deal of
72: uncertainty still remains regarding the earliest, so-called
73: Class 0 and Class I, stages of the star formation process. In the
74: conventional nomenclature, Class 0 stars are the protostellar
75: cloud cores in the very beginning of their collapse, while
76: Class I are the protostars embedded within an envelope of circumstellar
77: material that is infalling, accumulating in the centrifugally
78: supported disk, and being accreted by the protostars.
79:
80: At present, our knowledge exhibits a significant gap when
81: it comes to describing the actual buildup of the stellar
82: mass, from $M_\star=0~M_\odot$
83: in the Class 0 phase to $M_\star\sim 1~M_\odot$ in the end of
84: Class I phase. From the observational point of view the major
85: reasons for this
86: are (1) the heavy obscuration provided by the increased
87: densities in the central part of the infalling protostellar
88: core and the molecular cloud as a whole and (2) the
89: difficulty in deriving the spectra of the
90: central objects, namely distinguishing between the
91: intrinsic protostellar and accretion luminosities.
92: At the same time, our ignorance concerns not only the
93: history of {\it mass}
94: accumulation by the protostars. Thermodynamical state of the
95: accumulated gas is also an important ingredient of the
96: picture. Stars that form out of material with high entropy,
97: in particular that processed through
98: the accretion shock, tend to have large sizes, while objects
99: formed out of the lower entropy gas should be more compact.
100: At the moment the uncertainty in the initial
101: thermodynamical state of protostellar objects
102: precludes us from
103: getting a good handle on the evolutionary tracks of the fully
104: assembled (in terms of mass) protostars in the first $1-10$
105: Myrs after their formation.
106: %(D'Antona \& Mazzitelli 1994;
107: %Palla \& Stahler 1999; Siess \etal 2000).
108: Beyond
109: about 10 Myrs, when the initial conditions become largely
110: forgotten, the evolution tracks calculated under
111: different assumptions about the initial conditions typically
112: converge (Baraffe \etal 2002). However, prior to this stage
113: there are significant discrepancies between the results of
114: different groups, and the uncertainty in the initial
115: conditions for such calculations is the prime suspect for
116: the difference.
117:
118: It is generally accepted that the conservation of the angular
119: momentum in the collapsing protostellar cloud results in
120: accumulation of the collapsed gas in a rotationally-supported
121: disk in the cloud center. Only a small fraction
122: of the cloud mass has low enough angular momentum to
123: collapse directly into the protostellar core. The majority of
124: stellar mass is most likely accumulated by accretion from the
125: disk. According to observations, Class I stars acquire most
126: of their mass on timescale of several $10^5$ yrs
127: which implies that disks around these objects must exhibit time averaged
128: mass accretion rates of $\dot M \sim 10^{-6}-10^{-5}$
129: M$_\odot$ yr$^{-1}$. Accretion luminosity released in processing
130: such large mass flux through the disk exceeds the intrinsic
131: luminosity of the protostar. This immediately raises
132: an issue of the possible non-trivial radiative coupling between
133: the protostar and its circumstellar disk.
134:
135: Effects of disk accretion on structure of young stars
136: have been investigated by Mercer-Smith \etal (1984),
137: Palla \& Stahler (1992), Siess \& Forestini (1996),
138: Hartmann \etal (1997), Siess \etal (1997, 1999).
139: %In particular, these studies have generally found that
140: %if accreted material is able to lose its thermal energy
141: %rapidly, before its full incorporation into star, then
142: %star contracts more rapidly than in the absence of
143: %accretion.
144: Some of these authors studies how the heat advected
145: into the star with the freshly accreted material affects
146: protostellar properties. However, none of these investigations
147: looked at the effect of heat deposited at the stellar
148: {\it surface} by radiation originating in the inner parts
149: of the circumstellar disk, where most of the accretion energy
150: is released (see Figure \ref{fig:scheme} for a schematic
151: representation). Given that accretion luminosity may easily
152: exceed the intrinsic stellar luminosity (luminosity derived from
153: gravitational contraction, cooling and, possibly, deuterium
154: burning in stellar interior), omission of this effect may not
155: be justified in many cases.
156:
157:
158: \begin{figure}
159: \plotone{f1.eps}
160: \caption{
161: Schematic representation of stellar illumination by the disk
162: (slant hashed). Filled region at the point where disk joins the star
163: marks the boundary layer where intense energy dissipation
164: takes place. Part of radiation emitted by the disk (arrows) gets
165: intercepted by the star which heating it to temperature $T_{irr}$,
166: higher than the temperature $T_0$ that a star would have in
167: the absence of irradiation. Photospheric temperature is
168: preserved at the level of $T_0$ only in the polar regions
169: of the star (marked with dashed lines) where disk illumination
170: is weak. An external radiative zone
171: (horizontally hashed) forms in the strongly irradiated parts
172: of the stellar surface.
173: \label{fig:scheme}}
174: \end{figure}
175:
176: In this paper we investigate stellar irradiation by the
177: circumstellar disk and address the importance of this effect
178: in determining the intrinsic luminosity of young stars. We
179: calculate the spatial distribution of the disk flux on
180: the stellar surface and determine when irradiation is
181: important in \S \ref{sect:T_dist}. The effect of irradiation
182: on stellar cooling is investigated locally in \S
183: \ref{sect:stellar_cool} and globally in \S
184: \ref{sect:total_cooling}. Finally, in \S \ref{sect:disc}
185: we discuss the applications of this study to some real
186: objects and its possible limitations.
187:
188:
189: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
190: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
191:
192: \section{Temperature distribution due to disk irradiation.}
193: \label{sect:T_dist}
194:
195: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
196:
197: We start by calculating the distribution on the stellar surface
198: of the radiative flux $F_{irr}$ produced by the disk.
199: We consider an axisymmetric geometrically thin disk accreting
200: onto a star with radius $R_\star$ and mass $M_\star$.
201: Flux $F_{irr}$ intercepted by the star is a function of $\theta$
202: -- the angle between the normal to the stellar surface and the
203: normal to the disk (coincident with the polar axis of the star,
204: assuming that disk lies in the stellar equatorial plane).
205: Polar regions of the star are exposed to the radiation
206: of only the distant, cool parts of the disk, while the equatorial
207: regions have a direct view to the innermost parts of the disk
208: where most of the energy is dissipated.
209: One can easily show that a disk extending all the way to the
210: stellar surface gives rise to irradiation flux $F_{irr}(\theta)$
211: given by (Adams \& Shu 1986; Popham 1997)
212: \ba
213: && F_{irr}(\theta)=2\frac{R_\star\cos\theta}{\pi}
214: \int\limits_{R_{in}}^\infty F_d(R)R dR\nonumber\\
215: && \times\int\limits_0^{\phi_c} d\phi\frac{R\sin\theta\cos\phi-R_\star}
216: {(R^2+R_\star^2-2R_\star R\sin\theta\cos\phi)^2}
217: \label{eq:irr_flux}
218: \ea
219: where $R$ is the cylindrical radius,
220: $\cos\phi_c=R_\star/(r\sin\theta)$, $R_{in}=R_\star/\cos\theta$,
221: and $F_d(R)$ is the energy radiated by the unit surface area
222: of the disk per unit of time. In Appendix A we demonstrate
223: that this expression can be reduced to a one-dimensional
224: integral which is easier to analyze than equation
225: (\ref{eq:irr_flux}).
226:
227: To find the explicit dependence of $F_{irr}$ on $\theta$ one
228: needs to know $F_d(R)$ which is determined by the viscous
229: dissipation in the disk. Studies of steady-state thin accretion disks
230: have generally found that
231: \ba
232: F_d(R)=\frac{3}{8\pi}\frac{G M_\star \dot M}{R^3}f(R)
233: \label{eq:vis_dissip}
234: \ea
235: where $\dot M$ is a mass accretion rate and the function $f(R)$,
236: embodying the details of the disk emissivity near the stellar
237: surface, behaves as $f\to 1$ when $R\gg R_\star$. With $F_d$
238: given by (\ref{eq:vis_dissip}) one finds
239: \ba
240: F_{irr}(\theta)=\frac{G M_\star \dot M}{R_\star^3}g(\theta),
241: \label{eq:irr_flux_mod}
242: \ea
243: where the dimensionless function $g(\theta)$ is given by equation
244: (\ref{eq:g}).
245:
246: A standard disk with zero torque at the stellar surface
247: (situation appropriate for accretion onto
248: black holes) has (Shakura \& Sunyaev 1973)
249: $f(R)=1-(R_\star/R)^{1/2}$. The total viscous dissipation in
250: such a disk is $\dot E_d = (1/2)GM_\star \dot M/R_\star$
251: and the gas at the inner edge of the disk rotates at the local Keplerian
252: velocity. This is inappropriate in our case since the gas speed
253: has to match the velocity of the stellar surface at $R=R_\star$
254: (for simplicity assumed to be zero in our case).
255: As a result a boundary layer must form near the stellar surface in
256: which the azimuthal velocity of the gas is lowered by the viscous
257: torque from the local Keplerian value to the stellar rotation speed.
258: Viscous dissipation dramatically
259: increases gas temperature in this layer creating an additional
260: source of radiative flux very close to the stellar surface.
261: Irradiation by the boundary layer emission boosts up the
262: stellar surface temperature in a narrow belt at the equator (with
263: the thickness in $\theta$-direction comparable to the thickness
264: of the boundary layer) above that expected from the irradiation
265: by the more distant parts of the disk, outside of the boundary layer.
266: Thus, the existence of the boundary layer significantly modifies
267: disk structure and emissivity near the stellar surface
268: (Popham \etal 1993; Popham \& Narayan 1995) complicating
269: the calculation of $f(R)$.
270:
271: Fortunately, it will be shown later in \S
272: \ref{sect:total_cooling} that cooling of irradiated stars
273: depends only weakly on the behavior of $f(R)$ at $R\sim R_\star$
274: and is thus relatively insensitive to the structure of the
275: boundary layer. For the mass
276: accretion rates considered in this work ($10^{-6}-10^{-5}$
277: M$_\odot$ yr$^{-1}$) the geometric thickness of the boundary
278: layer is rather small, $\lesssim 0.15$ R$_\star$ (Popham \etal
279: 1993), so that the fraction of the stellar surface covered by
280: the boundary
281: layer and affected by the energy dissipation in it is rather small.
282: For this reason we will further assume for simplicity
283: that\footnote{Such assumption
284: results in $\dot E_d = (3/2)GM_\star \dot M/R_\star$, larger
285: than what can be provided by the change of the potential energy of
286: disk material, but this inconsistency is not going to strongly
287: affect our results.} $f(R)\approx 1$. In this case
288: $F_d$ keeps increasing all the way to the stellar surface [unlike
289: the zero inner torque case in which $F_d(R_\star)\to 0$]
290: thus roughly mimicking the contribution of the boundary layer to
291: the disk flux. We plot the behavior of function $g(\theta)$ in Figure
292: \ref{fig:irr_flux} for both $f(R)=1$ and $f(R)=1-(R_\star/R)^{1/2}$.
293:
294: \begin{figure}
295: \plotone{f2.eps}
296: \caption{
297: Irradiation flux absorbed by the stellar surface in units of
298: $GM_\star \dot M/R_\star^3$ as a function of $\theta$. Solid
299: curve corresponds to
300: $F_d(R)\propto R^{-3}$, dashed curve corresponds to
301: $F_d(R)\propto R^{-3}[1-(R_\star/R)^{1/2}]$, while the dotted
302: curve represents asymptotic behavior (\ref{eq:as}).
303: \label{fig:irr_flux}}
304: \end{figure}
305:
306: Irrespective of the complications related to the existence of
307: the boundary layer one can derive useful results for
308: $F_{irr}(\theta)$ in two asymptotic regimes. In particular,
309: using equation (\ref{eq:1D}) one finds that
310: as $\theta\to\pi/2$
311: \ba
312: F_{irr}\to\frac{F_d(R_\star)}{2},~~~~~g\approx\frac{3}{16\pi},
313: \label{eq:T_pi_2}
314: \ea
315: a result that is easy to
316: understand since any point at the stellar equator receives
317: disk radiation with uniform temperature corresponding to local
318: disk flux $F_d(R_\star)$
319: from $\pi$ steradian and reemits it into $2\pi$ steradian.
320: In the other limit
321: of $\theta\to 0$ one finds from equation (\ref{eq:1D}) that
322: \ba
323: %F_{irr}(\theta\to 0)\approx I_1\frac{G M_\star \dot M}
324: %{R_\star^3}\sin^5\theta,
325: g\approx I_1\sin^5\theta,
326: \label{eq:as}
327: \ea
328: where constant $I_1$ is given by equation (\ref{eq:I}). One can
329: see from Figure \ref{fig:irr_flux} that approximation (\ref{eq:as})
330: works
331: quite well (better than $22\%$ accuracy) for $\theta\lesssim 0.5$.
332: This asymptotic behavior is insensitive to the details of the
333: disk emissivity at $R\sim R_\star$ since polar regions
334: are irradiated only by parts of the disk at $R\gg R_\star$ where
335: $f(R)\approx 1$.
336:
337: Let us denote $T_0$ and $L_0=4\pi R_\star^2\sigma T_0^4$
338: the temperature and luminosity which a star with mass
339: $M_\star$ and radius $R_\star$ would have in the absence of
340: irradiation ($\sigma$ is a Stephan-Boltzmann constant).
341: To characterize the importance of irradiation we introduce
342: {\it irradiation parameter} $\Lambda$:
343: \ba
344: \Lambda\equiv\frac{GM_\star\dot M}{\sigma T_0^4 R_\star^3}\approx
345: 1.6~M_1\dot M_{-9}R_{11}^{-3}T_{3.5}^{-4},
346: \label{eq:Lambda}
347: \ea
348: where $T_n\equiv T_0/10^n$ K, $R_n\equiv R_\star/10^n$ cm,
349: $M_1\equiv M_\star/M_\odot$, and
350: $\dot M_n\equiv \dot M/(10^n~M_\odot~ {\rm yr}^{-1})$.
351: By construction, $\Lambda$ is roughly the ratio of accretion
352: luminosity and the stellar luminosity $L_0$ in the absence
353: of irradiation.
354:
355: When irradiation is
356: allowed for the photospheric temperature of the star $T_{ph}$
357: is a function of $\theta$ since
358: energy balance in steady state requires
359: \ba
360: \sigma T_{ph}^4(\theta) = \sigma T_{irr}^4(\theta) + F_{in}
361: \label{eq:balance}
362: \ea
363: at each point on the stellar surface,
364: where $F_{in}$ is
365: the intrinsic energy flux coming from the stellar interior
366: (derived from cooling of the stellar interior,
367: gravitational contraction, and D burning) and
368: \ba
369: && T_{irr}(\theta)\equiv \left[\frac{F_{irr}(\theta)}
370: {\sigma}\right]^{1/4}=T_0\left(\Lambda g\right)^{1/4}\\
371: && \approx 3.5\times 10^4 ~\mbox{K}
372: \left(M_1\dot M_{-5}R_{11}^{-3}\right)^{1/4}g^{1/4}.
373: \nonumber
374: \label{eq:T_irr}
375: \ea
376: Equation (\ref{eq:balance}) assumes that all radiation intercepted
377: by the star gets fully absorbed by its surface and reflection is
378: negligible. Our discussion can be easily extended
379: for the case of non-zero stellar albedo.
380:
381: In the absence of irradiation $F_{in}=\sigma T_0^4$.
382: With irradiation the local flux emitted by the photosphere
383: $\sigma T_{ph}^4$ exceeds $\sigma T_0^4$, but the
384: intrinsic stellar flux $F_{in}$ derived from the gravitational
385: contraction and cooling of the stellar interior
386: actually becomes smaller than $\sigma T_0^4$ as we
387: demonstrate in \S \ref{sect:1D}.
388:
389:
390:
391: We define the regime of {\it weak} irradiation as that
392: corresponding to low $\dot M$ such that
393: \ba
394: T_{irr}(\theta)\lesssim T_0
395: \label{eq:weak_condition}
396: \ea
397: for any $\theta$ (i.e. $\Lambda g\ll 1$). As the irradiation
398: is more intense near the stellar equator, weak irradiation
399: at {\it any} point on the stellar surface requires
400: $T_{irr}(\pi/2)\lesssim T_0$ (or $\Lambda\lesssim 1$),
401: or accretion rates lower than
402: \ba
403: && \dot M_c\approx \frac{16\pi}{3}\frac{R_\star^3\sigma T_0^4}
404: {G M_\star}\nonumber\\
405: && \approx 10^{-8}T_{3.5}^4 R_{11}^3 M_1^{-1}~
406: {\rm M}_\odot ~{\rm yr}^{-1}.
407: \label{eq:crit_M_dot}
408: \ea
409: Energy dissipation in the equatorial boundary layer (which we
410: do not account for here) can heat equatorial region
411: above $T_0$ even at $\dot M\lesssim \dot M_c$, but this heating
412: does not spread very far from the equator and
413: does little to affect the large scale stellar structure.
414:
415: A regime of {\it strong} irradiation is defined as that
416: corresponding to $\dot M\gtrsim \dot M_c$ ($\Lambda\gg 1$)
417: so that at least some parts of the stellar surface have
418: \ba
419: T_{irr}(\theta)\gtrsim T_0
420: \label{eq:strong_condition}
421: \ea
422: (or $\Lambda g \gtrsim 1$).
423: Initially this condition is satisfied only near the stellar
424: equator where an irradiated belt with
425: $T_{irr}(\theta)\gtrsim T_0$ forms. As $\dot M$ increases
426: this belt expands in $\theta$-direction, although rather
427: slowly since $F_{irr}$ is a rapidly decreasing function of
428: $\theta$, see Figure \ref{fig:irr_flux}.
429: As will be shown in \S \ref{sect:1D}, in irradiated regions
430: the intrinsic energy flux $F_{in}$ coming from the stellar
431: interior is suppressed compared
432: to $\sigma T_0^4$ so that the effective temperature of the
433: irradiated part of the star can be well approximated by
434: \ba
435: T_{ph}\approx T_{irr}(\theta).
436: \label{eq:irr_temp}
437: \ea
438: Transition between the low-latitude irradiated region
439: and the high-latitude part of the stellar surface where
440: $T_{ph}\approx T_0$ takes place at $\theta_{irr}$ given
441: by (see eq. [\ref{eq:as}])
442: \ba
443: &&\sin\theta_{irr}\approx \left(\frac{R_\star^3\sigma T_0^4}
444: {I_1 G M_\star \dot M}\right)^{1/5}=(I_1\Lambda)^{-1/5}\nonumber\\
445: &&\approx 0.5~
446: T_{3.5}^{4/5} R_{11}^{3/5} M_1^{-1/5}\dot M_{-5}^{-1/5}.
447: \label{eq:theta_c}
448: \ea
449: According to this formula, at $\dot M=10^{-5}$ M$_\odot$
450: yr$^{-1}$ polar regions having temperature $T_0\approx 3000$ K
451: occupy about $15\%$ of the stellar surface. The rest of
452: the surface has $T_{ph}$
453: significantly modified by intense radiation coming
454: from the disk. At this $\dot M$ equatorial temperature
455: reaches $T_{ph}(\pi/2)\approx 1.8\times
456: 10^4$ K, much higher than $T_0\sim 3000$ K corresponding
457: to the typical Hayashi track of a young star.
458:
459:
460: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
461: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
462:
463: \section{Cooling of irradiated stellar surface.}
464: \label{sect:stellar_cool}
465:
466: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
467:
468: Young stars, brown dwarfs and giant planets are fully
469: convective objects. It is well known (Kippenhahn \& Weigert 1994)
470: that many characteristics of fully convective objects such as
471: their luminosity and effective temperature are determined
472: mainly by the properties (opacity behavior, ratio of specific
473: heats of the gas) of their outermost, near-photospheric layers
474: and are rather insensitive to the processes
475: occurring in the convective interior.
476: Given that irradiation changes the boundary conditions
477: on the surface of accreting fully convective object we may
478: also expect that it should affect the luminosity of such an
479: object (Arras \& Bildsten 2006).
480:
481: Intense heating some parts of the
482: stellar surface suppresses convection in the subsurface layers
483: and gives rise to a convectively stable radiative zone sandwiched
484: between the photosphere and convective interior, see Figure
485: \ref{fig:scheme} for illustration. Appearance
486: of this zone is analogous to the formation of a roughly isothermal
487: radiative layer in the outer parts of the close-in extrasolar giant planets
488: caused by the intense radiation of their parent stars
489: (Guillot \etal 1996; Burrows 2000). It is the structure of
490: this zone that we want to investigate in order
491: to assess an impact of irradiation on stellar cooling.
492: Here we assume that the radiative zone is
493: \begin{enumerate}
494: \item optically thick, as measured from its bottom
495: (convective-radiative boundary) to the photosphere, and
496: \item geometrically thin compared to $R_\star$.
497: \end{enumerate}
498: Validity of these assumptions is verified in \S \ref{sect:1D}.
499:
500: In the absence of internal energy sources radiation
501: transport in the optically thick radiative layer is
502: governed by
503: \ba
504: \nabla\cdot {\bf F}=0,~~~~{\bf F}=
505: -\frac{16}{3}\frac{\sigma T^3}{\kappa\rho}\nabla T,
506: \label{eq:rad_tran}
507: \ea
508: where $F$ is the radiative flux density, $\kappa$ is opacity
509: and $\rho$ is the gas density. Equation of hydrostatic equilibrium
510: reads $\nabla P = -\rho {\bf g}$, where $P$ is the gas pressure and
511: ${\bf g}$ is the local gravitational acceleration. These two equations
512: describe the radiative zone structure subject to the boundary condition
513: \ba
514: && T\Big|_{\tau=2/3}=T_{ph}(\theta),
515: \label{eq:Tboundary}
516: \ea
517: where $\tau$ is the optical depth. The radiative boundary condition
518: (\ref{eq:Tboundary}) coupled with (\ref{eq:balance}) is appropriate
519: here because most of the stellar surface is not obscured by
520: the accreting gas and is free to radiate energy into space.
521:
522:
523: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
524:
525: \subsection{1D approximation for the structure of the radiative layer.}
526: \label{sect:1D}
527:
528: In general one must solve equation (\ref{eq:rad_tran}) together with
529: the equation of hydrostatic equilibrium in two dimensions -- $r$ and
530: $\theta$. However, under the circumstances clarified in \S
531: \ref{sect:1D_validity} the $r$-component of the radiative
532: flux $F_r$ is much larger than its $\theta$-component $F_\theta$,
533: so that the latitudinal transport of energy can be neglected.
534: This leaves $r$ as the only independent variable in equation
535: (\ref{eq:rad_tran}) effectively making it one-dimensional.
536: A dependence on $\theta$ then appears only through the external
537: boundary condition, namely $T_{ph}(\theta)$. This is the limit that
538: we will focus on in this work.
539:
540: Because of the thinness of the radiative zone, $r$ varies
541: only weakly through the radiative zone so that one can neglect
542: the divergence of the radial component of $F_r$, thereby
543: reducing equation (\ref{eq:rad_tran}) to simply
544: $\partial F_r/\partial r =0$.
545: With these simplifications equation (\ref{eq:rad_tran}) can be
546: integrated once to find
547: \ba
548: F_{in}=-\frac{16}{3}\frac{\sigma T^3}
549: {\kappa\rho}\frac{\partial T}{\partial r},
550: \label{eq:flux}
551: \ea
552: where the integration constant on the left-hand side is
553: independent of $r$ and as such has to coincide with the
554: intrinsic flux $F_{in}$ coming from
555: the convective interior of the star. Determination of
556: $F_{in}$ is the goal of our calculation.
557:
558: Radial pressure gradients in the radiative zone
559: are much larger than the latitudinal gradients so that the
560: equation of hydrostatic equilibrium can be written as
561: \ba
562: \frac{\partial P}{\partial r} = -\rho g,
563: \label{eq:hydro_eq}
564: \ea
565: where $g=|{\bf g}|\approx GM_\star/R_\star^2$ is the gravitational
566: acceleration which is roughly constant within the thin radiative
567: layer (stellar rotation is neglected throughout this work).
568:
569: Subsequent consideration is very similar to the calculation of
570: the atmospheric structure for the protoplanetary core
571: immersed in a protostellar nebula, which can be found in Rafikov
572: (2006). We assume that $\kappa$ depends on
573: gas pressure and temperature as
574: \ba
575: \kappa=\tilde\kappa P^\alpha T^\beta,
576: \label{eq:kappa}
577: \ea
578: where $\alpha>0$ and $\beta$ are constants.
579: Equation (\ref{eq:kappa}) is a reasonable approximation to the
580: opacity behavior in some density and temperature intervals
581: typical for young stars. In particular,
582: at $2500$ K $<T\lesssim 5000$ K opacity is mainly
583: due to H$^-$ absorption with electrons supplied by
584: elements heavier than H with low ionization potentials.
585: Bell \& Lin (1994) have demonstrated that in this regime
586: $\kappa$ can be well fit by
587: \ba
588: \kappa\approx 6\times 10^{-14}P^{2/3}T^{7/3}.
589: \label{eq:low_T}
590: \ea
591: At $T\gtrsim 5000$ K electrons from partial hydrogen ionization
592: enhance H$^-$ opacity and hydrogenic absorption dominates. In this regime
593: (Bell \& Lin 1994)
594: \ba
595: \kappa\approx 2.4\times 10^{-39}P^{1/3}T^{29/3}.
596: \label{eq:high_T}
597: \ea
598:
599: Here we also assume that the equation of state (EOS)
600: of gas in the whole star, including the external radiative zone,
601: can be characterized by a single ratio of specific heats $\gamma$.
602: In other words, we assume that under adiabatic conditions gas behaves as
603: $P=K\rho^\gamma$, where $K$ is a constant set by the entropy of the gas
604: and $\gamma$ is fixed throughout the star. We adopt $\gamma=5/3$, which
605: should work fine in fully ionized, convective interiors of young low-mass
606: stars, although this approximation is not very accurate at the transition
607: between the outer radiative zone and the convective interior since
608: gas is only partly atomic there. Continuing
609: dissociation and ionization cause variations of $\gamma$ in this region
610: which may be quite important, see \S \ref{sect:complications}.
611: Nevertheless, to get a qualitative picture of the effect of irradiation
612: on stellar cooling and for making rough numerical estimates this
613: constant-$\gamma$ approximation should be sufficient.
614:
615: With $\kappa$ in the form (\ref{eq:kappa})
616: equation (\ref{eq:flux}) can be integrated using
617: (\ref{eq:hydro_eq}), (\ref{eq:kappa}) and the
618: ideal gas law:
619: \ba
620: \left(\frac{P}{P_{ph}}\right)^{1+\alpha}-1 = \frac{\nabla_0}{\nabla_{ph}}
621: \left[\left(\frac{T}{T_{ph}}\right)^{4-\beta}-1\right],
622: \label{eq:PofT}
623: \ea
624: where
625: \ba
626: && \nabla_0=\frac{1+\alpha}{4-\beta},
627: \label{eq:nabla_0}\\
628: && \nabla_{ph}= \frac{3}{16}\frac{F_{in}\kappa_{ph}P_{ph}}
629: {g\sigma T_{ph}^4},
630: \label{eq:nabla_ph}
631: \ea
632: and $P_{ph}$ and $\kappa_{ph}=\tilde \kappa P_{ph}^\alpha
633: T_{ph}^\beta$ are the values of pressure and opacity at the
634: photosphere.
635:
636: Solution (\ref{eq:PofT}) allows us to calculate temperature gradient
637: \ba
638: \nabla(T)\equiv\frac{\partial \ln T}{\partial \ln P}=\nabla_0
639: \left[1-\left(\frac{T_{ph}}{T}\right)^{4-\beta}
640: \left(1-\frac{\nabla_{ph}}{\nabla_0}\right)\right],
641: \label{eq:nabla}
642: \ea
643: which determines whether gas is stable against convection.
644: Note that at the photosphere $\nabla(T_{ph})=\nabla_{ph}$.
645: Everywhere inside the radiative zone
646: \ba
647: \nabla<\nabla_{ad}\equiv
648: (\gamma-1)/\gamma,
649: \label{eq:Schwar}
650: \ea
651: where $\nabla_{ad}$ is the adiabatic temperature gradient.
652: In convective regions $\nabla>\nabla_{ad}$. For our adopted
653: $\gamma=5/3$ one finds $\nabla_{ad}=2/5$.
654:
655: We also assume that at some depth an object under consideration
656: does become convective and determine what is necessary for this
657: transition to occur. If $\beta<4$ then equations (\ref{eq:nabla})
658: and (\ref{eq:Schwar}) demonstrate that convection sets
659: in only provided that
660: \ba
661: \nabla_0>\nabla_{ad}.
662: \label{eq:less4}
663: \ea
664: Situation described by equation (\ref{eq:less4}) is realized e.g.
665: for opacity given by equation (\ref{eq:low_T}), when $\nabla_0=1$
666: exceeds $\nabla_{ad}=2/5$, implying that radiative energy
667: transport does indeed change to convective at some depth,
668: as we have assumed.
669:
670: On the other hand, when opacity is
671: characterized by $\beta>4$ equation (\ref{eq:nabla}) guarantees
672: that transition to convection occurs at some depth, irrespective
673: of the exact value of either $\beta$ or $\nabla_0$. This situation
674: is appropriate for $\kappa$ given by equation (\ref{eq:high_T})
675: since in that case $\beta\approx 10$.
676:
677: Despite this difference, in both cases the temperature $T_{cb}$
678: and pressure $P_{cb}$ at the convective-radiative boundary are
679: given by\footnote{These results coincide with equations
680: (47) and (48) of Rafikov (2006) if we identify
681: $\nabla_{ph}=\nabla_\infty$ and assume $\nabla_{ph}\ll 1$.}
682: \ba
683: && T_{cb}=T_{ph}\left(\frac{\nabla_0-\nabla_{ph}}
684: {\nabla_0-\nabla_{ad}}\right)^{1/(4-\beta)},
685: \label{eq:T_cb}\\
686: && P_{cb}=P_{ph}\left(\frac{\nabla_{ad}}{\nabla_{ph}}\cdot
687: \frac{\nabla_0-\nabla_{ph}}
688: {\nabla_0-\nabla_{ad}}\right)^{1/(1+\alpha)},
689: \label{eq:P_cb}
690: \ea
691: which can be derived by setting $\nabla(T_{cb})=\nabla_{ad}$
692: and using equation (\ref{eq:PofT}). Using equations
693: (\ref{eq:hydro_eq}), (\ref{eq:PofT}), and (\ref{eq:P_cb})
694: one can also find that the optical depth at the
695: convective-radiative boundary
696: \ba
697: \tau_{cb}\sim \nabla_{ph}^{-1},
698: \label{eq:tau_cb}
699: \ea
700: while the radial extent of the outer radiative zone is
701: \ba
702: \Delta R_r\sim H_{ph}\ln\nabla_{ph}^{-1}.
703: \label{eq:dR_r}
704: \ea
705: where $H_{ph}=k_B T_{ph}/(\mu g)$ is the photospheric scale
706: height. Given that irradiation cannot heat
707: the star to a temperature comparable to its central temperature
708: (otherwise outer layers would be unbound) $H_{ph}$ should be
709: much smaller than $R_\star$ even under rather extreme irradiation.
710:
711: Both $\nabla_0$ and $\nabla_{ad}$ are constants of order unity.
712: This makes it clear from equation (\ref{eq:T_cb}) that the
713: temperature variation between the photosphere and the convective
714: zone boundary is rather small, $|T_{cb}-T_{ph}|\sim T_{ph}$.
715: In practice, we find that at $T\lesssim 5000$ K, when $\kappa$
716: is given by (\ref{eq:low_T}), convection sets in at
717: $T_{cb}\approx 1.36 T_{ph}$, while at higher temperatures, when
718: $\kappa$ is given by (\ref{eq:high_T}), $T_{cb}\approx 1.19 T_{ph}$.
719: At the same time, under strong irradiation the external radiative
720: zone should be deep enough for the pressure $P_{cb}$ at its bottom
721: to greatly exceed $P_{ph}$. In this case equation (\ref{eq:P_cb})
722: suggests that
723: \ba
724: \nabla_{ph}\ll 1,
725: \label{eq:nabla_cond}
726: \ea
727: a result that is verified in \S \ref{sect:isolated}, see
728: equation (\ref{eq:nabl_ph}).
729:
730: According to equation (\ref{eq:tau_cb}) smallness of
731: $\nabla_{ph}$ results in $\tau_{cb}\gg 1$, thus justifying our
732: assumption (1) about the radiative zone properties.
733: At the same time, because of rather weak (logarithmic) dependence
734: of $\Delta R_r$ on $\nabla_{ph}$, the thickness of the outer
735: radiative zone should not be much different from $H_{ph}\ll
736: R_\star$. As
737: a result, $\Delta R_r\ll R_\star$, verifying our assumption (2).
738: Thus, the condition (\ref{eq:nabla_cond}) can then be viewed as a
739: prerequisite for the formation of a geometrically thin, optically thick
740: radiative zone with roughly isothermal temperature profile
741: under the action of intense external irradiation. External radiative
742: layers with similar near-isothermal structure are expected to exist in
743: the envelopes of irradiated hot Jupiters (Guillot \etal 1996; Baraffe
744: \etal 2003; Chabrier \etal 2004) and
745: in the outer parts of the low-luminosity atmospheres of protoplanetary
746: cores immersed into the protoplanetary nebulae (Rafikov 2006).
747:
748: The value of $P_{ph}$
749: can be fixed in the following way. Above the photosphere gas is
750: roughly isothermal with temperature $T_{ph}$ -- an
751: approximation which is good enough for our purposes. Then at
752: height $z$ above the photosphere
753: $\rho(z)=\rho_{ph}\exp(-z/H_{ph})$, where $\rho_{ph}$ is the
754: photospheric gas density. Using this result and equation
755: (\ref{eq:kappa}) we find the photospheric optical depth
756: \ba
757: \frac{2}{3}=\int\limits_0^\infty\kappa\rho dz=
758: \frac{\kappa_{ph}P_{ph}}{(\alpha+1)g},
759: \label{eq:2_3}
760: \ea
761: from which it follows that
762: \ba
763: P_{ph}=\left[\frac{2(\alpha+1)}{3}\frac{g}{\tilde
764: \kappa T_{ph}^\beta}\right]^{1/(1+\alpha)}.
765: \label{eq:P_ph}
766: \ea
767: As a byproduct of relation (\ref{eq:2_3}) one can rewrite
768: equation (\ref{eq:nabla_ph}) as
769: \ba
770: \nabla_{ph}=\frac{\alpha+1}{8}\frac{F_{in}}{\sigma T_{ph}^4}.
771: \label{eq:nab_ph}
772: \ea
773: It then follows from equations (\ref{eq:nabla_cond})
774: and (\ref{eq:nab_ph}) that $F_{in}\ll \sigma T_{ph}^4$.
775:
776: We are now in position to evaluate $F_{in}$ and see how irradiation
777: affects cooling of convective objects. To do this we note that
778: the inner boundary of the radiative zone is also the outer boundary
779: of the convective interior. We assume that convective
780: transport is so efficient that entropy is constant\footnote{In
781: reality stellar envelope contains superadiabatic regions which
782: are not captured in our analysis and may affect its results.}
783: throughout the inner convective zone, so that the EOS can be
784: well represented by $P=K\rho^\gamma$, where $K$ is the adiabatic
785: constant. As a result, $P_{ph}$ and $T_{ph}$ must be
786: related via $(kT_{cb}/\mu)^\gamma=K P_{cb}^{\gamma-1}$ which,
787: coupled with equations (\ref{eq:nabla_ph}), (\ref{eq:T_cb}),
788: (\ref{eq:P_cb}), and condition (\ref{eq:nabla_cond}), yields
789: the following expression for $F_{in}$:
790: \ba
791: && F_{in}(\theta)=\frac{16\nabla_{ad}}{3}
792: \left(\frac{\nabla_0-\nabla_{ad}}{\nabla_0}\right)^{\nabla_0/\nabla_{ad}-1}
793: \frac{\sigma g}{\tilde \kappa}\nonumber\\
794: && \times\left(\frac{\mu K^{1/\gamma}}{k_B}\right)^{(1+\alpha)/\nabla_{ad}}
795: \left[T_{ph}(\theta)\right]^{4-\xi},
796: \label{eq:F_in}
797: \ea
798: where
799: \ba
800: \xi=\beta+(1+\alpha)/\nabla_{ad}.
801: \label{eq:xi}
802: \ea
803: Intrinsic stellar flux $F_{in}$ exhibits an explicit
804: latitudinal dependence because it is a function of
805: $T_{ph}(\theta)$.
806:
807: As discussed before, when $\beta<4$ and $\nabla_0>0$ a transition to
808: convection at some depth requires $\nabla_0>\nabla_{ad}$.
809: As a result,
810: \ba
811: 4-\xi=(4-\beta)\left(1-\frac{\nabla_0}{\nabla_{ad}}\right)<0.
812: \label{eq:4xi}
813: \ea
814: On the other hand, when
815: $\beta > 4$ one also finds $4-\xi<0$ because $\nabla_0<0$
816: in this case. Thus, in both situations
817: $F_{in}$ {\it decreases} as $T_{ph}$ increases. In other words,
818: irrespective of the opacity behavior external irradiation of the
819: stellar surface {\it suppresses}
820: stellar cooling, a result known from the studies of irradiated
821: giant planets (Guillot \etal 1996; Burrows 2000).
822:
823: Since external radiative zone is rather thin compared to $R_\star$
824: it must contain negligible amount of mass compared with $M_\star$.
825: Then the structure of fully convective inner region of the
826: star should be well described by the classical
827: theory of polytropic spheres (Landau \& Lifshitz 1984; Kippenhahn 1994).
828: In particular, adiabatic constant $K$ can be related to the stellar
829: mass and radius as
830: \ba
831: K=\zeta(\gamma) G M_\star^{2-\gamma}R_\star^{3\gamma-4},
832: \label{eq:K}
833: \ea
834: where $\zeta(\gamma)\sim 1$ is a parameter set by the equation
835: of state of the gas. In a particular case of convective young stars
836: with fully ionized interior characterized by $\gamma=5/3$ one
837: has $\zeta(5/3)=0.1286$ and
838: \ba
839: K=1.081\times 10^{14}~M_1^{1/3}R_{11}.
840: \label{eq:K_3_2}
841: \ea
842:
843: Equations (\ref{eq:F_in}) and (\ref{eq:K})
844: unambiguously determine cooling
845: of the star as a function of stellar parameters $R_\star$ and
846: $M_\star$, temperature distribution at the photosphere
847: $T_{ph}(\theta)$, and opacity behavior in the outer radiative zone.
848:
849:
850:
851: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
852:
853: \subsection{Comparison with the case of an isolated star.}
854: \label{sect:isolated}
855:
856: We now compare stellar cooling in the irradiated case with that
857: occurring in isolated stars, in the absence of external
858: illumination. In the latter case $T_{ph}=T_0$, $F_{in}=F_0=\sigma T_0^4$
859: and equation (\ref{eq:nab_ph}) gives
860: $\nabla_{ph}=\nabla_{eff}=(\alpha+1)/8\sim 1$. Substituting this result
861: into equations (\ref{eq:T_cb}), (\ref{eq:P_cb}), using
862: adiabatic relation at the convective-radiative boundary,
863: and equation (\ref{eq:P_ph}) we find
864: \ba
865: && T_0=\left[\frac{16\nabla_{ad}}{3}
866: \left(\frac{\nabla_0-\nabla_{eff}}{\nabla_0-\nabla_{ad}}
867: \right)^{1-\nabla_0/\nabla_{ad}}\right.
868: \nonumber\\
869: && \left.\times\frac{g}{\tilde \kappa}
870: \left(\frac{\mu K^{1/\gamma}}
871: {k_B}\right)^{(1+\alpha)/\nabla_{ad}}\right]^{1/\xi}.
872: \label{eq:T_0}
873: \ea
874: This expression sets
875: the effective temperature of the star and its cooling
876: rate $F_0$ as functions of $M_\star$, $R_\star$ and
877: opacity behavior. In particular, for $\kappa$ typical
878: at temperatures below $5000$ K one finds using equation
879: (\ref{eq:K_3_2}) that
880: \ba
881: T_0\approx 1200~\mbox{K}~M_1^{11/39}R_{11}^{1/13}.
882: \label{eq:Hayashi}
883: \ea
884: This is considerably smaller than $T_{eff}\approx 3000-4000$
885: K typical for an isolated fully convective star on the Hayashi
886: track that one obtains with detailed numerical stellar structure
887: calculations (Siess \etal 2000). We ascribe this difference
888: to our adoption of fixed $\gamma$ throughout the whole star and
889: the neglect of superadiabaticity in the outer parts of the
890: convective region, see \S \ref{sect:complications}. At the same
891: time equation (\ref{eq:Hayashi}) captures the
892: main property of the Hayashi track -- extremely weak
893: sensitivity of $T_0$ to $R_\star$ and, consequently, stellar
894: luminosity.
895:
896: If we now go back to equation (\ref{eq:F_in}) one can easily
897: see that it can be rewritten as
898: \ba
899: F_{in}\approx F_0\left[\frac{T_{ph}(\theta)}{T_0}
900: \right]^{4-\xi},
901: \label{eq:Fin}
902: \ea
903: or, with equation (\ref{eq:balance}), as
904: \ba
905: \left(\frac{F_{in}}{F_0}\right)^{4/(4-\xi)}=\frac{F_{in}}{F_0}
906: +\left(\frac{T_{irr}}{T_0}\right)^4.
907: \label{eq:Fin_alt}
908: \ea
909: This result together with (\ref{eq:4xi}) once again vividly
910: illustrates the inhibition of stellar cooling by external
911: irradiation and specifies the magnitude of this effect.
912:
913: Using equations (\ref{eq:nab_ph}) and (\ref{eq:Fin}) we can
914: also write
915: \ba
916: \nabla_{ph}\approx \frac{(\alpha+1)}{8}\left(\frac{T_{ph}}{T_0}
917: \right)^{-\xi},
918: \label{eq:nabl_ph}
919: \ea
920: which shows that $\nabla_{ph}\ll 1$ when stellar
921: surface is strongly irradiated ($T_{ph}\gtrsim T_0$), thus confirming
922: equation (\ref{eq:nabla_cond}). Note the
923: strong dependence of $\nabla_{ph}$ on $T_{ph}/T_0$: with our
924: power-law anzatz for opacity $\xi\approx 13/2$ and $\approx 13$
925: below and above $5000$ K correspondingly, see equations (\ref{eq:low_T})
926: and (\ref{eq:high_T}).
927:
928:
929: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
930:
931: \subsection{Conditions of validity of 1D approximation.}
932: \label{sect:1D_validity}
933:
934: In Appendix \ref{ap:1D_validity} we determine the circumstances
935: under which the results of \S \ref{sect:1D} hold true. We show
936: there that the condition of the validity of 1D approximation
937: can be expressed as
938: \ba
939: \left(\frac{H_{ph}}
940: {L_\theta}\right)^2\lesssim \nabla_{ph},
941: \label{eq:validity}
942: \ea
943: where $L_\theta$ is a characteristic scale in $\theta$
944: direction over which the external boundary condition
945: [in our case $T_{ph}(\theta)$] experiences variation. Equation
946: (\ref{eq:irr_flux}) and Figure \ref{fig:irr_flux}
947: demonstrate that in the case of irradiation
948: by accretion disk $L_\theta\sim R_\star$. Then equations
949: (\ref{eq:nab_ph}) and (\ref{eq:Fin}) allow us to rewrite the
950: condition (\ref{eq:validity}) as (assuming that $T_0$
951: and $T_{ph}$ are in the same opacity regime)
952: \ba
953: T_{ph}\lesssim T_0\left(\frac{R_\star}{H_0}
954: \right)^{2/(2+\xi)},
955: \label{eq:c1}
956: \ea
957: where $H_0=k_B T_0/(\mu g)$ is the photospheric scale height
958: in the absence of irradiation. Given that
959: \ba
960: \frac{R_\star}{H_0}\approx 5\times 10^3~ M_1 R_{11}T_{3.5}^{-1}
961: \label{eq:rat}
962: \ea
963: we may conclude that 1D approximation should be rather accurate
964: even if $T_{ph}$ exceeds $T_0$ by a factor of several
965: (e.g. $T_{ph}\lesssim 7 T_0$ for $T\lesssim 5000$ K).
966:
967: Whenever the condition (\ref{eq:validity}) is violated
968: the redistribution of energy in
969: $\theta$-direction within the radiative layer becomes
970: important. In this case one needs to solve the full
971: two-dimensional equation (\ref{eq:rad_tran}) without assuming
972: that radiative flux in $\theta$ direction is small.
973: A similar situation arises at stellar equator where
974: a lot of energy is released in a boundary layer
975: that is not very extended in $\theta$ direction
976: (Popham \etal 1993). As a result, at equator
977: $L_\theta\ll R_\star$ and the condition (\ref{eq:validity})
978: can be violated there even though at all other latitudes
979: 1D approximation works fine.
980:
981: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
982: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
983:
984: \section{Integrated stellar cooling.}
985: \label{sect:total_cooling}
986:
987: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
988:
989: We are now in position to calculate the integrated intrinsic
990: luminosity $L$ (due to stellar contraction and interior cooling)
991: of a convective star that is irradiated by a circumstellar disk:
992: \ba
993: L=4\pi R_\star^2\int\limits_0^{\pi/2}F_{in}(\theta)
994: \sin\theta d\theta,
995: \label{eq:L}
996: \ea
997: where $F_{in}(\theta)$ is given by the expression (\ref{eq:F_in})
998: in the irradiated part of the stellar surface, for
999: $\theta\gtrsim\theta_{irr}$, while $F_{in}(\theta)\approx
1000: \sigma T_0^4$ in the weakly irradiated polar regions, for
1001: $\theta\lesssim\theta_{irr}$.
1002:
1003: Convective objects can exhibit different modes of cooling which
1004: is best illustrated by considering the limit $\theta_{irr}\ll 1$
1005: ($\Lambda\gg 1$). In this limit the contribution of polar
1006: caps to the total luminosity is
1007: \ba
1008: L_{pc}\approx 4\pi F_0 R_\star^2(1-\cos\theta_{irr})\approx
1009: 4\pi F_0 R_\star^2 \theta_{irr}^2,
1010: \label{eq:L_pc}
1011: \ea
1012: while irradiated equatorial regions contribute
1013: \ba
1014: L_{er}\approx \frac{4\pi C F_0 R_\star^2}
1015: {[I\sin^5\theta_{irr}]^{(4-\xi)/4}}\int
1016: \limits_{\theta_{irr}}^{\pi/2}[g(\theta)]^{(4-\xi)/4}
1017: \sin\theta d\theta,
1018: \label{eq:L_er}
1019: \ea
1020: see equation (\ref{eq:Fin}).
1021:
1022: \begin{figure}
1023: \plotone{f3.eps}
1024: \caption{
1025: Plot of function $I_2(\xi)$. Solid line corresponds to $I_2$
1026: computed using $g(\theta)$ (see Figure \ref{fig:irr_flux}) and
1027: the dashed line is $I_2$ calculated
1028: using asymptotic representation (\ref{eq:as}).
1029: \label{fig:I_2}}
1030: \end{figure}
1031:
1032: Using equations (\ref{eq:as}), (\ref{eq:theta_c})
1033: it is easy to see that for
1034: $\theta_{irr}\ll 1$ the latter integral is dominated by
1035: $\theta\approx \theta_{irr}$ if
1036: \ba
1037: \xi>\frac{28}{5}.
1038: \label{eq:xi_cond}
1039: \ea
1040: In this case, according to equation (\ref{eq:as}),
1041: one may approximate $g(\theta)\approx I\sin^5\theta$ and
1042: find that
1043: \ba
1044: L_{er}\approx \frac{16\pi}{5(\xi-28/5)}
1045: F_0 R_\star^2 \theta_{irr}^2\sim L_{pc}.
1046: \label{eq:L_er1}
1047: \ea
1048: This results leads us to an interesting conclusion that as long
1049: as the condition (\ref{eq:xi_cond}) is fulfilled, an object
1050: cools predominantly through its polar caps and its
1051: integrated luminosity $L$ is almost independent of the details
1052: of opacity behavior in its outer layers. The latter point
1053: is easy to understand, since in this case $L\sim F_0 S$,
1054: where $S$ is the
1055: surface area of the polar caps. But according to equation
1056: (\ref{eq:theta_c}) the value of $S$ is determined only by
1057: irradiation and $T_0$. As a result, $L$ depends on $\kappa$
1058: only weakly, through the $L_{er}$ contribution.
1059:
1060: We call the regime of stellar cooling realized under
1061: the condition (\ref{eq:xi_cond}) the {\it high-latitude}
1062: cooling. This regime naturally occurs in irradiated
1063: young stars since $\xi>28/5$ for $\kappa$ given by either
1064: (\ref{eq:low_T}) or (\ref{eq:high_T}). Equations (\ref{eq:L_pc})
1065: and (\ref{eq:L_er1}) demonstrate that in this regime $L$
1066: is suppressed roughly by $\sim \theta^2_{irr}$
1067: which may be as low as $\sim 0.2-0.4$ according to the
1068: expression (\ref{eq:theta_c}). Thus, disk irradiation
1069: can substantially slow down cooling of
1070: young stars.
1071:
1072: In the opposite case of $\xi<28/5$ cooling is in the
1073: {\it low-latitude} regime, so that star loses most
1074: of its internal energy through the equatorial regions
1075: even though they are strongly irradiated. In this case
1076: one should use the full expression (\ref{eq:L_er}) to
1077: evaluate $L\approx L_{er}$. Stellar luminosity suppression
1078: for $\theta_{irr}\ll 1$ is given by
1079: \ba
1080: && L/L_0\approx C I_2(\sin\theta_{irr})^{5(\xi-4)/4},
1081: \label{eq:second_as}\\
1082: && I_2(\xi)=I_1^{(\xi-4)/4}\int
1083: \limits_{0}^{\pi/2}[g(\theta)]^{(4-\xi)/4}
1084: \sin\theta d\theta.\nonumber
1085: \ea
1086: Function $I_2(\xi)\sim 1$ is shown in Figure \ref{fig:I_2}.
1087: Knowing that $4<\xi<28/5$ in the low-latitude case one
1088: can easily see that the degree of luminosity
1089: suppression is smaller than in the high-latitude
1090: cooling regime.
1091:
1092: In Figure \ref{fig:supp} we plot $L/L_0$ -- the ratio of
1093: stellar luminosities in the irradiated and isolated cases
1094: -- as a function of irradiation parameter $\Lambda$,
1095: for different values of $\zeta$. This calculation does not
1096: explicitly make an assumption $\theta_{irr}\ll 1$ (or
1097: $\Lambda\gg 1$), although it
1098: covers this regime as well. Here $L/L_0$ is computed by the
1099: straightforward integration of $F_{in}$ over the stellar
1100: surface (including the polar caps where $F_{in}=F_0$), with
1101: the distribution of $T_{irr}(\theta)$ found from equation
1102: (\ref{eq:irr_flux_mod}) and $g(\theta)$ displayed in Figure
1103: \ref{fig:irr_flux}. We also indicate the asymptotic
1104: behavior of $L/L_0$ as given by equations (\ref{eq:L_pc})
1105: and (\ref{eq:L_er1}) for $\xi>28/5$ and equation
1106: (\ref{eq:second_as}) for $\xi<28/5$.
1107:
1108: One can easily see that,
1109: as expected, $L/L_0\propto\theta_{irr}^2\propto \Lambda^{-2/5}$
1110: as $\Lambda\gg 1$
1111: for $\xi=13$ and $6.5$ independent of the actual value of $\xi$
1112: (only the normalizations of the curves are different because of the
1113: different contributions produced by the near-polar cap regions) since
1114: for both $\xi>28/5$. Significant suppression of the stellar
1115: flux (by $\sim 2$) is found in this case already at
1116: $\Lambda\sim 10^2-10^3$. In the case $\xi<28/5$ asymptotic behavior
1117: for $\Lambda\gg 1$ agrees well with equation (\ref{eq:second_as}),
1118: $L/L_0\propto \Lambda^{(4-\xi)/4}$, and
1119: the degree of stellar flux suppression is weaker than in the
1120: high-latitude regime: $L/L_0\approx 0.5$ only at
1121: $\Lambda\approx 5\times 10^3$ for $\xi=5.3$
1122: and at $\Lambda\approx 2\times 10^5$ for $\xi=4.5$.
1123:
1124: Note that results presented in Figure \ref{fig:I_2} are calculated
1125: neglecting any additional heating that can be produced near the
1126: equator by the boundary layer dissipation. We will address this
1127: point in more detail in \S \ref{sect:complications}.
1128:
1129: \begin{figure}[t]
1130: \plotone{f4.eps}
1131: \caption{
1132: Plots of the intrinsic luminosity of an irradiated star L
1133: (in units of an isolated star luminosity $L_0$) as a function
1134: of irradiation parameter
1135: $\Lambda=GM_\star \dot M/(\sigma T_0^4 R_\star^3)$ for
1136: different values of the power law index $\xi$ defined by equation
1137: (\ref{eq:xi}). Dotted lines illustrate the corresponding asymptotic
1138: behaviors for $\Lambda\gg 1$: $L/L_0\propto \Lambda^{-2/5}$
1139: for $\xi>28/5$ and $L/L_0\propto \Lambda^{(4-\xi)/4}$
1140: for $\xi<28/5$.
1141: \label{fig:supp}}
1142: \end{figure}
1143:
1144: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1145: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1146:
1147: \section{Discussion.}
1148: \label{sect:disc}
1149:
1150: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1151:
1152: Luminosity suppression by disk irradiation is one of the most important
1153: results of this work. Analogous phenomenon has been previously
1154: found in studies of extrasolar giant planets in short-period orbits,
1155: where stellar irradiation is quite severe (Guillot \etal 1996;
1156: Burrows \etal 2000). In that case irradiation affects only one
1157: side of the plant which always faces the star. Some heat from
1158: the day side gets redistributed to the night side by atmospheric
1159: circulation (Menou \etal 2003; Dobbs-Dixon \& Lin 2007)
1160: which should complicate the calculation of the
1161: photospheric boundary conditions across the whole planetary
1162: surface. In our case irradiation is azimuthally symmetric which
1163: makes our calculation more robust. Analogous to the case of
1164: extrasolar giant planets we expect that luminosity
1165: suppression by irradiation would tend to retain heat inside the star
1166: and increase stellar radius above the value found in the absence of
1167: irradiation (Baraffe \etal 2003). Whether this radius increase is
1168: significant will be investigated in future work.
1169:
1170: Luminosity suppression may have some effect on the strength of the
1171: magnetic field that is generated by dynamo action in the convective
1172: interior of the star. Since in irradiated case convective eddies
1173: transport smaller energy flux to the stellar surface than in the
1174: case of an isolated star the speed of convective motions is expected
1175: to be smaller. This results in a less vigorous dynamo action and
1176: likely weaker magnetic field generated inside the star.
1177:
1178: Formation of an optically thick radiative zone near the stellar surface
1179: in irradiated regions is another result of this work which has
1180: important implications. As the star grows the hot gas in the vicinity
1181: of the boundary layer where disk meets the star
1182: gets advected into the convective interior thereby raising stellar
1183: entropy. This provides another way of slowing down stellar contraction,
1184: in addition to the luminosity suppression discussed above. Hartmann
1185: \etal (1997) argued that impact of the heat advection on stellar structure
1186: is not significant as long as the temperature of
1187: advected gas is much smaller than the central temperature of the star.
1188: In the irradiated case it is the properties of the external radiative
1189: zone that determine the temperature of the gas at the
1190: convective-radiative boundary, and thus the amount of thermal energy
1191: advected into the stellar interior. Indeed, hot gas sinking through the
1192: radiative zone will lose a significant fraction of its thermal
1193: energy by radiative diffusion in the latitudinal direction, so that
1194: the temperature at the convective-radiative boundary is likely to be
1195: lower than in the center of the boundary layer. It is important to build the
1196: detailed 2D model of the radiative transport in the vicinity of the
1197: boundary layer to quantify this effect and to verify the significance of
1198: heat advection (see also \S \ref{sect:1D_validity}).
1199:
1200: Presence of the radiative zone may also affect atmospheric opacity in
1201: accreting brown dwarfs and giant planets. Accreted gas brings in
1202: significant amount of dust into the object's atmosphere (dust can
1203: also form out of the gas phase under low-temperature conditions)
1204: which changes $\kappa$ and radiative properties of the star
1205: (Chabrier \etal 2000). However, if the object is fully
1206: convective all the way to its photosphere, vertical fluid motions
1207: quickly advect dust grains into the hot interior where grains get
1208: easily destroyed. This is not the case in irradiated regions of accreting
1209: objects since dust grains can hover in the radiative zone for a
1210: long time, as long as their gravitational settling is not too fast.
1211: Of course, for grains to exist in the outer radiative layer
1212: in the first place $T_{irr}$ must be lower than the sublimation
1213: temperature of dust material, which may be possible only in accreting
1214: brown dwarfs and giant planets.
1215:
1216: We expect that in the case of young stars it would be very difficult
1217: to obtain a direct observational confirmation of the luminosity
1218: suppression by irradiation. The major reason for this
1219: is that $L/L_0$ starts to deviate from unity only when $\dot M$ and,
1220: correspondingly, accretion luminosity are very large. At this stage
1221: the luminosity of a star+disk system is completely dominated by
1222: the direct emission from the disk and the disk flux intercepted
1223: and re-radiated by the stellar surface. Intrinsic stellar luminosity
1224: provides negligible contribution which would be almost impossible
1225: to distinguish. Besides, forming protostar should still be enshrouded
1226: in the dense veil of the residual gas collapsing onto the
1227: circumstellar disk. Reprocessing of star+disk emission in
1228: this infalling envelope would complicate things even more.
1229: Another potential way of detecting the luminosity suppression
1230: is indirect, through its effect on the stellar radius and luminosity
1231: as the star emerges as almost fully formed Class I object in the end
1232: of active accretion phase.
1233:
1234: Effects of disk accretion in star formation have been previously
1235: investigated by a number of authors. Adams \& Shu (1986) and
1236: Popham (1997) have calculated the amount of energy which is
1237: emitted by the disk and is intercepted by the star. Unlike us
1238: these authors were not primarily concerned in the details of
1239: the distribution of irradiation flux over the stellar surface.
1240: This is a crucial point of our study allowing us to identify
1241: the two different regimes of stellar cooling -- high- and
1242: low-latitude.
1243:
1244: Mercer-Smith \etal (1984) were the first to explore the effect
1245: of disk accretion on the stellar structure.
1246: They handled disk accretion by specifying mass addition
1247: rate and accretion luminosity as external boundary conditions.
1248: They find that the stars formed by disk accretion have {\it larger}
1249: radii than nonaccreting stars of the same mass. As mentioned in
1250: Hartmann \etal (1997) this outcome most likely results from allowing
1251: the accreted material to have very high entropy which leads to
1252: stellar swelling, see Prialnik \& Livio (1985). This approximation
1253: is unlikely to be valid in reality since disk
1254: material joining the stellar surface should have enough time to
1255: radiate most of its thermal energy before being fully incorporated
1256: into the star. Palla \& Stahler (1992) and Hartmann \etal (1997)
1257: in their studies of intermediate- and low-mass stars allowed the
1258: accreted material to have low entropy. They found that accretion
1259: {\it reduces} stellar size compared to the non-accreting case since
1260: in this case the addition of mass leads only to
1261: the increase of gravitational energy of the star and is not accompanied by
1262: the increase of thermal energy.
1263:
1264: All these studies have either ignored irradiation of star by
1265: the disk or accounted for it only in the averaged sense which may
1266: not be acceptable as our study demonstrates. To get a complete
1267: picture of protostellar evolution one needs to include the
1268: luminosity suppression by disk irradiation into account.
1269: Such calculation must necessarily
1270: allow for the spatial distribution of irradiation flux on the
1271: stellar surface since only in this way a proper estimate of
1272: the luminosity suppression can be obtained.
1273:
1274:
1275: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1276:
1277: \subsection{Applications to real systems.}
1278: \label{sect:applications}
1279:
1280: Here we apply our results to different classes of fully convective
1281: objects which may accrete through the disk at high $\dot M$. In doing
1282: our estimates, which require the knowledge of $R_\star$ and $T_0$,
1283: we use the radii and photospheric temperatures of
1284: corresponding objects determined in the absence of irradiation and mass
1285: inflow by accretion as proxies for $R_\star$ and $T_0$ that
1286: these objects would have if irradiation and accretion were properly
1287: accounted for. Needless to say, a truly accurate estimate of the
1288: effect of irradiation can be obtained only if $R_\star$ and
1289: $T_0$ are calculated self-consistently accounting for the
1290: effects of accretion and irradiation.
1291:
1292: \subsubsection{Young stars}
1293: \label{sect:youngstars}
1294:
1295: Young low-mass stars transitioning from Class 0 to Class I
1296: phase are fully convective and should be
1297: assembled by mass accretion from a circumstellar disk within
1298: several $10^5$ yrs. This implies very high accretion rate
1299: and we adopt $\dot M=5\times 10^{-6}$ M$_\odot$ yr$^{-1}$ for a
1300: simple estimate. An isolated $M_\star=0.5$ M$_\odot$ star
1301: at an age of $10^5$ yrs has a radius $R_\star=3.9$ R$_\odot$ and
1302: effective temperature $T_0=3760$ K (Siess \etal 2000).
1303: These parameters yield $\Lambda=10^2$ and $T_{irr}(\pi/2)
1304: \approx 6000$ K, although the latter is likely to be higher
1305: because of the boundary layer dissipation. At this $\Lambda$
1306: cooling in the equatorial region is suppressed but the size
1307: of the polar caps is reduced only weakly: according to Figure
1308: \ref{fig:irr_flux} $\theta_{irr}\approx 73^\circ$
1309: [$\theta_{irr}$ is given by an implicit
1310: relation $g(\theta_{irr})=\Lambda^{-1}$]. In Figure
1311: \ref{fig:lambdas} we present more general results for $\Lambda$
1312: calculated for stars of different masses using stellar
1313: parameters from Siess \etal (2000) and assuming
1314: constant $\dot M=M_\star/t_{acc}$, where $t_{acc}$ is the accretion
1315: time (assumed equal to the stellar age).
1316: One can see that the low-mass stars assembled within
1317: $3\times 10^5$ yrs generally have $\Lambda$ in the range of
1318: $30-10^2$, agreeing with our simple estimate.
1319:
1320: Since $\xi\approx 6$ for $T\lesssim 5000$ K, young stars cool
1321: mainly through the polar regions and we find from Figure
1322: \ref{fig:supp} that stellar luminosity is reduced by
1323: irradiation only by about $10\%$ for
1324: $\Lambda\approx 10^2$. On the other hand, if $\dot M$
1325: is not constant but increases as $M_\star$ grows one may expect
1326: values of $\Lambda$ larger by a factor of several. Also, to
1327: calculate $\Lambda$ we have used parameters of isolated stars
1328: while stars assembled by disk accretion of the low-entropy material
1329: have smaller $R_\star$,
1330: leading to larger $\Lambda$. All these factors may increase the
1331: importance of disk irradiation in determining $L$ of
1332: young accreting stars.
1333:
1334: \begin{figure}[b]
1335: \plotone{f5.eps}
1336: \caption{
1337: Irradiation parameter $\Lambda$ for stars of different mass
1338: and age $t_{acc}$ (indicated on the plot) assembled by disk accretion
1339: with constant $\dot M=M_\star/t_{acc}$. Stellar parameters from
1340: Siess \etal (2000) were used in this calculation.
1341: \label{fig:lambdas}}
1342: \end{figure}
1343:
1344: \subsubsection{Young stars in quasar disks}
1345: \label{sect:youngstars_QSOs}
1346:
1347: A very interesting mode of star formation is possible in the
1348: accretion disks around the supermassive black holes (SMBHs) in the
1349: centers of galaxies (Illarionov \& Romanova 1988;
1350: Goodman \& Tan 2004; Nayakshin 2006).
1351: It is currently known that our own Galactic Center harboring
1352: a SMBH of mass $M_{BH}\approx 3.7\times 10^6$ M$_\odot$
1353: (Ghez \etal 2005) contains
1354: a number of young ($\lesssim 6$ Myrs) massive
1355: ($M_\star\gtrsim 10$ M$_\odot$) stars that form
1356: two misaligned disks around the SMBH (Paumard \etal 2006).
1357: One of the most likely scenarios for the origin of these stars
1358: is a fragmentation of a gravitationally unstable gaseous disk
1359: (or disks) followed by the growth of fragments to their present
1360: masses by gas accretion from the residual disk (Levin 2006;
1361: Nayakshin 2006). Assuming that disk temperature is kept at
1362: the level of $50$ K by the radiation of nearby stars (Levin 2007)
1363: one finds that at $a=0.1$ pc from the SMBH
1364: (which is the typical dimension of the observed disks)
1365: surface mass density of $\Sigma\approx 27$ g cm$^{-2}$ is
1366: required for the disk to be Toomre unstable.
1367:
1368: Fragments formed as a result of instability at $0.1$ pc
1369: have a typical mass
1370: $M_\star\sim \Sigma h^2\approx 10^{-3}$ M$_\odot$ (approximately one
1371: Jupiter mass) where $h$ is a disk scale height. At formation
1372: the Hill radius of such an object $R_H=a(M_\star/M_{BH})^{1/3}$
1373: is already comparable to $h$ and as $M_\star$
1374: grows by accretion $R_H$ becomes larger than $h$. As a result,
1375: accretion onto fragment proceeds through the {\it sub-disk} that
1376: forms within the fragment's Hill sphere, presenting us with the
1377: setting investigated in this paper. Rate at which gas flows into
1378: the fragment's Hill sphere is the Hill accretion rate\footnote{Such
1379: high $\dot M$ is also typical for FU Orioni objects. As demonstrated by
1380: Popham \etal (1993) in the high-$\dot M$ regime the boundary layer
1381: is so thick that it covers a significant ($\sim 0.5$) fraction of
1382: the stellar surface slowing down interior cooling. Heat advection
1383: into the stellar interior may also become an issue (Popham 1997).}
1384: $\dot M_H\approx
1385: \Sigma\Omega R_H^2\approx 2\times 10^{-4}~M_1^{2/3}(a/0.1~\mbox{pc})^2$
1386: M$_\odot$ yr$^{-1}$. Note that $\dot M_H$ is smaller
1387: than the Eddington mass accretion rate
1388: $\dot M_{Edd}=4\pi cR_\star/\kappa_{es}=
1389: 1.4\times 10^{-3}R_{11}$ M$_\odot$ yr$^{-1}$ (here $c$ is the
1390: speed of light and $\kappa_{es}$ is the electron scattering opacity) at
1391: $a=0.1$ pc but may become comparable to $\dot M_{Edd}$ further out from
1392: the SMBH provided that $R_\star$ is not much larger than $R_\odot$.
1393:
1394: If gas in the disk is able to accrete at the
1395: same high rate $\dot M_H$ onto the stellar surface then
1396: \ba
1397: \Lambda\approx 3\times 10^5 M_1^{5/3}T_{3.5}^{-4}
1398: R_{11}^{-3}.
1399: \label{eq:Hill_Lambda}
1400: \ea
1401: At present we do not have a theory for the structure of
1402: stars formed by fragmentation of gravitationally unstable disks
1403: so the value of $R_\star$ is highly uncertain. If $R_\star \lesssim 10~R_\odot$
1404: then $\Lambda\gtrsim 10^3$ and luminosity suppression by irradiation
1405: should be quite important,
1406: reducing $L$ by a factor of $2-3$ compared to $L_0$, as Figure \ref{fig:supp}
1407: demonstrates for $\xi=6.5$. As the value of $R_\star$ itself is
1408: affected by the time history of $L$, irradiation
1409: should not be overlooked in studies of star formation in quasar
1410: disks.
1411:
1412: \subsubsection{Young brown dwarfs}
1413: \label{sect:BDs}
1414:
1415: Brown dwarf (BD) formation is likely to be a scaled down version of
1416: the low-mass star formation: one again expects a formation
1417: of a centrifugally supported disk around a fully convective
1418: object that grows by disk accretion. The biggest uncertainty in
1419: determining $\Lambda$ is again $R_\star$: $0.1$ Gyr old
1420: BDs have radii of $0.1-0.2$ R$_\odot$ (Baraffe \etal 2003)
1421: but accumulation of their mass
1422: (poorly investigated at present) likely takes less than
1423: $10^5$ yr, during which time their entropy is still
1424: quite high,
1425: resulting in considerably larger $R_\star$. Assuming
1426: that an object with $M_\star=0.03$ M$_\odot$ grows at constant
1427: $\dot M$ in time $t_{acc}=5\times 10^4$ yr and has $R_\star=0.5$
1428: R$_\odot$ and $T_0\approx 3000$ K we find $\Lambda\approx 800$.
1429: Provided that opacity can still be characterized by expression
1430: (\ref{eq:low_T}) (which is a somewhat questionable assumption)
1431: we conclude that irradiation may lead to order unity reduction
1432: in $L$. As the brown dwarf cools and contracts $\Lambda$ increases
1433: bringing down $L/L_0$ even more, provided that $\dot M$ could still be
1434: maintained at high level. Thus, young BDs may be affected by
1435: the disk irradiation which may have consequences
1436: for their subsequent thermal evolution.
1437:
1438: \subsubsection{Young giant planets}
1439: \label{sect:planets}
1440:
1441: Finally, we consider the situation arising during the late stages
1442: of giant planet formation via the so-called core instability. This
1443: scenario of planet formation assumes buildup of a
1444: $\sim 10$ M$_\oplus$ refractory core in the protoplanetary nebula by
1445: planetesimal agglomeration. The self-gravity of the core triggers an
1446: instability and leads to rapid gas accumulation (Mizuno 1980).
1447: While the initial stages of this process can be adequately described
1448: in the spherically-symmetric approximation, the later epoch of unstable
1449: gas accretion must have distinctly non-spherical morphology. Indeed, as
1450: mentioned in Rafikov (2006), as soon as the mass of a rapidly growing planet
1451: exceeds the so-called {\it transitional} mass $M_{tr}=c_s^3/\Omega G\approx
1452: 40 M_\oplus a_5^{3/4}$ (here $c_s$ is the gas sound speed in the nebula
1453: and $a_5\equiv a/5$ AU is the planetary semi-major axis scaled by 5 AU)
1454: the Hill radius of the planet $R_H$ becomes larger than the scale
1455: height of the disk $h$. As a result, protoplanet starts accreting
1456: gas from the surrounding nebula through the sub-disk that forms
1457: within its Hill sphere, thereby presenting a situation analogous to
1458: the star formation in the Galactic Center described in
1459: \S \ref{sect:youngstars_QSOs} (except that now the collapsing
1460: fragment of a gravitationally unstable disk is replaced by a growing
1461: planet). Here we assess how important can irradiation be for planetary
1462: cooling when $M_p\gtrsim M_{tr}$.
1463:
1464: The maximum $\dot M$ available to the planet is still likely given
1465: by the Hill rate $\dot M_H\approx 2\times 10^{-3}M_{p,2}^{2/3}a_5^{-1}$
1466: M$_{\rm J}$ yr$^{-1}$, where $M_{p,2}\equiv M_p/10^2$ M$_\oplus$ and
1467: we have adopted a surface density profile $\Sigma=270 a_5^{-3/2}$
1468: g cm$^{-2}$ typical for the Minimum-Mass Solar Nebula. This allows us
1469: to compute
1470: \ba
1471: \Lambda\approx 10^4 M_2^{5/3}a_5^{-1}
1472: \left(\frac{R_p}{5~\mbox{R}_J}\right)^{-3}
1473: \left(\frac{T_0}{10^3~\mbox{K}}\right)^{-4}.
1474: \label{eq:Lam_planet}
1475: \ea
1476: This estimate is rather uncertain because of poorly constrained
1477: $R_p$ and $T_0$ during the stage of active gas accretion by the
1478: planet. Here we adopt $R_p=5$ R$_J$ and $T_0=10^3$ K mainly for
1479: illustrative purposes.
1480:
1481: Dust is the major source of opacity in the outer layers of forming
1482: giant planets. Dust opacity scales as $\kappa\propto T^{\beta}$
1483: with $\beta\approx 0.5-2$ depending on dust grain composition,
1484: spectrum of grain sizes, etc. Here we adopt $\beta=1$ in which
1485: case $\xi=4.5$. This corresponds to cooling dominated by the
1486: equatorial regions, which is different from the stellar case. As a
1487: result, $L/L_0$ should be more sensitive
1488: to the structure of the boundary layer through which disk material
1489: accretes onto the planet, namely, $L/L_0$ should be {\it lower}
1490: than Figure
1491: \ref{eq:Lam_planet} implies. Forgetting about this complication
1492: for the moment and using $\xi=4.5$ and $\Lambda$ from
1493: equation (\ref{eq:Lam_planet}), we find from Figure \ref{fig:supp}
1494: that in the planetary case $L$ can be suppressed compared to
1495: $L_0$ by several tens of per cent. Given the existing uncertainties
1496: in modeling the late stages of planet formation this degree of
1497: luminosity suppression by irradiation may not seem like a
1498: serious issue.
1499: However, more massive, compact and cooler planets can easily have
1500: $\Lambda\sim 10^5-10^6$ in which case irradiation would reduce
1501: $L$ by a factor of several potentially affecting planetary evolution.
1502:
1503: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1504:
1505: \subsection{Additional complications.}
1506: \label{sect:complications}
1507:
1508: Here we address various complications that may arise when the
1509: results of this work are applied to real objects.
1510:
1511: All our derivations and estimates explicitly assumed opacity
1512: in the form (\ref{eq:kappa}). While
1513: this representation can be quite accurate within some temperature
1514: intervals one has to bear in mind that on the surface of a star
1515: irradiated by the disk temperature can vary appreciably
1516: between the equator and the poles. Indeed, the equatorial
1517: regions of a $M_\star=1$ M$_\odot$, $R_\star=2$ R$_\odot$ star
1518: accreting at $\dot M=10^{-5}$ M$_\odot$ yr$^{-1}$ are heated
1519: to $1.1\times 10^4$ K so that equation (\ref{eq:high_T}) applies,
1520: while polar caps still have $T\approx 3\times 10^3$ K so
1521: that equation (\ref{eq:low_T}) is more appropriate. Thus, in
1522: different regions of stellar surface $\kappa$ has
1523: different dependence on $P$ and $T$. In this case equation
1524: (\ref{eq:second_as}) becomes invalid and to properly
1525: compute the luminosity suppression one would need to take into account
1526: the latitudinal variation of not only $T_{irr}$ but also
1527: $\kappa(P,T)$.
1528:
1529: Even at a fixed latitude opacity can switch from one
1530: regime to another within the outer radiative zone. Although $\kappa$
1531: is much more sensitive to $T$ than to $P$ and the radiative zone
1532: is roughly isothermal, pressure at its bottom
1533: $P_{cb}$ is $\sim \nabla_{ph}^{-1/(1+\alpha)}\gg 1$ of the
1534: photospheric pressure $P_{ph}$, so that even a weak dependence
1535: of $\kappa$ on $P$ can lead to opacity transition within the outer
1536: radiative zone. In particular, this situation is likely to occur
1537: at latitudes where $T_{irr}\approx 5000$ K and $\kappa$ switches
1538: from (\ref{eq:low_T}) to (\ref{eq:high_T}). In this case
1539: calculation of the local radiative flux $F_{in}$ gets more complicated
1540: as the external radiative region splits into two layers characterized
1541: by different opacity behaviors.
1542:
1543: Our numerical estimates of the luminosity suppression strongly
1544: rely on the assumption of fixed $\gamma$ within the radiative
1545: zone. Equation (\ref{eq:Hayashi}) which assumes $\gamma=5/3$
1546: throughout the whole star fails to predict the correct photospheric
1547: temperature of an isolated star. The reason for
1548: this is the variation of $\gamma$ at low $T$
1549: caused by molecular dissociation.
1550: This results in $\nabla_{ad}\lesssim 0.1$ and leads to a
1551: smaller drop of temperature in the outer convective parts of
1552: the star. On the other hand, superadiabaticity of convection
1553: in the outer layers of convective zone counteracts this effect
1554: to some extent. Our
1555: simple estimate (\ref{eq:Hayashi}) underpredicts $T_0$ by a factor
1556: of $2-3$ which is primarily a consequence of our assumption of fixed
1557: $\gamma=5/3$. A proper calculation of $F_{in}(\theta)$ and
1558: $L/L_0$ must be able to account for the variation of $\gamma$
1559: with $P$ and $T$ inside the outer radiative zone.
1560:
1561: Our analysis is affected to some extent by the presence of
1562: the boundary layer through which disk material
1563: joins the star. Viscous dissipation in this
1564: layer heats accreting gas to very high temperature. Since
1565: this energy release takes place very close to the stellar surface
1566: most of the heat is likely to leak out and not get carried into
1567: the star with the accreted gas. However, some residual
1568: heat may still be accreted. Moreover, in addition to advective there could also
1569: be a radiative energy transfer from the boundary layer into the outer
1570: layers of the star (Popham 1997). Increase of $T$ in the
1571: external radiative zone driven by these processes acts to
1572: additionally slow down stellar cooling in the equatorial region, as
1573: equation (\ref{eq:F_in}) demonstrates. Thus, presence of the
1574: boundary layer reduces $L/L_0$ even more than our analysis
1575: predicts, and the results presented in Figure \ref{fig:supp}
1576: should be viewed as upper limits on $L/L_0$. This effect is
1577: likely not very important for young stars which cool predominantly
1578: through their polar regions, largely unaffected by the
1579: additional heat deposition at
1580: the equator. However, in the case of young giant planets which lose
1581: fair amount of energy through the low-latitude part of the surface
1582: (see \S \ref{sect:planets}) the reduction of intrinsic flux in
1583: the equatorial region may produce quite noticeable decrease of
1584: $L/L_0$ compared to the idealized case considered in this work.
1585:
1586: \begin{figure}[t]
1587: \plotone{f6.eps}
1588: \caption{
1589: Schematic representation of magnetically channeled accretion.
1590: Disk material travels along the magnetic field lines and
1591: sediments onto the magnetospheric columns (shaded) heated by shock
1592: dissipation and gravitational settling of material in them.
1593: Hot magnetospheric columns irradiate stellar surface near the
1594: magnetic poles.
1595: \label{fig:mag_pole}}
1596: \end{figure}
1597:
1598: When discussing the external radiative zone throughout this
1599: work we have been concerned only with the radiative energy
1600: transport. At the same time, it is well known that in the case
1601: of hot Jupiters advective transport in the form of
1602: atmospheric jets and winds can be quite important in
1603: redistributing heat across the planetary surface (Menou
1604: \etal 2003; Dobbs-Dixon \& Lin 2007).
1605: In our azimuthally-symmetric setting only
1606: meridional atmospheric motions can lead to energy exchange
1607: between the hot equatorial and cold polar regions. Fluid
1608: motions occurring on surfaces of constant effective potential
1609: (gravitational plus centrifugal) are unlikely to produce
1610: efficient equator-pole energy exchange (compared to the
1611: radiative transfer) because of rather
1612: fast rotation typical for objects formed by disk accretion. Rotation
1613: forces angular momentum conservation and prevents significant
1614: fluid motions in $\theta$-direction, suppressing this mode of
1615: advective transport. On the other hand, rotation tends
1616: to promote meridional circulation within the radiative layer
1617: (Kippenhahn \& Weigert 1994) whose impact on the energy
1618: transport in the outer radiative zone should be
1619: investigated in more detail.
1620:
1621: Finally, our basic assumption of direct mass accretion from
1622: the disk may be challenged if the growing star possesses magnetic
1623: field strong enough to disrupt accretion flow outside of
1624: $R_\star$ (K\"onigl 1991; Matt \& Pudritz 2005).
1625: In this case gas is channeled by the magnetic field and
1626: is deposition onto the stellar surface at magnetic
1627: poles rather than at stellar equator. This completely changes
1628: the topology of accretion flow but the major conclusions about
1629: the effect on stellar cooling are likely to hold.
1630: Indeed, the magnetically channeled gas travels towards
1631: the stellar surface at a good fraction of the free-fall velocity
1632: and at some point it must pass through the radiative shock, after
1633: which it accumulates at the top of the magnetospheric column
1634: of accreted material, as schematically indicated in Figure
1635: \ref{fig:mag_pole}. Total energy release within the shock
1636: and magnetospheric column is comparable to that occurring
1637: if the accretion disk were extending all the way to the stellar
1638: surface. This hot column of accreted material illuminates the
1639: surface of the star leading to the same suppression of intrinsic
1640: stellar flux as we discussed in this work. In this case, however,
1641: irradiation is strongest near the magnetic poles while
1642: the magnetic equator is likely to be the coolest part of the
1643: stellar surface\footnote{Illumination of the star by the distant parts of
1644: accretion disk, beyond the point where accretion flow is
1645: disrupted by the magnetic field, is
1646: unlikely to be very important given the rapid
1647: fall-off of $F_d$ with the distance from the star.}.
1648: Calculation of stellar irradiation and integrated luminosity in
1649: this case would involve constructing a model for the
1650: magnetospheric column structure and its radiative properties.
1651:
1652: The impact of these details on the structure and evolution
1653: of young stars should be addressed by future work.
1654:
1655: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1656: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1657:
1658: \section{Summary.}
1659: \label{sect:concl}
1660:
1661: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1662:
1663: Luminosity of young stars actively accreting from the
1664: circumstellar disk can be significantly affected by the
1665: radiation which is produced in the inner parts of the disk
1666: and is intercepted by the stellar surface. We showed that
1667: if a star gains its mass via disk accretion on timescale
1668: of several $10^5$ yr then the radiative flux caused
1669: by viscous dissipation in the disk is more than sufficient
1670: to increase the surface temperature of the star above the
1671: photospheric temperature that an isolated star with the
1672: same mass and radius would have. Irradiation
1673: by the disk is strongest in the equatorial regions and is
1674: almost negligible near the poles. An outer radiative zone
1675: of almost constant temperature forms above the fully convective
1676: interior in the strongly irradiated parts of the stellar
1677: surface. This leads to the local suppression of intrinsic
1678: energy flux escaping from the stellar interior.
1679:
1680: We have
1681: demonstrated that there are two distinct modes in which a
1682: fully convective object can cool: mainly through the
1683: cool high-latitude polar regions or predominantly through the
1684: low-latitude parts of the stellar surface. A particular
1685: regime of cooling in a given object is set by the opacity
1686: behavior and the adiabatic temperature gradient $\nabla_{ad}$ in the outer
1687: radiative zone. Accreting young stars and brown dwarfs cool
1688: mainly through the polar regions while forming giant planets
1689: cool through the whole surface.
1690:
1691: Integrated stellar luminosity in accreting case is suppressed
1692: compared to the case of an isolated object, by up to a factor
1693: of several in some classes of objects (actively accreting brown dwarfs
1694: and planets, stars forming in gravitationally unstable
1695: disks in galactic nuclei). This leads to larger radii of
1696: irradiated objects and may affect the initial conditions which
1697: are used to calculate the evolution of the low-mass objects
1698: on timescales of $\sim 10$ Myr after their formation.
1699: Existence of external radiative zone may facilitate retention
1700: of dust in the atmospheres of brown dwarfs and planets, and
1701: may affect the strength of magnetic field generated by
1702: internal dynamo in convective objects.
1703:
1704: Some of the results obtained in this work may be
1705: applicable to accreting white dwarf and neutron star systems.
1706:
1707: \acknowledgements
1708:
1709: I am grateful to Gilles Chabrier for careful reading of the
1710: manuscript and many useful suggestion.
1711: The financial support for this work is provided
1712: by the Canada Research Chairs program and a NSERC
1713: Discovery grant.
1714:
1715: \appendix
1716:
1717: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1718: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1719:
1720: \section{Irradiation flux.}
1721: \label{ap:irr_flux}
1722:
1723: Performing an integral over $\phi$ in eq.
1724: (\ref{eq:irr_flux}) we find
1725: \ba
1726: && F_{irr}(\theta)=\frac{2\cos\theta}{\pi}
1727: \int\limits_{x_{in}}^\infty \frac{F_d(xR_\star)q(x,\theta)x dx}{D^2},
1728: \label{eq:1D}\\
1729: && q(x,\theta)=\sqrt{x^2\sin^2\theta-1}-2\frac{x^2(1-2\sin^2\theta)+1}{D}
1730: \arctan\left(\frac{x^2+2x\sin\theta+1}{x\sin\theta+1}
1731: \sqrt{\frac{x^2\sin^2\theta-1}{D^2}}\right),
1732: \label{eq:h}
1733: \ea
1734: where $D^2=(x^2+1)^2-4x^2\sin^2\theta$ and $x_{in}=1/\sin\theta$.
1735: For $F_d(R)$ obeying (\ref{eq:vis_dissip}) equation (\ref{eq:1D})
1736: can be rewritten as eq. (\ref{eq:irr_flux_mod}) with
1737: \ba
1738: g(\theta)=\frac{3\cos\theta}{4\pi^2}
1739: \int\limits_{x_{in}}^\infty \frac{f(xR_\star)q(x,\theta)dx}{x^2 D^2}
1740: \label{eq:g}
1741: \ea
1742: Integral in (\ref{eq:g}) is dominated by
1743: $x\sim x_{in}$. Near equator, where $\theta\to \pi/2$ one can expand
1744: integrand in (\ref{eq:1D}) in terms of $\pi/2-\theta\ll 1$ and
1745: $x-1\ll 1$ which results in $F_d(\pi/2)=F_d(R_\star)/2$.
1746: Polar regions of the star ($\theta\to 0$) are illuminated only by
1747: distant parts of the disk, $R\gtrsim R_\star/\sin\theta\gg
1748: R_\star$, so that $x\gg 1$ (while $x\sin\theta\sim 1$)
1749: in equation (\ref{eq:1D}). Also, far from the star one can safely
1750: use equation (\ref{eq:vis_dissip}) with $f=1$ to finally arrive at the
1751: equation (\ref{eq:as}) with
1752: \ba
1753: I_1=\frac{3}{4\pi^2}\int\limits_1^\infty\frac{dt}{t^6}
1754: \left(\sqrt{t^2-1}-2\arctan\frac{\sqrt{t^2-1}}{1+t}\right)=\frac{1}{50\pi^2}
1755: \label{eq:I}
1756: \ea
1757: This result is independent of the structure of the boundary layer near
1758: the stellar surface since the polar regions of the star do not have direct
1759: sight lines to the boundary layer. This is evidenced by the convergence
1760: at $\theta \to 0$ of the two curves in Figure \ref{fig:irr_flux}
1761: calculated assuming $f(R)=1$ and $f(R)=1-(R_\star/R)^2$.
1762:
1763:
1764:
1765: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1766: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1767:
1768: \section{Validity of 1D approximation.}
1769: \label{ap:1D_validity}
1770:
1771: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1772:
1773: To determine the validity limits of the 1D solution for the
1774: structure of the radiative zone found in \S \ref{sect:1D} we evaluate
1775: the magnitude of the corrections arising when the latitudinal
1776: radiative transfer is accounted for. Considering 1D solution
1777: (\ref{eq:PofT}) as a zeroth-order approximation we plug it into
1778: the full equation (\ref{eq:rad_tran}) and carefully expand all
1779: $\theta$-derivatives, remembering that $P$ is almost independent
1780: of $\theta$ (latitudinal pressure gradients are small). Integrating
1781: the resultant expression once over $r$ we again arrive
1782: at the equation (\ref{eq:flux}) but with
1783: $F_{in}\to F_{in}+\delta F_{in}$ in the left-hand side, where
1784: \ba
1785: \delta F_{in}=\left(\frac{k_B T_{ph}^{4-\beta}}
1786: {\mu L_\theta^2}\right)^2 g^{-1}
1787: \int\limits_P^{P_{cb}}\frac{T^{\beta-2}}{P^{\alpha+2}}
1788: Z(P)dP.
1789: \label{eq:deltaF}
1790: \ea
1791: Here $Z(P)\sim 1$ is a weak function of pressure (varying by at
1792: most a factor $\sim 1$) and $L_\theta$
1793: is a characteristic scale of latitudinal variation of $T_{ph}$,
1794: $L_\theta = R_\star(\partial \ln T_{ph}/\partial \theta)^{-1}$.
1795: Our 1D approximation is justified if the correction
1796: to the 1D result $\delta F_{in}$ is small compared to $F_{in}$
1797: given by equation (\ref{eq:F_in}).
1798:
1799: Integral in (\ref{eq:deltaF}) attains its
1800: highest value at $P\sim P_{ph}$ (latitudinal radiation transfer
1801: is easiest in the upper, low density layers of the star just below
1802: the photosphere) and one can easily find using equations
1803: (\ref{eq:P_ph}), (\ref{eq:nab_ph}), and (\ref{eq:F_in}) that
1804: \ba
1805: \frac{\delta F_{in}}{F_{in}}\sim \left(\frac{H_{ph}}
1806: {L_\theta}\right)^2\nabla_{ph}^{-1},
1807: \label{eqflux_ratio}
1808: \ea
1809: where $H_{ph}$ is a photospheric scale height. This result makes
1810: it clear that the 1D solution for the structure of the radiative
1811: zone should be reasonable as long as the condition (\ref{eq:validity})
1812: is fulfilled.
1813:
1814:
1815:
1816: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1817:
1818: \begin{thebibliography}{}
1819:
1820: \harvarditem{}{}{}
1821: Arras, P. \& Bildsten, L. 2006, ApJ, 650, 394
1822:
1823: \harvarditem{}{}{}
1824: Baraffe, I., Chabrier, G., Allard, F., \& Hauschildt, P. H. 1998, A\&A, 337, 403
1825:
1826: \harvarditem{}{}{}
1827: Baraffe, I., Chabrier, G., Allard, F., \& Hauschildt, P. H. 2002, A\&A, 382, 563
1828:
1829: \harvarditem{}{}{}
1830: Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., \& Hauschildt, P. H. 2003, A\&A, 402, 701
1831:
1832: \harvarditem{}{}{}
1833: Bell, K. R. \& Lin, D. N. C. 1994, ApJ, 427, 987
1834:
1835: \harvarditem{}{}{}
1836: Burrows, A., Guillot, T., Hubbard, W. B., Marley, M. S., Saumon, D., Lunine, J. I., \& Sudarsky, D. 2000, ApJL, 534, L97
1837:
1838: \harvarditem{}{}{}
1839: Chabrier, G., Baraffe, I., Allard, F., \& Hauschildt, P. H. 2000, ApJ, 542, 464
1840:
1841: \harvarditem{}{}{}
1842: Chabrier, G., Barman, T., Baraffe, I., Allard, F., \& Hauschildt, P. H. 2004, ApJL, 603, L53
1843:
1844: \harvarditem{}{}{}
1845: D'Antona, F. \& Mazzitelli, I. 1994, ApJS, 90, 467
1846:
1847: \harvarditem{}{}{}
1848: Dobbs-Dixon, I. \& Lin, D. N. C. 2007, astro-ph/0704.3269
1849:
1850: \harvarditem{}{}{}
1851: Frank, A. C. \& Shu, F. H. 1986, ApJ, 308, 836
1852:
1853: \harvarditem{}{}{}
1854: Ferguson, J. W. \etal 2005, ApJ, 623, 585
1855:
1856: \harvarditem{}{}{}
1857: Ghez, A. M., Salim, S., Hornstein, S. D., Tanner, A., Lu, J. R., Morris, M., Becklin, E. E., Duchene, G. 2005, ApJ, 620, 744
1858:
1859: \harvarditem{}{}{}
1860: Guillot, T., Burrows, A., Hubbard, W. B., Lunine, J. I., \& Saumon, D. 1996, ApJL, 459, L35
1861:
1862: \harvarditem{}{}{}
1863: Hartmann, L., Cassen, P., \& Kenyon, S. J. 1997, ApJ, 475, 770
1864:
1865: \harvarditem{}{}{}
1866: Hayashi, C. 1981, Progr. Theor. Phys. Suppl., 70, 35
1867:
1868: \harvarditem{}{}{}
1869: Illarionov, A. F. \& Romanova, M. M. 1988, Sov. Astr., 32, 148
1870:
1871: \harvarditem{}{}{}
1872: Kippenhahn, R. \& Weigert, A. 1994, Stellar Structure and Evolution (Berlin: Springer-Verlag)
1873:
1874: \harvarditem{}{}{}
1875: K\"onigl, A. 1991, ApJL, 370, L39
1876:
1877: \harvarditem{}{}{}
1878: Landau, L. D. \& Lifshitz, E. M. 1984, Statistical Physics; Butterworth-Heinemann
1879:
1880: \harvarditem{}{}{}
1881: Levin, Y. 2007, MNRAS, 374, 515
1882:
1883: \harvarditem{}{}{}
1884: Matt, S. \& Pudritz, R. E. 2005, MNRAS, 356, 167
1885:
1886: \harvarditem{}{}{}
1887: Menou, K., Cho, J. Y.-K., Seager, S. \& Hansen, B. M. S. 2003, ApJL, 587, L113
1888:
1889: \harvarditem{}{}{}
1890: Mercer-Smith, J. A., Cameron, A. G. W., \& Epstein, R. I. 1984, ApJ, 279, 363
1891:
1892: \harvarditem{}{}{}
1893: Mizuno, H. 1980, Progr. Theor. Phys., 64, 544
1894:
1895: \harvarditem{}{}{}
1896: Palla, F. \& Stahler, S. W. 1992, ApJ, 392, 667
1897:
1898: \harvarditem{}{}{}
1899: Palla, F. \& Stahler, S. W. 1999, ApJ, 525, 772
1900:
1901: \harvarditem{}{}{}
1902: Popham, R. 1997, ApJ, 478, 734
1903:
1904: \harvarditem{}{}{}
1905: Popham, R., \& Narayan, R. 1995, ApJ, 442, 337
1906:
1907: \harvarditem{}{}{}
1908: Popham, R., Narayan, R., Hartmann, L., \& Kenyon, S. 1993, ApJL, 415, L127
1909:
1910: \harvarditem{}{}{}
1911: Prialnik, D. \& Livio, M. 1985, MNRAS, 216, 37
1912:
1913: \harvarditem{}{}{}
1914: Rafikov, R. R. 2006, ApJ, 648, 666
1915:
1916: \harvarditem{}{}{}
1917: Saumon, D., Chabrier, G., \& van Horn, H. M. 1995, ApJS, 99, 713
1918:
1919: \harvarditem{}{}{}
1920: Siess, L., Dufour, E., \& Forestini, M. 2000, A\& A, 358, 593
1921:
1922: \harvarditem{}{}{}
1923: Siess, L. \& Forestini, M. 1996, A\&A, 308, 472
1924:
1925: \harvarditem{}{}{}
1926: Siess, L., Forestini, M. \& Bertout, C. 1997, A\&A, 326, 1001
1927:
1928: \harvarditem{}{}{}
1929: Siess, L., Forestini, M. \& Bertout, C. 1999, A\&A, 342, 480
1930:
1931: \end{thebibliography}
1932:
1933:
1934: \end{document}
1935:
1936:
1937: To quantify the effect of the boundary layer on stellar
1938: cooling one must have a good description of its structure.
1939: Popham \etal (1993) have explored the boundary layer
1940: physics in the pre-main sequence accretion disks and found
1941: boundary layer effective temperatures at the level of
1942: $(1-1.5)\times 10^4$ K for $\dot M=10^{-6}-10^{-5}$ M$_\odot$
1943: yr$^{-1}$. Deep within the boundary layer $T$ should be even
1944: higher which may result in a significant radiative energy flow
1945: into the equatorial part of the outer radiative zone.
1946: Popham \etal (1993) found the geometric thickness of the
1947: boundary layer to be quite substantial, with $h/R_\star$
1948: reaching $0.15$ for $\dot M=10^{-6}-10^{-5}$ M$_\odot$
1949: yr$^{-1}$. These results may be affected by the boundary
1950: conditions adopted in Popham et al., namely that the
1951: angular velocity of the gas matches the angular speed of the
1952: star at some arbitrarily chosen radius. Siess \& Forestini (1996)
1953: have adopted a different approach by assuming that the
1954: boundary layer is neutrally stable with respect to mixing
1955: instabilities which led them to conclude that some kinetic
1956: energy of accreted gas is actually dissipated quite
1957: deep within the star. Clearly, more work is required to
1958: better characterize the structure of the boundary layer
1959: and its impact on stellar cooling.
1960: